Skip to main content

ORIGINAL RESEARCH article

Front. Oncol., 30 January 2023
Sec. Pediatric Oncology

Identify potential driver genes for PAX-FOXO1 fusion-negative rhabdomyosarcoma through frequent gene co-expression network mining

Xiaohui ZhanXiaohui Zhan1Yusong LiuYusong Liu2Asha Jacob JannuAsha Jacob Jannu3Shaoyang HuangShaoyang Huang4Bo YeBo Ye1Wei WeiWei Wei1Pankita H. PandyaPankita H. Pandya5Xiufen YeXiufen Ye2Karen E. PollokKaren E. Pollok5Jamie L. RenbargerJamie L. Renbarger5Kun HuangKun Huang3Jie Zhang*Jie Zhang6*
  • 1Department of Bioinformatics, School of Basic Medicine, Chongqing Medical University, Chongqing, China
  • 2College of Intelligent Systems Science and Engineering, Harbin Engineering University, Harbin, China
  • 3Department of Biostatistics and Health Data Science, Indiana University, School of Medicine, Indianapolis, IN, United States
  • 4Carmel High School, Indianapolis, IN, United States
  • 5Department of Pediatrics, Indiana University, School of Medicine, Indianapolis, IN, United States
  • 6Department of Medical and Molecular Genetics, Indiana University, School of Medicine, Indianapolis, IN, United States

Background: Rhabdomyosarcoma (RMS) is a soft tissue sarcoma usually originated from skeletal muscle. Currently, RMS classification based on PAX–FOXO1 fusion is widely adopted. However, compared to relatively clear understanding of the tumorigenesis in the fusion-positive RMS, little is known for that in fusion-negative RMS (FN-RMS).

Methods: We explored the molecular mechanisms and the driver genes of FN-RMS through frequent gene co-expression network mining (fGCN), differential copy number (CN) and differential expression analyses on multiple RMS transcriptomic datasets.

Results: We obtained 50 fGCN modules, among which five are differentially expressed between different fusion status. A closer look showed 23% of Module 2 genes are concentrated on several cytobands of chromosome 8. Upstream regulators such as MYC, YAP1, TWIST1 were identified for the fGCN modules. Using in a separate dataset we confirmed that, comparing to FP-RMS, 59 Module 2 genes show consistent CN amplification and mRNA overexpression, among which 28 are on the identified chr8 cytobands. Such CN amplification and nearby MYC (also resides on one of the above cytobands) and other upstream regulators (YAP1, TWIST1) may work together to drive FN-RMS tumorigenesis and progression. Up to 43.1% downstream targets of Yap1 and 45.8% of the targets of Myc are differentially expressed in FN-RMS vs. normal comparisons, which also confirmed the driving force of these regulators.

Discussion: We discovered that copy number amplification of specific cytobands on chr8 and the upstream regulators MYC, YAP1 and TWIST1 work together to affect the downstream gene co-expression and promote FN-RMS tumorigenesis and progression. Our findings provide new insights for FN-RMS tumorigenesis and offer promising targets for precision therapy. Experimental investigation about the functions of identified potential drivers in FN-RMS are in progress.

GRAPHICAL ABSTRACT
www.frontiersin.org

Graphical Abstract

1 Introduction

Rhabdomyosarcoma (RMS) is one of the most common soft tissue sarcomas in children. It accounts for about 5% of all childhood tumors and represents about 50% of pediatric soft tissue sarcomas cases (1, 2). RMS usually originates and develops from skeletal (striated) muscle cells (3). Traditionally, RMS contains two major subtypes: embryonal RMS (ERMS; accounts for ~60% of RMS cases) and alveolar RMS (ARMS; ~20% of RMS cases); each presents distinct histologic features, molecular alterations, and clinical outcome (2, 3). At the molecular level, ARMS could be further divided into two subtypes according to the presence or absence of the PAX3 or PAX7–FOXO1 gene fusion (1, 4), which refers to the chromosomal translocations t(2;13)(q35;q14) and t(1;13)(p36;q14). A part of PAX3 (on chromosome 2) or PAX7 (on chromosome 1) is fused to FOXO1 (on chromosome 13) gene (4). ARMS with PAX3/PAX7-FOXO1(together called PAX-FOXO1) gene fusion represents about 80% of ARMS cases (1, 4, 5), and the ones without fusion gene accounts for the rest 20%, which exhibit similar molecular patterns and clinical behaviors to ERMS. This indicates that the fusion status can better classify patient outcomes than the histologic subtype (1, 4). Thus, this molecular level classification with respect to PAX–FOXO1 fusion status has been gradually adopted. In this work, we refer to PAX-FOXO1 fusion-positive cases as FP-RMS and the others as FN-RMS. FP-RMS contains relatively low-overall somatic mutation burden (1); therefore, a number of studies have been carried out to investigate whether PAX-FOXO1 gene fusion functions as the driver for the tumorigenesis and development of FP-RMS. Ren et al. (6) confirmed that PAX-FOXO1 fusion genes commit mesenchymal stem cells to a myogenic lineage by inhibiting terminal differentiation and contributing to ARMS formation. Another study showed that PAX-FOXO1 fusion activates transcription factor (TF) MYOD and myogenin and transforms mesenchymal progenitor cells to the skeletal muscle lineage, leading to malignant formation resembling ARMS (6). Rarer gene fusion of PAX3 to NCOA1 (7) or NCOA2 (8) have been reported, but because they are much rarer and NCOA1/2 functions similarly as FOXO1, those fusions can be treated as special cases of PAX-FP RMS.

However, compared with relatively clear understanding of tumorigenesis and driving force in FP-RMS, the potential driver genes of the more heterogeneous FN-RMS, which represent more than 80% of the RMS cases, are still unclear. Previously, several SNVs for FN-RMS have been identified such as RAS as a potential driver gene. RAS mutations has been detected in 5–30% of all RMS patients (9). However, they are only significantly associated with a subset of ERMS patients (75% of high-risk and 45% of intermediate-risk cases had RAS mutations, whereas low-risk ERMS had no mutations) (10), and not with all of the FN-RMS patients. The understanding about the mechanisms of FN-RMS as a whole is still incomplete, and further efforts to look for other potential driver genes of this group are still needed. Moreover, while FN-RMS harbor many more SNVs, these SNVs are barely linked to clinic due to limited recurrent SNVs across RMS specimens, lack of ability to target a cancer-driving mutation, and incorrect assumptions about the relevance of individual aberrations (1115). Accumulated studies have pointed out that DNA copy number variations (CNVs) are common and more highly recurrent in FN RMS (15, 16). In previous genomic analysis, ERMS (all considered as FN-RMS) demonstrates a high frequency of LOH on chromosomes 11p, 11q, and 16q (1) and tends to have more CNV than ARMS (1). However, a more systematic comparative analysis of FN- versus FP-RMS is still needed to pinpoint the driving force behind FN-RMS tumorigenesis and development, as well as to identify promising therapeutic targets for FN-RMS.

Gene co-expression network mining identifies groups of genes possessing highly correlated/anti-correlated expression profiles across samples or disease conditions. These so-called gene co-expression network modules (GCN modules) are generally enriched in specific pathways or biological functions. Some module’s co-expression is the result of common upstream regulators such as TFs. Therefore, detecting GCN modules specific to a disease condition can lead to quick identification of potential driving regulators. Alternatively, module co-expression can also result from the chromosomal CNVs. Genes in such modules reside on the same fragment of a chromosome, and the CN changes cause their expression levels to vary synchronously. Therefore, GCN mining can also serve as a powerful tool to identify both functional CNVs and driving regulators. This approach has been successfully applied to multiple adult cancer studies by our group on colorectal, breast, lung, and kidney cancers to identify potential driver genes and CN changes (1721).

In this study, we aim to identify the driving mechanism of FN-RMS through the coexpression network mining, which can quickly pinpoint down the potential CNV changes as well as the pathway-level changes in the FN-RMS including ERMS. We systematically investigated the molecular mechanisms and the potential upstream regulators of FN-RMS with the gene co-expression network mining approach. First, we applied our previously developed frequent co-expression network (fGCN) mining algorithm to identify frequent (or consensus) co-expression network modules (called fGCN modules) across multiple RMS gene expression datasets and obtained 50 fGCN modules. Second, we explored the distinctive co-expressed modules between FP- and FN-RMS, and five modules (Modules 2, 18, 34, 41, and 46) enriched with genes differentially expressed between the two groups were detected. For these five modules, we observed that four major types of biological processes (namely, extracellular matrix organization, cell morphology, neuron development, and muscle structure related functions) were highly enriched. Specifically, 117 of 504 genes in Module 2 are concentrated on several cytobands on chromosome 8; Module-enriched upstream regulators such as Myc, Yap1, and Twist1 presented distinctive expression patterns between two groups. Moreover, we also found that up to 24% of the Myc target genes and 27% of the Yap1 (through TEA/ATTS activators) target genes are differentially expressed in the FN versus FP-RMS in three datasets. More importantly, up to 45.8% of Myc target genes and up to 43.1% of Yap1/TEA/ATTS targets are differentially expressed in FN-RMS versus normal comparisons (Table 5). These observations promoted us to hypothesize that the CN changes on these enriched cytobands and the upstream regulators work together to generate the co-expression pattern in FN-RMS and contribute to tumor development. To verify our hypothesis, we performed differential CN and gene expression analyses on a separate RMS dataset from St. Jude Children’s hospital. We observed the consistent CN amplification and overexpression patterns comparing FN-RMS with FP-RMS at both discovery datasets and St. Jude validation dataset. Meanwhile, we also detected the same patterns for three transcriptional regulators (MYC, YAP1, and TWIST1). Notably, MYC is located on chr8q24, one of the identified cytobands by fGCN analysis, whereas YAP1 and TWIST1 were previously reported RMS oncogenes or muscle development genes.

In summary, we identified specific cytobands CN amplification on chromosome 8 as well as the enriched upstream transcriptional regulators through GCN mining and validated the findings through genomic and transcriptomic analysis. We believe that those upregulated gene expression through CN amplification combined with those transcriptional regulators work together to induce the gene co-expression and subsequently may drive the FN-RMS tumorigenesis and development. Our findings not only provide new insights for tumorigenesis and development of fusion-negative RMS but also offer promising directions for FN-RMS driving force study and new potential targets for precision therapy. Experimental validations work is currently going to further validate the roles of the potential driver genes in the FN-RMS samples and cell lines.

2 Materials and methods

2.1 Data source

Five RMS datasets including GSE66533 (22), GSE108022 (23), GSE114621 (15), GSE28511 (24), and St. Jude Children’s hospital dataset (https://www.stjude.cloud/ ) were used in our analysis. Among the five, the first three datasets (GSE66533, GSE108022, and GSE114621) served as the discovery datasets and the last two datasets (GSE28511 and St. Jude dataset) were served as the validation dataset. For the discovery datasets and GSE28511 validation dataset, both gene expression data and corresponding clinical data were obtained from the NCBI Gene Expression Omnibus Database (https://www.ncbi.nlm.nih.gov/geo/). For GSE66533 and GSE108022 data, the PAX–FOXO1 fusion status for each sample are available. In GSE108022, the three samples marked as “other fusion” types were excluded from the downstream comparative analysis and the five samples marked as “normal” were used for validation analysis. For GSE114621, only 11 samples were known for their fusion status and seven of the 11 were included in GSE108022. Therefore, GSE114621 dataset was excluded from the further comparative analysis between FP- and FN-RMS samples. For GSE28511, skeletal muscle, tumor adjacent skeletal muscle, and ERMS and ARMS samples were available. Here, the eight ERMS samples (all FN-RMS) and six normal samples (skeletal muscle and tumor adjacent skeletal muscle samples) were used for targets validation. For the St. Jude dataset, matched CN data, expression data, and clinical data were obtained. The CN data were downloaded from UCSC Xena database (http://xena.ucsc.edu ), and the gene expression data and matched clinical data were obtained from St. Jude Cloud (https://www.stjude.cloud ). Here, the 12 samples taken from initial diagnosis were chosen for analysis. The PAX–FOXO1 fusion status of each sample was manually checked using IGV browser and confirmed by Chen et al. (12). In addition, the target genes of the upstream regulators Myc and TEAD family were obtained from the TF Target Gene Database (25). More details about the samples are summarized in Table 1.

TABLE 1
www.frontiersin.org

Table 1 Demographic of RMS samples and their PAX-FOXO1 fusion status in the cohorts used in the study.

2.2 Analysis workflow and statistical analyses

The main objective of this study was to investigate the molecular basis and the potential driver genes of PAX–FOXO1 fusion-negative RMS. The whole workflow is shown in Figure 1. First, fGCN mining was applied to three RMS gene expression datasets (GSE66533, GSE108022, and GSE114621, Table 1) using our developed lmQCM (local maximal quasi-clique merger) to identify frequent (consensus) GCN (fGCN) modules across the datasets. Second, the differential expression analysis was performed between different fusion status samples, and modules with significantly enriched differential co-expressed genes were further investigated. Third, Gene Ontology (GO) enrichment analysis and cytoband enrichment analysis were carried out for all the fGCN modules to obtain the associated biological processes and chromosomal regions with ToppGene (26). In addition, we also explored the potential upstream regulators for all the fGCN modules with Ingenuity Pathway Analysis (IPA) software (27). Fourth, the target genes for identified TFs were checked for their differential expression with respect to fusion status as well as between FN-RMS and normal samples. Finally, we validated our findings on both CN variation level and gene expression level using the independent validation dataset from St. Jude Children’s hospital.

FIGURE 1
www.frontiersin.org

Figure 1 Study workflow: fGCN, frequent co-expression network minging; DE w.r.t. normal samples, Differential expression analysis with respect to normal tissue.

2.3 Frequent co-expression network mining and identification of fusion-negative RMS significant modules

We applied the frequent co-expression network (fGCN) mining tool lmQCM (17) to identify frequent co-expression network modules (fGCN modules) across multiple RMS gene expression datasets (i.e., GSE66533, GSE108022, and GSE114621). This tool has been extensively used to mine tightly connected fGCN modules from multiple datasets in cancers and other diseases (1720) and is capable of generating small and densely connected modules with highly enriched GO terms. To remove noise, genes with a value of zero in more than 50% of the samples in each dataset were excluded, and protein-coding genes were selected for future analysis. To identify the fGCN modules, we firstly calculated the Pearson correlation coefficient (PCC) between each pair of genes in each dataset separately. The gene pairs with top 5 percentile |PCC| among each dataset were chosen for subsequent analysis. Then, the frequency of gene pairs appeared in the top 5 percentile lists were computed among multiply datasets. Finally, by using the frequency as weight, we identified the fGCN modules. The parameters setting were γ = 0.7, Λ = 1.0, t = 1.0, β = 0.4, minimum size of cluster = 10, and adopted PCC to calculate gene-wised correlations.

To identify key modules associating with fusion status, differential expression analysis for each module gene was performed using R package Limma-voom (28). Only genes with both Benjamini–Hochberg adjusted q-values < 0.05 and |fold change| > 1.5 were considered as differential expressed genes (DEGs). Fisher’s exact test was further applied to identify the modules enriched with DEGs between FP-RMS and FN-RMS with the cutoff p-value < 0.05.

2.4 Cytoband and functional enrichment analysis

To identify the associated biological functions or chromosome locations of a particular fGCN module, we performed GO enrichment analysis and cytoband enrichment analysis using ToppGene (https://topgene.cchmc.org) (26) with the above fGCN module genes. Only GO terms or cytoband with false discovery rate (FDR) < 0.05 were considered to be significantly to the particular fGCN module.

2.5 Identification of enriched upstream transcriptional regulators for fGCN modules

To identify the potential upstream regulators of a particular module, we performed the “Core” and “Upstream Regulator” analysis with IPA (27) on identified fGCN module genes. First, we explored the potential upstream regulators for each module. Notably, we only limited to the experimentally validated interactions between upstream regulators and their target genes that have been manually curated in IPA’s Knowledge Base. Next, we further selected the potential upstream regulators for each module as follows: (1) The predicted upstream regulator itself is a transcription factor (TF); (2) The predicted upstream regulator is significantly enriched (Fisher’s exact test with p-value < 0.01); (3) The predicted upstream regulator affected over 10% of the examined module genes. To further investigate the transcriptional factors’ effect, target genes of Myc and Yap1 (through the interaction with TEAD family protein) were obtained from Transcription Factor Target Gene Database (http://http://tfbsdb.systemsbiology.net/) and further examined for their differential expression with respect to normal tissues in GSE108022 and GSE28511, which are the only dataset contains normal samples. The DEGs were called with the same package and thresholds as described in section 2.4.

2.6 Differential expression analysis for the target genes of the identified TFs

To check the effect of upstream regulators on FN-RMS, we obtained the target gene list from Transcription Factor Target Gene Database for Myc and Yap1 (through TEA/ATTS family activators). For each target to Myc or Yap1, differential expression analysis was performed to examine their molecular patterns with respect to fusion status. Furthermore, differential expression analysis between FN-RMS and normal was also carried out within GSE108022 and GSE28511 validation dataset. Moreover, the cytoband enrichment analysis for DEGs within Module 2 was performed to evaluate the influence of specific cytoband CN alteration on FN-RMS. DEGs with both Benjamini–Hochberg adjusted q-values < 0.05 and |fold change| >1.5 were selected.

2.7 Validation of the CNV and differential gene expression on St. Jude RMS samples

We examined the molecular patterns of identified five fGCN modules at both CN level and transcriptomic level in St. Jude RMS data. Particularly, we identified the modules that are enriched with differentially expressed genes and differential CNVs between different fusion status samples. In the CN level, we first performed differential CN analysis on the gene level for each module gene between FP- and FN-RMS. Wilcoxon rank-sum test method was used and p-values less than 0.05 was considered to be significantly different. Then, we applied Fisher’s exact test to identify fGCN modules exhibiting significant CNV difference with respect to fusion status. CN information was directly downloaded from St. Jude UCSC Xena Database. Differential CN with p < 0.05 were selected. Similarly, to identify the modules enriched with differential gene expression, we performed the same DEG analysis as for the two discovery datasets described in the above section. DEGs with both Benjamini–Hochberg adjusted q-values < 0.05 and |fold change| > 1.5 were selected. Then, Fisher’s exact test was further applied to check whether such kind of module was significantly different with respect to fusion status. P-values less than 0.05 were considered to be significantly different.

2.8 Statistical analysis software

Except where noted above, all statistical analyses were performed in R version 3.5.1.

3 Results

3.1 Identification of key modules associated with fusion status in RMS

Gene co-expression network mining has been proven as an effective and efficient approach to infer tumor mechanisms with transcriptomic data for potential driver genes or drug targets (1721). To obtain robust and consensus co-expression modules in RMS, we performed frequent co-expression network (fGCN) analysis across multiple RMS gene expression datasets and 50 consensus gene expression modules were generated (Supplementary Table S1). Considering the distinctions exist with regard to tumor progression and prognosis between FP and FN-RMS, we focused on identify the modules enriched with genes transcriptomics different between the fusion status. Therefore, we extracted the modules enriched with differentially expressed genes using transcriptomic data from GSE66533 and GSE108022 individually (GSE114621 was omitted as the fusion status of most samples are not available). In GSE66533, six fGCN modules are enriched with DEGs between different fusion status samples (Supplementary Table S2), whereas in GSE108022, seven fGCN modules are enriched with such DEGs (Supplementary Table S2). Specifically, among them, five modules, which are fGCN Modules 2, 18, 34, 41, and 46, are shared between the two datasets. As seen from Table 2, the majority of the five fGCN module genes are differentially expressed between different fusion status samples in both GSE66533 dataset and GSE108022 dataset. 58.53% (295/504) genes in Module 2, 70% (21/30) genes in Module 18, 61.54% (8/13) genes in Module 34, 66.67% (8/12) genes in Module 41, and 81.82% (9/11) genes in Module 46 showed significant differential expressions between FN and FP samples in both GSE66533 and GSE108022 (Table 2). Consistent with this, the heatmaps of these modules also indicate such expression difference in GSE66533 and GSE108022 data (Figure 2). All these results suggested the five consensus modules present consistent expression within them but distinctive molecular patterns between different fusion statuses. Therefore, these five modules were further examined with GO enrichment analysis using ToppGene (26) and IPA (27).

TABLE 2
www.frontiersin.org

Table 2 Modules associated with fusion status in both discovery datasets.

FIGURE 2
www.frontiersin.org

Figure 2 Heatmaps of the five modules significantly associated with fusion status. For each module, left is the heatmap with all module genes and right is the heatmap with differential expressed module genes. (A) is for GSE66533 dataset and (B) is for GSE108022 dataset.

3.2 The functional enrichment of identified modules

For five consensus fGCN modules that are significantly differentially expressed between FN- and FP-RMS samples, we observed that four major types of biology processes, that is, extracellular matrix organization, cell morphology, neuron development, and muscle structure related functions are highly enriched for those modules, suggesting that they may play important roles in RMS development and progression (Table 3 and Supplementary Table S3). Extracellular matrix organization (FDR = 2.05E-05), neuron development (FDR = 3.67E-05), and cell morphogenesis (FDR = 3.30E-04) related biological processes were highly enriched for Module 2 genes. Muscle structure-related biological processes were significantly enriched in Module 18 and 46 genes (FDR = 2.78E-05 and 1.35E-05, separately; Table 3 and Supplementary Table S3). For GO enrichment analysis results on all the 50 modules, please see the Supplementary Table S3.

TABLE 3
www.frontiersin.org

Table 3 The summary of GO-terms and upstream regulator enrichment analysis of the five fusion status associated fGCN modules.

3.3 The potential upstream regulators for the consensus fGCN modules

As consensus fGCN modules are commonly associated with CNVs or co-regulated by common regulators such as TFs, we first checked the chromosome co-localization of the module genes. Cytoband enrichment analysis was performed based on the module genes using ToppGene (Supplementary Table S3). Interestingly, 117 of 504 Module 2 genes are significantly enriched on chr8 cytobands of 8q24.3 (23 genes), 8p21.3 (12), 8p11.2 (8), 8q24.13 (7), 8q22.1 (7), and so forth (Figure 3).

FIGURE 3
www.frontiersin.org

Figure 3 The enriched cytobands of Module 2 genes. Black dots are module 2 genes enriched on chromosome 8 and red dots are transcription factor (i.e. MYC and GATA4) located near or on the identified Module 2-enriched cytobands.

Next, to systematically explore how regulators such as the TFs may play their roles in the consensus modules, for each module, we carried out the TF enrichment analysis using IPA Upstream Regulator Analysis (27) on the fGCN genes and the potential upstream regulators for each module were obtained (Table 3). We found that several consensus modules were regulated by several common upstream regulator genes YAP1, KLF11, TWIST1, MYC, GATA4, MYOG, MYOD1, MYCN and so forth (Table 3). Among them, most are differentially expressed between different fusion status in both GSE66533 and GSE108022. They are either previously known to be oncogenic regulators (GATA4, MYC, and YAP1) or myogenic or muscle differentiation factors (GATA4, MYOG, MYOD1, and TWIST1) (Table 3). More specifically, YAP1, KLF11, TWIST1, and MYC are upregulated, whereas MYCN, MYOD1, MYOG, and GATA4 are downregulated in FN- versus FP-RMS samples (Table 4). MYC and TWIST1 are the upstream regulators of Module 41, and YAP1 is the upstream regulator of Module 18 (Table 3). Most notably, two such regulators are actually located near or on the identified Module 2–enriched cytobands (MYC on chr8q24.21; GATA4 on chr8p23.1).

TABLE 4
www.frontiersin.org

Table 4 The differential expression and copy number change for the enriched upstream regulators for the identified modules.

Furthermore, to confirm the upstream regulator activity changes actually caused downstream a cascade of gene expression changes, we checked the relationships between upstream regulator and their downstream targets within different fusion status. We observed that up to 24% of the MYC target genes and 27% of the YAP1/TEAD target genes (YAP1 acts on genes through TEA/ATTS family transcription activators) are differentially expressed in the FN- versus FP-RMS in the discovery datasets (Table 5). To check whether the same trend also exists for the tumor versus normal sample, we examined the same set of target genes expression in FN-RMS versus normal tissue comparisons and found that 18.9–45.8% of the Myc target genes are differentially expressed in FN versus normal samples. For Yap1/TEA/ATTS family targets, the percentage of targets with altered expression is 15.2–43.1% (Table 5). Meanwhile, we observed that for Module 2, 25% genes in GSE28511 and 52.9% genes in GSE108022 were differential expressed when comparing FN-RMS with normal tissues (Supplementary Figure S1). These DEGs were also significantly enriched in several cytobands on Chr8 (Supplementary Table S4).

TABLE 5
www.frontiersin.org

Table 5 Target genes differential expression in multiply datasets.

Because over 23% (117 of 504) genes from Module 2 are enriched on specific cytobands, we hypothesized that the strong co-expression may not only results from common regulators but also results from the CN alteration on those cytobands.

3.4 CN changes play important roles in fusion-negative RMS

To test our hypothesis, we obtained CN information for a separate RMS dataset from St. Jude Children’s hospital and performed differential CN and gene expression analyses for the fusion status related modules (Modules 2, 18, 34, 41, and 46). The St. Jude data contain eight FN-RMS and four FP-RMS samples with matching RNA-seq and CN data. Overall, a visual examination of the entire genome shows that FN samples contain much more CN amplification than FP samples, especially in the chr8 region, except for one FN sample (Figure 4). For the modules we focused, Module 2 showed significantly differential CNV patterns between FP and FN-RMS (adjusted p-value 1.24E-50) (Supplementary Table S5). At transcriptomic level, Module 2 (adjusted p-value 2.56E-66), Module 18 (adjusted p-value 0.013198578), Module 34 (adjusted p-value 7.82E-10), and Module 41 (adjusted p-value 3.67E-07) presented differential expression patterns between FP-RMS and FN-RMS (Supplementary Table S5).

FIGURE 4
www.frontiersin.org

Figure 4 CNV distribution in FN-RMS vs. FP-RMS of St. Jude dataset. (A) is for entire genome,(B) is for chromosome 8, and (C) is for enriched cytoband genes of chr 8 for Module 2 from discovery datasets.

For Module 2 genes, most of the expressions were elevated in the FN compared with FP samples (Figure 5B), which are consistent with the CN amplification in those regions in FN samples (Figure 5A). More specifically, 45.4% (229/504) of them display higher CN in FN-RMS versus FP-RMS samples. At the transcriptome level, 27.4% (138/504) of Module 2 genes are differential expressed between FN and FP-RMS cases and most of these DEGs were higher expressed in FN-RMS. Moreover, 59 of 229 of the genes discovered with CN alteration show significant changes on mRNA expression (Figure 5). Twenty-eight of 59 (47.5%) genes were also concentrated on the previously identified chr 8 cytobands (Supplementary Table S6). All these results indicated that Module 2 genes are highly altered at both genome and transcriptome level in FN samples, and the CN amplifications of chr 8 contribute significantly to the gene co-expression of Module 2, causing a large portion of the genes to be higher expressed (including MYC) and could further induce other downstream biological changes in FN-RMS.

FIGURE 5
www.frontiersin.org

Figure 5 Heatmaps of Module 2 genes with consistent copy number and expression changes in FN-RMS vs FP-RMS in St. Jude dataset. (A) is for CNV; (B) is for mRNA.

Next, we further checked previously identified common regulators (i.e., YAP1, TWIST1, KLF11, GATA4, MYC, MYCN, MYOG, and MYOD1) in comparative analysis using St. Jude data, and five of eight TFs (YAP1, MYC, TWIST1, GATA4, and MYOD1) are significantly altered in both genomic or transcriptomic levels (Table 4). YAP1 is located on chr11q22.1 and is the common upstream regulator of Module 18. It shows consistent higher CN and gene expression in FN-RMS versus FP-RMS cases in both discovery and St Jude datasets (Table 4). MYC is located on chr8q24.21, which is adjacent to the enriched cytobands we identified through co-expression, the same as YAP1, it presents consistent higher CN and mRNA expression (significantly amplified [p-value = 0.0174] in CN data and near significant upregulation expression data [p-value = 0.058]) in FN- versus FP-RMS cases of St. Jude data (Table 4). TWIST1 is located on 7p21.1 with no CN change in St. Jude dataset but shows consistent higher expression in FN-RMS cases in both discovery and St Jude dataset (Table 4). MYOD1 is located on 11p15.1. It exhibits lower gene expression in FN- versus FP-RMS cases in discovery datasets but shows CN amplification in St. Jude dataset with no significant gene expression change (Table 4). GATA4 is located on chr8p23.1. It exhibits lower gene expression in FN- versus FP-RMS cases in discovery datasets but shows CN amplification in St Jude dataset with near significant downregulation (p-value = 0.057) in St. Jude expression dataset in FN- versus FP-RMS cases (Table 4). KLF11 is located on 2p25.1. It exhibits no CN change in St. Jude dataset and near significant higher expression in FN-RMS cases, which is consistent with the changes of discovery datasets. As for the other two upstream regulators (MYCN and MYOG), no significant differences are observed between two groups at the both genome and transcriptome level in St. Jude dataset (Table 4), possibly due to the small sample size. We also checked MYC, YAP1, and TWIST1 expression changes in FN-RMS with respect to normal tissue in GSE108022 and GSE28511. However, the two datasets showed inconsistent expression change for the three genes; therefore, their expression changes are inconclusive at this point. We expect more normal samples and larger cohort data in the future may give us more definitive answers on this aspect.

Taken together, these results confirmed that the CN amplification of chr 8 and the common regulators (YAP1, MYC, and TWIST1) work together to affect the downstream gene co-expression and contribute FN-RMS progression.

4 Discussion

RMS has been traditionally classified into embryonal RMS (ERMS) and alveolar RMS (ARMS) based on the histopathological features (1). Compared with histological classification, molecular classification based on the presence or absence of PAX3– or PAX7–FOXO1(PAX–FOXO1) fusion can more accurately capture the molecular biology patterns and cytological features of RMS and prognose patient outcome to guide clinical therapy (2), therefore has gained more and more popularity in the RMS research community in recent years. However, compared with the relatively clear fusion-driven oncogenic cascade events in FP-RMS, where the PAX3/7 gene fusion to FOXO1 creates a new fused transcriptional factor and completely altered the downstream targets’ expression profile, therefore serves as one of the driving mutations of FP-RMS, little is known about the molecular mechanisms for the tumorigenesis of FN-RMS which accounts for over 80% RMS cases. There is an urgent need to understand the disease etiology and develop new and effective therapeutic targets for this subtype.

In this study, we applied the fGCN analysis approach to explore the molecular mechanisms behind tumorigenesis in FN-RMS. fGCN analysis identified five consensus modules (Modules 2, 18, 34, 41, and 46), which present distinct expression patterns between FN versus FP RMS samples. Go enrichment analysis reveals that these modules were enriched in four major types of biological processes, that is, extracellular matrix organization, cell morphology, neuron development, and muscle structure processes. Davicioni et al. have demonstrated that neuronal and muscle structure related biological process were significantly altered between FP-RMS and FN-RMS (29), and our results confirmed that. As RMS is a skeletal muscle−derived soft tissue tumor, it is not surprising that muscle differentiation and its dysregulation are major contributors to FN-RMS tumorigenesis and progression, which are indicated by the differentially expressed fGCN modules 18 and 46. Cell morphology refers to the shape and size of cells and pathologists performed diagnosis and prognosis of disease based on the changes of cell morphology. In addition, the interaction of extracellular matrix (ECM) with tumor cells plays important roles in tumor invasion and metastasis. There is no doubt that the dysregulation of cell morphology and extracellular matrix related biological process promote FN-RMS development.

CN amplification of specifical cytobands has been observed in FN-RMS samples, although their roles in tumorigenesis were not further explored (15, 16, 30). Previous studies also discovered that ERMS samples (FN-RMS) displayed chromosomal region gains, including regions on chr 8 (30, 31). In our analysis, we observed the same CN amplifications on FN samples (Figure 4), and Module 2 genes were highly enriched on chr 8 with CN amplification and the consistent gene overexpression in FN-RMS samples (Figure 5) from our validation cohort. Fifty-nine genes from Module2 located on the identified chr 8 cytobands show CN amplification and elevated gene expression. They participate in biological processes such as protein transport and vesicle trafficking and are components of synapse, endoplasmic reticulum membrane, and Golgi apparatus. Not surprisingly, among them are EGFR (32, 33), NCOA2 (34, 35), DERL1 (36, 37), EXT1 (38, 39), PLAG1 (39), COPS5 (40, 41), ASAP1 (41, 42), CHRNA6 (43, 44), and CHRNB3 (44, 45), which have been associated with multiple types of cancers in previous studies. Therefore, the CN amplification not only explains the co-expression of this group of genes but also provides a potential driving force for FN-RMS and generates new potential therapeutic targets for this subtype.

We also discovered that multiple potential oncogenic transcriptional factors MYC, YAP1, and TWIST1 with concordant CN and expression changes may contribute to FN-RMS tumorigenesis as well. MYC, which is located on chr8q24.21 and right next to the identified cytobands, showing consistent CN amplification and overexpression changes in FN-RMS samples (Figure 3) (Table 4). MYC is also the upstream regulator of Module 41 and 9.2–24% of the targets are differentially expressed compared FN-RMS with FP-RMS in the discovery datasets (Table 5). More importantly, as high as 45.8% of Myc target genes are differentially expressed in FN-RMS versus normal samples (Table 5). Accumulated evidences have suggested that MYC was an important oncogenic factor in RMS (46, 47), and its upregulation is closely related to tumor aggression and poor clinical outcome (4651). Durbin et al. have highlighted that MYC could pathogenically subvert a myogenic core regulatory circuitry (CRC; the essential TFs of RMS) to promote RMS tumorigenesis and progression (52). They have also pointed that MYC was closely associated with the CN amplification (52). MYC higher expression were observed in ERMS (FN RMS) samples and FN-ARMS (53, 54). Moreover, several studies demonstrated that the depletion of c-MYC resulted the decrease of metastatic, invasive, and angiogenic-related markers (53, 5557). Importantly, the inhibition of MYC was observed to decrease tumorigenicity of ERMS and myogenic differentiation (54). All these results agreed with our observation that the CN amplification of MYC contribute to the FN-RMS tumorigenesis and progression and MYC could serve as an attractive target for FN-RMS clinical treatment. YAP1 also presents concordant CN amplification and overexpression in FN-RMS in both discovery datasets and St. Jude data (Table 4). This observation is consistent with another report showing CN amplification and overexpression of YAP1 in ERMS (FN-RMS) compared with FP-ARMS (58). Meanwhile, we noticed that YAP1 is the upstream regulator of Module 18, which presenting distinct expression patterns between FN- and FP-RMS are enriched with muscle differentiation and development genes, and about up to 27% of the YAP1/TEAD target genes are differential expressed in the FN- versus FP-RMS in the discovery datasets. Therefore, we proposed that YAP1 CN amplification may also contribute to the different molecular patterns between two groups. Tremblay et al. found that reducing YAP1 expression could decrease the expression of mature skeletal muscle differentiation related genes (58). Moreover, they have also highlighted that the hyperactivity of YAP1 was highly related to the increased tumor stage and poor prognosis in ERMS and FN-ARMS (FN-RMS) and the knockdown of YAP1 could reduce ERMS (FN-RMS) tumorigenicity. Although we did not observe consistent upregulation of YAP1 in FN-RMS versus normal samples among different cohorts, we did observe that up to 43.1% YAP1 target genes are significantly differentially expressed in this comparison. All these evidences suggest that YAP1 is a potential oncogenic driver in FN-RMS. For another TF TWIST1 we identified through fGCN mining, we observed consistent overexpression patterns in FN-RMS samples compared with FP-RMS at both discovery datasets and St. Jude dataset (Table 4). Maestro et al. have showed that TWIST1 CN amplification and overexpression play important roles in RMS development (59). Furthermore, TWIST1 play a key role in myogenesis and genetically or pharmacologically inhibition of TWIST1 could curb cancer-driven muscle cachexia and reduce morbidity and mortality (60). Tremblay et al. have also discovered that YAP1 could upregulate TWIST1 to contribute ERMS (FN-RMS) development (58). This is consistent with our identification that both YAP1 and TWIST1 show higher expressed in FN-RMS. All these evidences support that TWIST1 is another potential driver gene in RMS. MYOD1 is a myogenic regulatory factor and is necessary for muscle fiber differentiation (61). However, its role in FN-RMS samples that we examined shows inconsistent results. In the discover datasets, we observed lower expression of MYOD1 in FN-RMS samples compared with FP-RMS ones. While in the St. Jude dataset, MYOD1 CN amplification is observed in FN-RMS, but no significant changes in term of gene expression between two fusion status subtypes. Zibat et al. identified that MYOD1 is significantly lower expressed in FN-RMS (ERMS and FN-ARMS) than FN-RMS(PAX-FOXO1 ARMS) (62), and lacking of MYOD1 is associated with poor prognosis in ERMS (61). In addition, TWIST1 is a known MYOD1 inhibitor (63) and Tremblay et al. have observed that YAP1 and TWIST1 could inhibit MYOD1 activity in ERMS (58). This agrees with our finding that YAP1, TWIST1, and MYOD1 showed inverse expression patterns (in the discovery dataset). As in the St. Jude dataset, we observed MYOD1 CN amplification in FN-RMS than FP-RMS and no significantly expression difference, this may due to a relatively small sample size in this cohort and no normal skeletal muscles included for analysis. A large cohort with different fusion subtypes and normal skeletal muscles are needed for a definitive answer. For GATA4, it is right on the identified cytoband chr8p23.1 and shows downregulation in discovery datasets and St. Jude dataset transcriptomic level. It is involved in embryogenesis and in myocardial differentiation and function and is a common regulator for fusion status-associated Module 18 and 46. However, we observed CN amplification in St. Jude dataset FN versus FP RMS (Table 4). Therefore, GATA4 role in term of tumorigenesis may require a large cohort data to verify.

In summary, all these findings demonstrate that the amplified CNs on chr8 genes and the overexpression of MYC, YAP1, and TWIST1 may contribute to FN-RMS development, and they can serve as promising targets for FN-RMS therapeutic drug targets. Interestingly, our study also indicated that different molecular mechanisms in terms of genomics alteration exist among distinct RMS subtypes, which induce similar downstream transcriptomic changes and biological processes to manifest the similar tumor development in the skeletal muscle tissues.

Despite the extensive observations and consistent results generated from our analysis, some limitations of this study should be noticed as well. First, while the molecular mechanisms and the potential drivers have been largely defined, further experimental validation are still needed to confirm their function in FN-RMS. Second, as the two datasets contain very few normal skeletal muscles samples (five for GSE108022 and six for GSE28511), further larger scale comparative study between FN-RMS and normal skeletal muscles are still needed to verify the contribution of CN alteration and the three TF changes in FN-RMS tumorigenesis and progression. Third, even though the driving force of CNV has been revealed in FN-RMS, not all FN-RMS samples shown that CN changes. Other driving mechanism may also exist or co-exist in the FN-RMS patients. As FN-RMS is still a quite heterogeneous RMS subtype, the driving force can be multiple factors in multiple levels that work concordantly toward the tumorigenesis in different FN samples. It is very possible that they may possess other important somatic mutations and SNVs such as the ones in RAS pathway. As of now, we do not have the somatic mutation or SNV data for the samples that we investigated in our validation cohort. In our future work, when these data are available, we will separate the RAS pathway SNV and mutations from the wild-type samples, see how much CNV effects contribute to each subtype, and focus more on the RAS wild-type samples, which CNV may exert strong driving forces. Last but not the least, although MYC, YAP1, and TWIST1 can be promising therapeutic targets, both pharmacologically and clinical trials assessment are still needed before clinical application.

5 Conclusions

In conclusion, we identified the potential driver genes and investigate their molecular mechanism for PAX3/PAX7-FOXO1 fusion-negative RMS based on fGCN mining, which can also serve as promising therapeutical drug targets in FN-RMS. We confirmed that the CN amplification of specific cytobands (mostly on chromosome 8) and the elevated expression of the upstream regulators (such as MYC, YAP1, and TWIST1) work together to regulate the downstream gene differential expression and subsequent induce extracellular matrix organization, cell morphology, neuron development, and muscle structure–related biological processes dysfunction, which promoting FN-RMS tumorigenesis and development. Our findings not only helped to improve the understanding of the mechanisms of FN-RMS but also offered promising potential targets for FN-RMS personalized therapeutic intervention. Future works of manipulating these driver genes on FN-RMS cell lines are undergoing.

Data availability statement

The datasets presented in this study can be found in online repositories. The names of the repository/repositories and accession number(s) can be found in the article/Supplementary Material.

Author contributions

Conceptualization, JZ and KH; methodology, XZ, YL, AJ, and JZ; validation, XZ, AJ, SH, WW, and BY; formal analysis, YL; investigation, XZ; resources, PP and KP; data curation, XZ; writing—original draft preparation, XZ; writing—review and editing, XZ and JZ; visualization, XZ; supervision, JZ, KH, JR, and XY; project administration, JZ, KH, and JR; funding acquisition, XZ, JZ, KH, JR, and XY. All authors contributed to the article and approved the submitted version.

Funding

This work was funded by Indiana University Precision Health Initiative to JZ and KH; The National Natural Science Foundation of China (grant number No. 32100530) to XZ, 2021 Chongqing Doctoral “Through Train” Scientific Research Project (grant number No. CSTB2022BSXM-JCX0065) to XZ and Youth project of science and technology research program of Chongqing Education Commission of China (grant No. KJQN202200455) to XZ.

Acknowledgments

We thank Indiana University High Performance Computing Center for data storage and management and St Jude Children’s hospital for RMS data sharing.

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.

Publisher’s note

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.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fonc.2023.1080989/full#supplementary-material

Supplementary Table 1 | The 50 consensus modules. xlsx

Supplementary Table 2 | Modules associated with fusion status in discovery datasets. xlsx

Supplementary Table 3 | The enriched biological processes and cytobands of the 50 fGCN modules. xlsx

Supplementary Table 4 | Cytoband enrichment for Module 2 DEGs between FN-RMS and normal tissue.

Supplementary Table 5 | The molecular patterns of the five fusion status associated modules in St. Jude dataset. xlsx.

Supplementary Table 6 | Chromosome location of St. Jude Module 2 genes with changes in both CNV level and mRNA level. xlsx.

Supplementary Figure 1 | Heatmap of Module 2 differential-expressed genes with respect to normal tissue.

References

1. Sun X, Guo W, Shen JK, Mankin HJ, Hornicek FJ, Duan Z. Rhabdomyosarcoma: Advances in molecular and cellular biology. Sarcoma (2015) 2015:232010. doi: 10.1155/2015/232010

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Kirk R. Pediatric oncology: Fusion-gene status crucial in RMS. Nat Rev Clin Oncol (2010) 7:356. doi: 10.1038/nrclinonc.2010.88

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Ramadan F, Fahs A, Ghayad SE, Saab R. Signaling pathways in rhabdomyosarcoma invasion and metastasis. Cancer Metastasis Rev (2020) 39:287–301. doi: 10.1007/s10555-020-09860-3

PubMed Abstract | CrossRef Full Text | Google Scholar

4. El Demellawy D, McGowan-Jordan J, de Nanassy J, Chernetsova E, Nasr A. Update on molecular findings in rhabdomyosarcoma. Pathology (2017) 49:238–46. doi: 10.1016/j.pathol.2016.12.345

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Williamson D, Missiaglia E, de Reyniès A, Pierron G, Thuille B, Palenzuela G, et al. Fusion gene-negative alveolar rhabdomyosarcoma is clinically and molecularly indistinguishable from embryonal rhabdomyosarcoma. J Clin Oncol (2010) 28(13):2151–8. doi: 10.1200/JCO.2009.26.3814

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Ren Y-X, Finckenstein FG, Abdueva DA, Shahbazian V, Chung B, Weinberg KI, et al. Mouse mesenchymal stem cells expressing PAX-FKHR form alveolar rhabdomyosarcomas by cooperating with secondary mutations. Cancer Res (2008) 68:6587–65. doi: 10.1158/0008-5472.CAN-08-0859

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Wachtel M, Dettling M, Koscielniak E, Stegmaier S, Treuner J, Simon-Klingenstein K, et al. Gene expression signatures identify rhabdomyosarcoma subtypes and detect a novel t(2;2)(q35;p23) translocation fusing PAX3 to NCOA1. Cancer Res (2004) 64:5539–45. doi: 10.1158/0008-5472.CAN-04-0844

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Mosquera JM, Sboner A, Zhang L, Kitabayashi N, Chen C-L, Sung YS, et al. Recurrent NCOA2 gene rearrangements in congenital/infantile spindle cell rhabdomyosarcoma. Genes Chromosom Cancer (2013) 52:538—550. doi: 10.1002/gcc.22050

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Nakagawa N, Kikuchi K, Yagyu S, Miyachi M, Iehara T, Tajiri T, et al. Mutations in the RAS pathway as potential precision medicine targets in treatment of rhabdomyosarcoma. Biochem Biophys Res Commun (2019) 512:524–30. doi: 10.1016/j.bbrc.2019.03.038

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Dolgikh N, Hugle M, Vogler M, Fulda S. NRAS-mutated rhabdomyosarcoma cells are vulnerable to mitochondrial apoptosis induced by coinhibition of MEK and PI3Kα. Cancer Res (2018) 78:2000–13. doi: 10.1158/0008-5472.CAN-17-1737

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Heuckmann JM, Thomas RK. A new generation of cancer genome diagnostics for routine clinical use: overcoming the roadblocks to personalized cancer medicine. Ann Oncol Off J Eur Soc Med Oncol (2015) 26:1830–7. doi: 10.1093/annonc/mdv184

CrossRef Full Text | Google Scholar

12. Chen X, Stewart E, Shelat AA, Qu C, Bahrami A, Hatley M, et al. Targeting oxidative stress in embryonal rhabdomyosarcoma. Cancer Cell (2013) 24:710–24. doi: 10.1016/j.ccr.2013.11.002

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Seki M, Nishimura R, Yoshida K, Shimamura T, Shiraishi Y, Sato Y, et al. Integrated genetic and epigenetic analysis defines novel molecular subgroups in rhabdomyosarcoma. Nat Commun (2015) 6:7557. doi: 10.1038/ncomms8557

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Shern JF, Chen L, Chmielecki J, Wei JS, Patidar R, Rosenberg M, et al. Comprehensive genomic analysis of rhabdomyosarcoma reveals a landscape of alterations affecting a common genetic axis in fusion-positive and fusion-negative tumors. Cancer Discov (2014) 4:216–31. doi: 10.1158/2159-8290.CD-13-0639

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Xu L, Zheng Y, Liu J, Rakheja D, Singleterry S, Laetsch TW, et al. Integrative Bayesian analysis identifies rhabdomyosarcoma disease genes. Cell Rep (2018) 24:238–51. doi: 10.1016/j.celrep.2018.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Paulson V, Chandler G, Rakheja D, Galindo RL, Wilson K, Amatruda JF, et al. High-resolution array CGH identifies common mechanisms that drive embryonal rhabdomyosarcoma pathogenesis. Genes Chromosomes Cancer (2011) 50:397–408. doi: 10.1002/gcc.20864

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Zhang J, Huang K. Normalized lmQCM: An algorithm for detecting weak quasi-cliques in weighted graph with applications in gene Co-expression module discovery in cancers. Cancer Inform (2014) 13 Suppl 3:137–46. doi: 10.4137/CIN.S14021

CrossRef Full Text | Google Scholar

18. Zhang J, Huang K. Pan-cancer analysis of frequent DNA co-methylation patterns reveals consistent epigenetic landscape changes in multiple cancers. BMC Genomics (2017) 18 Suppl 1:1045. doi: 10.1186/s12864-016-3259-0

CrossRef Full Text | Google Scholar

19. Xiang S, Huang Z, Wang T, Han Z, Yu CY, Ni D, et al. Condition-specific gene co-expression network mining identifies key pathways and regulators in the brain tissue of alzheimer’s disease patients. BMC Med Genomics (2018) 11 Suppl 6:115. doi: 10.1186/s12920-018-0431-1

CrossRef Full Text | Google Scholar

20. Shroff S, Zhang J, Huang K. Gene Co-expression analysis predicts genetic variants associated with drug responsiveness in lung cancer. AMIA Jt Summits Transl Sci Proc AMIA Jt Summits Transl Sci (2016) 2016:32–41.

Google Scholar

21. Xiang Y, Zhang J, Huang K. Mining the tissue-tissue gene co-expression network for tumor microenvironment study and biomarker prediction. BMC Genomics (2013) 14 Suppl 5 Suppl 5:S4. doi: 10.1186/1471-2164-14-S5-S4

CrossRef Full Text | Google Scholar

22. Sun W, Chatterjee B, Wang Y, Stevenson HS, Edelman DC, Meltzer PS, et al. Distinct methylation profiles characterize fusion-positive and fusion-negative rhabdomyosarcoma. Mod Pathol an Off J United States Can Acad Pathol Inc. (2015) 28:1214–24. doi: 10.1038/modpathol.2015.82

CrossRef Full Text | Google Scholar

23. Hayes MN, McCarthy K, Jin A, Oliveira ML, Iyer S, Garcia SP, et al. Vangl2/RhoA signaling pathway regulates stem cell self-renewal programs and growth in rhabdomyosarcoma. Cell Stem Cell (2018) 22:414–27.e6. doi: 10.1016/j.stem.2018.02.002

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Aaron Sarver. endothelial and alveolar rhabdomyosarcoma mRNA expression, arrayexpress-repository, V1 (2013). Available at: https://www.ebi.ac.uk/arrayexpress/experiments/E-GEOD-28511.

Google Scholar

25. Plaisier CL, O’Brien S, Bernard B, Reynolds S, Simon Z, Toledo CM, et al. Causal mechanistic regulatory network for glioblastoma deciphered using systems genetics network analysis. Cell Syst (2016) 3:172–86. doi: 10.1016/j.cels.2016.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Chen J, Xu H, Aronow BJ, Jegga AG. Improved human disease candidate gene prioritization using mouse phenotype. BMC Bioinf (2007) 8:392. doi: 10.1186/1471-2105-8-392

CrossRef Full Text | Google Scholar

27. Krämer A, Green J, Pollard JJ, Tugendreich S. Causal analysis approaches in ingenuity pathway analysis. Bioinformatics (2014) 30:523–30. doi: 10.1093/bioinformatics/btt703

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Law CW, Chen Y, Shi W, Smyth GK. Voom: Precision weights unlock linear model analysis tools for RNA-seq read counts. Genome Biol (2014) 15:R29. doi: 10.1186/gb-2014-15-2-r29

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Davicioni E, Anderson JR, Buckley JD, Meyer WH, Triche TJ. Gene expression profiling for survival prediction in pediatric rhabdomyosarcomas: A report from the children’s oncology group. J Clin Oncol (2010) 28:1240–6. doi: 10.1200/JCO.2008.21.1268

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Emami A, Sepehri Z, Gordon JW, Ghavami S. Biologic and clinical aspects of rhabdomyosarcoma. Int J Basic Sci Med (2017) 2:1–4. doi: 10.15171/ijbsm.2017.01

CrossRef Full Text | Google Scholar

31. Kommoss FKF, Stichel D, Mora J, Esteller M, Jones DTW, Pfister SM, et al. Clinicopathologic and molecular analysis of embryonal rhabdomyosarcoma of the genitourinary tract: evidence for a distinct DICER1-associated subgroup. Mod Pathol an Off J United States Can Acad Pathol Inc (2021) 34:1558–69. doi: 10.1038/s41379-021-00804-y

CrossRef Full Text | Google Scholar

32. Wan Z, Huang S, Mo F, Yao Y, Liu G, Han Z, et al. CSN5 controls the growth of osteosarcoma via modulating the EGFR/PI3K/Akt axis. Exp Cell Res (2019) 384:111646. doi: 10.1016/j.yexcr.2019.111646

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Ranganathan S, Ningappa M, Ashokkumar C, Higgs BW, Min J, Sun Q, et al. Loss of EGFR-ASAP1 signaling in metastatic and unresectable hepatoblastoma. Sci Rep (2016) 6:38347. doi: 10.1038/srep38347

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Olsen JV, Blagoev B, Gnad F, Macek B, Kumar C, Mortensen P, et al. Global, in vivo, and site-specific phosphorylation dynamics in signaling networks. Cell (2006) 127:635–48. doi: 10.1016/j.cell.2006.09.026

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Shi Y, Tao Y, Jiang Y, Xu Y, Yan B, Chen X, et al. Nuclear epidermal growth factor receptor interacts with transcriptional intermediary factor 2 to activate cyclin D1 gene expression triggered by the oncoprotein latent membrane protein 1. Carcinogenesis (2012) 33:1468–78. doi: 10.1093/carcin/bgs171

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Kennedy SA, Jarboui M-A, Srihari S, Raso C, Bryan K, Dernayka L, et al. Extensive rewiring of the EGFR network in colorectal cancer cells expressing transforming levels of KRAS(G13D). Nat Commun (2020) 11:499. doi: 10.1038/s41467-019-14224-9

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Oresic K, Ng CL, Tortorella D. TRAM1 participates in human cytomegalovirus US2- and US11-mediated dislocation of an endoplasmic reticulum membrane glycoprotein. J Biol Chem (2009) 284:5905–14. doi: 10.1074/jbc.M807568200

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Wolf EJ, Miles A, Lee ES, Nabeel-Shah S, Greenblatt JF, Palazzo AF, et al. MKRN2 physically interacts with GLE1 to regulate mRNA export and zebrafish retinal development. Cell Rep (2020) 31:107693. doi: 10.1016/j.celrep.2020.107693

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Kim J-J, Lee H-I, Park T, Kim K, Lee J-E, Cho NH, et al. Identification of 15 loci influencing height in a Korean population. J Hum Genet (2010) 55:27–31. doi: 10.1038/jhg.2009.116

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Hutchins JRA, Toyoda Y, Hegemann B, Poser I, Hériché J-K, Sykora MM, et al. Systematic analysis of human protein complexes identifies chromosome segregation proteins. Science (2010) 328:593–9. doi: 10.1126/science.1181348

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Emdal KB, Pedersen A-K, Bekker-Jensen DB, Tsafou KP, Horn H, Lindner S, et al. Temporal proteomics of NGF-TrkA signaling identifies an inhibitory role for the E3 ligase cbl-b in neuroblastoma cell differentiation. Sci Signal (2015) 8:ra40. doi: 10.1126/scisignal.2005769

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Guo Z, Neilson LJ, Zhong H, Murray PS, Zanivan S, Zaidel-Bar R. E-cadherin interactome complexity and robustness resolved by quantitative proteomics. Sci Signal (2014) 7:rs7. doi: 10.1126/scisignal.2005473

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Dash B, Lukas RJ. Modulation of gain-of-function α6*-nicotinic acetylcholine receptor by β3 subunits. J Biol Chem (2012) 287:14259–69. doi: 10.1074/jbc.M111.322610

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Song Y, Wang Y, Xu L, Ma J, Chen E, Zang R, et al. A genetic variant in CHRNB3-CHRNA6 increases risk of esophageal squamous cell carcinoma in Chinese populations. Carcinogenesis (2015) 36:538–42. doi: 10.1093/carcin/bgv019

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Zhang X-B, Zhao Z-H, Chen H-Y, Wang J-C, Qian J, Yang Y-J, et al. [Human chromosome 8p11 (CHRNB3-CHRNA6) region gene polymorphisms and susceptibility to lung cancer in Chinese han population]. Yi Chuan = Hered (2011) 33:886–94. doi: 10.3724/SP.J.1005.2011.00886

CrossRef Full Text | Google Scholar

46. Dang CV, Resar LM, Emison E, Kim S, Li Q, Prescott JE, et al. Function of the c-myc oncogenic transcription factor. Exp Cell Res (1999) 253:63–77. doi: 10.1006/excr.1999.4686

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Dang CV. MYC on the path to cancer. Cell. (2012) 149:22–35. doi: 10.1016/j.cell.2012.03.003

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Dang CV. C-myc oncoprotein function. Biochim Biophys Acta (1991) 1072:103–13. doi: 10.1016/0304-419X(91)90009-A

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Lee LA, Dang CV. Myc target transcriptomes. Curr Top Microbiol Immunol (2006) 302:145–67. doi: 10.1007/3-540-32952-8_6

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Tang H-Y, Goldman AR, Zhang X, Speicher DW, Dang CV. Measuring MYC-mediated metabolism in tumorigenesis. Methods Mol Biol (2021) 2318:231–9. doi: 10.1007/978-1-0716-1476-1_11

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Kato GJ, Dang CV. Function of the c-myc oncoprotein. FASEB J Off Publ Fed Am Soc Exp Biol (1992) 6:3065–72. doi: 10.1096/fasebj.6.12.1521738

CrossRef Full Text | Google Scholar

52. Durbin AD, Kugener G, Zimmerman MW, Yan C, Dharia NV, Frank ES, et al. Abstract B10: Rhabdomyosarcoma requires MYC family genomic events to pathogenically subvert core-regulatory circuitry. Cancer Res (2020) 80 14_Supplement:B10–0. doi: 10.1158/1538-7445.PEDCA19-B10

CrossRef Full Text | Google Scholar

53. Sims D, Maranyane HM, Damerell V, Govender D, Isaacs AW, Peres J, et al. The c-Myc/AKT1/TBX3 axis is important to target in the treatment of embryonal rhabdomyosarcoma. Cancers (Basel) (2020) 12:501. doi: 10.3390/cancers12020501

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Missiaglia E, Selfe J, Hamdi M, Williamson D, Schaaf G, Fang C, et al. Genomic imbalances in rhabdomyosarcoma cell lines affect expression of genes frequently altered in primary tumors: an approach to identify candidate genes involved in tumor development. Genes Chromosomes Cancer (2009) 48:455–67. doi: 10.1002/gcc.20655

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Marampon F, Ciccarelli C, Zani BM. Down-regulation of c-myc following MEK/ERK inhibition halts the expression of malignant phenotype in rhabdomyosarcoma and in non muscle-derived human tumors. Mol Cancer (2006) 5:31. doi: 10.1186/1476-4598-5-31

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Gravina GL, Festuccia C, Popov VM, Di Rocco A, Colapietro A, Sanità P, et al. C-myc sustains transformed phenotype and promotes radioresistance of embryonal rhabdomyosarcoma cell lines. Radiat Res (2016) 185:411–22. doi: 10.1667/RR14237.1

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Zhang J, Song N, Zang D, Yu J, Li J, Di W, et al. C-myc promotes tumor proliferation and anti−apoptosis by repressing p21 in rhabdomyosarcomas. Mol Med Rep (2017) 16:4089–94. doi: 10.3892/mmr.2017.7101

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Tremblay AM, Missiaglia E, Galli GG, Hettmer S, Urcia R, Carrara M, et al. The hippo transducer YAP1 transforms activated satellite cells and is a potent effector of embryonal rhabdomyosarcoma formation. Cancer Cell (2014) 26:273–87. doi: 10.1016/j.ccr.2014.05.029

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Maestro R, Dei Tos AP, Hamamori Y, Krasnokutsky S, Sartorelli V, Kedes L, et al. Twist is a potential oncogene that inhibits apoptosis. Genes Dev (1999) 13:2207–17. doi: 10.1101/gad.13.17.2207

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Parajuli P, Kumar S, Loumaye A, Singh P, Eragamreddy S, Nguyen TL, et al. Twist1 activation in muscle progenitor cells causes muscle loss akin to cancer cachexia. Dev Cell (2018) 45:712–25.e6. doi: 10.1016/j.devcel.2018.05.026

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Ahmed AA, Habeebu S, Farooqi MS, Gamis AS, Gonzalez E, Flatt T, et al. MYOD1 as a prognostic indicator in rhabdomyosarcoma. Pediatr Blood Cancer. (2021) 68:e29085. doi: 10.1002/pbc.29085

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Zibat A, Missiaglia E, Rosenberger A, Pritchard-Jones K, Shipley J, Hahn H, et al. Activation of the hedgehog pathway confers a poor prognosis in embryonal and fusion gene-negative alveolar rhabdomyosarcoma. Oncogene (2010) 29:6323–30. doi: 10.1038/onc.2010.368

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Yang Z, MacQuarrie KL, Analau E, Tyler AE, Dilworth FJ, Cao Y, et al. MyoD and e-protein heterodimers switch rhabdomyosarcoma cells from an arrested myoblast phase to a differentiated state. Genes Dev (2009) 23:694–707. doi: 10.1101/gad.1765109

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: fusion-negative RMS (FN-RMS), fusion-positive RMS (FP-RMS), frequent co-expression network (fGCN), copy number alteration, upstream regulator

Citation: Zhan X, Liu Y, Jannu AJ, Huang S, Ye B, Wei W, Pandya PH, Ye X, Pollok KE, Renbarger JL, Huang K and Zhang J (2023) Identify potential driver genes for PAX-FOXO1 fusion-negative rhabdomyosarcoma through frequent gene co-expression network mining. Front. Oncol. 13:1080989. doi: 10.3389/fonc.2023.1080989

Received: 26 October 2022; Accepted: 12 January 2023;
Published: 30 January 2023.

Edited by:

Atsushi Makimoto, Tokyo Metropolitan Children’s Medical Center, Japan

Reviewed by:

Kayoko Tao, National Cancer Center Hospital, Japan
Marielle Elizabeth Yohe, National Institutes of Health (NIH), United States

Copyright © 2023 Zhan, Liu, Jannu, Huang, Ye, Wei, Pandya, Ye, Pollok, Renbarger, Huang and Zhang. 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: Jie Zhang, jizhan@iu.edu

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.