Comprehensive N6-methyladenosine (m6A) epitranscriptomic profiling of primary tumors remains largely uncharted. Here, we profiled the m6A epitranscriptome of 10 nonneoplastic lung tissues and 51 lung adenocarcinoma (LUAD) tumors, integrating the corresponding transcriptomic, proteomic, and extensive clinical annotations. We identified distinct clusters and genes that were exclusively linked to disease progression through m6A modifications. In comparison with nonneoplastic lung tissues, we identified 430 transcripts to be hypo-methylated and 222 to be hyper-methylated in tumors. Among these genes, EML4 emerged as a novel metastatic driver, displaying significant hypermethylation in tumors. m6A modification promoted the translation of EML4, leading to its widespread overexpression in primary tumors. Functionally, EML4 modulated cytoskeleton dynamics by interacting with ARPC1A, enhancing lamellipodia formation, cellular motility, local invasion, and metastasis. Clinically, high EML4 protein abundance correlated with features of metastasis. METTL3 small-molecule inhibitor markedly diminished both EML4 m6A and protein abundance and efficiently suppressed lung metastases in vivo.

Significance: Our study reveals a dynamic and functional epitranscriptomic landscape in LUAD, offering a valuable resource for further research in the field. We identified EML4 hypermethylation as a key driver of tumor metastasis, highlighting a novel therapeutic strategy of targeting EML4 to prevent LUAD metastasis.

Lung cancer remains the most commonly diagnosed cancer globally and the leading cause of cancer-related death in both men and women (1). Non–small cell lung cancer (NSCLC), which accounts for approximately 80% to 85% of all cases, is predominantly represented by lung adenocarcinoma (LUAD). LUAD is particularly notable for its poor survival rates (1), underscoring the urgent need for improved understanding of the molecular mechanisms and therapeutic strategies. Recent advancements in molecular technologies, including genomics, transcriptomics, epigenomics, and proteomics, have unveiled some of the genetic, genomic, and epigenetic alterations underlying LUAD lethality (2, 3). In contrast, posttranscriptional mechanisms contributing to tumor biology remain underexplored. Epitranscriptomics, often referred to as the “epigenetics” of RNA, encompasses myriad RNA modifications that refine gene expression (4). N6-methyladenosine (m6A), the most prevalent modification in eukaryotic RNAs, governs mRNA export, splicing, translation, and degradation (5, 6). A large number of m6A modifications are dynamically catalyzed (i.e., added or removed) by writer complex METTL3/METTL14/WTAP and removed by erasers FTO and ALKBH5 (7). These modifications occur in more than 25% of human transcripts (8), often near the stop codons of mRNA transcripts with the consensus motif RRACH (R = A or G; H = A, U, or C; refs. 5, 8), influencing RNA fate and gene expression through recruitment of m6A readers like YTHDF1/2/3, YTHDC1/2, and IGF2BP1/2/3 (5, 6, 9).

There is growing evidence that underscores the significant dysregulation of m6A modifications in tumorigenesis and metastasis (7, 10). m6A influences cancer hallmarks such as proliferation, stemness, metastasis, and immune evasion. Most m6A regulators (writers, erasers, and cofactors) and readers have been implicated in modulating tumor biology via regulating the RNA stability or protein translation of key oncogenes and tumor suppressors during cancer onset and progression (7, 10). Notably, while many of these studies have explored the role of m6A by modulating the abundance of one or multiple m6A regulators in vitro or in vivo, the composite m6A methylome in primary tumors, a product of the orchestrated dysregulation of m6A regulators, remains uncharacterized.

m6A immunoprecipitation followed by high-throughput sequencing (m6A MeRIP-seq) offers a robust method for analyzing transcriptome-wide m6A modifications (8, 11). Employing a previously developed low-input m6A MeRIP-seq protocol (12), we profiled the m6A epitranscriptomes of 10 nonneoplastic lung (NL) tissue samples and 51 primary LUAD tumors. This comprehensive dataset allowed us to assess the overarching consequences of the disrupted m6A regulators on m6A epitranscriptome in LUAD tumors. Our analysis discerned three gene clusters, each of which was characterized by unique genomic features and underwent methylation via either canonical or noncanonical manners. By integrating this epitranscriptomic data with clinicopathologic and patient outcome metrics, we identified distinctive m6A patterns and site-specific modifications that were exclusively associated with patient outcome, bypassing any mRNA and protein correlation. We pinpointed hypermethylation of EML4 as a new driver of metastasis. Pharmacologic inhibition of METTL3 profoundly reduced EML4 m6A modification and protein abundance, leading to a significant reduction in tumor metastases. This study not only provides an extensive m6A profile resource for LUAD research but also lays the groundwork for further studies on m6A regulatory dynamics and novel therapeutic targets.

Low-input m6A MeRIP-Seq Effectively Profiles the Epitranscriptome of Lung Tumors and NL Tissues

Applying an optimized m6A MeRIP-seq approach previously developed by our group (12), we profiled 10 NL tissues and 53 primary LUAD tumors, using 2 μg of total RNA for each sample (Fig. 1A; Supplementary Fig. S1A; Supplementary Table S1). Primary lung tumors were sourced from a subset of patients accrued to the University Health Network Patient-derived Lung Cancer Model Establishment (UHN-PLCME) project, with transcriptomic and proteomic data and rich clinical annotation available (1315). NL tissues were histologically nonneoplastic lung tissues collected from normal-appearing lung of lobectomy specimens from patients undergoing lung cancer resection. Escherichia coli (E. coli) K-12 spike-in RNA was used for library size normalization. On average, more than 80% of the MeRIP IP reads were uniquely mapped, with 85.8% originating from protein-coding genes and 10.5% from long intergenic noncoding RNA (lincRNA) genes (Supplementary Table S1). Notably, only 0.06% of the uniquely mapped IP reads were from ribosomal RNA (rRNA)/small nuclear RNA (snRNA)/small nucleolar RNA (snoRNA). Principal component analysis conducted on read counts from annotated genes revealed distinct separation between NL tissues and tumors in both m6A immunoprecipitated RNA (IP) and the total RNA (INPUT) libraries (Fig. 1B). Two tumor samples were identified as outliers with high percentages of spike-in reads in the INPUT libraries (Fig. 1B; Supplementary Table S1), indicating either poor RNA quality or inaccurate RNA quantification. Consequently, these two tumor samples were excluded from downstream analysis. m6A peak numbers in NL and tumor samples varied between 14,170 and 15,806 and between 8,141 and 15,878, respectively. While m6A peaks were relatively evenly distributed across the gene bodies of lincRNAs (Supplementary Fig. S1B), m6A peaks in protein-coding genes predominantly clustered near the stop codon for both NL and tumor samples (Fig. 1C). The motif “RRACH” was consistently enriched in the majority of samples (Supplementary Table S1), in accordance with previous reports (8, 11). The quantity and distribution of peaks, along with the detection of the consensus motif, underscore the effectiveness of the low-input m6A MeRIP-seq method in profiling the m6A epitranscriptome from patient specimens.

Figure 1.

m6A epitranscriptomic profiling in nonneoplastic lung (NL) and lung adenocarcinoma (LUAD) tissues. A, Overview of m6A epitranscriptomic profiling and multiomics data integration in NL and LUAD tissues. B, Principal component analysis plots of m6A MeRIP-seq IP (left) and INPUT (right) libraries in NL and tumor samples postnormalization. The colors denote different experimental batches. C, Density distribution of m6A peaks across the 5′-UTR, CDS, and 3′-UTR regions of mRNA transcripts. The dashed line indicates the mean density, and the shaded area indicates SD. D, Count of methylated and unmethylated genes at indicated abundance levels in the IP and INPUT libraries. The inserted pie chart illustrates the gene-type proportions for the 8,030 methylated genes, of which IP and INPUT Reads Per Kilobase Million (RPKM) were more than 1 in at least half of NL or tumor samples. E, Whole-transcript m6A IP/INPUT ratios for genes with and without m6A peaks in NL and tumor samples (one-sided Z-test: both P < 2.2 × 10−16). F, Correlation between m6A percentages as measured by m6A-LAIC-seq in H1-hESC cells and log2(mean m6A level) derived from m6A MeRIP-seq in lung samples. Methylated genes with log2(mean m6A level) in the same range in both NL and tumor samples were included for the analysis.

Figure 1.

m6A epitranscriptomic profiling in nonneoplastic lung (NL) and lung adenocarcinoma (LUAD) tissues. A, Overview of m6A epitranscriptomic profiling and multiomics data integration in NL and LUAD tissues. B, Principal component analysis plots of m6A MeRIP-seq IP (left) and INPUT (right) libraries in NL and tumor samples postnormalization. The colors denote different experimental batches. C, Density distribution of m6A peaks across the 5′-UTR, CDS, and 3′-UTR regions of mRNA transcripts. The dashed line indicates the mean density, and the shaded area indicates SD. D, Count of methylated and unmethylated genes at indicated abundance levels in the IP and INPUT libraries. The inserted pie chart illustrates the gene-type proportions for the 8,030 methylated genes, of which IP and INPUT Reads Per Kilobase Million (RPKM) were more than 1 in at least half of NL or tumor samples. E, Whole-transcript m6A IP/INPUT ratios for genes with and without m6A peaks in NL and tumor samples (one-sided Z-test: both P < 2.2 × 10−16). F, Correlation between m6A percentages as measured by m6A-LAIC-seq in H1-hESC cells and log2(mean m6A level) derived from m6A MeRIP-seq in lung samples. Methylated genes with log2(mean m6A level) in the same range in both NL and tumor samples were included for the analysis.

Close modal

A gene was defined as m6A methylated if it contains an m6A peak(s) in at least two samples, either NL or tumor. In total, 11,409 such genes were detected across NL and tumor samples, with an average of 1.8 peaks per gene (Fig. 1D). To reliably quantify m6A level, we further excluded genes with an abundance below Reads Per Kilobase Million of 1 in both INPUT and IP libraries for more than half of NL or tumor samples. This filtration yielded 6,140 unmethylated and 8,030 methylated genes, inclusive of 184 methylated lincRNAs (Fig. 1D). We then quantified the m6A level of each gene with three distinct methods: (i) the maximum ratio of each peak’s IP signal to the corresponding INPUT signal (max peak), (ii) the ratio of the sum of all peaks’ IP/INPUT (sum peak), and (iii) the ratio of whole-transcript IP/INPUT (sum gene). We cross-referenced each method using an external dataset with matched m6A MeRIP-seq and m6A-LAIC-seq in human embryonic stem cell line H1-hESC (16, 17). m6A-LAIC-seq, a technology that extracts both m6A methylated and unmethylated transcripts simultaneously (17), serves as a benchmark for m6A quantification at the gene level.

All three methods were strongly correlated in H1-hESC cells (Supplementary Fig. S1C) and exhibited similar trends in both NL and tumor samples (Supplementary Fig. S1D). The whole-transcript IP/INPUT ratio showed the strongest correlation with the percentage of methylated transcripts as determined by m6A-LAIC-seq (Supplementary Fig. S1C; r = 0.67; P < 2.2 × 10−16), making it the preferred method for subsequent analyses. Whole-transcript IP/INPUT ratios (m6A level hereafter) of the methylated genes were significantly higher compared with the unmethylated genes in both NL and tumor samples (Fig. 1E), further supporting its efficacy as the optimal method for m6A quantification from m6A MeRIP-seq data. In both NL and tumor samples, m6A levels were highly correlated with those observed in H1-hESC (Supplementary Fig. S1E), in agreement with previous findings that the m6A level for a given gene is highly conserved across varied cell types (8, 18). This consistency allowed us to project m6A levels in our dataset to the percentage of methylation as characterized by m6A-LAIC-seq. Notably, log2(m6A level) ranges of −6 to 0, 0 to 2, and 2 to 6 corresponded to methylation percentages of 20% to 40%, 40% to 60%, and 60% to 80%, respectively (Fig. 1F).

Distinct Patterns of Association Between m6A Levels and m6A Regulators

The m6A level is dynamically regulated by m6A writers and erasers (7). Thus, we sought to systematically assess the correlation between the mRNA abundance of m6A regulators and m6A levels for each methylated gene in our cohort. We included three genes (METTL3, METTL14, and WTAP) in the catalytic core components and three essential regulatory adaptors (VIRMA, CBLL1, and ZC3H13) in the writer complex, along with two erasers (ALKBH5 and FTO) in our analysis. First, we examined the mRNA abundance of these regulators in our cohort and the TCGA LUAD cohort (Supplementary Fig. S2A). Both METTL3 and VIRMA showed significant upregulation, while WTAP and FTO were significantly downregulated in tumors compared with NL tissues in both cohorts (Supplementary Fig. S2A). This observation suggests a complex dysregulation of m6A in tumors.

Next, we conducted clustering analysis based on the correlations between mRNA abundance of the regulators and m6A levels of the 8,030 methylated genes in tumor samples. The writer complexes and erasers were clearly separated and clustered into two distinct groups (Fig. 2A). These 8,030 methylated genes formed three groups (G1, G2, and G3) based on their correlation patterns (Fig. 2A; Supplementary Table S2). The m6A levels of most G1 genes displayed a positive correlation with writer complexes and a negative correlation with erasers (Fig. 2A). G2 genes were featured by positive correlation with METTL3 and METTL14 but a negative correlation with CBL11, VIRMA, and the two erasers. The majority of G3 genes exhibited a negative correlation with writer complexes and positive correlation with erasers (Fig. 2A). This diverse association between m6A levels and regulator abundances aligns with previous reports in which m6A levels in certain genes increased, while others decreased after knocking down m6A writer complexes (19). In A549 cells, the proportions of the binding target genes of m6A regulators in the G3 group were significantly lower than in the G1 and G2 groups (Fig. 2B; Supplementary Table S2; Pearson’s χ2 test). Similarly, G3 genes exhibited significantly decreased proportions of target genes of m6A regulators compared with G1 genes in HeLa and HEK-293 cells (Supplementary Fig. S2B and S2C; Pearson’s χ2 test; see Supplementary Methods). About the binding targets of m6A readers, we collected the public data of the target genes of YTHDF, YTHDC2, and IGF2BP proteins identified in A549, HeLa, and HEK-293 cells (see Supplementary Methods). We found that the targets of YTHDC2 and IGF2BP proteins were significantly depleted in the G3 group, whereas the proportions of the binding target genes of YTHDF proteins in the G2 group were significantly higher than those in the G1 and G3 groups (Supplementary Fig. S2D and S2E; Pearson’s χ2 test). Collectively, G3 genes exhibited distinct binding patterns of m6A regulators and readers, with evident lower binding by m6A regulators and readers. The m6A levels of G3 genes were likely influenced by indirect effects or other uncharacterized regulatory mechanisms.

Figure 2.

Categorization of methylated genes based on their correlations between m6A levels and m6A regulator abundances. A, Heatmap illustrating the correlations between m6A levels of methylated genes and the mRNA abundances of m6A regulators. Hierarchical clustering identified three gene groups: G1, G2, and G3. The name of m6A regulators is indicated at the bottom of each column. The types of m6A regulators (writers and erasers) are indicated at the top of each column. The color in the heatmap indicates the correlation for each methylated gene across tumor samples: red denotes positive correlation; blue denotes negative correlation. B, Percentages of genes bound by the corresponding m6A regulators in G1, G2, and G3 gene groups in A549 cells characterized by RIP-seq. C, Mean of m6A levels (left) and RNA abundances (RPKM; right) of the genes in G1, G2, and G3 groups. D, Density distribution of m6A peaks across the 5′-UTR, CDS, and 3′-UTR regions of the genes in G1, G2, and G3 groups (Kolmogorov–Smirnov test). E, Comparisons of the number of the exons (left, Wilcoxon rank-sum one-sided test: G1 vs. G2: P < 2.2 × 10−16; G2 vs. G3: P = 3.45 × 10−14; G1 vs. G3: P < 2.2 × 10−16) and the length of the last CDS (right, Wilcoxon rank-sum one-sided test: G1 vs. G2: P < 2.2 × 10−16; G2 vs. G3: P = 5.22 × 10−4; G1 vs. G3: P < 2.2 × 10−16) among the G1, G2, and G3 groups. F, m6A signals and gene structures of representative genes in each group. G, Top two enriched Kyoto Encyclopedia of Genes and Genomes pathways in each group. Compared with the indicated group, **, P < 0.01; ***, P < 0.001; n.s., not significant.

Figure 2.

Categorization of methylated genes based on their correlations between m6A levels and m6A regulator abundances. A, Heatmap illustrating the correlations between m6A levels of methylated genes and the mRNA abundances of m6A regulators. Hierarchical clustering identified three gene groups: G1, G2, and G3. The name of m6A regulators is indicated at the bottom of each column. The types of m6A regulators (writers and erasers) are indicated at the top of each column. The color in the heatmap indicates the correlation for each methylated gene across tumor samples: red denotes positive correlation; blue denotes negative correlation. B, Percentages of genes bound by the corresponding m6A regulators in G1, G2, and G3 gene groups in A549 cells characterized by RIP-seq. C, Mean of m6A levels (left) and RNA abundances (RPKM; right) of the genes in G1, G2, and G3 groups. D, Density distribution of m6A peaks across the 5′-UTR, CDS, and 3′-UTR regions of the genes in G1, G2, and G3 groups (Kolmogorov–Smirnov test). E, Comparisons of the number of the exons (left, Wilcoxon rank-sum one-sided test: G1 vs. G2: P < 2.2 × 10−16; G2 vs. G3: P = 3.45 × 10−14; G1 vs. G3: P < 2.2 × 10−16) and the length of the last CDS (right, Wilcoxon rank-sum one-sided test: G1 vs. G2: P < 2.2 × 10−16; G2 vs. G3: P = 5.22 × 10−4; G1 vs. G3: P < 2.2 × 10−16) among the G1, G2, and G3 groups. F, m6A signals and gene structures of representative genes in each group. G, Top two enriched Kyoto Encyclopedia of Genes and Genomes pathways in each group. Compared with the indicated group, **, P < 0.01; ***, P < 0.001; n.s., not significant.

Close modal

Among the three groups, G1 genes have the lowest m6A levels (Wilcoxon rank-sum one-sided test: G1 vs. G2 and G1 vs. G3: P values < 2.2 × 10−16; G2 vs. G3: P = 0.49) and highest mRNA abundance (Wilcoxon rank-sum one-sided test: P values < 2.2 × 10−16; Fig. 2C). Notably, G1 genes exhibited a canonical peak distribution pattern, primarily enriched in 3′-UTR and near stop codon (Fig. 2D). In contrast, G2 and G3 genes had a relatively higher number of m6A peaks in their coding sequence (CDS) region compared with G1 (Fig. 2D; Kolmogorov–Smirnov test: G1 vs. G2: P = 1.06 × 10−4; G1 vs. G3: P = 4.38 × 10−3; G2 vs. G3: P = 6.38 × 10−2). We further assessed the genomic characteristics of these genes because they might influence the m6A signal distribution. Specifically, because fragmented RNA for m6A MeRIP is approximately 200 nucleotides (nt), peak distribution in meta-gene analysis over the genomic regions will shift toward the 5′ end of the gene in cases with extremely short CDS. There were negligible differences in the length of 5′-UTR, whole CDS region, and 3′-UTR between the three groups (Supplementary Fig. S2F), indicating that the noncanonical peak distribution of G2 and G3 genes is not a technical artifact. G1 genes contained the highest number of exons with the shortest length of the last CDS, whereas G3 genes harbored the fewest number of exons with the most extended length of the last CDS (Fig. 2E). This was exemplified by genes RBM22, PLA2G15, and GIMAP4 in the G1, G2, and G3 groups, respectively (Fig. 2F). The Kyoto Encyclopedia of Genes and Genomes pathway analysis revealed that G1 genes were linked to terms related to cancer cell survival and metastasis, such as “focal adhesion,” “adherens junction,” “hypoxia-inducible factor-1 (HIF-1),” and “NOTCH signaling pathways” (Fig. 2G; Supplementary Table S2). Interestingly, the top enriched signature in both G2 and G3 genes was herpes simplex virus 1 (HSV-1) infection, which mainly consisted of zinc finger (ZNF) genes (Fig. 2G; Supplementary Table S2). The ZNF genes are a large class of putative transcription factors featured by tandem ZNF structures (20), displaying a more profound shift of peak distribution toward the 5′ of CDS (Supplementary Fig. S2G). There were no statistically significant differences in CDS length between ZNF genes and the remaining G2 and G3 genes (Supplementary Fig. S2H), further reinforcing the notion that the noncanonical distribution was not a technical artifact. The unique association pattern with m6A regulators, combined with the noncanonical m6A peak distribution, suggests that this gene group might undergo a yet-to-be-identified regulation by m6A.

A Unique m6A Subtype Is Associated with Favorable Clinical Outcome

To explore the heterogeneity of m6A modification in lung patient tumors, we performed consensus clustering based on m6A levels from the top 20% of methylated genes, ranked by interquartile range (IQR). This led to the identification of five m6A subtypes (P1–5; Fig. 3A). We applied a similar consensus clustering approach to RNA and protein levels. Consistent with the positive correlations between RNA and protein abundance for a significant proportion of genes, there was significant overlap between the RNA and protein subtypes (Supplementary Fig. S3A). However, no significant overlap was found between the m6A subtypes and those of RNA or protein (Supplementary Fig. S3A; Supplementary Table S3). Furthermore, two RNA expression-based LUAD classification methods were compared with the m6A subtypes (21, 22). In the four-cluster (C1–C4) LUAD classification system, C1 and C2 were characterized by high expression of proliferation-associated genes and similarities with small-cell lung cancer. In contrast, the C4 subclass was akin to normal lung and associated with better survival. The C3 subclass displayed features of both C4 and normal lung (21). In the “bronchioid/magnoid/squamoid” classification system, the bronchioid subclass was associated with early-stage tumors and favorable outcome (22).

Figure 3.

m6A epitranscriptome-based LUAD subtyping. A, Consensus clustering using m6A levels from the top 20% of methylated genes by IQR ranking in tumors identified five patient subtypes, labeled P1 through P5. Clinical variables for each patient are depicted in the covariate bar to the right. B, Mean values of Z-transformed m6A levels from the top 20% methylated genes by IQR ranking are shown for each m6A level-derived patient group (Wilcoxon rank-sum two-sided test). C, Comparison of overall survival between the P2 group and the rest of the cohort. D, A volcano plot illustrates the statistical significance against the magnitude of changes in RNA abundances between the P2 group and the rest of the cohort. FC, fold change; FDR, false discovery rate. The differentially expressed genes (DEG) are labeled with darker colors. E, Top enriched GO/BP terms for the downregulated genes in the P2 group. Compared with the indicated group, ***, P < 0.001.

Figure 3.

m6A epitranscriptome-based LUAD subtyping. A, Consensus clustering using m6A levels from the top 20% of methylated genes by IQR ranking in tumors identified five patient subtypes, labeled P1 through P5. Clinical variables for each patient are depicted in the covariate bar to the right. B, Mean values of Z-transformed m6A levels from the top 20% methylated genes by IQR ranking are shown for each m6A level-derived patient group (Wilcoxon rank-sum two-sided test). C, Comparison of overall survival between the P2 group and the rest of the cohort. D, A volcano plot illustrates the statistical significance against the magnitude of changes in RNA abundances between the P2 group and the rest of the cohort. FC, fold change; FDR, false discovery rate. The differentially expressed genes (DEG) are labeled with darker colors. E, Top enriched GO/BP terms for the downregulated genes in the P2 group. Compared with the indicated group, ***, P < 0.001.

Close modal

As shown in Supplementary Fig. S1A, the UHN-PLCME cohort comprises 71.7% early-stage tumors. In line with the characteristics of the aforementioned classifications, the majority of the tumors in the UHN-PLCME cohort was categorized into the C3/C4 subclasses (92.2%; 47 out of 51) and the bronchioid subclass (58.8%; 30 out of 51; Supplementary Fig. S3B and S3C; Supplementary Table S3). These classifications demonstrated significant associations with our RNA and protein subtypings (Supplementary Fig. S3B and S3C). In contrast, when comparing these classifications with the m6A subtypes, we observed no significant overlap (Supplementary Fig. S3B and S3C). This lack of correlation highlights the uniqueness of m6A subtyping, underscoring potential distinct molecular mechanisms that might be overlooked by conventional mRNA expression-based classifications.

Notably, patients in the P2 group demonstrated the lowest m6A levels compared with other groups. This observation held true whether considering the top 20% of methylated genes by IQR ranking (Fig. 3A and 3B) or all methylated genes (Supplementary Fig. S3D). The P2 subtype exhibited relatively low RNA and protein abundances of the core writer complex component WTAP and relatively high protein levels of eraser ALKBH5, along with multiple reader complex proteins (Supplementary Fig. S3E). Intriguingly, the P2 subtype was associated with a significantly better survival rate compared with other groups (Fig. 3C; Supplementary Fig. S3F and S3G). A comparative analysis of the RNA expression profiles between the P2 subtype and the rest of the groups revealed 246 upregulated and 199 downregulated genes (Fig. 3D). A functional enrichment analysis of these differentially expressed genes identified “cell cycle” as the top enriched Gene Ontology/Biological Process (GO/BP) and Kyoto Encyclopedia of Genes and Genomes pathways for the upregulated genes in the P2 group (Supplementary Table S3), whereas the downregulated genes were enriched in immune-related pathways (Fig. 3E; Supplementary Table S3). The majority of the downregulated genes contributing to the enrichment of immune-related pathways were immunoglobulin molecules (Supplementary Table S3). Taken together, we identified a unique m6A subtype associated with a favorable clinical outcome.

BLVRA m6A Levels Predict Clinical Outcome in Lung Adenocarcinoma, Independent of Its mRNA and Protein Levels

To understand the clinical implications of individually methylated genes, we examined the correlations between mRNA, protein, and m6A levels and various clinicopathologic features. These features, which include sex, smoking status, and tumor stage, are documented for all the 51 tumors studied. Each of the transcriptomic, proteomic, and epitranscriptomic datasets contained genes uniquely correlated with a specific feature (Supplementary Fig. S4A and S4B). The m6A levels of TBRG4, SEC24B, and TXLNA were most significantly associated with sex, smoking status, and tumor stage, respectively (Supplementary Fig. S4A and S4B). These associations were generally independent of the mRNA and protein levels of corresponding genes (Supplementary Fig. S4B). We then extended the analysis to clinical outcomes. The cohort included 45 patients with clinical follow-up data, 82% of whom had early-stage cancer (stages I and II; Supplementary Table S1). A total of 635 methylated genes with complementary RNA and protein data available in all the 45 patients were included for the analysis. The m6A levels of 40 genes were associated with overall survival (P < 0.05), with BLVRA being the most significant (Fig. 4A and 4B; log-rank test: P = 3.4 × 10−4). Notably, neither the RNA nor the protein abundances of BLVRA served as prognostic indicators in this cohort (Fig. 4B). Consistently, BLVRA mRNA level was not associated with patient survival in the TCGA LUAD cohort (Supplementary Fig. S4C). In line with these findings, BLVRA m6A level was not correlated with its mRNA or protein abundance (Fig. 4C). However, BLVRA protein and mRNA levels were highly correlated (Supplementary Fig. S4D), indicating that the steady-state abundance of BLVRA protein was mainly driven by its mRNA.

Figure 4.

BLVRA m6A level is independently associated with patient survival outcome. A, Analysis of the associations between overall survival and genes’ m6A, mRNA, and protein levels. The P values from log-rank tests were plotted to illustrate these associations. B, Kaplan–Meier survival curves showing overall survival stratified by BLVRA m6A, mRNA, and protein levels in the UHN-PLCME cohort. The P values from log-rank tests are indicated at the bottom left side. C, Correlation between BLVRA m6A levels with its mRNA (left) and protein abundances (right) in the UHN-PLCME cohort. D, Correlation between BLVRA m6A levels and its m6A peak intensities (peak region IP/INPUT ratio). E, m6A IP and INPUT MeRIP-seq signals of three tumor samples representing low, medium, and high m6A levels in the UHN-PLCME cohort. The predicted motif close to the peak summit is indicated at the bottom. F, SELECT-based quantification of m6A levels at the predicted A site within the motif for the three tumor samples (Student t test). G, Kaplan–Meier survival curves showing overall survival rates stratified by the m6A levels at the A site as determined by SELECT (left) and BLVRA mRNA abundances as measured by microarray (right) in the validation cohort. H, Correlation between BLVRA m6A and mRNA levels in the validation cohort. All histogram data are presented as mean ± SD. Compared with the indicated group, ****, P < 0.0001.

Figure 4.

BLVRA m6A level is independently associated with patient survival outcome. A, Analysis of the associations between overall survival and genes’ m6A, mRNA, and protein levels. The P values from log-rank tests were plotted to illustrate these associations. B, Kaplan–Meier survival curves showing overall survival stratified by BLVRA m6A, mRNA, and protein levels in the UHN-PLCME cohort. The P values from log-rank tests are indicated at the bottom left side. C, Correlation between BLVRA m6A levels with its mRNA (left) and protein abundances (right) in the UHN-PLCME cohort. D, Correlation between BLVRA m6A levels and its m6A peak intensities (peak region IP/INPUT ratio). E, m6A IP and INPUT MeRIP-seq signals of three tumor samples representing low, medium, and high m6A levels in the UHN-PLCME cohort. The predicted motif close to the peak summit is indicated at the bottom. F, SELECT-based quantification of m6A levels at the predicted A site within the motif for the three tumor samples (Student t test). G, Kaplan–Meier survival curves showing overall survival rates stratified by the m6A levels at the A site as determined by SELECT (left) and BLVRA mRNA abundances as measured by microarray (right) in the validation cohort. H, Correlation between BLVRA m6A and mRNA levels in the validation cohort. All histogram data are presented as mean ± SD. Compared with the indicated group, ****, P < 0.0001.

Close modal

BLVRA m6A profile demonstrated a canonical signal distribution, with a single prominent peak near the stop codon (8), suggesting that this singular peak likely influences BLVRA m6A level as a prognostic factor. Indeed, the m6A peak intensity around the stop codon (peak region IP/INPUT ratio) was highly correlated with the m6A level of the entire transcript (Fig. 4D), and this association was also significant about survival (Supplementary Fig. S4E). Near the peak summit, we identified a consensus motif (Fig. 4E). To validate this site, we selected three tumor samples representing low, medium, and high m6A levels, to quantify the m6A level at the “A” site within the motif using the single-base elongation- and ligation-based quantitative polymerase chain reaction (qPCR) amplification method (termed “SELECT”; Fig. 4D–4F; ref. 23). Methylation levels at this “A” site aligned well with that at both peak and whole-transcript levels in these tumor samples (Fig. 4F).

Using the SELECT method, we further assessed m6A levels at the “A” site in a validation cohort of 80 patients with early-stage LUAD with transcriptomic profiling and clinical follow-up data available (24). Consistent with the UHN-PLCME cohort, only the high levels of m6A, but not the mRNA levels of BLVRA, predicted poor survival in the validation cohort (Fig. 4G). No significant correlation was observed between the m6A and mRNA levels of BLVRA (Fig. 4H). In a multivariate Cox regression analysis, after adjustment for age, sex, smoking status, and tumor stage, high BLVRA m6A level remained significantly associated with increased risk of cancer-related death [UHN-PLCME cohort: hazard ratio (HR) (95% confidence interval (CI)): 10.47 (1.02–107.7), P = 0.048; validation cohort: HR (95% CI): 2.48 (1.08–5.70), P = 0.033; Supplementary Table S4]. Therefore, BLVRA’s m6A level, being the first of its kind identified, emerges as a potential prognostic factor for patients with LUAD.

To understand why high BLVRA m6A levels correlate with shorter survival in patients with LUAD, we conducted a differential RNA expression analysis on tumor samples dichotomized by high and low BLVRA m6A levels in our UHN-PLCME cohort. Gene set enrichment analysis identified that epithelial-to-mesenchymal transition (EMT) was the most significantly enriched signature in genes upregulated in tumors with high BLVRA m6A levels (Supplementary Fig. S5A). This is particularly relevant as EMT is known to be associated with drug resistance, immune evasion, metastasis, and poor outcome during cancer progression (2527). Next, we analyzed BLVRA m6A levels under two well-established EMT stimuli, including transforming growth factor beta (TGF-β) and hepatocyte growth factor (HGF; ref. 28). BLVRA m6A modification was significantly elevated upon HGF treatment with negligible changes in its mRNA or protein levels in NSCLC cells, whereas TGF-β had no impact on BLVRA m6A levels (Supplementary Fig. S5B and S5C). The dCas13b-ALKBH5 fusion protein system was utilized to specifically remove the m6A modification from BLVRA transcripts in HGF-treated A549 cells (29). Despite a significant reduction in m6A levels (Supplementary Fig. S5D), we did not observe substantial changes in the abundance of BLVRA mRNA or protein (Supplementary Fig. S5D and S5E), suggesting that while BLVRA m6A level may serve as a biomarker for EMT stimuli and clinical outcome in LUAD, it does not substantially influence RNA stability or translation efficiency.

Prevalent m6A Hypomethylation in Patient Tumors

To identify the unique characteristics of m6A modifications in LUAD between tumor and NL samples, we analyzed the m6A MeRIP-seq data from both sample types. The median number of m6A peaks detected in NL tissues was 15,442, significantly higher than the 11,618 in tumor samples [Fig. 5A (left); Wilcoxon rank-sum one-sided test: P = 1.16 × 10−6]. Every RNA category, including protein-coding RNAs and lincRNAs, displayed significantly higher numbers of m6A peaks in NL tissues (Supplementary Fig. S6A). A much higher variation in peak number was observed in tumors compared with NL tissues [Fig. 5A (left); Levene’s test: P = 3.5 × 10−3]. To further investigate the variation, a pair-wise intragroup comparison was performed in NL and tumor samples. High overlapping rate (>68.6%) of m6A peaks was observed in NL tissues, whereas that of tumor samples was significantly lower, ranging from 38.4% to 81.4% [Fig. 5A (right); Wilcoxon rank-sum one-sided test: P < 2.2 × 10−16], indicating high inter-tumor heterogeneity. The majority of m6A peaks (12,562) were detected in both NL and tumor samples, which was defined as “common peaks” (Fig. 5B; hypergeometric test: P < 2.2 × 10−16). There were 1,607 and 6,519 m6A peaks exclusively detected in NL and tumor samples, which were termed as “NL-specific peaks” and “tumor-specific peaks,” respectively (Fig. 5B). The preponderance of tumor-specific peaks likely stems from the larger sample size in tumor and the noted high intertumor heterogeneity. Indeed, only 10% of the tumor-specific peaks appeared in >20% of tumor samples, whereas half of the NL-specific and 80% of common peaks were observed in >20% of their respective samples (Supplementary Fig. S6B). A negative correlation between m6A and mRNA levels was observed consistently within each sample, with correlation coefficients in NL ranging from −0.45 to −0.39 (mean = −0.42; Supplementary Fig. S6C, left), similar to that observed in human embryonic stem cells (17). In contrast, correlations in tumor samples were less consistent, showing greater variability [Supplementary Fig. S6C, left; −0.33 (−0.45, −0.19); Wilcoxon rank-sum one-sided test: P = 2.44 × 10−5]. Inverse correlations were evident in NL samples when analyzing individual genes across the cohort, but this pattern was nearly absent in tumor samples (Supplementary Fig. S6C, right).

Figure 5.

EML4 upregulation through m6A hypermethylation in LUAD. A, Number of m6A peaks identified in NL and tumor samples (left) and intragroup pair-wise comparison of the overlapping percentage of m6A peaks (right). Peaks are considered overlapping if they share at least one nucleotide. B, Diagram illustrating the distribution of NL-specific, tumor-specific, and common peaks. C, Mean and coefficient of variation of m6A levels in NL and tumor samples. D, Volcano plot of statistical significance and changes of RNA abundances for hypo- and hyper-methylated genes. FC, fold change; FDR, false discovery rate. The DEGs are labeled with darker colors. E, Heatmap of hypo- and hyper-methylated genes. The color represents the Z score of m6A level in each sample. Sample types are indicated on top of the heatmap. Light blue, NL; coral, tumor. F, The statistical significance of m6A level difference was plotted against log2 transformed fold change of m6A level between tumor and NL samples for the 447 DMGs without significant changes at the RNA level. Red denotes hyper-methylated genes; blue denotes hypo-methylated genes. FC, fold change. G, m6A IP and INPUT MeRIP-seq signals near the stop codon of EML4 in representative NL and tumor samples in the UHN-PLCME cohort. The predicted motif nearest to the peak summit is indicated at the bottom. H, Mean of EML4 m6A peak intensities (peak region IP/INPUT ratio) in NL and tumor samples. I,EML4 mRNA levels (left, Student t test), EML4 m6A levels (middle left, peak level and middle right, single A site level; Student t test), METTL3 and EML4 protein levels (middle right), and EML4 translation efficiency (right, Student t test) upon METTL3 knockdown in A549 cells transfected with the indicated siRNAs. Forty-eight hours after siRNA transfection, cells were harvested for analysis. For EML4 m6A analysis, EML4 m6A levels around the peak region and at the “A” site within the predicted motif nearest to the peak summit were determined by m6A MeRIP qPCR and SELECT analysis in METTL3-knockdown and control A549 cells, respectively. TE, translation efficiency. J,EML4 mRNA levels (left), EML4 m6A levels at the “A” site within the predicted motif nearest to the peak summit (middle), and EML4 translation efficiency (right) in A549 cells transfected with dCas13b-ALKBH5 combined with control guide RNA (sgCtrl) or EML4-targeting guide RNA (sgEML4) plasmids for 48 hours (Student t test). TE, translation efficiency. K, Western blot analysis (top) and quantification of EML4 protein levels (bottom) in A549 cells transfected with dCas13b-ALKBH5 combined with sgCtrl or sgEML4 plasmids for 48 hours. The protein levels of EML4 were normalized to β-ACTIN. L,EML4 mRNA (left) and protein (right) abundances in 99 paired tumors and adjacent normal tissues in the CPTAC LUAD cohort. M, Representative images of EML4 protein expression in LUAD tumor and adjacent normal tissues by IHC (left). Protein levels of EML4 in 55 matched LUAD tumors and adjacent normal tissues assessed by IHC score (right). All histogram data are presented as mean ± SD. Compared with the indicated group, **, P < 0.01; ***, P < 0.001; ****, P < 0.0001; n.s., not significant.

Figure 5.

EML4 upregulation through m6A hypermethylation in LUAD. A, Number of m6A peaks identified in NL and tumor samples (left) and intragroup pair-wise comparison of the overlapping percentage of m6A peaks (right). Peaks are considered overlapping if they share at least one nucleotide. B, Diagram illustrating the distribution of NL-specific, tumor-specific, and common peaks. C, Mean and coefficient of variation of m6A levels in NL and tumor samples. D, Volcano plot of statistical significance and changes of RNA abundances for hypo- and hyper-methylated genes. FC, fold change; FDR, false discovery rate. The DEGs are labeled with darker colors. E, Heatmap of hypo- and hyper-methylated genes. The color represents the Z score of m6A level in each sample. Sample types are indicated on top of the heatmap. Light blue, NL; coral, tumor. F, The statistical significance of m6A level difference was plotted against log2 transformed fold change of m6A level between tumor and NL samples for the 447 DMGs without significant changes at the RNA level. Red denotes hyper-methylated genes; blue denotes hypo-methylated genes. FC, fold change. G, m6A IP and INPUT MeRIP-seq signals near the stop codon of EML4 in representative NL and tumor samples in the UHN-PLCME cohort. The predicted motif nearest to the peak summit is indicated at the bottom. H, Mean of EML4 m6A peak intensities (peak region IP/INPUT ratio) in NL and tumor samples. I,EML4 mRNA levels (left, Student t test), EML4 m6A levels (middle left, peak level and middle right, single A site level; Student t test), METTL3 and EML4 protein levels (middle right), and EML4 translation efficiency (right, Student t test) upon METTL3 knockdown in A549 cells transfected with the indicated siRNAs. Forty-eight hours after siRNA transfection, cells were harvested for analysis. For EML4 m6A analysis, EML4 m6A levels around the peak region and at the “A” site within the predicted motif nearest to the peak summit were determined by m6A MeRIP qPCR and SELECT analysis in METTL3-knockdown and control A549 cells, respectively. TE, translation efficiency. J,EML4 mRNA levels (left), EML4 m6A levels at the “A” site within the predicted motif nearest to the peak summit (middle), and EML4 translation efficiency (right) in A549 cells transfected with dCas13b-ALKBH5 combined with control guide RNA (sgCtrl) or EML4-targeting guide RNA (sgEML4) plasmids for 48 hours (Student t test). TE, translation efficiency. K, Western blot analysis (top) and quantification of EML4 protein levels (bottom) in A549 cells transfected with dCas13b-ALKBH5 combined with sgCtrl or sgEML4 plasmids for 48 hours. The protein levels of EML4 were normalized to β-ACTIN. L,EML4 mRNA (left) and protein (right) abundances in 99 paired tumors and adjacent normal tissues in the CPTAC LUAD cohort. M, Representative images of EML4 protein expression in LUAD tumor and adjacent normal tissues by IHC (left). Protein levels of EML4 in 55 matched LUAD tumors and adjacent normal tissues assessed by IHC score (right). All histogram data are presented as mean ± SD. Compared with the indicated group, **, P < 0.01; ***, P < 0.001; ****, P < 0.0001; n.s., not significant.

Close modal

The m6A levels in tumors were significantly lower than those in NL tissues (one-sided Z-test: P = 3.04 × 10−8) and displayed greater variance (Kolmogorov–Smirnov test: P < 2.2 × 10−16; Fig. 5C; Supplementary Fig. S6D), consistent with observation at the peak level (Fig. 5A). Differential methylation analysis identified 652 differentially methylated genes (DMG), including 430 hypo-methylated (20 lincRNAs) and 222 hyper-methylated (4 lincRNAs) genes in tumors (Fig. 5D and 5E; Supplementary Fig. S6E and S6F; Supplementary Table S5). To ensure the robustness of these findings, we performed a validation analysis by randomly subsampling 10 tumors from the 51 LUAD samples 100 times and comparing these subsets with the 10 NL samples. This comparison demonstrated that, on average, more than 70% of the hyper- and hypo-methylated genes identified in the full dataset were consistently observed across the subsampled comparisons (Supplementary Fig. S6G). Additionally, targeted validation was performed on three representative hypo-methylated (AP1M1, KLRG2, and ZNF688) and three representative hyper-methylated (MAT2A, EML4, and R3HDM2) genes, using the m6A MeRIP qPCR assay in an independent cohort comprising 15 LUAD tumor and 15 NL samples. We observed significantly lower m6A levels of the hypo-methylated genes and higher levels of the hyper-methylated genes within tumor samples compared with NL samples (Supplementary Fig. S6H).

Notably, 91.2% of hypo-methylated genes contained common peaks, whereas the proportion of hypermethylated genes with common peaks was only 59% (Supplementary Fig. S6I). Across the 51 tumor samples, more than 2,600 tumor-specific methylated genes were identified, which were specifically methylated in a subgroup of tumors (Supplementary Table S5), in line with the high inter-tumor heterogeneity observed at peak levels (Fig. 5A). These genes were most strongly enriched for pathways related to immunology, including “immunoglobulin receptor binding” and “antigen binding” (Supplementary Fig. S6J; Supplementary Table S5). Although the lung tissue samples used in this study consisted mainly of lung epithelial cells with high purity (>60% tumor cellularity), the bulk tissue samples were infiltrated with other cell types, such as immune cells. To test whether the enriched immunology signatures were related with tumor-specific immune infiltration, we assessed the abundance of six major immune cell subtypes (B cells, CD4 T cells, CD8 T cells, macrophages, neutrophils, and dendritic cells) in both NL and tumor samples in the UHN-PLCME and the TCGA LUAD cohorts (Supplementary Fig. S6K). B cells were the only cell type that showed significantly higher infiltration in tumors compared with NL tissues in both cohorts (Supplementary Fig. S6K; Wilcoxon rank-sum test: UHN-PLCME: P = 3.2 × 10−5, TCGA LUAD: P = 3.5 × 10−2). Consistently, tumor samples with high B-cell infiltration had significantly higher proportion of B-cell-specific genes with m6A peaks (Supplementary Fig. S6L; Wilcoxon rank-sum one-sided test: P = 1.2 × 10−2). Considering that immunoglobulin molecules primarily originate from B lineage cells (30), the elevated B-cell infiltration in the tumor microenvironment likely influences the observed tumor-specific m6A peaks.

Hypomethylation tends to occur in transcripts with low RNA abundance but high m6A levels, whereas hyper-methylation predominantly emerges in transcripts with high RNA abundance but low m6A levels (Supplementary Fig. S6M; Wilcoxon rank-sum one-sided test: P < 2.2 × 10−16). Neither hypo- nor hypermethylated genes showed notable RNA abundance variations between NL and tumor samples (Supplementary Fig. S6M; Wilcoxon rank-sum two-sided test: hypo: P = 0.18; hyper: P = 0.17). Specifically, 68.1% of hypomethylated (293/430) and 69.4% of hyper-methylated (154/222) genes did not show a significant difference at RNA levels between NL and tumor samples (Fig. 5D). These data indicate that the vast majority of the methylation differences between NL and tumor samples do not result in significant alteration in RNA abundance in tumors.

m6A Hypermethylation Enhances EML4 Translation and Promotes the Metastatic Potential of LUAD

Among the 447 DMGs with no significant changes at RNA levels, EML4 is of particular interest: (i) It ranked second in hypermethylated genes (Fig. 5F) and has a single m6A peak near the stop codon (Fig. 5G), ideal for further functional validation, and (ii) EML4 is well-known for its role in the EML4-ALK fusion gene, which produces a transforming fusion tyrosine kinase, an oncogenic driver mutation present in approximately 5% of NSCLC cases (31). However, the independent function of EML4 in cancers, including LUAD, has largely been overlooked. Consistent with the UHN-PLCME cohort, there was no obvious upregulation in EML4 mRNA abundance in tumors compared with adjacent normal samples in the TCGA LUAD cohort [Wilcoxon rank-sum one-sided test: log2(FC (tumor/normal)) = 0.53, P < 1.55 × 10−8; Supplementary Fig. S7A]. Aligning with the m6A level of the entire transcript, EML4 m6A peak intensity (peak region IP/INPUT ratio) in tumor samples was markedly elevated compared with NL tissues (Fig. 5H; Wilcoxon rank-sum one-sided test: P = 2.03 × 10−4).

While knocking down the key writer of m6A, METTL3, had a minimal effect on EML4 mRNA abundance, it resulted in a notable decline in methylation levels around the peak region and at the “A” site within the motif predicted to be closest to the peak summit, along with reduced EML4 protein abundance and translation efficiency (Fig. 5I). Accordingly, the site-specific removal of m6A from this “A” site on EML4 transcript, facilitated by the dCas13b-ALKBH5 fusion protein system, led to a significant reduction in EML4 protein abundance and translation efficiency, with a negligible effect on its mRNA level compared with the controls (Fig. 5J and 5K; ref. 29). Proteomic and transcriptomic data of 99 paired LUAD tumors and adjacent normal tissues from the Clinical Proteomic Tumor Analysis Consortium (CPTAC) LUAD cohort revealed a more pronounced upregulation of EML4 protein abundance in tumor samples than of mRNA abundance [protein: FC (tumor/normal) = 2.78, Wilcoxon rank-sum one-sided test: P = 1.27 × 10−12; mRNA: FC (tumor/normal) = 1.40, Wilcoxon rank-sum one-sided test: P = 7.35 × 10−9; Fig. 5L]. Moreover, 77.8% of LUAD tumors (77 of 99) exhibited an upregulation of EML4 protein in comparison to their corresponding adjacent normal tissues (Supplementary Table S6; ref. 3). Immunohistochemistry (IHC) also underscored the enhanced expression of EML4 protein in primary LUAD (n = 55, paired Student t test: P < 0.0001; Fig. 5M; Supplementary Fig. S7B).

We further set out to assess which m6A reader(s) mediate EML4 protein translation. Given that translation occurs in the cytoplasm, we examined six major m6A readers with cytoplasmic localization, including YTHDF1, YTHDF2, YTHDF3, IGF2BP1, IGF2BP2, and IGF2BP3 (5). Knockdown of IGF2BP1 and IGF2BP3 led to a significant decrease in EML4 protein abundance and translation efficiency, with negligible changes on EML4 mRNA levels (Supplementary Fig. S7C–S7E). In contrast, the depletion of other m6A readers did not significantly impact the abundance or translation efficiency of EML4 (Supplementary Fig. S7C–S7E). RNA-binding protein immunoprecipitation (RIP) qPCR assay confirmed the binding of IGF2BP1 and IGF2BP3 to EML4 mRNA (Supplementary Fig. S7F). Similar regulation patterns in EML4 translation efficiency, as measured by ribosome profiling (Ribo-seq; see Supplementary Methods), were observed in HeLa cells. Specifically, EML4 translation efficiency was significantly reduced upon METTL3 knockdown, whereas knockdown of YTHDF family proteins had no obvious effect [log2(METTL3 knockdown/control FC) = −1.07; Supplementary Fig. S7G].

The protein-to-mRNA ratio served as a proxy for the translation rate, under the assumption that translation efficiency is the dominant posttranscriptional mechanism regulating protein abundance (32). In line with the translation-promoting function of EML4 m6A in cell line models, both EML4 protein and protein-to-mRNA ratio levels were significantly higher in tumors with high EML4 m6A levels compared with those with low levels in the UHN-PLCME cohort (Supplementary Fig. S7H). This reinforces the role of EML4 m6A in enhancing protein translation in LUAD. Taken together, these data suggest that m6A hypermethylation of EML4, catalyzed by classic METTL3-METTL14 methyltransferase complex and recognized by m6A readers IGF2BP1 and IGF2BP3, leads to enhanced protein translation and pervasive overexpression in primary LUAD.

We sought to delineate the functional roles of EML4 in LUAD tumorigenesis. Knockdown of EML4 using siRNA substantially decreased NSCLC cell (A549 and H358) migration and invasion in vitro, as assessed by a Transwell assay, with no notable changes in cell proliferation (Fig. 6A and 6B; Supplementary Fig. S8A and S8B). Similarly, compromised migration abilities were evident in EML4 CRISPR knockout (KO) cells, confirming the specificity of the effects of siRNAs and shRNAs (Fig. 6A–6C; Supplementary Fig. S8A, S8C, and S8D). Statistically significant yet modest effects on cell proliferation were observed with EML4 overexpression, with increased growth in H358 cells and decreased growth in A549 cells (Supplementary Fig. S8A, right). Nevertheless, significantly enhanced tumor cell migration abilities were observed in both EML4-overexpressing A549 and H358 cells (Fig. 6A and 6D; Supplementary Fig. S8E). Wound healing assay and live cell imaging further confirmed the influence of EML4 on cell motility (Fig. 6D and 6E; Supplementary Fig. S8F; Supplementary Videos S1–S5). We thus further investigated the role of EML4 in metastasis in vivo using mouse lung metastases models, established by tail vein injection of A549 and H358 cells into immunodeficient mice. EML4 knockdown strikingly reduced lung metastasis and prolonged survival (Fig. 6E–6G), whereas its overexpression resulted in increased tumor metastasis in the lungs of these mice (Fig. 6H). Collectively, these in vitro and in vivo results indicate that EML4 primarily promotes tumor cell migration, invasion, and ultimately metastasis in LUAD.

Figure 6.

EML4 promotes NSCLC cell migration and invasion in vitro and metastasis in vivo. A, Western blot analysis of EML4 protein expression in EML4-knockdown (left), EML4-overexpressing (middle), and KO (right) cells. For EML4-knockdown assay, 48 hours post siRNA transfection, A549 and H358 cells were harvested for analysis. B, Cell migration (left) and invasion (right) abilities of EML4-knockdown and control A549 and H358 cells. C, Cell migration abilities of EML4 KO and wild-type (WT) A549 and H358 cells. D, Cell migration abilities and migration rates of EML4-overexpressing and control A549 and H358 cells. E, Western blot analysis of EML4 protein expression upon shRNA-mediated EML4 knockdown in A549 and H358 cells. F, Impact of EML4 knockdown on lung metastasis in a tail vein metastasis model. Representative images of the formalin-fixed NOD/SCID mouse lungs (left) and hematoxylin and eosin (H&E) staining of lung tissues (middle) approximately 8 weeks after tail vein injection of 1 × 106EML4-knockdown and control A549 and H358 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (right). G, Survival of NOD/SCID gamma (NSG) mice bearing EML4-knockdown and control H358 lung metastases (log-rank test). H,EML4 overexpression enhanced lung metastasis in a tail vein metastasis model. Representative images of the formalin-fixed NOD/SCID mouse lungs (left) and H&E staining of lung tissues (top middle) approximately 8 weeks after tail vein injection of 1 × 106EML4-overexpressing and control H358 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (right). Representative IHC staining of EML4 (bottom middle) and quantification of EML4 protein expression (right) by IHC score in EML4-overexpressing and control lung metastases. I, F-actin staining in EML4 KO, EML4-overexpressing, and the corresponding control A549 and H358 cells. Nuclei were visualized with 4',6-diamidino-2-phenylindole (DAPI). J, IF staining of EML4 and α-tubulin in EML4-overexpressing and control cells. Nuclei were visualized with DAPI. K, Top enriched CC terms in LUAD tumors with high EML4 protein expression vs. LUAD tumors with low EML4 protein expression from the CPTAC cohort (top 50% vs. bottom 50%; left) and in EML4-overexpressing vs. control H358 cells (right). L, Co-IP assays showing the interactions of EML4 with ARPC1A in Flag-tagged EML4-overexpressing NSCLC cells. M, IF staining of Flag (Flag-tagged EML4) and ARPC1A in Flag-tagged EML4-overexpressing cells. Nuclei were visualized with DAPI. N, IF staining of Flag and ARPC1A in Flag-tagged EML4-overexpressing and control cells (left). Nuclei were visualized with DAPI. Quantification of the percentages of cells with ARPC1A enriched at the edge of lamellipodia in EML4-overexpressing and control cells (right). Protein expression levels of EML4 and ARPC1A (O), F-actin staining (P), cell migration abilities (Q, left), and the relative lamellipodia lengths (Q, right) of EML4-overexpressing and control cells transfected with the indicated siRNAs. Forty-eight hours after siRNA transfection, cells were harvested for the above analysis. All histogram data are presented as mean ± SD. Compared with the indicated group, ***, P < 0.001; ****, P < 0.0001.

Figure 6.

EML4 promotes NSCLC cell migration and invasion in vitro and metastasis in vivo. A, Western blot analysis of EML4 protein expression in EML4-knockdown (left), EML4-overexpressing (middle), and KO (right) cells. For EML4-knockdown assay, 48 hours post siRNA transfection, A549 and H358 cells were harvested for analysis. B, Cell migration (left) and invasion (right) abilities of EML4-knockdown and control A549 and H358 cells. C, Cell migration abilities of EML4 KO and wild-type (WT) A549 and H358 cells. D, Cell migration abilities and migration rates of EML4-overexpressing and control A549 and H358 cells. E, Western blot analysis of EML4 protein expression upon shRNA-mediated EML4 knockdown in A549 and H358 cells. F, Impact of EML4 knockdown on lung metastasis in a tail vein metastasis model. Representative images of the formalin-fixed NOD/SCID mouse lungs (left) and hematoxylin and eosin (H&E) staining of lung tissues (middle) approximately 8 weeks after tail vein injection of 1 × 106EML4-knockdown and control A549 and H358 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (right). G, Survival of NOD/SCID gamma (NSG) mice bearing EML4-knockdown and control H358 lung metastases (log-rank test). H,EML4 overexpression enhanced lung metastasis in a tail vein metastasis model. Representative images of the formalin-fixed NOD/SCID mouse lungs (left) and H&E staining of lung tissues (top middle) approximately 8 weeks after tail vein injection of 1 × 106EML4-overexpressing and control H358 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (right). Representative IHC staining of EML4 (bottom middle) and quantification of EML4 protein expression (right) by IHC score in EML4-overexpressing and control lung metastases. I, F-actin staining in EML4 KO, EML4-overexpressing, and the corresponding control A549 and H358 cells. Nuclei were visualized with 4',6-diamidino-2-phenylindole (DAPI). J, IF staining of EML4 and α-tubulin in EML4-overexpressing and control cells. Nuclei were visualized with DAPI. K, Top enriched CC terms in LUAD tumors with high EML4 protein expression vs. LUAD tumors with low EML4 protein expression from the CPTAC cohort (top 50% vs. bottom 50%; left) and in EML4-overexpressing vs. control H358 cells (right). L, Co-IP assays showing the interactions of EML4 with ARPC1A in Flag-tagged EML4-overexpressing NSCLC cells. M, IF staining of Flag (Flag-tagged EML4) and ARPC1A in Flag-tagged EML4-overexpressing cells. Nuclei were visualized with DAPI. N, IF staining of Flag and ARPC1A in Flag-tagged EML4-overexpressing and control cells (left). Nuclei were visualized with DAPI. Quantification of the percentages of cells with ARPC1A enriched at the edge of lamellipodia in EML4-overexpressing and control cells (right). Protein expression levels of EML4 and ARPC1A (O), F-actin staining (P), cell migration abilities (Q, left), and the relative lamellipodia lengths (Q, right) of EML4-overexpressing and control cells transfected with the indicated siRNAs. Forty-eight hours after siRNA transfection, cells were harvested for the above analysis. All histogram data are presented as mean ± SD. Compared with the indicated group, ***, P < 0.001; ****, P < 0.0001.

Close modal

In line with its role in regulating cell motility, EML4 knockdown, KO, and overexpression induced pronounced cellular morphology changes in NSCLC cells (Fig. 6I; Supplementary Fig. S8G and S8H), suggesting a role in cytoskeleton regulation. Specifically, EML4 knockdown and KO resulted in impaired tumor cell spreading [Fig. 6I (left); Supplementary Fig. S8G], whereas cells with EML4 overexpression exhibited more expansive shapes compared with control cells [Fig. 6I (right)]. Interestingly, axon-like cellular protrusions were induced upon EML4 overexpression in a large proportion of H358 cells [Fig. 6I (right); Supplementary Fig. S8H]. As a member of the echinoderm microtubule-associated protein-like family, EML4 has been reported to bind directly to microtubules, colocalizing with microtubules in interphase in HeLa and U2OS cells and throughout all mitosis phases in HeLa cells (3335). Yet, immunofluorescence (IF) staining revealed diffuse EML4 localization in the cytoplasm of interphase NSCLC cells (Fig. 6J; Supplementary Fig. S8I). EML4 was only colocalized with α-tubulin in the mitotic spindle during mitosis and in the midbody during cytokinesis in A549 cells (Supplementary Fig. S8J). This suggests that EML4 promotes lung tumor cell motility without directly binding to microtubules. Moreover, in a portion of NSCLC cells, EML4 showed membrane localization, which increased upon its overexpression (Fig. 6J; Supplementary Fig. S8I and S8K).

To further elucidate the molecular mechanism underlying the pro-metastatic role of EML4, we performed RNA-sequencing (RNA-seq) in EML4-modified and control A549 and H358 cells (Fig. 6K; Supplementary Fig. S8L; Supplementary Table S7). We also compared the gene expression profiles of LUAD tumors with high and low EML4 protein expression levels from the CPTAC cohort (Fig. 6K; Supplementary Table S7). Pathway enrichment analysis revealed that “membrane” and “cell periphery” were common top ranked terms. They were depleted upon EML4 knockdown and enriched upon EML4 overexpression in both cell lines tested and also enriched in LUAD tumors with high EML4 protein expression (Fig. 6K; Supplementary Fig. S8L; Supplementary Table S7). Other commonly enriched terms identified in both cell line and clinical samples related to cellular protrusions and neuronal synapses (e.g., synapse, axon, and dendrite; Fig. 6K; Supplementary Fig. S8L; Supplementary Table S7). Those terms were closely associated with cell shape control and cell motility and migration, suggesting that the transcriptomic profiles mirror the phenotypic changes induced by EML4. However, an overlapping analysis of differentially expressed genes from the RNA-seq data revealed only a few common differentially expressed genes (Supplementary Table S7), such as synaptotagmin 12 (SYT12) and serpin family E member 1 (SERPINE1; Supplementary Fig. S8M)–both known to promote tumorigenesis and metastasis in multiple types of cancers including NSCLC (3638). The consistency at pathway level, but not at individual gene level, suggests that EML4 may promote cell motility through mechanisms other than direct transcriptional regulation.

To further understand the mechanisms underlying EML4 function, we performed co-immunoprecipitation coupled with mass spectrometry (MS) (Co-IP/MS) in A549 cells, identifying 26 potential interacting proteins unique to EML4 and not found in IgG control precipitates (Supplementary Table S8). Of particular interest is APRC1A, a component of the actin-related protein 2/3 (Arp2/3) complex known for its crucial role in cell motility and metastasis and its cytoplasmic and membrane localization as EML4 (39). The Arp2/3 protein complex is a central modulator of actin cytoskeleton and controls the nucleation of actin filament branches and actin polymerization, thereby regulating a variety of cellular structures and functions that involve the actin cytoskeleton (39, 40). The Arp2/3 complex is crucial for the formation and dynamics of lamellipodia, a major type of membrane protrusions and the crucial force leading to mesenchymal cell migration (41). Overactivation of the Arp2/3 complex drives metastasis and is associated with aggressiveness and tumor progression in varieties of cancers (40, 41). Western blot analysis against ARPC1A using immunoprecipitated Flag-tagged EML4-associated protein complex confirmed the interaction between EML4 and ARPC1A in NSCLC cells (Fig. 6L). IF staining demonstrated colocalization of EML4 and ARPC1A in the cytoplasm as well as at the edge of lamellipodia (Fig. 6M). Functional studies further revealed that overexpression of EML4 significantly enhanced ARPC1A enrichment at the edge of lamellipodia (Fig. 6N). In line with the known function of the Arp2/3 complex, lamellipodia formation was significantly reduced in EML4 KO cells (Supplementary Fig. S9A), whereas EML4 overexpression significantly increased lamellipodia formation (Supplementary Fig. S9B). Reciprocally, silencing of ARPC1A phenocopied the effect of EML4 loss, inducing similar morphology changes and inhibiting both cell migration and lamellipodia formation (Supplementary Fig. S9C–S9F). Furthermore, the cell migration and lamellipodia formation enhancements induced by EML4 overexpression were significantly reduced upon ARPC1A depletion (Fig. 6O–6Q; Supplementary Fig. S9G). These data collectively suggest that EML4 enhances the formation of lamellipodia and the migration capabilities of tumor cells through interacting with ARPC1A in LUAD.

EML4 Protein Expression is Associated with Metastatic Features in LUAD

The m6A-mediated widespread overexpression of EML4 protein in primary LUAD tumors, combined with the pro-metastatic functions of EML4, prompted us to explore the clinical correlation between EML4 protein abundance and pathological characteristics related to metastasis. Elevated EML4 protein expression in primary LUAD tumors significantly correlated with lymph node status in both the CPTAC cohort (Pearson’s χ2 test; P = 0.049) and our internal cohort (Pearson’s χ2 test; P = 0.029; Supplementary Table S6). In contrast, no significant association between EML4 mRNA expression in primary LUAD tumors and lymph node status was observed (Pearson’s χ2 test; P = 0.331; Supplementary Table S6). Furthermore, EML4 protein abundance in metastases (lymph node metastases: n = 36; distant metastases: n = 3) was significantly higher than that in matched primary LUAD tumors (n = 39, paired Student t test; P < 0.0001; Fig. 7A). An increase in EML4 protein level was evident in 79.5% of metastases (31 of 39) compared with their matched primary tumors (Supplementary Table S6). EML4 protein abundance in both primary tumors and lymph node metastases was significantly correlated with the number of positive lymph nodes and the number of positive lymph node stations (Supplementary Fig. S10A and S10B). In addition, EML4 protein expression in widespread lymph node metastases (N2) was significantly higher than that in limited lymph node metastases (N1; Supplementary Fig. S10C). However, METTL3 protein abundance was not significantly elevated in metastases compared with primary tumors as EML4 (Supplementary Fig. S10D and S10E; Supplementary Table S6). There was also no significant correlation between METTL3 and EML4 protein abundance either in primary LUAD tumors (the CPTAC cohort: Supplementary Fig. S10F, left; the internal cohort: Supplementary Fig. S10F, middle) or in metastases (Supplementary Fig. S10F, right), suggesting that EML4 m6A hypermethylation is not caused by METTL3 protein upregulation. Taken together, these analyses in clinical specimens underscore EML4 as a novel driver of LUAD metastatic progression.

Figure 7.

Targeting EML4 with METLL3 small-molecule inhibitor reduces metastasis in vivo. A, Representative IHC staining of EML4 in primary tumor and matched lymph node metastases from one LUAD patient (left). Protein levels of EML4 in 39 matched LUAD tumors and metastases by IHC score (right). B and C, Impact of METTL3 knockdown on lung metastasis in a tail vein metastasis model. Representative images of the NSG mouse lungs (B, top) and H&E staining of lung tissues (B, bottom) approximately 7 weeks after tail vein injection of 1 × 106METTL3-knockdown and control A549 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (C, left). Quantification of METTL3 (C, middle) and EML4 protein expression (C, right) by IHC score in METTL3-knockdown and control lung metastases. D, Relative m6A levels near the stop codon of EML4 gene in A549 cells with the indicated treatment, determined by m6A MeRIP qPCR. E, Rescue assay of STM2457 treatment. Western blot analysis of EML4 protein expression in control and EML4-overexpressing A549 cells with the indicated treatment (E, left). Cell migration abilities of control and EML4-overexpressing A549 cells with the indicated treatment (E, right). F, Rescue assay of METTL3 knockdown. Western blot analysis of METTL3 and EML4 protein expression in control and EML4-overexpressing A549 cells transfected with the indicated siRNAs (F, left). Cell migration abilities of control and EML4-overexpressing A549 cells transfected with the indicated siRNAs (F, right). Forty-eight hours after siRNA transfection, cells were harvested for Western blot and migration analysis. G, Experimental design of STM2457 treatment in the lung metastasis model (top). Representative images of the formalin-fixed NOD/SCID mouse lungs harvested 60 days after treatment with STM2457 (bottom). H, Representative images of H&E staining of mouse lung tissues (left), quantification of the numbers of metastatic nodules in mouse lungs based on H&E staining (middle), and body weight of the mice (right) at the experimental endpoint in the lung metastasis model. All histogram data are presented as mean ± SD. Compared with the indicated group, **, P < 0.01; ****, P < 0.0001; n.s., not significant.

Figure 7.

Targeting EML4 with METLL3 small-molecule inhibitor reduces metastasis in vivo. A, Representative IHC staining of EML4 in primary tumor and matched lymph node metastases from one LUAD patient (left). Protein levels of EML4 in 39 matched LUAD tumors and metastases by IHC score (right). B and C, Impact of METTL3 knockdown on lung metastasis in a tail vein metastasis model. Representative images of the NSG mouse lungs (B, top) and H&E staining of lung tissues (B, bottom) approximately 7 weeks after tail vein injection of 1 × 106METTL3-knockdown and control A549 cells. Quantification of the proportions of tumor metastases in mouse lungs based on H&E staining (C, left). Quantification of METTL3 (C, middle) and EML4 protein expression (C, right) by IHC score in METTL3-knockdown and control lung metastases. D, Relative m6A levels near the stop codon of EML4 gene in A549 cells with the indicated treatment, determined by m6A MeRIP qPCR. E, Rescue assay of STM2457 treatment. Western blot analysis of EML4 protein expression in control and EML4-overexpressing A549 cells with the indicated treatment (E, left). Cell migration abilities of control and EML4-overexpressing A549 cells with the indicated treatment (E, right). F, Rescue assay of METTL3 knockdown. Western blot analysis of METTL3 and EML4 protein expression in control and EML4-overexpressing A549 cells transfected with the indicated siRNAs (F, left). Cell migration abilities of control and EML4-overexpressing A549 cells transfected with the indicated siRNAs (F, right). Forty-eight hours after siRNA transfection, cells were harvested for Western blot and migration analysis. G, Experimental design of STM2457 treatment in the lung metastasis model (top). Representative images of the formalin-fixed NOD/SCID mouse lungs harvested 60 days after treatment with STM2457 (bottom). H, Representative images of H&E staining of mouse lung tissues (left), quantification of the numbers of metastatic nodules in mouse lungs based on H&E staining (middle), and body weight of the mice (right) at the experimental endpoint in the lung metastasis model. All histogram data are presented as mean ± SD. Compared with the indicated group, **, P < 0.01; ****, P < 0.0001; n.s., not significant.

Close modal

Small-Molecule Inhibitor of METTL3 Suppresses EML4 Protein Expression and Metastasis In Vivo

Our findings suggest EML4 as a promising therapeutic target for preventing LUAD metastases. Because silencing METTL3 efficiently reduced EML4’s m6A modification and protein abundance in vitro (Fig. 5I), we hypothesized that suppressing METTL3 would reduce EML4 m6A modification and protein abundance, thereby inhibiting metastasis in LUAD. Consistent with our expectations, METTL3 knockdown significantly decreased tumor cell migration in vitro (Supplementary Fig. S10G) and metastasis in vivo (Fig. 7B and 7C; Supplementary Fig. S10H). Similarly, the METTL3 inhibitor STM2457 markedly reduced both EML4 m6A level and its protein expression (Fig. 7D and 7E; Supplementary Fig. S10I; ref. 42). This reduction coincided with a decline in tumor cell migration (Fig. 7E; Supplementary Fig. S10J). Importantly, overexpression of EML4 significantly restored the cell migration capability in METTL3-knockdown and STM2457-treated cells (Fig. 7E and 7F; Supplementary Fig. S10J). These data suggest that the cell migration-inhibitory effects of METTL3 knockdown and STM2457 were primarily mediated through the modulation of EML4 m6A and protein abundance.

Next, we assessed the efficacy of STM2457 in preventing metastases in vivo. We developed a mouse lung metastasis model by intravenously injecting 1 × 106 A549 cells into each nonobese diabetic/severe combined immunodeficiency (NOD/SCID) mouse. Starting from the day of tumor cell injection, mice were randomly assigned to receive either a control (vehicle) or STM2457 treatment. This treatment was administered daily for an initial five consecutive days, followed by every 3 days for a duration of 36 days (Fig. 7G). We evaluated mouse lung metastases 2 months post the tail vein injection. Our data showed that STM2457 treatment significantly alleviated lung metastases compared with the vehicle group (P < 0.01; Fig. 7H). STM2457 treatment demonstrated no obvious toxicity as evaluated by mice body weight and hematology parameters (Fig. 7H; Supplementary Fig. S11A and S11B). Thus, targeting EML4 through METTL3 inhibition may offer a feasible and promising therapeutic strategy for suppressing LUAD metastases.

Utilizing the low-input m6A MeRIP-seq protocol, we generated the epitranscriptome of 10 NL tissues and 51 LUAD tumors from a well-characterized clinical cohort. This dataset afforded an unparalleled avenue to investigate genome-wide m6A regulation and in relation to clinicopathologic characteristics. Notably, tumors displayed a global decrease and higher variance of m6A levels compared with NL tissues. While the m6A methyltransferase METTL3 is significantly upregulated in LUAD, the methyltransferase core subunit WTAP is significantly downregulated. Previous studies indicated that disruption of WTAP impacts the m6A levels to a greater extent compared with that of METTL3 and METTL14 (19). Thus, the downregulation of WTAP might contribute to the global hypomethylation in tumor samples. Another study reported that histone modification H3K36me3 recruits the m6A methyltransferase complex to deposit m6A on nascent RNAs co-transcriptionally in the nucleus. SET domain containing 2 (SETD2), the histone H3K36me3 methyltransferase, is mutated in 22% of LUAD in the TCGA cohort (2, 43), which might further contribute to m6A hypomethylation in tumors.

m6A methylation is a finely tuned process governed by a robust mechanism. The cornerstone of this mechanism is the m6A methyltransferase complex, comprising the core component METTL3/METTL14/WTAP and adaptors like VIRMA, CBLL1, and ZC3H13). This complex is responsible for the deposition of m6A, while erasers such as ALKBH5 and FTO remove these modifications. A hallmark of m6A is its typical enrichment near the stop codon. While the G1 gene group aligns well with this model, group G3 deviates, suggesting secondary effects or even uncharacterized methylation/demethylation mechanisms. Deepening the complexity, various transcription factors, such as ZFP217, SMAD2/3, and CEBPZ, and exon junction complex can either recruit or repel the m6A methyltransferase complex to specific chromatin sites under specific conditions, thereby shaping m6A topologies (4446). In addition, writers and erasers may have crosstalk. A previous study discovered a mutual regulatory relationship between METTL14 and ALKBH5 (47). Furthermore, because FTO preferentially demethylates nuclear m6A, the spatial regulation may result in different accessibility to its target m6A (48). Thus, these identified and potentially uncharacterized regulatory pathways further complicate the regulatory network, leading to disparities in m6A levels in specific genes when compared with the abundance of their regulators.

By integrating multiomics and survival data, we discovered BLVRA m6A level is associated with patient survival independent of the gene’s mRNA and protein levels. We showed that BLVRA m6A was induced by HGF treatment an EMT-associated stimuli in LUAD cells. Other environmental stimuli may also exist to induce BLVRA m6A. However, BLVRA m6A does not substantially influence RNA stability or translation efficiency, two major functions of m6A modifications. Further study is warranted to investigate the functional impact of BLVRA m6A.

We observed negative correlations between m6A and RNA levels at sample levels in both NL and tumor samples and at gene levels in NL samples. Whether this phenomenon is caused by the RNA-destabilizing function of m6A warrants further investigation (49). Our differential methylation analysis spotlighted EML4 as a notably hyper-methylated gene, underlining its significant pro-metastatic capacity in lung cancer cells. While EML4 is well-known for its EML4-ALK translocation (31), emphasis has largely been on the role of ALK tyrosine kinase in driving the proliferation and growth of the tumor cells, and the function of EML4 remains largely unexplored. One reason is the minimal variation in EML4 mRNA levels observed between tumors and normal samples across various cancer types, including LUAD. Our study revealed an overexpression of EML4 protein in primary LUAD tumors, a process driven by robust m6A modification, promoting its translation. Data from both the CPTAC LUAD cohort and our in-house IHC revealed that EML4 protein levels in primary tumors were approximately three- to fourfold higher than in normal samples. Despite its varied expression in primary tumors, mirroring the inherent intratumoral heterogeneity in LUAD (50), metastases predominantly showed a consistent higher EML4 protein expression. Furthermore, EML4 protein abundance correlates with the degree of lymph node metastases, hinting at its role in tumor aggression. This pattern suggests that cells with elevated EML4 protein might be the primary source in invasion and metastatic spread. EML4 exhibited a distinct membrane localization in LUAD, which was enhanced upon EML4 overexpression along with boosted cell migration abilities. Transcriptomic profiling of EML4-modulated cells pinpointed pathways related to cellular membranes and membrane-formed structures, such as cell protrusions. We further characterized that the migration-promoting effect by EML4 was dependent on an EML4-interacting protein, ARPC1A, which is one of seven subunits of the Arp2/3 protein complex. The Arp2/3 protein complex governs the organization of actin cytoskeleton and drives mesenchymal cell migration by modulating the formation and dynamics of lamellipodia (3941). The molecular mechanism of how EML4 regulates the activity of the Arp2/3 protein complex such as actin polymerization and nucleation of actin filaments warrants further investigation.

Metastasis is the leading cause of cancer-related deaths. Therefore, preventing tumor progression and metastasis is a pivotal strategy to bolster NSCLC outcome. We have pinpointed EML4 as a novel metastatic driver in LUAD, advocating for therapeutic interventions targeting EML4. Our in vitro and in vivo studies confirmed the viability of such an approach, in which METTL3 small-molecule inhibitor mitigated EML4 m6A marking, effectively suppressing its protein translation and potently reducing tumor metastasis in A549 cells with relatively high EML4 m6A modification. Whereas existing research underscores the therapeutic promise of METTL3 inhibitors for tumor growth reduction (51, 52), our findings suggest that these inhibitors could be particularly effective against LUAD metastasis, especially in cases with prominent EML4 m6A modification. Whereas silencing of EML4 has very minimal, if any, impact on cell proliferation, inhibition of METTL3 results in noticeable reduction of lung cancer cell growth. The profound reduction of tumor metastases upon METTL3 knockdown or pharmacologic inhibition is likely because of combinatorial effects on both tumor growth and metastasis. Our study provides strong rationale for developing strategies, such as small-molecule inhibitors, to directly target wild-type EML4 in patients with high EML4 m6A and protein levels to prevent or reduce tumor metastasis. In addition, METTL3 small-molecule inhibitor has been reported to regulate the differentiation and stemness of cancer cells in an m6A-dependent manner (42, 53). Both differentiation and stemness play critical roles in metastasis (54). Therefore, while we demonstrate EML4 as a primary target of METTL3 in driving tumor migration and metastasis in our experimental research models, the additional effects of METTL3 inhibition on tumor biology contributing to metastasis in LUAD warrant further investigation.

As an antibody-based immunoprecipitation method, m6A MeRIP-seq is sensitive to the quality of the antibody used. Cautions must be taken to mitigate potential batch effects. The utility of current nonantibody-based genome-wide m6A quantification methods, such as glyoxal- and nitrite-mediated deamination of unmethylated adenosines (GLORI), m6A-selective allyl chemical labeling and sequencing (m6A-SAC-seq), and evolved TadA-assisted N6-methyladenosine sequencing (eTAM-seq; ref. 55), has yet to be tested in patient tumors with limited RNA material. Therefore, there is significant interest in the continuous development of robust and precise methods for m6A measurement.

The tumor heterogeneity, including intratumoral heterogeneity, tumor microenvironment heterogeneity, and inter-tumor heterogeneity, of the epitranscriptomic marks is underexplored (56). In our study, the tumor-specific m6A methylated genes were enriched in immune-related pathways. Some of these genes, such as PAX5 and CD79B, were specifically expressed in immune cells, prompting the question of whether m6A modifications in those genes are due to increased immune cell infiltration in LUAD. On the other hand, some of these genes, such as HLA genes, were strongly expressed in both LUAD and immune cells. Whether those tumor-specific m6A modifications originated from LUAD or immune cells could not be adequately addressed with bulk-level m6A profiling. Furthermore, it is possible that LUAD subclones exist that possess unique pro-tumorigenic or pro-metastatic epitranscriptomic modifications and are more prone to malignant transformation, metastasis, drug resistance, and immune evasion. These epitranscriptomic modifications were probably underestimated in bulk-level analysis when the quantity of these malignant subclones is small. With the aid of single-cell technologies for the analyses of the epitranscriptomic marks (57, 58), these questions could be addressed in the near future.

In summary, our work pioneered an m6A epitranscriptomic blueprint in LUAD. By aligning it with corresponding transcriptomic, proteomic, and clinical data, we offer an enriched dataset that unearths the intricate dynamics of m6A modifications in a clinical milieu, paving the way for novel therapeutic insights.

Cell Lines

The NSCLC A549 (Cat. #CCL-185) and H358 (Cat. #CRL-5807) cell lines were obtained from the ATCC and maintained in RPMI 1640 medium (Cat. #350-000-CL, Wisent), supplemented with 10% FBS (Cat. #080150, Wisent) and 1% penicillin-streptomycin (Cat. #450-201-EL, Wisent). All cell lines were at early passages (n < 10) and tested for mycoplasma contamination once a month by the MycoAlert Mycoplasma Detection Kit (Cat. #LT07-118, Lonza). All cell lines were authenticated by short tandem repeat every 3 months by Wuhan Pricella Biotechnology Co., Ltd. (China).

Mouse Lung Metastasis Models

For EML4 knockdown, METTL3 knockdown, and EML4 overexpression assays, A549 or H358 cells (overexpression: 5 × 105 cells per mouse; knockdown: 1 × 106 cells per mouse) were injected intravenously through the tail vein into each 4- to 6-week-old male NOD/SCID (Strain #: 001303, The Jackson Laboratory) or NOD/SCID gamma (NSG) mouse (Cat. NO. NM-NSG-001, Shanghai Model Organisms Center, Inc., China). The humane intervention points were closely monitored. The mice were sacrificed ∼8 weeks after tail vein injection. For survival assay, EML4-knockdown and control H358 cells were injected intravenously through the tail vein into each 5-week-old male NSG mouse (1 × 106 cells per mouse). The health conditions of the mice were closely monitored. Mice were euthanized when the humane intervention points were met (e.g., suppressed activity, being unresponsive to touch, marked hunched posture, dehydration, rough hair coat, dyspnea, and weight loss of 30% or more from maximum body weight). For STM2457 treatment assay, the mouse lung metastasis model was first established by tail vein injection of 1 × 106 A549 cells into each ∼5-week-old male NOD/SCID mouse. On the same day of tumor cell injection, the mice were randomly assigned for treatment with vehicle control or 50 mg/kg STM2457 (Cat. #S9870, Selleck) daily for five consecutive days and then every 3 days for 36 days. STM2457 was freshly prepared at each time of treatment. To prepare STM2457 working solution, 10 mg STM2457 was dissolved by adding each solvent one by one: 100 μL DMSO, 800 μL PEG300, 100 μL Tween-80, and 1 mL H2O. The mice were intraperitoneally injected with 50 mg/kg of STM2457 or the same volume of the solvent mix (vehicle group). The mice were sacrificed 8 weeks after tail vein injection. All the five lobes of mouse lungs were harvested and embedded in paraffin. Lung metastasis was assessed either by the number of pulmonary tumor nodules or the proportion of tumor metastases in the mouse lung when there were too many tumor nodules so that single nodule was indistinguishable based on hematoxylin and eosin staining. All mice experimental procedures were approved by the Animal Ethics Committee of the University Health Network (UHN, Canada) or Sun Yat-sen University (China).

Patient Samples

LUAD snap-frozen specimens were collected and banked from patients who underwent surgical resection at the UHN using study protocols approved by the UHN Human Research Ethics Board (protocols 08-0230 and 09-0510). Informed written consent was obtained from each patient. Histology slides were reviewed, and cases were selected for >60% tumor cellularity. The 53 tumors were collected from the UHN-PLCME project during 2005 to 2010 (annotated as PLC samples; Supplementary Table S1; refs. 1315). The validation cohort included a subset of 80 cases from a prognostic gene classifier validation project collected during 1996 to 2005 (annotated as MB samples; Supplementary Table S4; ref.24); these included only cases with residual mRNA still available after previous study. All patients with LUAD received no prior treatment for their disease (targeted therapy, immunotherapy, chemotherapy, or radiotherapy) before surgical resection. Quality assurance protocol included tumor histology review prior to RNA isolation. Diagnoses were based on the WHO classification for lung cancer (59). Patient and tumor demographic features and survival were retrieved from electronic patient records, as per approved protocols mentioned above. NL tissues were collected from patients with early-stage LUAD (IA) with >1 cm resection margin distance from the tumor edge. Fifteen NL tissue samples and 15 primary LUAD tissue samples used for DMG validation were obtained from the First Affiliated Hospital of Sun Yat-sen University. LUAD IHC specimens were obtained from patients with LUAD who underwent surgery at the First Affiliated Hospital of Sun Yat-sen University (Supplementary Table S6). The study protocol was approved by the Clinical Research Ethics Committee of the First Affiliated Hospital of Sun Yat-sen University. Written informed consent was obtained from each patient. All studies were conducted in accordance with the Declaration of Helsinki.

Inducible EML4 KO by CRISPR

To generate stable Cas9-expressing cells, A549 cells were transduced with Cas9 lentivirus, which was generated by transfecting 293T cells with lentiCas9-Blast (Cat. #52962, Addgene) and packaging plasmids. After blasticidin selection, Cas9-expressing A549 cells were transduced with doxycycline-inducible sgRNA lentivirus and enriched by puromycin selection. EML4 knockdown was induced by adding 1 μg/mL doxycycline to Cas9-inducible sgRNA-expressing cells. After 72 hours, cells were harvested for Western blot analysis and cell migration assays.

EML4 KO by CRISPR

Two EML4 sgRNAs (EML4 sgRNA-1: 5′-CUGUUAAGAUAUGGUUAUCG-3′, Cat. #LM2301206-A; EML4 sgRNA-2: 5′-CUAUUUUCCCGGUCGGAAGA-3′, Cat. #LM2301206-B; Supplementary Table S9) were purchased from Beijing Tsingke Biotech Co., Ltd. (China). Cas9 protein (GenCrispr NLS-Cas9-NLS Nuclease, Cat. #Z03469) was purchased from GenScript Biotech (China). Cas9/sgRNA ribonucleoprotein complexes were prepared by mixing Cas9 protein with two EML4 sgRNAs at 1:2.5:2.5 ratio and transfected into A549 and H358 cells by electroporation using Neon NxT Electroporation System (Thermo Fisher Scientific). Two EML4 KO clones (EML4-KO-1 and EML4-KO-2) were then isolated from the transfected cells by single-cell sorting and validated by Western blot analysis.

RNA-seq of NSCLC Cells

Total RNAs were purified with RNeasy Mini Kit (Cat. #74106; QIAGEN) or RNAprep Pure Cell/Bacteria Kit (Cat. #4992235/DP430, TIANGEN, China). Approximately 5 μg of total RNA was used to deplete rRNA using Ribo-Zero Gold rRNA Removal Kit (Cat. #MRZG12324, Illumina). rRNA-depleted RNAs were fragmented using NEBNext Magnesium RNA Fragmentation Module (Cat. #E6150S, NEB) and then reverse-transcribed by SuperScript II Reverse Transcriptase (Cat. #18064071, Thermo Fisher Scientific). Next, U-labeled second-stranded DNAs were synthetized with DNA Polymerase I (E. coli; Cat. #M0209L, NEB), RNase H (Cat. #M0297L, NEB) and dUTP Solution (Cat. #R0133, Thermo Fisher Scientific). After adding adapters, the fragments were purified with AMPure XP beads, were treated with Uracil-DNA Glycosylase (Cat. #M0280, NEB), and were then amplified by PCR. RNA-seq libraries were subjected for 2 × 150 bp paired-end sequencing (PE150) in duplicates at ∼40 million reads per library on an Illumina NovaSeq 6000 (LC-Bio Technology CO., Ltd., Hangzhou, China) following the vendor’s recommended protocol.

IF and F-Actin Staining

Cells were seeded on glass coverslips coated with fibronectin (Cat. #F0895, MilliporeSigma) in six-well plates. Approximately 16 to 24 hours after seeding, cells were fixed with 4% paraformaldehyde (Cat. #P6148, MilliporeSigma)/PBS for 20 minutes and then permeabilized by 0.2%Triton X-100 (Cat. #T9284, MilliporeSigma)/PBS for 5 minutes. Nonspecific binding sites were blocked with 3% FBS/PBS for 1 hour. Primary antibodies (EML4: Cat. #12156S, Cell Signaling Technology; α-tubulin: Cat. #3873S, Cell Signaling Technology; ARPC1A: Cat. #PA5-82149, Thermo Fisher Scientific; Flag: Cat. #F1804, MilliporeSigma) were applied to the cells at room temperature for 2 hours. After PBS wash, secondary antibodies with desired fluorescence probes were applied. Then cells were washed and mounted with ProLong Gold Antifade Mountant with DAPI (4′, 6-diamidino-2-phenylindole; Thermo Fisher Scientific). To stain F-actin, cells were fixed, permeabilized, blocked, and stained with Alexa Fluor Plus 555 phalloidin solution (Cat. #A30106; Thermo Fisher Scientific; dissolved in DMSO) for 30 minutes.

In vitro STM2457 Treatment

STM2457 was purchased from MedChemExpress (Cat. #HY-134836) and dissolved in DMSO as the stock solution. Twenty-four hours after seeding, A549 cells were treated with STM2457 (60 μmol/L) or DMSO for 3 days and harvested for Western blot analysis, cell growth analysis, and cell migration analysis.

Correlation Analysis and Clustering

To assess the correlation between genes’ m6A levels and mRNA abundances of m6A regulators, we calculated the Pearson correlation coefficient between log-transformed m6A levels and mRNA abundances of m6A regulators for individual genes (Fig. 2A). Hierarchical clustering was performed to group the methylated genes according to the correlation coefficient between genes’ m6A levels and mRNA abundances of m6A regulators. Consensus clustering was employed to identify subtypes of patients based on the m6A, RNA, and protein levels of the top 20% of the genes by IQR ranking after filtering out genes with missing measurements (Fig. 3A). The optimal number of clusters was chosen based on elbow and silhouette methods. The classification of previously reported RNA-based LUAD subtypes, including C1–C4 subtypes (21) and bronchioid/magnoid/squamoid subtypes (22), was determined by comparing the similarities between RNA levels of the reported marker genes from each UHN-PLCME sample and the centroid of the maker genes for each subtype (21, 22). The similarity was measured using the Pearson correlation coefficient.

Epitranscriptome-, Transcriptome-, and Proteome-Wide Association Analysis

To assess the correlations with clinicopathologic features, we focused on the genes with quantitative data for its mRNA, protein, and m6A levels. To minimize the sample size effects, especially for proteomic data, we used a more stringent cutoff that requires the protein to be detected in at least 40 samples, resulting in a dataset of 1,011 genes. Then, using the set of clinical characteristics including sex, smoking status, and tumor stage, an ANOVA was conducted for each gene and each of the three types of omics data. For survival analysis, we further filtered our dataset to 635 genes with all three types of omics data in 45 patient samples with documented survival information. We classified the patients into two even groups by median dichotomizing the abundance of mRNA, protein, and m6A levels and conducted a log-rank test for each gene.

Statistic Methods

All statistical tests were performed using R. P <0.05 was considered significant. Data are presented as mean ± SD. The Shapiro–Wilk test and Levene’s test were used to assess the normality and equality of variances, respectively. The Student t test (for small sample size), Z-test (for large sample size), and Wilcoxon rank-sum test (also known as the Mann–Whitney U test) were used to assess the difference between two samples. The ANOVA test and the Kruskal–Wallis test were used for comparison of more than two groups. The Kolmogorov–Smirnov (K-S) test was applied to compare two distributions. Pearson’s χ2 test or Fisher’s exact test were used for enrichment analysis or independence analysis. Log-rank test was performed for survival analysis. The Benjamini–Hochberg correction was used to correct for multiple hypothesis testing.

Data Availability

m6A MeRIP-seq data is available upon request through EGA with accession number EGAS00001005524 (https://ega-archive.org/studies/EGAS00001005524). RNA-seq and RIP-seq data of NSCLC cells have been deposited in NCBI’s Gene Expression Omnibus (GEO) database under accession number GSE263727 (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?&acc=GSE263727). The code and processed files to reproduce all analyses and figures are available at https://github.com/HansenHeLab/Epitranscriptomic_m6A_Profiling_in_LUAD_Data_and_Codes.

No disclosures were reported.

H.H. He, M.S. Tsao, J. Yu, S. Wang, and Y. Zeng designed studies; S. Wang performed the majority of the experiments; Y. Zeng and S. Wang performed data analysis; L. Zhu, M. Zhang, L. Zhou, W. Luo, L. Wang, Y. Liu, H. Zhu, X. Xu, P. Su, X. Zhang, M. Ahmed, W. Chen, M. Chen, S. Chen, M. Slobodyanyuk, Z. Xie, J. Guan, P.C. Boutros, Z. Ke, and Z. Cai assisted experiments or analysis; W. Zhang, A.A. Khan, S. Sakashita, N.Liu, N.-A. Pham, M.F. Moran, W. Yang, C. Cheng, and M.S. Tsao generated patient cohort coupled with transcriptome and proteomic dataset; S. Wang, Y. Zeng, H. Zhu, and H.H. He wrote the first draft of the manuscript; all authors revised and approved manuscript.

This work was supported by the Princess Margaret Cancer Foundation (886012001223 to H.H. He), Canada Foundation for Innovation and Ontario Research Fund (CFI32372 to H.H. He), NSERC discovery grants (498706 to H.H. He; 06305 to M.F. Moran), Canadian Cancer Society innovation grants (703800 to H.H. He), IMPACT grant (701595 to M.S. Tsao), CIHR operating grants (142246, 152863, 152864, and 159567 to H.H. He; 364778 to M.F. Moran), Foundation grant (FDN-148395 to M.S. Tsao), and Terry Fox New Frontiers Program Project Grant (1090 and 1124 to H.H. He). H.H. He was supported by TFRI New Investigator Awards and CIHR New Investigator Awards. H.H. He holds an OMIR Early Researcher Award. Y. Zeng was supported by a postdoctoral fellowship by Princess Margaret Cancer Foundation. We thank Jonathan R. Krieger (SPARC BioCentre, Hospital for Sick Children, Toronto Canada) for help with MS data analysis.

NoteSupplementary data for this article are available at Cancer Discovery Online (http://cancerdiscovery.aacrjournals.org/).

1.
Bray
F
,
Ferlay
J
,
Soerjomataram
I
,
Siegel
RL
,
Torre
LA
,
Jemal
A
.
Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries
.
CA Cancer J Clin
2018
;
68
:
394
424
.
2.
Cancer Genome Atlas Research Network
.
Comprehensive molecular profiling of lung adenocarcinoma
.
Nature
2014
;
511
:
543
50
.
3.
Gillette
MA
,
Satpathy
S
,
Cao
S
,
Dhanasekaran
SM
,
Vasaikar
SV
,
Krug
K
, et al
.
Proteogenomic characterization reveals therapeutic vulnerabilities in lung adenocarcinoma
.
Cell
2020
;
182
:
200
25.e35
.
4.
Zhao
BS
,
Roundtree
IA
,
He
C
.
Post-transcriptional gene regulation by mRNA modifications
.
Nat Rev Mol Cell Biol
2017
;
18
:
31
42
.
5.
Shi
H
,
Wei
J
,
He
C
.
Where, when, and how: context-dependent functions of RNA methylation writers, readers, and erasers
.
Mol Cell
2019
;
74
:
640
50
.
6.
Zaccara
S
,
Ries
RJ
,
Jaffrey
SR
.
Reading, writing and erasing mRNA methylation
.
Nat Rev Mol Cell Biol
2019
;
20
:
608
24
.
7.
Huang
H
,
Weng
H
,
Chen
J
.
m6A modification in coding and non-coding RNAs: roles and therapeutic implications in cancer
.
Cancer Cell
2020
;
37
:
270
88
.
8.
Dominissini
D
,
Moshitch-Moshkovitz
S
,
Schwartz
S
,
Salmon-Divon
M
,
Ungar
L
,
Osenberg
S
, et al
.
Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq
.
Nature
2012
;
485
:
201
6
.
9.
Meyer
KD
,
Jaffrey
SR
.
Rethinking m6A readers, writers, and erasers
.
Annu Rev Cell Dev Biol
2017
;
33
:
319
42
.
10.
Deng
X
,
Qing
Y
,
Horne
D
,
Huang
H
,
Chen
J
.
The roles and implications of RNA m6A modification in cancer
.
Nat Rev Clin Oncol
2023
;
20
:
507
26
.
11.
Meyer
KD
,
Saletore
Y
,
Zumbo
P
,
Elemento
O
,
Mason
CE
,
Jaffrey
SR
.
Comprehensive analysis of mRNA methylation reveals enrichment in 3' UTRs and near stop codons
.
Cell
2012
;
149
:
1635
46
.
12.
Zeng
Y
,
Wang
S
,
Gao
S
,
Soares
F
,
Ahmed
M
,
Guo
H
, et al
.
Refined RIP-seq protocol for epitranscriptome analysis with low input materials
.
PLoS Biol
2018
;
16
:
e2006092
.
13.
John
T
,
Kohler
D
,
Pintilie
M
,
Yanagawa
N
,
Pham
N-A
,
Li
M
, et al
.
The ability to form primary tumor xenografts is predictive of increased risk of disease recurrence in early-stage non-small cell lung cancer
.
Clin Cancer Res
2011
;
17
:
134
41
.
14.
Wang
D
,
Pham
N-A
,
Tong
J
,
Sakashita
S
,
Allo
G
,
Kim
L
, et al
.
Molecular heterogeneity of non-small cell lung carcinoma patient-derived xenografts closely reflect their primary tumors
.
Int J Cancer
2017
;
140
:
662
73
.
15.
Wang
D
,
Pham
NA
,
Freeman
TM
,
Raghavan
V
,
Navab
R
,
Chang
J
, et al
.
Somatic alteration burden involving non-cancer genes predicts prognosis in early-stage non-small cell lung cancer
.
Cancers (Basel)
2019
;
11
:
1009
.
16.
Batista
PJ
,
Molinie
B
,
Wang
J
,
Qu
K
,
Zhang
J
,
Li
L
, et al
.
m(6)A RNA modification controls cell fate transition in mammalian embryonic stem cells
.
Cell Stem Cell
2014
;
15
:
707
19
.
17.
Molinie
B
,
Wang
J
,
Lim
KS
,
Hillebrand
R
,
Lu
Z-X
,
Van Wittenberghe
N
, et al
.
m(6)A-LAIC-seq reveals the census and complexity of the m(6)A epitranscriptome
.
Nat Methods
2016
;
13
:
692
8
.
18.
Liu
J
,
Li
K
,
Cai
J
,
Zhang
M
,
Zhang
X
,
Xiong
X
, et al
.
Landscape and regulation of mA and mAm methylome across human and mouse tissues
.
Mol Cell
2020
;
77
:
426
40.e6
.
19.
Schwartz
S
,
Mumbach
MR
,
Jovanovic
M
,
Wang
T
,
Maciag
K
,
Bushkin
GG
, et al
.
Perturbation of m6A writers reveals two distinct classes of mRNA methylation at internal and 5' sites
.
Cell Rep
2014
;
8
:
284
96
.
20.
Cassandri
M
,
Smirnov
A
,
Novelli
F
,
Pitolli
C
,
Agostini
M
,
Malewicz
M
, et al
.
Zinc-finger proteins in health and disease
.
Cell Death Discov
2017
;
3
:
17071
.
21.
Bhattacharjee
A
,
Richards
WG
,
Staunton
J
,
Li
C
,
Monti
S
,
Vasa
P
, et al
.
Classification of human lung carcinomas by mRNA expression profiling reveals distinct adenocarcinoma subclasses
.
Proc Natl Acad Sci U S A
2001
;
98
:
13790
5
.
22.
Wilkerson
MD
,
Yin
X
,
Walter
V
,
Zhao
N
,
Cabanski
CR
,
Hayward
MC
, et al
.
Differential pathogenesis of lung adenocarcinoma subtypes involving sequence mutations, copy number, chromosomal instability, and methylation
.
PLoS One
2012
;
7
:
e36530
.
23.
Xiao
Y
,
Wang
Y
,
Tang
Q
,
Wei
L
,
Zhang
X
,
Jia
G
.
An elongation- and ligation-based qPCR amplification method for the radiolabeling-free detection of locus-specific N6 -methyladenosine modification
.
Angew Chem Int Ed Engl
2018
;
57
:
15995
6000
.
24.
Der
SD
,
Sykes
J
,
Pintilie
M
,
Zhu
C-Q
,
Strumpf
D
,
Liu
N
, et al
.
Validation of a histology-independent prognostic gene signature for early-stage, non-small-cell lung cancer including stage IA patients
.
J Thorac Oncol
2014
;
9
:
59
64
.
25.
Xiao
D
,
He
J
.
Epithelial mesenchymal transition and lung cancer
.
J Thorac Dis
2010
;
2
:
154
9
.
26.
Heerboth
S
,
Housman
G
,
Leary
M
,
Longacre
M
,
Byler
S
,
Lapinska
K
, et al
.
EMT and tumor metastasis
.
Clin Transl Med
2015
;
4
:
6
.
27.
Terry
S
,
Savagner
P
,
Ortiz-Cuaran
S
,
Mahjoubi
L
,
Saintigny
P
,
Thiery
J-P
, et al
.
New insights into the role of EMT in tumor immune escape
.
Mol Oncol
2017
;
11
:
824
46
.
28.
Sistigu
A
,
Di Modugno
F
,
Manic
G
,
Nisticò
P
.
Deciphering the loop of epithelial-mesenchymal transition, inflammatory cytokines and cancer immunoediting
.
Cytokine Growth Factor Rev
2017
;
36
:
67
77
.
29.
Li
J
,
Chen
Z
,
Chen
F
,
Xie
G
,
Ling
Y
,
Peng
Y
, et al
.
Targeted mRNA demethylation using an engineered dCas13b-ALKBH5 fusion protein
.
Nucleic Acids Res
2020
;
48
:
5684
94
.
30.
Qiu
X
,
Zhu
X
,
Zhang
L
,
Mao
Y
,
Zhang
J
,
Hao
P
, et al
.
Human epithelial cancers secrete immunoglobulin g with unidentified specificity to promote growth and survival of tumor cells
.
Cancer Res
2003
;
63
:
6488
95
.
31.
Sabir
SR
,
Yeoh
S
,
Jackson
G
,
Bayliss
R
.
EML4-ALK variants: biological and molecular properties, and the implications for patients
.
Cancers (Basel)
2017
;
9
:
118
.
32.
Wilhelm
M
,
Schlegl
J
,
Hahne
H
,
Gholami
AM
,
Lieberenz
M
,
Savitski
MM
, et al
.
Mass-spectrometry-based draft of the human proteome
.
Nature
2014
;
509
:
582
7
.
33.
Pollmann
M
,
Parwaresch
R
,
Adam-Klages
S
,
Kruse
M-L
,
Buck
F
,
Heidebrecht
H-J
.
Human EML4, a novel member of the EMAP family, is essential for microtubule formation
.
Exp Cell Res
2006
;
312
:
3241
51
.
34.
Chen
D
,
Ito
S
,
Yuan
H
,
Hyodo
T
,
Kadomatsu
K
,
Hamaguchi
M
, et al
.
EML4 promotes the loading of NUDC to the spindle for mitotic progression
.
Cell Cycle
2015
;
14
:
1529
39
.
35.
O’Regan
L
,
Barone
G
,
Adib
R
,
Woo
CG
,
Jeong
HJ
,
Richardson
EL
, et al
.
EML4-ALK V3 oncogenic fusion proteins promote microtubule stabilization and accelerated migration through NEK9 and NEK7
.
J Cell Sci
2020
;
133
:
jcs241505
.
36.
Suo
H
,
Xiao
N
,
Wang
K
.
Potential roles of synaptotagmin family members in cancers: recent advances and prospects
.
Front Med
2022
;
9
:
968081
.
37.
Kang
J
,
Kim
W
,
Kwon
T
,
Youn
H
,
Kim
JS
,
Youn
B
.
Plasminogen activator inhibitor-1 enhances radioresistance and aggressiveness of non-small cell lung cancer cells
.
Oncotarget
2016
;
7
:
23961
74
.
38.
Morrow
GB
,
Mutch
NJ
.
Past, present, and future perspectives of plasminogen activator inhibitor 1 (PAI-1)
.
Semin Thromb Hemost
2023
;
49
:
305
13
.
39.
Molinie
N
,
Gautreau
A
.
The Arp2/3 regulatory system and its deregulation in cancer
.
Physiol Rev
2018
;
98
:
215
38
.
40.
Zheng
S
,
Qin
F
,
Yin
J
,
Li
D
,
Huang
Y
,
Hu
L
, et al
.
Role and mechanism of actin-related protein 2/3 complex signaling in cancer invasion and metastasis: a review
.
Medicine
2023
;
102
:
e33158
.
41.
Innocenti
M
.
New insights into the formation and the function of lamellipodia and ruffles in mesenchymal cell migration
.
Cell Adh Migr
2018
;
12
:
401
16
.
42.
Yankova
E
,
Blackaby
W
,
Albertella
M
,
Rak
J
,
De Braekeleer
E
,
Tsagkogeorga
G
, et al
.
Small-molecule inhibition of METTL3 as a strategy against myeloid leukaemia
.
Nature
2021
;
593
:
597
601
.
43.
Rosell
R
,
Karachaliou
N
.
Co-mutations in EGFR driven non-small cell lung cancer
.
EBioMedicine
2019
;
42
:
18
9
.
44.
Aguilo
F
,
Zhang
F
,
Sancho
A
,
Fidalgo
M
,
Di Cecilia
S
,
Vashisht
A
, et al
.
Coordination of m(6)A mRNA methylation and gene transcription by ZFP217 regulates pluripotency and reprogramming
.
Cell Stem Cell
2015
;
17
:
689
704
.
45.
Bertero
A
,
Brown
S
,
Madrigal
P
,
Osnato
A
,
Ortmann
D
,
Yiangou
L
, et al
.
The SMAD2/3 interactome reveals that TGFβ controls m6A mRNA methylation in pluripotency
.
Nature
2018
;
555
:
256
9
.
46.
Yang
X
,
Triboulet
R
,
Liu
Q
,
Sendinc
E
,
Gregory
RI
.
Exon junction complex shapes the m6A epitranscriptome
.
Nat Commun
2022
;
13
:
7904
.
47.
Panneerdoss
S
,
Eedunuri
VK
,
Yadav
P
,
Timilsina
S
,
Rajamanickam
S
,
Viswanadhapalli
S
, et al
.
Cross-talk among writers, readers, and erasers of m6A regulates cancer growth and progression
.
Sci Adv
2018
;
4
:
eaar8263
.
48.
Wei
J
,
Liu
F
,
Lu
Z
,
Fei
Q
,
Ai
Y
,
He
PC
, et al
.
Differential m6A, m6Am, and m1A demethylation mediated by FTO in the cell nucleus and cytoplasm
.
Mol Cell
2018
;
71
:
973
85.e5
.
49.
Wang
X
,
Lu
Z
,
Gomez
A
,
Hon
GC
,
Yue
Y
,
Han
D
, et al
.
N6-methyladenosine-dependent regulation of messenger RNA stability
.
Nature
2014
;
505
:
117
20
.
50.
Jamal-Hanjani
M
,
Wilson
GA
,
McGranahan
N
,
Birkbak
NJ
,
Watkins
TBK
,
Veeriah
S
, et al
.
Tracking the evolution of non-small-cell lung cancer
.
N Engl J Med
2017
;
376
:
2109
21
.
51.
Fiorentino
F
,
Menna
M
,
Rotili
D
,
Valente
S
,
Mai
A
.
METTL3 from target validation to the first small-molecule inhibitors: a medicinal chemistry journey
.
J Med Chem
2023
;
66
:
1654
77
.
52.
Wang
L
,
Hui
H
,
Agrawal
K
,
Kang
Y
,
Li
N
,
Tang
R
, et al
.
m6A RNA methyltransferases METTL3/14 regulate immune responses to anti-PD-1 therapy
.
EMBO J
2020
;
39
:
e104514
.
53.
Bueno-Costa
A
,
Piñeyro
D
,
García-Prieto
CA
,
Ortiz-Barahona
V
,
Martinez-Verbo
L
,
Webster
NA
, et al
.
Remodeling of the m6A RNA landscape in the conversion of acute lymphoblastic leukemia cells to macrophages
.
Leukemia
2022
;
36
:
2121
4
.
54.
Welch
DR
,
Hurst
DR
.
Defining the hallmarks of metastasis
.
Cancer Res
2019
;
79
:
3011
27
.
55.
Chen
X
,
Xu
H
,
Shu
X
,
Song
C-X
.
Mapping epigenetic modifications by sequencing technologies
.
Cell Death Differ
2023 Sep 1
[
Epub ahead of print
].
56.
Casado-Pelaez
M
,
Bueno-Costa
A
,
Esteller
M
.
Single cell cancer epigenetics
.
Trends Cancer
2022
;
8
:
820
38
.
57.
Tegowski
M
,
Flamand
MN
,
Meyer
KD
.
scDART-seq reveals distinct m6A signatures and mRNA methylation heterogeneity in single cells
.
Mol Cell
2022
;
82
:
868
78.e10
.
58.
Li
Y
,
Wang
Y
,
Vera-Rodriguez
M
,
Lindeman
LC
,
Skuggen
LE
,
Rasmussen
EMK
, et al
.
Single-cell m6A mapping in vivo using picoMeRIP-seq
.
Nat Biotechnol
2024
;
42
:
591
6
.
59.
Travis
WD
,
Brambilla
E
,
Nicholson
AG
,
Yatabe
Y
,
Austin
JHM
,
Beasley
MB
, et al
.
The 2015 world health organization classification of lung tumors: impact of genetic, clinical and radiologic advances since the 2004 classification
.
J Thorac Oncol
2015
;
10
:
1243
60
.
This open access article is distributed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International (CC BY-NC-ND 4.0) license.

Supplementary data