Skip to main content

The role of m6A methylation in therapy resistance in cancer

Abstract

Cancer therapy resistance is the main cause of cancer treatment failure. The mechanism of therapy resistance is a hot topic in epigenetics. As one of the most common RNA modifications, N6-methyladenosine (m6A) is involved in various processes of RNA metabolism, such as stability, splicing, transcription, translation, and degradation. A large number of studies have shown that m6A RNA methylation regulates the proliferation and invasion of cancer cells, but the role of m6A in cancer therapy resistance is unclear. In this review, we summarized the research progress related to the role of m6A in regulating therapy resistance in cancers.

Introduction

Cancer is a major public health problem in the world. According to the data from the National Cancer Center (China) published in March 2022, there were approximately 4 million new cancer cases and 2.4 million cancer-related deaths in China in 2016 [1]. In the United States, approximately 0.59 million people died of cancer in 2019, accounting for 21% of all deaths. Cancer has become the second leading cause of death in the United States after heart disease [2]. In addition to traditional treatment methods including surgery, radiotherapy and chemotherapy, the main emerging treatment methods for cancer include immunotherapy, targeted therapy, and gene therapy. Currently, comprehensive and individualized therapy is regarded as the most common and promising method for cancer treatment [3, 4].

The main reason for cancer treatment failure is that patients are resistant to treatments such as chemotherapy, which leads to tumor recurrence, progression, or distant metastasis. Chemotherapy drugs can effectively kill fast-growing tumor cells, which are widely used in the treatment of various tumors. However, there are still patients with a poor prognosis, and approximately 90% of cases of treatment failure are related to drug resistance [5]. Patients can present resistance to a specific drug or to multiple drugs with different structures and mechanisms, which is defined as multidrug resistance (MDR). Chemotherapy drug resistance is divided into intrinsic resistance and acquired resistance [6]. Intrinsic resistance refers to innate resistance that exists before patients are exposed to drugs. It is related to inherent genetic mutations in tumors. For example, patients suffering from gastric cancer with HER2 (human epidermal growth factor receptor 2) upregulation have an inferior response to cisplatin [7]. Acquired resistance refers to diminishing response to drugs after treatment, and secondary mutations in drug targets is one of the explanations. For instance, BCR-ABL fusion gene is a target for imatinib, a tyrosine kinase inhibitor, which is widely used in the treatment of chronic myeloid leukemia. If threonine 315 in its kinase domain is mutated, the binding ability of imatinib to BCR-ABL will decrease, thus significantly reducing drug efficacy. Approximately 20–30% of patients do not have a complete cytogenetic response following completion of imatinib [8]. In addition, cancer cell resistance to chemotherapy occurs at other levels, including increased drug efflux; decreased drug influx; cancer steam cells (CSCs); autophagy and so on(Fig. 1).

Fig. 1
figure 1

The mechanisms of chemical drugs resistance

Increased drug efflux and decreased drug influx; Secondary mutation in drug target; Cell cycle arrest; Gene mutation and DNA damage repair; Changes in metabolism and signaling pathway; Generation of cancer steam cells; Inducing autophagy

Apart from drug resistance, radiation resistance is another cause of treatment failure. Radiation resistance refers to the adaptation of tumor cells or tissues to radiotherapy. It is a complex process involving multiple genes and various mechanisms. The main reasons are as follows: DNA damage repair: DNA damage occurs after cells are irradiated, which is one of the reasons why radiation therapy is applied for tumor treatment. DNA damage repair is one of the causes of radiation resistance, which involves several pathways. For example, the PI3K (Phosphoinositide 3-kinase) signaling pathway is an important signaling pathway that facilitates DNA double-strand breaks (DSBs) repair by regulating nonhomologous end joining (NHEJ) and homologous recombination (HR) [9]. PI-103, the PI3K inhibitor, enhances radiation-induced cell death significantly [10]. Cell cycle arrest: After the occurrence of DNA damage, cell cycle arrest might occur, thus providing time for repair. Radiation induces G2/M phase arrest, accompanied by accumulation of a large number of cells in S phase, and the cells in S phase are resistant to radiation [11,12,13,14]. Generation of CSCs: CSCs are undifferentiated cancer cells with high tumorigenic activity, self-renewal ability and multidirectional differentiation potential. It has been suggested that this mechanism may be related to the vascular wall. In brain tumors, endothelial cells can secrete certain factors that keep CSCs in a self-renewing and undifferentiated state and protect them from radiation damage, thus promoting tumor radiation resistance [15]. Compared with tumors in wild-type mice, tumors in mice with altered endothelial cells are resistant to apoptosis induction by radiation over 20 Gy [16]. Changes in the tumor microenvironment (TME) and hypoxia. TME consists of a chemical microenvironment (characterized by pH, NO and metabolites, such as glucose) and a cellular microenvironment including tumor cells, stromal cells and extracellular matrix (ECM) [17]. Cancer-associated fibroblasts (CAFs) are dynamic components of the TME that affect the occurrence and development of tumors by producing ECM proteins, secreting growth factors and regulating the epithelial-mesenchymal transition [18]. Hypoxia promotes tumor progression by regulating CAFs function. Hypoxia can activate hypoxia-inducible factor-1 (HIF-1) and stimulate the expression of transforming growth factor-β (TGF-β), promoting fibroblast activation. Hypoxia also increases the expression of vascular endothelial growth factor in CAFs, which produces endothelial cells radioresistance and promotes the proliferation and regeneration of tumor vessels [19, 20]. Autophagy: Autophagy, which occurs in almost all eukaryotic cells, is a highly evolutionarily conserved physiological process. By degrading and making use of long-lived proteins and cytoplasmic organelles, autophagy plays an important role in maintaining intracellular metabolic homeostasis [21]. Under stresses such as hypoxia, inadequate growth factors, radiation or chemical drugs, tumor cells can escape from death via autophagy-mediated inhibition of the apoptotic pathway, thereby inducing the development of drug resistance [22]. Moreover, autophagy inhibits the production of reactive oxygen species to enhance radiation resistance of cells [23](Fig. 2).

Fig. 2
figure 2

The mechanisms of radiotherapy resistance

Nonhomologous end joining (NHEJ) and homologous recombination (HR); G2/M arrest and the accumulation of a large number of cells in S phase; Autophagy inhibits the production of reactive oxygen species; Changes in the tumor microenvironment and hypoxia; Generation of cancer stem cells

Cancer therapy resistance is a complex process, and understanding its molecular mechanisms can help us overcome it. Epigenetics plays an important role in cancer treatment resistance. Progress in research on DNA methylation, histone modification, chromatin remodeling and RNA modification has led to a better understanding of therapy resistance. For example, DNA demethylation in the promoter region of oncogenes increases their expression, leading to drug resistance. Thymosin β4 (Tβ4) is abnormally expressed after DNA demethylation and histone H3 modification in the promoter region. Overexpression of Tβ4 enables hepatoma carcinoma cells to acquire cancer stem cell-like abilities and causes resistance to sorafenib [24]. Although research on RNA modification began as early as the 1970s, it has been stalled by technical problems. In 2012, a novel approach combining RNA immunoprecipitation with next-generation sequencing was developed and allows further study of RNA modifications [25]. RNA methylation accounts for more than 60% of RNA modifications. The 5’ cap and 3’ poly-A modifications in eukaryotic cell mRNA play a key role in transcriptional regulation, while the internal modifications of mRNA usually maintain mRNA stability, such as N6-adenylate methylation and N1-adenylate methylation [26]. N6-methyladenosine (m6A) is one of the most common internal modifications in eukaryotic cells and has an important impact on mRNA splicing, transport, translation and other processes [27].

The m6A methylation is a dynamic reversible process regulated by three factors. Methyltransferases, called “writers”, include METTL3, METTL14, WTAP, RBM15, ZC3H13, and KIAA1429 (VIRMA). METTL3 is a core subunit with catalytic activity and METTL14 has a substrate recognition function. WTAP is responsible for recruiting METTL3 and METTL14 and combining with other components to form hybrids. “Erasers”, including FTO and ALKBH5, can demethylate RNA. FTO has the similar structure with ALKBH5 in core domain and is closely related to obesity and cancer. Methylated RNA is recognized by different enzymes to perform specific biological functions. The proteins that recognize methylation sites are called “readers”, that include YTHDC1, YTHDC2, YTHDF1, YTHDF2, and HNRNPC. Proteins with the YTH domain bind to methylated mRNA specifically and regulate downstream translation and degradation(Fig. 3).

Fig. 3
figure 3

The process of m6A methylation

The “writers” add methyl groups to RNA. METTL3 and METTL14 have a substrate recognition function. WTAP is responsible for recruiting METTL3, METTL14 and other components. The m6A modification is removed by “erasers” (FTO and ALKBH5). The “readers” recognize m6A and affect various functions of RNA, and they mainly include members of the YTH domain-containing family (YTHDC1/2 and YTHDF1/2) which can degrade mRNA.

A growing number of studies suggest that m6A methylation is a broad regulatory mechanism of gene expression. However, the role of m6A in the mechanism of cancer treatment resistance is unclear. Herein, we focus on the mechanisms of therapeutic resistance associated with m6A RNA methylation and potential approaches to overcome it.

Chemoresistance and m6A methylation

Cisplatin

The antibacterial activity of cisplatin was first discovered in 1965. Subsequently, it was observed to have a powerful antitumor effect, cross-linking with DNA to disrupt its normal function. In 1978, cisplatin was approved as an antitumor drug for clinical use, and approval of carboplatin and oxaliplatin were followed. Although the current theme of cancer treatment is precision targeting and immunotherapy, platinum-based drugs are still one of the most widely used antitumor drugs because of its significant and broad spectrum of antitumor activity [28]. Many studies have indicated that m6A is associated with the development of cisplatin resistance in tumors. HOXA10 is a member of the HOX gene family. The “erasers” ALKBH5 and HOXA10 form a positive feedback loop, which eliminates the m6A modification of JAK2 mRNA and promotes cisplatin resistance in epithelial ovarian cancer by activating the JAK2/STAT3 signaling pathway. Cell sensitivity to cisplatin was rescued by ALKBH5 and HOXA10 knockdown or inhibition of the JAK2/STAT3 signaling pathway [29]. Similarly, ALKBH5 reduces the stability of casein kinase (CK) 2α mRNA via demethylation, resulting in a decrease in CK2α protein expression. Casein kinase 2 (CK2) is a serine/threonine kinase that affects glycolysis in cancer cells. Research has reported that ALKBH5 was downregulated in bladder cancer cells. Knockdown of ALKBH5 promotes the proliferation and migration of bladder cancer cells, which may be achieved by increasing glucose utilization, lactate production and intracellular ATP levels in cancer cells. Moreover, ALKBH5 overexpression can enhance the sensitivity of bladder cancer cells to cisplatin through the CK2α-mediated m6A-dependent glycolysis pathway [30, 31]. Cisplatin resistance caused by ALKBH5 is not only associated with metabolic pathways but is also with CSCs. It has been found that the RNA helicase DDX3 mediated cisplatin resistance in oral squamous cell carcinoma (OSCC) by regulating the expression of the CSCs transcription factors FOXM1 and NANOG through ALKBH5. Pharmacological (ketorolac salt) inhibition of DDX3 restored cisplatin-mediated cell death and reduced CSCs population. A combination regimen of ketorolac salt with cisplatin may rescue OSCC resistance to chemotherapy [32]. It suggested that the development of cisplatin resistance is caused by the eraser-mediated demethylation process that destabilizes the mRNA of certain key enzymes. In fact, “writers” as well contribute to this process. One study showed that circ0008399 binding to WTAP promoted TNFα-induced protein 3 (TNFAIP3) expression by increasing its mRNA stability in an m6A-dependent manner, reducing the chemosensitivity of bladder cancer cells to cisplatin. The targeted regulation of the circ0008399/WTAP/TNFAIP3 axis could enhance cisplatin efficacy [33]. In nasal-type natural killer/T-cell lymphoma (NKTCL), WTAP stabilizes DUSP6 mRNA and promotes DUSP6 protein expression, which results in cisplatin resistance [34]. DUSP6, a dual-specificity phosphatase, is a negative regulator of the RAS-ERK signaling pathway and plays a key role in tumor progression, invasion, metastasis and chemoresistance [35]. In addition to WTAP, METTL3 is related to cisplatin resistance. Chen et al. found that METTL3 induced autophagy by regulating ATG5 expression, which enhanced seminoma resistance to cisplatin. The use of autophagy inhibitor 3-methyladenine (3-MA) and knockdown of ATG5 could overcome the cisplatin resistance in seminoma [36]. Wei et al. found that METTL3 rendered seminoma insensitive to cisplatin by stabilizing TFAP2C mRNA and activating the DNA repair-gene BRCA1 [37]. KIAA1429 (VIRMA) regulated the response of seminoma to cisplatin by disturbing DNA damage, and tumor sensitivity to cisplatin was increased after KIAA1429 knockdown [38]. In pancreatic cancer, METTL3 modulates the MAPK cascade to increase the resistance of pancreatic cancer cells to chemotherapy and radiotherapy [39]. Moreover, “readers” have also contributed to the development of cisplatin resistance. It has been reported that YTHDF1 promoted the production of GLS1, a key enzyme in glutamine metabolism, and regulated the glutamine metabolic pathway in colon cancer cells to make them resistant to cisplatin [40]. In ovarian cancer, YTHDF1 enhances the resistance of ovarian cancer cells to cisplatin, which may result from the maintenance of ovarian cancer CSCs through interaction with TRIM29. Hao et al. found that knockdown of YTHDF1 significantly reduced TRIM29 expression and suppressed stem cell-like features of ovarian cancer cells [41]. The CDKN1B gene encodes p27 protein, a cyclin-dependent kinase inhibitor, which arrests cell cycle in the G1 phase. One study showed that p27 can enhance the DNA damage response [42]. The cytotoxic effects of cisplatin are also dependent on DNA damage. YTHDF2 promotes the progression of intrahepatic cholangiocarcinoma and decreases its sensitivity to cisplatin by decreasing CDKN1B mRNA expression [43]. Wu et al. found that YTHDF2 knockdown inhibited the epithelial-mesenchymal transition process of cervical cancer cells and that YTHDF2 and AXIN1 contributed to cisplatin resistance in cervical cancer cells [44].

Anthracyclines

Since the 1970s, anthracyclines have been regarded as the most effective chemotherapy drugs for breast cancer and are widely used in neoadjuvant therapy or combination therapy [45]. However, drug resistance to anthracyclines frequently occurs, leading to the recurrence and metastasis of tumors. Research has shown that m6A methylation is associated with the development of chemoresistance in breast cancer. miR-221-3p is a miRNA involved in tumorigenesis, metastasis, and drug resistance. METTL3 methylates pri-miR-221-3p mRNA to increase miR-221-3p expression and negatively regulate HIPK2, a tumor suppressor that can be activated by doxorubicin. Therefore, METTL3 may make breast cancer cells resistant to doxorubicin through the miR-221-3p/HIPK2 axis. And miR-221-3p inhibition was confirmed to negate the METTL3-induced breast cancer cell resistance to doxorubicin [46]. The latest study showed that METTL3 promoted HR by regulating the EGF/RAD51 axis, leading to increased doxorubicin resistance in breast cancer cells. Moreover, knockdown of the reader protein YTHDC1 reversed the METTL3-mediated upregulation of epidermal growth factors (EGF) and DNA repair proteins (RAD51). YTHDC1 binds to m6A-modified EGF mRNA and promotes EGF synthesis, suggesting that METTL3 and YTHDC1 together enhance HR and cell survival during doxorubicin treatment, leading to the development of drug resistance in breast cancer [47]. Li et al. found another pathway by which METTL3 contributes to drug resistance in breast cancer. METTL3 increases MALAT1 protein levels in an m6A modification-dependent manner, recruits E2F1 and activates downstream AGR2 transcription to enhance the resistance of breast cancer cells to doxorubicin [48]. In addition, Wu et al. found that ALKBH5 removed the m6A modification of BRCA1 (DNA repair protein), stabilized BRCA1 mRNA and further enhanced DNA repair capacity, which results in reduced efficacy of doxorubicin in breast cancer. Additionally, protein arginine methyltransferase 5 (PRMT5) could enhance the nuclear translocation and translation of ALKBH5. Tadalafil was identified as a novel PRMT5 inhibitor that could increase doxorubicin sensitivity in breast cancer [49]. Wang et al. indicated that FTO activated signal transducer and activator of transcription 3 (STAT3) signaling in breast cancer cells, and both factors worked together to mediate breast cancer resistance to doxorubicin. The knockdown of FTO or STAT3 decreased doxorubicin resistance, which could be a potential strategy in the treatment of breast cancer [50].

5-fluorouracil (5-FU) and gemcitabine

5-Fluorouracil (5-FU) and the pyrimidine analog gemcitabine are anti-nucleotide metabolism drugs that are widely used in the treatment of various tumors, especially gastrointestinal tract cancers. Resistance to these drugs is common. There are few studies about the association between drug resistance and m6A methylation, and those that exist mainly focused on pancreatic cancer and colorectal cancer (CRC). The Wnt/β-catenin signaling pathway is genetically conserved and widely involved in embryonic development and tumor development. Tang et al. found that the eraser ALKBH5 removed the m6A modification in the 3’UTR of WIF-1 mRNA to promote its transcription and inhibited Wnt signaling by upregulating Wnt inhibitory factor 1 (WIF-1) instead of regulating β-catenin expression, which makes pancreatic ductal adenocarcinoma (PDAC) cells less sensitive to gemcitabine. ALKBH5 overexpression sensitized PDAC cells to chemotherapy [51]. In addition, Zhang et al. found that the writer METTL14 was associated with gemcitabine resistance in pancreatic cancer cells [52]. They reported that METTL14 was overexpressed in gemcitabine-resistant pancreatic cancer cells and p65 (a transcription factor) promoted the expression of METTL14 and subsequently upregulated cytidine deaminase (CDA), a gemcitabine inhibitor. Silencing METTL14 increased the sensitivity of pancreatic cancer cells to gemcitabine. Another study indicated that METTL3-mediated m6A methylation decreased the expression of lncRNA DBH-AS1 in pancreatic cancer, while DBH-AS1 increased the sensitivity of pancreatic cancer cells to gemcitabine through the miR-3163/USP44 axis. This suggested that METTL3 and DBH-AS1 may be involved in the development of gemcitabine resistance in pancreatic cancer. Targeted regulation of the DBH-AS1 could enhance the efficacy of gemcitabine [53]. ZC3H13 knockdown reduced the translation of PHF10 in a YTHDF1-mediated manner. Dysregulation of PHF10 increased the number of DSBs (DNA double-strand breaks) and inhibited HR. Targeted regulation of ZC3H13 and PHF10 by fisetin inhibited DNA damage repair and enhanced the sensitivity of pancreatic cancer cells to gemcitabine [54]. Liu et al. found that the expression of the oncogene Sect. 62 was regulated by METTL3. Section 62 can maintain the stemness of CRC cells and lead to 5-FU resistance by activating the Wnt/β-catenin pathway. Thus, m6A modification-Sect. 62-β-catenin molecular axis could act as therapeutic targets in improving treatment of CRC [55]. In addition, c-Myc promotes the expression of YTHDF1 in CRC, and knockdown of YTHDF1 enhances the sensitivity of CRC cells to 5-FU [56]. Another study indicated that METTL3-mediated m6A modification increased the expression of miR-181d-5p. This miRNA is contained in exosomes produced by CAFs and can decrease the sensitivity of CRC cells to 5-FU [57].

Kinase inhibitor (KI)

Targeted therapy, represented by kinase inhibitors, is an indispensable part of current cancer treatment. For example, tyrosine kinase inhibitors (TKIs), such as erlotinib and imatinib, have shown effects in the treatment of lung cancer and leukemia, but the problems of drug resistance and disease recurrence are inevitable. Understanding the mechanism of drug resistance and improving sensitivity to kinase inhibitors are currently a focus. Li et at. found that Notch signaling activation and TUSC7 inhibition occur in erlotinib-resistant lung adenocarcinoma cells [58]. The evolutionarily conserved Notch signaling pathway is related to the survival and proliferation of cancer stem cells [59]. TUSC7 is a long noncoding RNA tumor suppressor whose overexpression can inhibit the proliferation and invasion of tumor cells [60]. They showed that METTL3 persistently activated miR-146a/Notch signaling, while YTHDF2 inhibited TUSC7. Both effects promoted the formation of drug resistance in lung adenocarcinoma cells. The combinational use of TKIs and Notch specific inhibitory lncRNA may pave the way to improve targeted therapy in lung cancer. A previous study found that TKI resistance in leukemia was associated with reduced m6A due to FTO overexpression. In mice, cells with mRNA m6A hypomethylation and FTO upregulation had higher growth rates and were more tolerant to TKIs, while FTO inactivation re-sensitized resistant cells to TKIs [61]. Interestingly, saikosaponin-d (SsD), extracted from Radix Bupleuri, restored the sensitivity of resistant leukemia cells to nilotinib by reversing FTO-mediated m6A hypomethylation, which may provide a new approach for the treatment of acute myeloid leukemia [62]. Another study reported that downregulation of METTL3 and METTL14 decreased imatinib resistance in chronic myelogenous leukemia (CML), thus, inhibitors of METTL3/METTL14 complexes could be a new approach to rescue TKI resistance [63]. In malignant melanoma, METTL3 activates the RAF/MEK/ERK pathway by upregulating EGFR, inducing resistance to PLX4032, a BFAF (V600E) kinase inhibitor. Targeted treatment of METTL3 may improve the efficacy of PLX4032 in patients with advanced melanoma [64]. Chidamide is a histone deacetylase inhibitor (HDACI) that downregulates WTAP and METTL3 to induce c-MET mRNA hypomethylation, reducing the expression of c-MET, which helps to enhance the sensitivity to the ALK/ROS1/c-MET kinase inhibitor crizotinib in the treatment of non-small cell lung cancer (NSCLC) with high expression of c-MET [65]. Significantly, approximately 40% of lung cancer tissues overexpress the MET gene, whereas only 4–6% of lung adenocarcinoma patients have ALK mutations [66, 67]. Sorafenib is a multi-kinase inhibitor that inhibits various kinases, including VEGFR and BRAF. A study has shown that circRNA-SORE induced sorafenib resistance through competitive activation of the Wnt/β-catenin pathway, and the level of circRNA-SORE in sorafenib-resistant hepatocellular carcinoma (HCC) cells was regulated by m6A methylation. Silencing circRNA-SORE could effectively reverse the acquired sorafenib resistance and retard tumor progression [68]. Lin et al. found that METTL3 promoted FOXO3 mRNA stability in a YTHDF1-dependent manner. In sorafenib-resistant HCC cells, the downregulation of METTL3 led to FOXO3 degradation, indicating that sorafenib resistance is mediated by autophagy. Overexpression of FOXO3 rescued the m6A-depedent sorafenib sensitivity in HCC by inhibiting autophagy [69].

Immunotherapy

Because of the discovery of cancer immune checkpoints and the successful development of immune checkpoint inhibitors, immunotherapy has opened a new era of anticancer treatment that shows promising therapeutic prospects [70]. At present, there are two main types of immune checkpoint inhibitors in clinical application, PD-1/PD-L1 inhibitors and CTLA-4 inhibitors. Despite the unprecedented efficacy of immunotherapy, a large proportion of patients do not benefit from the treatment because of primary and acquired drug resistance. The mechanisms of resistance to immunotherapy are complex and diverse. Recently, the role of m6A modification has attracted attention. Most of the studies on m6A related to immune checkpoint inhibitor resistance have focused on PD-1/PD-L1 inhibitors. METTL3 promotes the back-splicing and circularization of circIGF2BP3 through YTHDC1. This m6A methylation event not only upregulates circIGF2BP3 but also increases the expression of the downstream target gene PKP3. The immunosuppressive effect of PKP3 depends on the deubiquitination of PD-L1 mediated by OTUB1 (a deubiquitinase), which protects PD-L1 from degradation. The accumulated PD-L1 on the surface of NSCLC cells interacts with PD-1 on T cells to decrease T-cell activity. Ultimately, the circIGF2BP3/PKP3/PD-L1 axis allows NSCLC cells to evade the killing effect of CD8+ T cells. The inhibition of circIGF2BP3/PKP3 enhanced the treatment efficacy of anti-PD-1 therapy in NSCLC [71]. In addition, inhibition of JNK signal transduction can downregulate METTL3, thereby affecting the stability of PD-L1 mRNA and decreasing its expression. Therefore, the JNK/METTL3/PD-L1 axis is critical for bladder cancer cells to resist death mediated by CD8+ T cells. Knockdown of JNK1 or administration of a JNK inhibitor maybe a potential strategy enhancing enhanced immune effect of in bladder cancer [72]. Wang et al. showed that deletion of METTL3 and METTL14 increased the stability of Stat1 mRNA and Irf1 mRNA in a YTHDF2-dependent manner and enhanced IFN-γ-Stat1-Irf1 signaling. It promoted the secretion of IFN-γ, Cxcl9 and Cxcl10 and the recruitment of CD8+ T cells in the tumor microenvironment. These cytokines and chemokines enhance the response of pMMR CRC and melanoma to anti-PD-1 therapy, providing a new approach for the combination of immune checkpoint inhibitors and methyltransferase inhibitors in the treatment of CRC and melanoma [73]. In addition to “writers”, “erasers” also affect immunotherapy sensitivity. Yang et al. found that m6A methylation reduced the proliferation and survival rates of melanoma cells, thus exerting a tumor-suppressive effect. FTO-mediated m6A demethylation promoted melanoma growth and reduced the response to anti-PD-1 immunotherapy. Mechanistically, FTO overexpression promoted melanoma development by increasing the key oncogenic genes (PD-1, CXCR4 and SOX10) in mRNA and protein levels and SOX10, and inhibiting IFN-γ-induced melanoma cell death. FTO knockdown increased the sensitivity of melanoma cells to anti-PD-1 therapy [74]. Unlike FTO, which is involved in the regulation of IFN-γ-mediated cytokine and chemokine pathways, ALKBH5 mainly affects the recruitment of immune cells in tumor microenvironment. Li et al. found that ALKBH5 changes the content of lactate in the tumor microenvironment by altering mRNA splicing and the expression of the target gene Mct4/Slc16a3, thereby affecting the recruitment of regulatory T cells (Tregs) and myeloid-derived suppressor cells (MDSCs). Loss of ALKBH5 can enhance the sensitivity of malignant melanoma to anti-PD-1 therapy [75].

Radiotherapy resistance

As previously mentioned, DSBs repair is one of the causes of radiation resistance. METTL3-mediated m6A modification plays a critical role in the maintenance of glioma stem-like cells (GSCs) and glioma cell dedifferentiation. Silencing METTL3 reduced DSBs repair and increased sensitivity to γ-radiation in GSCs [76]. Another study indicated that the overexpression of ALKBH5 in GSCs enhanced radio-resistance by regulating HR. Knockdown of ALKBH5 significantly reduced the expression of several key genes involved in HR, such as Rad51, XRCC2, BRCA2 and EXO1, while the expression of key genes involved in NHEJ (including Ku70, Ku80 and DNA-PKs) was not affected. This finding demonstrated that ALKBH5 inhibition could be a novel radiosensitizer [77]. In pancreatic cancer, METTL3 modulates the MAPK cascade to increase the resistance of pancreatic cancer cells to chemotherapy and radiotherapy, while knockdown of METTL3 enhances the radiosensitivity of pancreatic cancer cells [39]. In cervical cancer, FTO-mediated demethylation regulates β-catenin expression and promotes ERCC1 activity, which makes cervical squamous cell carcinoma resistant to chemoradiotherapy. ERCC1 (nucleotide excision repair cross-complementation group 1) is a key member of the nucleotide excision repair (NER) system, and its enhanced activity is one of the main reasons for platinum-based drug resistance [78]. In hypopharyngeal squamous cell carcinoma (HPSCC), METTL3 stabilizes the expression of circCUX1 via m6A methylation, while circCUX1 binds to Caspase-1 and inhibits its expression. Caspase-1, also known as interleukin-1β convertase, activates IL-1β and IL-18 and releases them into the extracellular environment to regulate the tumor microenvironment. When tumor cells exist in an inflammatory microenvironment, Caspase-1 actively induces the programmed death of tumor cells. METTL3 reduces tumor cell death through the circCUX1/caspase 1 axis and confers radiation resistance to HPSCC. Knockdown circCUX1 promotes the sensitivity of HPSCC cells to radiotherapy by increasing the release of inflammatory factors [79]. In nasopharyngeal carcinoma (NPC), YTHDC2 increased the translation efficiency and expression of IGF1R mRNA, thereby activating the downstream PI3K-AKT/S6 pathway, inhibiting tumor cell apoptosis, stimulating protein synthesis, and promoting radiotherapy resistance. Knockdown of YTHDC2 promotes the radiotherapy effect in NPC [80].

Causes of m6A methylation changes due to therapies

Why do these therapies cause alteration of m6A methylation? There are two possible reasons for this. On the one hand, drugs regulate the expression of some m6A genes, thereby affecting methylation levels (direct effect). Sunitinib (TKI) is widely used in the treatment of different types of solid and blood tumors. Ma et al. found that Sunitinib downregulated the expression of FTO, while upregulated the expressions of MELLT14. These changes increased the m6A methylation in vivo [81]. On the other hand, drugs act on a targeted gene directly, then regulate the downstream m6A gene to alter the m6A methylation levels (indirect effect). For instance, Wu et al. found that doxorubicin increased the expression of H2AX and activated m6A modification in breast cancer cells. Doxorubicin enhanced RNA m6A levels through DNA damage [49]. The reason for the changes of m6A methylation levels caused by various therapies is unclear. Further study is needed to understand the detailed mechanisms.

Role of m6A in cancer

The role of m6A in cancer is reflected in the change of several tumor-related genes. There are two main aspects—one involves the expression of oncogenes, while the other contains the regulation of tumor suppressor genes. Many studies have demonstrated that m6A is critical to tumor initiation, progression and metastasis. He et al. summarized these mechanisms in their review study [82]. On the one hand, m6A as a tumor promoter plays an important role in the development of cancer by promoting the expression of oncogenes and inhibiting the expression of tumor suppressor genes. On the other hand, the suppressing effect of m6A on tumor is reflected in the inhibition of oncogenes and the promotion of tumor suppressor genes. Here we elaborate on recent advances in research of the role of m6A in cancer. Du et al. showed that m6A modification of circ MDK improved its RNA stability, which resulting in the activation of PI3K/AKT/mTOR signaling pathway to promote HCC cell proliferation, migration and invasion [83]. Cui et al. reported that m6A demethylation of LINC00022 by FTO promoted tumor growth of esophageal squamous cell carcinoma [84]. Wang et al. found that YTHDF1 promoted CRC tumorigenesis and metastasis through upregulation of ARHGEF2 translation and protein expression. ARHGEF2 functions to activate RhoA signaling as an oncogene [85]. METTL14 promotes prostate tumorigenesis by inhibiting THBS1 (tumor suppressor gene) expression via an m6A-YTHDF2-dependent manner [86]. YTHDF2 promotes bladder cancer progression by suppressing RIG-I expression, a tumor suppressor related to immune response [87]. Moreover, METTL14 suppresses proliferation and metastasis of CRC by down-regulating oncogenic long non-coding RNA XIST [88]. Li et al. found that METTL3 increased the ZNF677 mRNA stability and promoted its expression. ZNF677 plays a tumor suppressor role in renal cell carcinoma (RCC) through transcriptionally repressing its downstream target CDKN3. METTL3/ZNF677/CDKN3 axis might provide new insight into the potential mechanism of the pathogenesis and development of RCC [89].

Methods of determining the m6A methylation

A growing number of studies have shown that m6A methylation serve as a potential biomarker for therapy resistant cancer cells [90,91,92]. Methods of determining the m6A methylation includes LC-MS/MS, Colorimetric and MeRIP-seq. LC-MS/MS can detect the overall m6A level of mRNA by the quantitative and qualitative analysis of the bases. Firstly, Total RNA is isolated from tissues by TRIzol reagent and mRNA and then purified using rRNA depletion kit. Secondly, mRNA is digested by nuclease P1, followed by the addition of NH4HCO3 and alkaline phosphatase with incubation. Lastly, the total amount of m6A in RNA is measured using LC-MS/MS [93]. Colorimetric is a method of measuring the total amount of m6A in RNA, like LC-MS/MS. Compared with LC-MS/MS, colorimetric is more sensitive and convenient by the use of m6A methylation quantification kit [84]. MeRIP-seq identified m6A methylation levels in human with a wide range and high throughput manner conveniently and economically. The m6A methylated mRNA fragments are enriched by immunomagnetic beads with m6A antibody. Then the methylated RNA is purified for further MeRIP sequencing [94].

Animal model in the m6A research

There are various animal models for m6A research in cancer drug resistance. CDX (Cell-line-derived xenograft) models transplant tumor cells cultured in vitro into mouse and have the advantages of convenience and cost effective. The CDX model mainly used BALB/C-nude, nu/nu, SCID, and NOD-scid mice. The patient-derived xenografts (PDX) models, using human cell lines injected into immunocompromised hosts such as athymic nude mice, are the most widely used models for evaluating cytotoxic therapies against cancer [95]. The specimens used for transplantation are directly derived from human tumor tissue without cultured in vitro, retaining the heterogeneity and growing environment of tumor. Moreover, the common cell lines for exploring the roles of m6A in cancer resistance includes the RCC cell lines (OS-RC-2 and 786–0) [96], NSCLC cell lines (A549 and H1299) [97], ESCC cell lines (Eca109, EC9706, Kyse150, Kyse410, and TE-1) [98], and melanoma cell lines (WM35 and WM115) [74].

Conclusion

In conclusion, tumor therapeutic resistance is a complex process involving multiple genes, factors and mechanisms that are related to oncogene activation, DNA damage repair, cancer stem cells, hypoxia, tumor microenvironment changes, autophagy, and metabolism. The m6A methylation affects the development of cancer therapy resistance in the above aspects by altering the stability of transcription products of certain key genes and activating or inhibiting certain signaling pathways (Fig. 4; Table 1). Therefore, upregulation or downregulation of certain m6A-related genes and activation or inhibition of certain m6A regulators can enhance the sensitivity of tumors to treatment. Understanding the key role of m6A modification in cancer therapy resistance provides new ideas for the development of drugs or combination therapies.

Fig. 4
figure 4

The role of m6A methylation in therapy resistance

The m6A regulators promote tumor cell resistance to radiation via stabilizing mRNA, regulating HR and inhibiting signaling pathway. The m6A regulators promote the stability of target gene mRNA and affect the recruitment of immune cell and the secretion of cytokines. The various pathway of m6A regulators in the resistance to chemical drugs (Cisplatin, Gemcitabine, Kinase inhibitor). The red arrows indicate “promote” and the blue arrows indicate “inhibit”

Table 1 The functions of m6A regulators in therapy resistance

Data availability

Not applicable.

Abbreviations

5-FU:

5-Fluorouracil

ALKBH5:

AlkB homologue 5

ATG5:

Autophagy related 5

BRCA1:

Breast Cancer 1

CAFs:

Cancer-associated fibroblasts

CDA:

Cytidine deaminase

CDKN1B:

Cyclin-dependent kinase N1B

CK 2:

Casein kinase 2

CML:

Chronic myelogenous leukemia

CSCs:

Cancer stem cells

DSBs:

DNA double-strand breaks

DUSP6:

Dual-specificity phosphatase 6

ECM:

Extracellular matrix

EGF:

Epidermal growth factor

ERCC1:

Excision repair cross-complementing group 1

FOXM1:

Forkhead box protein M1

FTO:

Fat mass and obesity-related protein

GBM:

Glioblastoma

GLS1:

Glutaminase 1

GSCs:

Glioma stem-like cells

HCC:

Hepatocellular carcinoma

HDACI:

Histone deacetylase inhibitor

HER2:

Human epidermal growth factor receptor 2

HIF-1:

Hypoxia-inducible factor-1

HIPK2:

Homeodomain-Interacting Protein Kinase 2

HNRNPC:

Heterogeneous nuclear ribonucleoprotein C

HOXA10:

Homeobox A10

HPSCC:

Hypopharyngeal squamous cell carcinoma

HR:

Homologous recombination

KIAA1429(VIRMA):

Vir-like m6A methyltransferase-associated

KI:

Kinase inhibitors

m6A:

N6-methyladenosine

MALAT1:

Metastasis Associated Lung Adenocarcinoma Transcript 1

MDR:

Multidrug resistance

MDSCs:

Myeloid-derived suppressor cells

METTL14:

Methyltransferase-like 14

METTL3:

Methyltransferase-like 3

NER:

Nucleotide excision repair

NHEJ:

Nonhomologous end joining

NKTCL:

Natural killer/T-cell lymphoma

NPC:

Nasopharyngeal carcinoma

NSCLC:

Non-small cell lung cancer

OTUB1:

OTU domain, ubiquitin aldehyde binding 1

PD-1:

Programmed cell death 1

PD-L1:

Programmed cell death 1 ligand 1

PI3K:

Phosphoinositide 3-kinase

RBM15:

RNA binding motif protein 15

SsD:

Saikosaponin-d

STAT3:

Signal transducer and activator of transcription 3

TFAP2C:

Transcription factor AP-2 γ

TGF-β:

Transforming growth factor-β

TKIs:

Tyrosine kinase inhibitors

TME:

Tumor microenvironment

TNFAIP3:

TNFα-induced protein 3

Tregs:

Regulatory T cells

TRIM29:

Tripartite motif 29

TUSC7:

Tumor suppressor candidate 7

Tβ4:

Thymosin β4

VEGFR:

Vascular endothelial growth factor receptor

WIF-1:

Wnt inhibitory factor 1

WTAP:

Wilms tumor 1 associated protein

YTHDC1:

YTH domain containing 1

YTHDC2:

YTH domain-containing 2

YTHDF1:

YTH domain-containing family 1

YTHDF2:

YTH domain-containing family 2

ZC3H13:

Zinc finger CCCH-type containing 13

References

  1. Zheng R, Zhang S, Zeng H, Wang S, Sun K, Chen R, et al. Cancer incidence and mortality in China, 2016. J Natl Cancer Cent. 2022;2(1):1–9. https://doi.org/10.1016/j.jncc.2022.02.002.

    Article  Google Scholar 

  2. Siegel RL, Miller KD, Fuchs HE, Jemal A. Cancer statistics, 2022. CA Cancer J Clin. 2022;72(1):7–33. https://doi.org/10.3322/caac.21708.

    Article  PubMed  Google Scholar 

  3. Urruticoechea A, Alemany R, Balart J, Villanueva A, Viñals F, Capellá G. Recent advances in cancer therapy: an overview. Curr Pharm Des. 2010;16(1):3–10. https://doi.org/10.2174/138161210789941847.

    Article  CAS  PubMed  Google Scholar 

  4. Baskar R, Lee KA, Yeo R, Yeoh KW. Cancer and radiation therapy: current advances and future directions. Int J Med Sci. 2012;9(3):193–9. https://doi.org/10.7150/ijms.3635.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Mansoori B, Mohammadi A, Davudian S, Shirjang S, Baradaran B. The different mechanisms of Cancer Drug Resistance: a brief review. Adv Pharm Bull. 2017;7(3):339–48. https://doi.org/10.15171/apb.2017.041.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lippert TH, Ruoff HJ, Volm M. Intrinsic and acquired drug resistance in malignant tumors. The main reason for therapeutic failure. Arzneimittelforschung. 2008;58(6):261–4. https://doi.org/10.1055/s-0031-1296504.

    Article  CAS  PubMed  Google Scholar 

  7. Huang D, Duan H, Huang H, Tong X, Han Y, Ru G, et al. Cisplatin resistance in gastric cancer cells is associated with HER2 upregulation-induced epithelial-mesenchymal transition. Sci Rep. 2016;6:20502. https://doi.org/10.1038/srep20502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Quintás-Cardama A, Kantarjian HM, Cortes JE. Mechanisms of primary and secondary resistance to imatinib in chronic myeloid leukemia. Cancer Control. 2009;16(2):122–31. https://doi.org/10.1177/107327480901600204.

    Article  PubMed  Google Scholar 

  9. Kumar A, Fernandez-Capetillo O, Carrera AC. Nuclear phosphoinositide 3-kinase beta controls double-strand break DNA repair. Proc Natl Acad Sci U S A. 2010;107(16):7491–6. https://doi.org/10.1073/pnas.0914242107.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Jang NY, Kim DH, Cho BJ, Choi EJ, Lee JS, Wu HG, et al. Radiosensitization with combined use of olaparib and PI-103 in triple-negative breast cancer. BMC Cancer. 2015;15:89. https://doi.org/10.1186/s12885-015-1090-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Kastan MB, Bartek J. Cell-cycle checkpoints and cancer. Nature. 2004;432(7015):316–23. https://doi.org/10.1038/nature03097.

    Article  CAS  PubMed  Google Scholar 

  12. Landsverk KS, Patzke S, Rein ID, Stokke C, Lyng H, De Angelis PM, et al. Three independent mechanisms for arrest in G2 after ionizing radiation. Cell Cycle. 2011;10(5):819–29. https://doi.org/10.4161/cc.10.5.14968.

    Article  CAS  PubMed  Google Scholar 

  13. Vlatkovic T, Veldwijk MR, Giordano FA, Herskind C. Targeting cell cycle checkpoint kinases to overcome intrinsic Radioresistance in Brain Tumor cells. Cancers (Basel). 2022;14(3). https://doi.org/10.3390/cancers14030701.

  14. Deckbar D, Jeggo PA, Löbrich M. Understanding the limitations of radiation-induced cell cycle checkpoints. Crit Rev Biochem Mol Biol. 2011;46(4):271–83. https://doi.org/10.3109/10409238.2011.575764.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Calabrese C, Poppleton H, Kocak M, Hogg TL, Fuller C, Hamner B, et al. A perivascular niche for brain tumor stem cells. Cancer Cell. 2007;11(1):69–82. https://doi.org/10.1016/j.ccr.2006.11.020.

    Article  CAS  PubMed  Google Scholar 

  16. Garcia-Barros M, Paris F, Cordon-Cardo C, Lyden D, Rafii S, Haimovitz-Friedman A, et al. Tumor response to radiotherapy regulated by endothelial cell apoptosis. Science. 2003;300(5622):1155–9. https://doi.org/10.1126/science.1082504.

    Article  CAS  PubMed  Google Scholar 

  17. Roy S, Kumaravel S, Sharma A, Duran CL, Bayless KJ, Chakraborty S. Hypoxic tumor microenvironment: implications for cancer therapy. Exp Biol Med (Maywood). 2020;245(13):1073–86. https://doi.org/10.1177/1535370220934038.

    Article  CAS  PubMed  Google Scholar 

  18. LeBleu VS, Kalluri R. A peek into cancer-associated fibroblasts: origins, functions and translational impact. Dis Model Mech. 2018;11(4). https://doi.org/10.1242/dmm.029447.

  19. Ammirante M, Shalapour S, Kang Y, Jamieson CA, Karin M. Tissue injury and hypoxia promote malignant progression of prostate cancer by inducing CXCL13 expression in tumor myofibroblasts. Proc Natl Acad Sci U S A. 2014;111(41):14776–81. https://doi.org/10.1073/pnas.1416498111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Harada H, Kizaka-Kondoh S, Li G, Itasaka S, Shibuya K, Inoue M, et al. Significance of HIF-1-active cells in angiogenesis and radioresistance. Oncogene. 2007;26(54):7508–16. https://doi.org/10.1038/sj.onc.1210556.

    Article  CAS  PubMed  Google Scholar 

  21. Levine B, Klionsky DJ. Development by self-digestion: molecular mechanisms and biological functions of autophagy. Dev Cell. 2004;6(4):463–77. https://doi.org/10.1016/s1534-5807(04)00099-1.

    Article  CAS  PubMed  Google Scholar 

  22. Kondo Y, Kanzawa T, Sawaya R, Kondo S. The role of autophagy in cancer development and response to therapy. Nat Rev Cancer. 2005;5(9):726–34. https://doi.org/10.1038/nrc1692.

    Article  CAS  PubMed  Google Scholar 

  23. Chen X, Wang P, Guo F, Wang X, Wang J, Xu J, et al. Autophagy enhanced the radioresistance of non-small cell lung cancer by regulating ROS level under hypoxia condition. Int J Radiat Biol. 2017;93(8):764–70. https://doi.org/10.1080/09553002.2017.1325025.

    Article  CAS  PubMed  Google Scholar 

  24. Ohata Y, Shimada S, Akiyama Y, Mogushi K, Nakao K, Matsumura S, et al. Acquired Resistance with epigenetic alterations under long-term antiangiogenic therapy for Hepatocellular Carcinoma. Mol Cancer Ther. 2017;16(6):1155–65. https://doi.org/10.1158/1535-7163.Mct-16-0728.

    Article  CAS  PubMed  Google Scholar 

  25. 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(7):1635–46. https://doi.org/10.1016/j.cell.2012.05.003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Yang B, Wang JQ, Tan Y, Yuan R, Chen ZS, Zou C. RNA methylation and cancer treatment. Pharmacol Res. 2021;174:105937. https://doi.org/10.1016/j.phrs.2021.105937.

    Article  CAS  PubMed  Google Scholar 

  27. Wang T, Kong S, Tao M, Ju S. The potential role of RNA N6-methyladenosine in Cancer progression. Mol Cancer. 2020;19(1):88. https://doi.org/10.1186/s12943-020-01204-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Rottenberg S, Disler C, Perego P. The rediscovery of platinum-based cancer therapy. Nat Rev Cancer. 2021;21(1):37–50. https://doi.org/10.1038/s41568-020-00308-y.

    Article  CAS  PubMed  Google Scholar 

  29. Nie S, Zhang L, Liu J, Wan Y, Jiang Y, Yang J, et al. ALKBH5-HOXA10 loop-mediated JAK2 m6A demethylation and cisplatin resistance in epithelial ovarian cancer. J Exp Clin Cancer Res. 2021;40(1):284. https://doi.org/10.1186/s13046-021-02088-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Yu H, Yang X, Tang J, Si S, Zhou Z, Lu J, et al. ALKBH5 inhibited cell proliferation and sensitized bladder Cancer cells to cisplatin by m6A-CK2α-Mediated glycolysis. Mol Ther Nucleic Acids. 2021;23:27–41. https://doi.org/10.1016/j.omtn.2020.10.031.

    Article  CAS  PubMed  Google Scholar 

  31. Zhang X, Yang X, Yang C, Li P, Yuan W, Deng X, et al. Targeting protein kinase CK2 suppresses bladder cancer cell survival via the glucose metabolic pathway. Oncotarget. 2016;7(52):87361–72. https://doi.org/10.18632/oncotarget.13571.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Shriwas O, Priyadarshini M, Samal SK, Rath R, Panda S, Das Majumdar SK, et al. DDX3 modulates cisplatin resistance in OSCC through ALKBH5-mediated m(6)A-demethylation of FOXM1 and NANOG. Apoptosis. 2020;25(3–4):233–46. https://doi.org/10.1007/s10495-020-01591-8.

    Article  CAS  PubMed  Google Scholar 

  33. Wei W, Sun J, Zhang H, Xiao X, Huang C, Wang L, et al. Circ0008399 Interaction with WTAP promotes Assembly and Activity of the m(6)a methyltransferase complex and promotes cisplatin resistance in bladder Cancer. Cancer Res. 2021;81(24):6142–56. https://doi.org/10.1158/0008-5472.Can-21-1518.

    Article  CAS  PubMed  Google Scholar 

  34. Ma H, Shen L, Yang H, Gong H, Du X, Li J. m6A methyltransferase Wilms’ tumor 1-associated protein facilitates cell proliferation and cisplatin resistance in NK/T cell lymphoma by regulating dual-specificity phosphatases 6 expression via m6A RNA methylation. IUBMB Life. 2021;73(1):108–17. https://doi.org/10.1002/iub.2410.

    Article  CAS  PubMed  Google Scholar 

  35. Gao Y, Li H, Han Q, Li Y, Wang T, Huang C, et al. Overexpression of DUSP6 enhances chemotherapy-resistance of ovarian epithelial cancer by regulating the ERK signaling pathway. J Cancer. 2020;11(11):3151–64. https://doi.org/10.7150/jca.37267.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Chen H, Xiang Y, Yin Y, Peng J, Peng D, Li D, et al. The m6A methyltransferase METTL3 regulates autophagy and sensitivity to cisplatin by targeting ATG5 in seminoma. Transl Androl Urol. 2021;10(4):1711–22. https://doi.org/10.21037/tau-20-1411.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Wei J, Yin Y, Zhou J, Chen H, Peng J, Yang J, et al. METTL3 potentiates resistance to cisplatin through m(6) a modification of TFAP2C in seminoma. J Cell Mol Med. 2020;24(19):11366–80. https://doi.org/10.1111/jcmm.15738.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Miranda-Gonçalves V, Lobo J, Guimarães-Teixeira C, Barros-Silva D, Guimarães R, Cantante M, et al. The component of the m(6)a writer complex VIRMA is implicated in aggressive tumor phenotype, DNA damage response and cisplatin resistance in germ cell tumors. J Exp Clin Cancer Res. 2021;40(1):268. https://doi.org/10.1186/s13046-021-02072-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Taketo K, Konno M, Asai A, Koseki J, Toratani M, Satoh T, et al. The epitranscriptome m6A writer METTL3 promotes chemo- and radioresistance in pancreatic cancer cells. Int J Oncol. 2018;52(2):621–9. https://doi.org/10.3892/ijo.2017.4219.

    Article  CAS  PubMed  Google Scholar 

  40. Chen P, Liu XQ, Lin X, Gao LY, Zhang S, Huang X. Targeting YTHDF1 effectively re-sensitizes cisplatin-resistant colon cancer cells by modulating GLS-mediated glutamine metabolism. Mol Ther Oncolytics. 2021;20:228–39. https://doi.org/10.1016/j.omto.2021.01.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Hao L, Wang JM, Liu BQ, Yan J, Li C, Jiang JY, et al. m6A-YTHDF1-mediated TRIM29 upregulation facilitates the stem cell-like phenotype of cisplatin-resistant ovarian cancer cells. Biochim Biophys Acta Mol Cell Res. 2021;1868(1):118878. https://doi.org/10.1016/j.bbamcr.2020.118878.

    Article  CAS  PubMed  Google Scholar 

  42. Iadevaia V, Wouters MD, Kanitz A, Matia-González AM, Laing EE, Gerber AP. Tandem RNA isolation reveals functional rearrangement of RNA-binding proteins on CDKN1B/p27(Kip1) 3’UTRs in cisplatin treated cells. RNA Biol. 2020;17(1):33–46. https://doi.org/10.1080/15476286.2019.1662268.

    Article  CAS  PubMed  Google Scholar 

  43. Huang CS, Zhu YQ, Xu QC, Chen S, Huang Y, Zhao G, et al. YTHDF2 promotes intrahepatic cholangiocarcinoma progression and desensitises cisplatin treatment by increasing CDKN1B mRNA degradation. Clin Transl Med. 2022;12(6):e848. https://doi.org/10.1002/ctm2.848.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Wu M, Chen G, Liao X, Xiao L, Zheng J. YTHDF2 interference suppresses the EMT of cervical cancer cells and enhances cisplatin chemosensitivity by regulating AXIN1. Drug Dev Res. 2022. https://doi.org/10.1002/ddr.21942.

    Article  PubMed  Google Scholar 

  45. Fisusi FA, Akala EO. Drug combinations in breast Cancer therapy. Pharm Nanotechnol. 2019;7(1):3–23. https://doi.org/10.2174/2211738507666190122111224.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Pan X, Hong X, Li S, Meng P, Xiao F. METTL3 promotes adriamycin resistance in MCF-7 breast cancer cells by accelerating pri-microRNA-221-3p maturation in a m6A-dependent manner. Exp Mol Med. 2021;53(1):91–102. https://doi.org/10.1038/s12276-020-00510-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Li E, Xia M, Du Y, Long K, Ji F, Pan F, et al. METTL3 promotes homologous recombination repair and modulates chemotherapeutic response in breast cancer by regulating the EGF/RAD51 axis. Elife. 2022;11. https://doi.org/10.7554/eLife.75231.

  48. Li S, Jiang F, Chen F, Deng Y, Pan X. Effect of m6A methyltransferase METTL3 -mediated MALAT1/E2F1/AGR2 axis on adriamycin resistance in breast cancer. J Biochem Mol Toxicol. 2022;36(1):e22922. https://doi.org/10.1002/jbt.22922.

    Article  CAS  PubMed  Google Scholar 

  49. Wu Y, Wang Z, Han L, Guo Z, Yan B, Guo L, et al. PRMT5 regulates RNA m6A demethylation for doxorubicin sensitivity in breast cancer. Mol Ther. 2022. https://doi.org/10.1016/j.ymthe.2022.03.003.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Wang Y, Cheng Z, Xu J, Lai M, Liu L, Zuo M, et al. Fat mass and obesity-associated protein (FTO) mediates signal transducer and activator of transcription 3 (STAT3)-drived resistance of breast cancer to doxorubicin. Bioengineered. 2021;12(1):1874–89. https://doi.org/10.1080/21655979.2021.1924544.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Tang B, Yang Y, Kang M, Wang Y, Wang Y, Bi Y, et al. M(6)a demethylase ALKBH5 inhibits pancreatic cancer tumorigenesis by decreasing WIF-1 RNA methylation and mediating wnt signaling. Mol Cancer. 2020;19(1):3. https://doi.org/10.1186/s12943-019-1128-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zhang C, Ou S, Zhou Y, Liu P, Zhang P, Li Z, et al. M(6)a methyltransferase METTL14-Mediated Upregulation of Cytidine Deaminase promoting Gemcitabine Resistance in Pancreatic Cancer. Front Oncol. 2021;11:696371. https://doi.org/10.3389/fonc.2021.696371.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Ye X, Wang LP, Han C, Hu H, Ni CM, Qiao GL, et al. Increased m(6)a modification of lncRNA DBH-AS1 suppresses pancreatic cancer growth and gemcitabine resistance via the miR-3163/USP44 axis. Ann Transl Med. 2022;10(6):304. https://doi.org/10.21037/atm-22-556.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Huang C, Zhou S, Zhang C, Jin Y, Xu G, Zhou L, et al. ZC3H13-mediated N6-methyladenosine modification of PHF10 is impaired by fisetin which inhibits the DNA damage response in pancreatic cancer. Cancer Lett. 2022;530:16–28. https://doi.org/10.1016/j.canlet.2022.01.013.

    Article  CAS  PubMed  Google Scholar 

  55. Liu X, Su K, Sun X, Jiang Y, Wang L, Hu C, et al. Section 62 promotes stemness and chemoresistance of human colorectal cancer through activating Wnt/β-catenin pathway. J Exp Clin Cancer Res. 2021;40(1):132. https://doi.org/10.1186/s13046-021-01934-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Nishizawa Y, Konno M, Asai A, Koseki J, Kawamoto K, Miyoshi N, et al. Oncogene c-Myc promotes epitranscriptome m(6)a reader YTHDF1 expression in colorectal cancer. Oncotarget. 2018;9(7):7476–86. https://doi.org/10.18632/oncotarget.23554.

    Article  PubMed  Google Scholar 

  57. Pan S, Deng Y, Fu J, Zhang Y, Zhang Z, Qin X. N6–methyladenosine upregulates miR–181d–5p in exosomes derived from cancer–associated fibroblasts to inhibit 5–FU sensitivity by targeting NCALD in colorectal cancer. Int J Oncol. 2022;60(2). https://doi.org/10.3892/ijo.2022.5304.

  58. Li K, Peng ZY, Gao S, Wang QS, Wang R, Li X, et al. M6A associated TSUC7 inhibition contributed to Erlotinib resistance in lung adenocarcinoma through a notch signaling activation dependent way. J Exp Clin Cancer Res. 2021;40(1):325. https://doi.org/10.1186/s13046-021-02137-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. BeLow M, Osipo C. Notch signaling in breast Cancer: a role in Drug Resistance. Cells. 2020;9(10). https://doi.org/10.3390/cells9102204.

  60. Li N, Shi K, Li W. TUSC7: a novel tumor suppressor long non-coding RNA in human cancers. J Cell Physiol. 2018;233(9):6401–7. https://doi.org/10.1002/jcp.26544.

    Article  CAS  PubMed  Google Scholar 

  61. Yan F, Al-Kali A, Zhang Z, Liu J, Pang J, Zhao N, et al. A dynamic N(6)-methyladenosine methylome regulates intrinsic and acquired resistance to tyrosine kinase inhibitors. Cell Res. 2018;28(11):1062–76. https://doi.org/10.1038/s41422-018-0097-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Sun K, Du Y, Hou Y, Zhao M, Li J, Du Y, et al. Saikosaponin D exhibits anti-leukemic activity by targeting FTO/m(6)a signaling. Theranostics. 2021;11(12):5831–46. https://doi.org/10.7150/thno.55574.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Ianniello Z, Sorci M, Ceci Ginistrelli L, Iaiza A, Marchioni M, Tito C, et al. New insight into the catalytic -dependent and -independent roles of METTL3 in sustaining aberrant translation in chronic myeloid leukemia. Cell Death Dis. 2021;12(10):870. https://doi.org/10.1038/s41419-021-04169-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Bhattarai PY, Kim G, Poudel M, Lim SC, Choi HS. METTL3 induces PLX4032 resistance in melanoma by promoting m(6)A-dependent EGFR translation. Cancer Lett. 2021;522:44–56. https://doi.org/10.1016/j.canlet.2021.09.015.

    Article  CAS  PubMed  Google Scholar 

  65. Ding N, You A, Tian W, Gu L, Deng D. Chidamide increases the sensitivity of non-small cell Lung Cancer to Crizotinib by decreasing c-MET mRNA methylation. Int J Biol Sci. 2020;16(14):2595–611. https://doi.org/10.7150/ijbs.45886.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Takeuchi K, Soda M, Togashi Y, Suzuki R, Sakata S, Hatano S, et al. RET, ROS1 and ALK fusions in lung cancer. Nat Med. 2012;18(3):378–81. https://doi.org/10.1038/nm.2658.

    Article  CAS  PubMed  Google Scholar 

  67. Ma PC, Tretiakova MS, MacKinnon AC, Ramnath N, Johnson C, Dietrich S, et al. Expression and mutational analysis of MET in human solid cancers. Genes Chromosomes Cancer. 2008;47(12):1025–37. https://doi.org/10.1002/gcc.20604.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Xu J, Wan Z, Tang M, Lin Z, Jiang S, Ji L, et al. N(6)-methyladenosine-modified CircRNA-SORE sustains sorafenib resistance in hepatocellular carcinoma by regulating β-catenin signaling. Mol Cancer. 2020;19(1):163. https://doi.org/10.1186/s12943-020-01281-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Lin Z, Niu Y, Wan A, Chen D, Liang H, Chen X, et al. RNA m(6) a methylation regulates sorafenib resistance in liver cancer through FOXO3-mediated autophagy. Embo j. 2020;39(12):e103181. https://doi.org/10.15252/embj.2019103181.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Sharma P, Hu-Lieskovan S, Wargo JA, Ribas A. Primary, adaptive, and Acquired Resistance to Cancer Immunotherapy. Cell. 2017;168(4):707–23. https://doi.org/10.1016/j.cell.2017.01.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Liu Z, Wang T, She Y, Wu K, Gu S, Li L, et al. N(6)-methyladenosine-modified circIGF2BP3 inhibits CD8(+) T-cell responses to facilitate tumor immune evasion by promoting the deubiquitination of PD-L1 in non-small cell lung cancer. Mol Cancer. 2021;20(1):105. https://doi.org/10.1186/s12943-021-01398-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ni Z, Sun P, Zheng J, Wu M, Yang C, Cheng M, et al. JNK Signaling promotes bladder Cancer Immune escape by regulating METTL3-Mediated m6A modification of PD-L1 mRNA. Cancer Res. 2022;82(9):1789–802. https://doi.org/10.1158/0008-5472.Can-21-1323.

    Article  CAS  PubMed  Google Scholar 

  73. Wang L, Hui H, Agrawal K, Kang Y, Li N, Tang R, et al. M(6) a RNA methyltransferases METTL3/14 regulate immune responses to anti-PD-1 therapy. Embo j. 2020;39(20):e104514. https://doi.org/10.15252/embj.2020104514.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Yang S, Wei J, Cui YH, Park G, Shah P, Deng Y, et al. M(6)a mRNA demethylase FTO regulates melanoma tumorigenicity and response to anti-PD-1 blockade. Nat Commun. 2019;10(1):2782. https://doi.org/10.1038/s41467-019-10669-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Li N, Kang Y, Wang L, Huff S, Tang R, Hui H, et al. ALKBH5 regulates anti-PD-1 therapy response by modulating lactate and suppressive immune cell accumulation in tumor microenvironment. Proc Natl Acad Sci U S A. 2020;117(33):20159–70. https://doi.org/10.1073/pnas.1918986117.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Visvanathan A, Patil V, Arora A, Hegde AS, Arivazhagan A, Santosh V, et al. Essential role of METTL3-mediated m(6)a modification in glioma stem-like cells maintenance and radioresistance. Oncogene. 2018;37(4):522–33. https://doi.org/10.1038/onc.2017.351.

    Article  CAS  PubMed  Google Scholar 

  77. Kowalski-Chauvel A, Lacore MG, Arnauduc F, Delmas C, Toulas C, Cohen-Jonathan-Moyal E, et al. The m6A RNA demethylase ALKBH5 promotes Radioresistance and Invasion Capability of Glioma Stem cells. Cancers (Basel). 2020;13(1). https://doi.org/10.3390/cancers13010040.

  78. Zhou S, Bai ZL, Xia D, Zhao ZJ, Zhao R, Wang YY, et al. FTO regulates the chemo-radiotherapy resistance of cervical squamous cell carcinoma (CSCC) by targeting β-catenin through mRNA demethylation. Mol Carcinog. 2018;57(5):590–7. https://doi.org/10.1002/mc.22782.

    Article  CAS  Google Scholar 

  79. Wu P, Fang X, Liu Y, Tang Y, Wang W, Li X, et al. N6-methyladenosine modification of circCUX1 confers radioresistance of hypopharyngeal squamous cell carcinoma through caspase1 pathway. Cell Death Dis. 2021;12(4):298. https://doi.org/10.1038/s41419-021-03558-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. He JJ, Li Z, Rong ZX, Gao J, Mu Y, Guan YD, et al. M(6)a reader YTHDC2 promotes Radiotherapy Resistance of Nasopharyngeal Carcinoma via activating IGF1R/AKT/S6 Signaling Axis. Front Oncol. 2020;10:1166. https://doi.org/10.3389/fonc.2020.01166.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Ma Y, Liu X, Bi Y, Wang T, Chen C, Wang Y, et al. Alteration of N(6)-Methyladenosine mRNA methylation in a human stem cell-derived cardiomyocyte model of tyrosine kinase Inhibitor-Induced Cardiotoxicity. Front Cardiovasc Med. 2022;9:849175. https://doi.org/10.3389/fcvm.2022.849175.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. He L, Li H, Wu A, Peng Y, Shu G, Yin G. Functions of N6-methyladenosine and its role in cancer. Mol Cancer. 2019;18(1):176. https://doi.org/10.1186/s12943-019-1109-9.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Du A, Li S, Zhou Y, Disoma C, Liao Y, Zhang Y, et al. M6A-mediated upregulation of circMDK promotes tumorigenesis and acts as a nanotherapeutic target in hepatocellular carcinoma. Mol Cancer. 2022;21(1):109. https://doi.org/10.1186/s12943-022-01575-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Cui Y, Zhang C, Ma S, Li Z, Wang W, Li Y, et al. RNA m6A demethylase FTO-mediated epigenetic up-regulation of LINC00022 promotes tumorigenesis in esophageal squamous cell carcinoma. J Exp Clin Cancer Res. 2021;40(1):294. https://doi.org/10.1186/s13046-021-02096-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Wang S, Gao S, Zeng Y, Zhu L, Mo Y, Wong CC, et al. N6-Methyladenosine reader YTHDF1 promotes ARHGEF2 translation and RhoA signaling in Colorectal Cancer. Gastroenterology. 2022;162(4):1183–96. https://doi.org/10.1053/j.gastro.2021.12.269.

    Article  CAS  PubMed  Google Scholar 

  86. Wang Y, Chen J, Gao WQ, Yang R. METTL14 promotes prostate tumorigenesis by inhibiting THBS1 via an m6A-YTHDF2-dependent mechanism. Cell Death Discov. 2022;8(1):143. https://doi.org/10.1038/s41420-022-00939-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Zhang L, Li Y, Zhou L, Zhou H, Ye L, Ou T, et al. The m6A reader YTHDF2 promotes bladder cancer progression by suppressing RIG-I-mediated immune response. Cancer Res. 2023. https://doi.org/10.1158/0008-5472.Can-22-2485.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Yang X, Zhang S, He C, Xue P, Zhang L, He Z, et al. METTL14 suppresses proliferation and metastasis of colorectal cancer by down-regulating oncogenic long non-coding RNA XIST. Mol Cancer. 2020;19(1):46. https://doi.org/10.1186/s12943-020-1146-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Li A, Cao C, Gan Y, Wang X, Wu T, Zhang Q, et al. ZNF677 suppresses renal cell carcinoma progression through N6-methyladenosine and transcriptional repression of CDKN3. Clin Transl Med. 2022;12(6):e906. https://doi.org/10.1002/ctm2.906.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Chen C, Guo Y, Guo Y, Wu X, Si C, Xu Y, et al. m6A modification in non-coding RNA: the role in Cancer Drug Resistance. Front Oncol. 2021;11:746789. https://doi.org/10.3389/fonc.2021.746789.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Liu H, Lyu H, Jiang G, Chen D, Ruan S, Liu S, et al. ALKBH5-Mediated m6A demethylation of GLUT4 mRNA promotes glycolysis and resistance to HER2-Targeted therapy in breast Cancer. Cancer Res. 2022;82(21):3974–86. https://doi.org/10.1158/0008-5472.Can-22-0800.

    Article  CAS  PubMed  Google Scholar 

  92. Li H, Wang C, Lan L, Yan L, Li W, Evans I, et al. METTL3 promotes oxaliplatin resistance of gastric cancer CD133 + stem cells by promoting PARP1 mRNA stability. Cell Mol Life Sci. 2022;79(3):135. https://doi.org/10.1007/s00018-022-04129-0.

    Article  CAS  PubMed  Google Scholar 

  93. Wang X, Wu R, Liu Y, Zhao Y, Bi Z, Yao Y, et al. M(6)a mRNA methylation controls autophagy and adipogenesis by targeting Atg5 and Atg7. Autophagy. 2020;16(7):1221–35. https://doi.org/10.1080/15548627.2019.1659617.

    Article  CAS  PubMed  Google Scholar 

  94. Li T, Hu PS, Zuo Z, Lin JF, Li X, Wu QN, et al. METTL3 facilitates tumor progression via an m(6)A-IGF2BP2-dependent mechanism in colorectal carcinoma. Mol Cancer. 2019;18(1):112. https://doi.org/10.1186/s12943-019-1038-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Olson B, Li Y, Lin Y, Liu ET, Patnaik A. Mouse models for Cancer Immunotherapy Research. Cancer Discov. 2018;8(11):1358–65. https://doi.org/10.1158/2159-8290.Cd-18-0044.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Chen Y, Lu Z, Qi C, Yu C, Li Y, Huan W, et al. N(6)-methyladenosine-modified TRAF1 promotes sunitinib resistance by regulating apoptosis and angiogenesis in a METTL14-dependent manner in renal cell carcinoma. Mol Cancer. 2022;21(1):111. https://doi.org/10.1186/s12943-022-01549-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Xie H, Yao J, Wang Y, Ni B. Exosome-transmitted circVMP1 facilitates the progression and cisplatin resistance of non-small cell lung cancer by targeting miR-524-5p-METTL3/SOX2 axis. Drug Deliv. 2022;29(1):1257–71. https://doi.org/10.1080/10717544.2022.2057617.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Liu Z, Wu K, Gu S, Wang W, Xie S, Lu T, et al. A methyltransferase-like 14/miR-99a-5p/tribble 2 positive feedback circuit promotes cancer stem cell persistence and radioresistance via histone deacetylase 2-mediated epigenetic modulation in esophageal squamous cell carcinoma. Clin Transl Med. 2021;11(9):e545. https://doi.org/10.1002/ctm2.545.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Thanks to Danping Xiong for her encouragement and support.

Funding

This work was supported by the National Natural Science Foundation of China (82073476 and 81773226), the National Key R&D Program of China (2022YFC2503700, 2022YFC2503703), Jiangsu Provincial Medical Key Discipline (ZDXK202235), Innovation Research Project of Medical and Industrial Cooperation in Suzhou (SLJ2021005), the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions, Wuxi double hundred young and middle-aged medical and health top-notch talent project (No. 202014, China), and Youth Scientific Research Project of Wuxi Municipal Health Commission(No. Q202236, China).

Author information

Authors and Affiliations

Authors

Contributions

Lili Wang, Yang Jiao and Jian Wang designed and revised the manuscript. Hengzhao Zhuang, Bo Yu and Dan Tao drafted the manuscript. Xiaoyan Xu and Yijun Xu participated in the the procedures. All authors reviewed the manuscript and approved the submitted version.

Corresponding authors

Correspondence to Jian Wang, Yang Jiao or Lili Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interest

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhuang, H., Yu, B., Tao, D. et al. The role of m6A methylation in therapy resistance in cancer. Mol Cancer 22, 91 (2023). https://doi.org/10.1186/s12943-023-01782-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-023-01782-2

Keywords