Next Article in Journal
Tumor Colonization and Therapy by Escherichia coli Nissle 1917 Strain in Syngeneic Tumor-Bearing Mice Is Strongly Affected by the Gut Microbiome
Previous Article in Journal
The Emerging Role of uORF-Encoded uPeptides and HLA uLigands in Cellular and Tumor Biology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Biological Roles and Molecular Mechanisms of Long Non-Coding RNA MEG3 in the Hallmarks of Cancer

1
Key Laboratory of Biorheological Science and Technology, Ministry of Education, College of Bioengineering, Chongqing University, Chongqing 400044, China
2
The 111 Project Laboratory of Biomechanics and Tissue Repair, College of Bioengineering, Chongqing University, Chongqing 400044, China
3
Chongqing Key Laboratory of Translational Research for Cancer Metastasis and Individualized Treatment, Chongqing University Cancer Hospital, Chongqing University, Chongqing 400030, China
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(24), 6032; https://doi.org/10.3390/cancers14246032
Submission received: 25 November 2022 / Accepted: 6 December 2022 / Published: 7 December 2022

Abstract

:

Simple Summary

MEG3 is a class of lncRNA, which is considered a tumor suppressor. It is lost or decreased in different biological processes of various human tumors and is closely related to various diseases. MEG3 can modulate the expression of target genes through transcription, translation, post-translational modification and epigenetic regulation. Studies have shown that MEG3 dysfunction has been linked to a poor prognosis and drug resistance. MEG3 mediates the hallmarks of cancer through a variety of mechanisms, acting as a tumor suppressor to limit tumor growth. Hence, MEG3 is a potential prognostic marker and antitumor therapeutic target.

Abstract

Long non-coding RNAs (lncRNAs) are critical regulators in various biological processes involved in the hallmarks of cancer. Maternally expressed gene 3 (MEG3) is lncRNA that regulates target genes through transcription, translation, post-translational modification, and epigenetic regulation. MEG3 has been known as a tumor suppressor, and its downregulation could be found in various cancers. Furthermore, clinical studies revealed that impaired MEG3 expression is associated with poor prognosis and drug resistance. MEG3 exerts its tumor suppressive effect by suppressing various cancer hallmarks and preventing cells from acquiring cancer-specific characteristics; as it could suppress tumor cells proliferation, invasion, metastasis, and angiogenesis; it also could promote tumor cell death and regulate tumor cell metabolic reprogramming. Hence, MEG3 is a potential prognostic marker, and overexpressing MEG3 might become a potential antitumor therapeutic strategy. Herein, we summarize recent knowledge regarding the role of MEG3 in regulating tumor hallmarks as well as the underlying molecular mechanisms. Furthermore, we also discuss the clinical importance of MEG3, as well as their potential in tumor prognosis and antitumor therapeutic strategies.

1. Introduction

Cancer is the main cause of death globally and a significant impediment to extending life expectancy. In 2020, there was an estimated 19.3 million new cases of cancer and nearly 10 million cancer-related mortality globally [1]. While cancers that are accessible for early identification are slowing down, other prevalent malignancies are making significant progress [2]. Hence, there is an urgent need to find novel prognostic biomarkers and tumor therapeutic targets to combat cancer.
Tumorigenesis as well as the malignant transformation from benign tumors to malignant cancers is a complex process due to aberrant gene expressions. Distinct from normal cells, tumor cells have gained special characteristics, which are known as “hallmarks of cancer”, including sustaining proliferative signaling, resisting cell death, inducing angiogenesis, activating invasion and metastasis, and metabolic reprogramming [3,4]. Besides mutations in protooncogenes and tumor suppressor genes, impaired gene expression regulatory pathways such as transcriptional, translational, post-translational, or epigenetic regulations are also the main reasons for tumorigenesis and malignant transformation [5,6]. In the last two decades, numerous studies have revealed that non-coding RNA (ncRNA), such as microRNA (miRNA), long non-coding RNA (lncRNA), circular RNA (circRNA), small interfering RNA (siRNA), and RNA interacting with piwi proteins (piRNA), are crucial regulators of gene expression. These ncRNAs could exert their regulatory functions by regulating various steps of gene expression, that is, transcription, post-transcriptional modifications, translation, post-translational modifications, chromatin remodeling, and signal transduction [7,8].
Long non-coding RNAs (lncRNAs) are a class of non-coding RNAs more than 200 nucleotides in length. LncRNAs can act as competing endogenous RNA (ceRNA) that sponges and blocks the effect of miRNAs, a class of ncRNA that suppresses target genes expression at their translational level by binding to their 3′ untranslated regions (3′ UTR) [9]. Furthermore, lncRNAs could interact with DNA, RNA, protein molecules and/or their complexes, acting as an essential regulator in transcriptional, post-transcriptional, and chromatin remodeling regulations [10]. Interestingly, recent studies found that some lncRNAs contain short open reading frames (sORFs), which can encode small proteins or micropeptides to exert their physiological roles [11,12]. Aberrant lncRNA expression, which could be caused by single nucleotide polymorphism (SNPs), copy number alterations, and mutations, has been found in various tumors such as colorectal cancer (CRC), thyroid cancer, and ovarian cancer (OC), and is closely related with cancer hallmarks and malignant transformation [13,14,15]. Thus, lncRNAs have gained attention as novel biomarkers for tumors and as targets for antitumor therapeutic strategies [16].
Maternally expressed gene 3 (MEG3) is an imprinted gene with an approximate length of 35 kbp and found at the DLK1-MEG3 locus on human chromosome 14q32.3 [17]. Its mouse homolog, gene trap loci 2 (Gtl2), is located on mouse distal chromosome 12 [18]. MEG3 is transcribed by RNA polymerase 2 and spliced into 10 exons containing five key structural motifs (M-I to M-V) [17,19,20,21]. Mature MEG3 RNA, which is 1.6 kbp in length, is polyadenylated at its 3′ ends, and is located both in the nucleus and cytoplasm [22,23] (Figure 1). MEG3 could impact various diseases, including ischemic neuronal death, atherosclerosis and type 2 diabetes mellitus [24,25,26]. Recent studies revealed that MEG3 expression decreased in a wide variety of tumors, playing a crucial role as a tumor suppressor [27,28]. The first study regarding the role of MEG3 in tumors was reported by Zhang et al. They found the defect of MEG3 expression in pituitary adenomas, and that ectopic expression of this gene suppressed tumor cell growth [29]. More recently, Moradi et al. reported that MEG3 could function as a ceRNA that interact directly with multiple genes or proteins, including p53, enhancer of zeste homologue 2 (EZH2), and nuclear factor-kappa B (NF-κB). MEG3 exerts its tumor-suppressive effect by regulating various cancer hallmarks, as it could inhibit tumor cell proliferation, induce cell death, reduce invasion and metastasis, prevent angiogenesis, and inhibit tumor cells’ metabolic reprogramming.
In this review, we summarized the current knowledge regarding the expression level, functions, as well as the molecular mechanism underlying MEG3 regulation on cancer hallmarks. Furthermore, we highlighted the clinical significance of MEG3 as a biomarker for cancer prognosis, as well as a novel therapeutic strategy for cancer therapy.

2. Mechanism of MEG3 Regulations

2.1. MEG3 Could Sponge miRNAs

LncRNAs can function as ceRNAs that bind to target miRNAs like a sponge and prevent miRNA from binding to its target mRNA, thus affecting the mRNA abundance of the target gene and their protein levels [30,31]. Numerous studies have reported that MEG3 can function as ceRNA by sponging and sequestering miRNAs, such as miR-21, miR-181a, and miR-421, from their target genes [32,33,34,35,36,37,38,39,40,41,42]. Similar to the targets of the miRNAs it regulates, MEG3 possesses microRNA response elements (MREs). Through these MREs, MEG3 binds to the miRNA binding sites competitively with the corresponding target mRNAs, thereby removing the target mRNAs and eliminating the inhibitory effect of miRNA on them. The lncRNA-miRNA-mRNA forms a complex network of action, whose homeostasis is crucial for maintaining normal physiological conditions. Meanwhile, disruption of this homeostasis is closely related to diseases including cancers [43].

2.2. MEG Regulations on Target Genes Transcription

Besides as a ceRNA, MEG3 can regulate its targets through transcriptional as well as post-translational regulations (Figure 2). For example, MEG3 could promote p53 expression by promoting its transcriptional activity and post-translational modification. MEG3 could enhance p53 transcriptional activity, thereby increasing p53 expression level and negatively regulating the cell cycle [44,45] Furthermore, MEG3 could also decrease the level of murine double minute 2 (MDM2), an E3 ubiquitin ligase that enhances p53 ubiquitination/proteasomal degradation, leading to p53 protein stabilization and transcriptional activation of p53 downstream targets [46]. Meanwhile, Weng et al. also found that MEG3, through its 732-1174 nucleic acid region, binds directly to Clusterin (CLU) protein and impedes CLU’s interactions with its target proteins, such as vascular endothelial growth factor (VEGF) or matrix metalloproteinase (MMP-9) [47]. MEG3 affects the stability of proteins by regulating their post-translational modifications. Zhang et al. showed that MEG3 could suppress the accumulation of the phosphorylated signal transducer and activator of transcription 3 (p-STAT3) protein by recruiting ubiquitination enzymes and thus directing pSTAT3 into ubiquitin/proteasomal degradation pathway without affecting its phosphorylation. This in turn suppresses the p-STAT3/c-Myc axis, and subsequently, leads to a decrease in cell proliferation potential [48].
Several studies have reported that lncRNAs are involved in chromatin remodeling by directing the recruitment of chromatin modifiers to target gene sites, for example, by associating with polycomb repressive complex 2 (PRC2) and inducing the trimethylation of histone H3 lysine 27 (H3K27me3) [49,50]. MEG3 could function as a molecular scaffold linking different proteins and forming large complexes that regulate chromatin structure and gene expression. By interacting with the RNA binding domain of Jumonji and AT-rich interaction domain containing 2 (JARID2), MEG3 stimulates PRC2 and JARID2 assembly, thereby enhancing H3K27me3 recruitment and suppressing the transcription of E-cadherin and miR-200 family [51].
MEG3 could also induce H3K27me3 by interacting with EZH2, a catalytic subunit of the PRC2 complex. Through this regulation, MEG3 induces the deposition of H3K27me3 in the distal regulatory region (DRE) of the transforming growth factor-β (TGF-β) gene, thereby inhibiting TGF-β gene transcription in trans, and subsequently, the transcription of TGF-β pathway genes transforming growth factor beta receptor 1 (TGFBR1), transforming growth factor beta 1 (TGFB1), and SMAD family member 2 (SMAD2) [52]. Similarly, MEG3 inhibits engrailed-2 (EN-2) expression by EZH2-mediated H3K27me3 [53]. Interestingly, MEG3 could also interact with EZH2 protein and stimulates its ubiquitination/proteasomal degradation; EZH2 could in turn suppress MEG3 through its interaction with DNA methyltransferase 1 (DNMT1) and histone deacetylase 1 (HDAC1), thereby suppressing MEG3 transcription by inducing DNA methylation. Hence, the regulation of MEG3 on EZH2 forms a concerted negative feedback loop [54].

3. MEG3 Regulates Various Hallmarks of Cancer

Recent studies revealed that MEG3 is associated with hallmarks of cancer, including proliferation, cell death, invasion and metastasis, metabolic reprogramming, and angiogenesis, by regulating various pathways (Table 1). MEG3 inhibits cancer progression through different mechanisms. MEG3 is involved in tumor progression in two ways, such as acting as a sponge for miRNA and regulating its targets through transcriptional as well as post-translational regulations. For example, MEG3 is closely related to the expression level of p53, a tumor suppressor whose mutation could be found in more than 50% of cancer patients [55]. MEG3 can directly interact with the DNA binding domain of p53 thereby enhancing the transcription of numerous p53 target genes [56]. MEG3 can also regulate p53 expression level indirectly by decreasing MDM2 protein level, leading to the decrease in MDM2-mediated p53 ubiquitination/proteasomal degradation, thereby stabilizing p53 protein levels [57,58].

3.1. MEG3 Inhibits Tumor Cell Proliferation

Abnormal, uncontrolled cell growth due to the dysregulation of cell proliferation is the most fundamental cause of tumorigenesis. Aberrant MEG3 expression has been observed in various tumors and is closely linked with tumor cell proliferation [76]. MEG3 could up-regulate OTU deubiquitinase 4 (OTUD4) and RNA binding motif single-stranded interacting protein 3 (RBMS3) by sponging miR-494 and miR-141-3p, respectively, thereby suppressing breast cancer cells proliferation [59,77]. MEG3 could inhibit the growth and proliferation of T-cell lymphoblastic lymphoma by sponging miR-214, thereby activating apoptosis-inducing factor mitochondrion-associated 2 (AIFM2) expression [61]. By sponging miR-494 and miR-374a-5p, MEG3 can up-regulate phosphatase and tensin homolog (PTEN), resulting in cell growth inhibition in bladder cancer and pancreatic ductal adenocarcinoma [78,79]. MEG3 also can inhibit cholangiocarcinoma proliferation and invasion by inhibiting the major components of the PRC1 complex, B lymphoma Mo-MLV insertion region 1 (Bmi1), and RING finger protein 2 (RNF2) [80].
The cell cycle is an important process that regulates cell proliferation. MEG3 could induce cell-cycle arrest in G0/G1 phase, thereby suppressing cell proliferation and ultimately inducing cell apoptosis. MEG3 could induce G0/G1 cell cycle arrest in glioma and ovarian cancer cells by inactivating the Wnt/β-catenin signaling pathway and upregulating PTEN expression, respectively [60,81]. Furthermore, by sponging its target miRNAs, such as miR-10a-5p, MEG3 can cause G0/G1 cell cycle arrest and enhance the expression of PTEN, Bcl-2-associated X (Bax), and p53 protein in hepatocellular carcinoma (HCC) [82], or by sponging miR-7 and miR-376, leading to the downregulation of miR-7/RAS like family 11 member B (RASL11B) and miR-376/protein kinase D1 (PKD1) axis [62,63].
MEG3 could also inactivate the PI3K/Akt and ERK pathways, which are crucial for cell proliferation [3]. In renal cell carcinoma, MEG3 upregulates β-galactoside α-2,3-sialyltransferase 1 (ST3Gal1) through its interaction with transcription factor c-Jun, leading to the decrease in epithelial growth factor receptor (EGFR) phosphorylation and PI3K/Akt pathway inactivation [83]. Meanwhile, in gliomas and hemangiomas, MEG3 could inactivate PI3K/Akt pathway by sponging miR-93 and miR-494, respectively [84,85]; in pancreatic neuroendocrine tumors, MEG3 downregulates brain protein I3 (BRI3) expression by sponging miR-183, leading to the inactivation of p38/ERK/Akt and Wnt/β-catenin signaling pathways [64].
Cancer stem cells (CSCs) are a small population of tumor cells that are usually in the dormant stage and have been assumed to be the main reason for tumorigenesis potential, tumor metastasis, recurrence, and drug resistance [86]. Targeting CSCs has been considered a potential therapeutic strategy for eradicating cancers; however, they are significantly less sensitive to current chemotherapy- and radiotherapy-based antitumor therapeutic strategies, as these strategies target proliferative cells [87]. MEG3 can repress CSC self-renewal ability and decrease cancer stemness phenotype in oral CSCs by blocking miR-421 [88]. Furthermore, by sponging miR-708, MEG3 enhances SOCS3 expression, thereby decreasing colorectal CSCs stemness by suppressing STAT3 signaling [89].

3.2. MEG3 Induces Cell Death

Apoptosis is a programmed cell death controlled by a signaling cascade to maintain a stable internal environment. The elimination of cancer cells by apoptosis has been a key cue in clinical cancer treatment [90]. Apoptosis could be divided into intrinsic and extrinsic apoptotic pathways. The intrinsic apoptotic pathway, also known as mitochondria-mediated apoptosis, is regulated by pro-apoptotic B-cell lymphoma 2 (Bcl-2) proteins, anti-apoptotic Bcl-2 proteins, and BH3-only proteins, which triggers the activation of executor caspases 3 and 7 by activating caspase 8. Meanwhile, the extrinsic apoptotic pathway is regulated by death receptors, such as the tumor necrosis factor (TNF) receptor, which promotes the cleavage of initiator caspase, caspase 9, subsequently activating executor caspases [91,92].
Previous reports have shown that MEG3 could enhance intrinsic apoptosis by various mechanisms. In prostate cancer, osteosarcoma, urinary tract epithelial cancer and pituitary tumor cells, MEG3 could directly bind to miR-361-5p, miR-96, and miR-376B-3p, leading to the promotion of forkhead box M1 (FoxM1) and tropomyosin 1 (TPM1) expression while repressing oncogene high mobility group AT-hook 2 (HMGA2) expression. This results in the reduction in Bcl-2 and the rising of Bax protein levels, as well as the increase in caspases-3 and -9 cleavages, thereby inducing tumor cell apoptosis [93,94,95,96]. In oral squamous cell carcinoma (OSCC) and CML, MEG3 promoted apoptosis by sponging miR-548d-3p and miR-147, thereby promoting suppressor of cytokine signaling 5 (SOCS5) and suppressor of cytokine signaling 6 (SOCS6) expression while inhibiting the JAK-STAT signaling pathway [65,66].
MEG3 could also trigger apoptosis in ESCC and CRC by increasing endoplasmic reticulum (ER) stress-related proteins, including glucose-regulated protein 78 (GRP78), activating transcription factor 6 (ATF6), protein kinase R-like endoplasmic reticulum kinase (PERK), and C/EBP-homologous protein (CHOP), leading to enhanced caspases-9 and -3 cleavages [97,98]. Furthermore, MEG3 could activate apoptotic cascade in laryngeal cancer by sponging miR-23a and promotes apoptotic protease activating factor-1 (APAF-1) expression [99], and in gallbladder cancer by promoting EZH2 ubiquitination/proteasomal degradation. This in turn suppressed the expression level of its downstream target tumor suppressor large tumor suppressor 2 (LATS2), thus increasing the levels of cleaved PARP, Bax, and Bcl-2 [100]. Moreover, it could also downregulate miR-21-5p, leading to an increase in p53 and caspase 3 cleavage protein cleavage [67].
Besides intrinsic apoptosis, MEG3 could also trigger extrinsic apoptosis pathways. In cholangiocarcinoma and gallbladder cancer, MEG3 stimulated NF-κB signaling pathway and triggered apoptosis by sponging miR-361-5p expression and activating TNF receptor-associated factor 3 (TRAF3) [101,102].
Autophagy is an intracellular self-destructive form of cell death that transfers cytoplasmic proteins or organelles to the lysosome to fulfill the metabolic and self-renewal needs of organelles and the cell itself [103,104,105]. Previous studies have reported that MEG3 could attenuate autophagy by suppressing the forkhead box O1 (FOXO1) expression, leading to the decrease in autophagy-related proteins microtubule-associated protein light chain 3 II (LC3 II), beclin 1, autophagy related 3 (ATG3), autophagy related 5 (ATG5), and autophagy related 12 (ATG12), as well as the increase in the autophagy substrate p62 [106]. Hence, MEG3 regulation on autophagy needs further investigation.

3.3. MEG3 Negatively Regulates Tumor Cells Invasion and Metastasis Potentials

Metastasis is a complex process that includes epithelial-mesenchymal transition (EMT), invasion, intravasation, cell survival in circulation, extravasation, and metastatic colonization [107,108]. EMT is the first, initiative event in cancer metastasis in which epithelial cells gained mesenchymal characteristics such as decreased intercellular adhesion and increased motility, while losing epithelial characteristics [109,110]. MEG3 could suppress GC cells’ EMT and metastasis potential by sponging miR-21, leading to the increase in the expression of epithelial marker E-cadherin, and a decrease in mesenchymal markers such as N-cadherin, Snail, and β-catenin as well as cell migration markers such as matrix metalloproteinase-2 (MMP-2), matrix metalloproteinase-3 (MMP-3), and MMP-9 [34,111] and by inhibiting the binding between miR-665 and its target, cytokine signaling 3 (SOCS3), thereby enhancing SOCS3 expression and suppressing FAK/Src pathway [112]. Meanwhile, by sponging miR-216a, MEG3 enhances programmed death-1 (PD-1) expression while suppressing EMT inducer myeloid cell leukemia-1 (MCL-1) in endometrial cancer cells [113]. In OC cells, MEG3 could inhibit tumor cells migration and invasion potentials by sponging up miR-219a-5p and miR-30e-3p, resulting in the downregulation of EGFR and increase in laminin subunit alpha 4 (LAMA4), respectively [68,114]. Meanwhile, by sponging miR-19a, MEG3 enhanced PTEN expression, thereby suppressing glioma cell migration and invasion potentials [69]. Furthermore, in HCC, MEG3 could inhibit metastasis by sponging miR-544b and miR-5195-3p, thereby upregulating target genes B-cell translocation gene (BTG2) and FOXO1 expression [70,115]. Moreover, MEG3 could also suppress EMT by blocking the phosphoserine aminotransferase 1 (PSAT1)-dependent glycogen synthase kinase (GSK)-3β/Snail signaling [116].
The link between MEG3 and metastasis has also been confirmed by clinical samples from thyroid cancer (TC) patients showing that MEG3 downregulation was associated with lymph node metastasis. MEG3 could suppress TC cell migration and invasion by downregulating Rac family small GTPase 1 (Rac1) expression by targeting its 3′ UTR [117]. Furthermore, MEG3 competitively interacts with miR-27a as the ceRNA of PH domain and leucine-rich repeat protein phosphatase 2 (PHLPP2) mRNA, promoting PHLPP2 protein translation and inhibiting c-Jun phosphorylation and c-Jun-mediated c-Myc mRNA transcription, thereby impairing invasion and lung metastasis of bladder cancer cells [71].

3.4. MEG3 Regulation on Tumor Cells Metabolic Reprogramming

Metabolic alteration is a characteristic of tumor cells crucial for supporting their rapid cell growth [3]. Unlike normal cells, which mainly depend on glycolysis followed by oxidative phosphorylation, tumor cells prefer inefficient aerobic glycolysis with a significantly higher turnover rate compared to normal cells even under adequate oxygen availability. This phenomenon is known as the Warburg effect [118,119]. The reprogrammed metabolic network generates intermediates, such as those involved in the glycolysis or tricarboxylic acid (TCA) cycle processes, which benefit cancer cells by helping them meet their energy needs as well as anabolic and redox and building blocks demands in the early stages of cancer development [120]. MEG3 activated by vitamin D can inhibit aerobic glycolysis and lactic acid production in CRC cells by inducing ubiquitin-dependent c-Myc degradation, thereby inhibiting c-Myc target genes expression involved in the glycolysis pathway, such as lactate dehydrogenase A (LDHA), pyruvate kinase muscle 2 (PKM2) and hexokinase 2 (HK2) [72]. Furthermore, MEG3 can promote succinate dehydrogenase (SDH) expression by sponging miR-361-5p, leading to an increase in succinate, a key TCA metabolite, thereby suppressing OSCC progression [73].

3.5. MEG3 Suppresses Tumor Angiogenesis

Formation of new blood vessels in tumor tissues from existing blood vessels is crucial for supplying tumor cells with oxygen and nutrient, for adapting to the fluctuating oxygen pressure in their microenvironment, as well as for metastasis [121]. This process involved many angiogenic factors, including vascular endothelial growth factor A (VEGFA), basic fibroblast growth factor (bFGF), and angiogenin. These factors increase endothelial cell development and vascular permeability, resulting in the formation of new blood vessels [122]. The role of MEG3 in tumor angiogenesis remains intriguing. Zhang et al. reported that MEG3 can suppress angiogenesis-related gene VEGFA, placental growth factor (PGF), bFGF, transforming growth factor β1 (TGF-β1) and MMP-9 expression by decreasing phosphorylated levels of AKT and inhibiting AKT pathway, ultimately suppressing angiogenesis in breast cancer [74]. However, Li et al. demonstrated that MEG3 could promote angiogenesis in lung carcinoma, as it could significantly increase the expression of angiogenesis-related factors VEGFA, vascular endothelial growth factor B (VEGFB), bFGF, stromal cell-derived factor-1 (SDF-1), transforming growth factor β (TGF-β), angiogenin, and MMP-9 [75]. The reasons underlying this discrepancy need further investigation.

4. Clinical Significance of lncRNA MEG3

4.1. MEG3 Is a Potential Biomarker for Tumor Prognosis

Decreased expression of MEG3 was associated with poor prognosis in a variety of human malignancies [123]. As shown in Table 2, MEG3 has been proven to have anti-tumor effects, and potential prognostic and clinical significance in various human cancers [60,76,123,124,125,126,127,128,129,130].
Analysis of MEG3 expression in glioma patients showed that low expression of MEG3 was associated with poor overall survival rates, advanced WHO grade, low Karnofsky performance score (KPS), isocitrate dehydrogenase (IDH) wild-type, and tumor recurrence [60,125]. Xu et al. revealed that the copy number variation (CNV) levels of MEG3 were positively associated with overall survival and progression-free survival compared to the wild-type in low-grade glioma [123]; Gao et al. revealed, using 63 patients with retinoblastoma, that hypermethylation of MEG3 promoter was highly associated with poor survival, further confirming that MEG3 expression level is negatively correlated with poor prognosis [128]. Meanwhile, using 58 clinical ESCC tissues, Ma et al. found that low MEG3 expression was correlated with tumor size, lymph node metastasis, clinical stage, and poor prognosis [126]. These results were in accordance with other studies involving 48 CRC cases [129]. Furthermore, a negative correlation between MEG3 expression and short overall survival, relapse-free survival, and poor prognosis has also been found in breast cancer, NSCLC, and glioblastoma [76,127,130]. Together, these results show a negative correlation between MEG3 and tumor progression as well as prognosis, indicating the potential of using MEG3 as a biomarker for tumor prognosis.

4.2. MEG3 Is a Potential Target for Tumor Therapy

Anti-tumor therapies have been evolving and improving in recent years, yet resistance to chemotherapy, radiotherapy, targeted therapy, and immunotherapy remains a major problem [131]. Cytotoxic anti-tumor drugs such as cisplatin, paclitaxel, and doxorubicin, as well as targeted medicines such as imatinib, have been used for clinical cancer treatment. However, the persistent rise of drug resistance seriously undermines their efficacies [132]. MEG3 can facilitate chemotherapeutic drug sensitivity and radiosensitivity by altering key signaling pathways, making it a novel therapeutic strategy for cancer treatment (Table 3).
Assessment using 90 peritoneal biopsies of high-grade serous OC showed that MEG3 expression is associated with sensitivity to platinum-based chemotherapy [133]. MEG3 can act as an agonist of cisplatin in suppressing triple-negative breast cancer (TNBC) growth and metastasis potentials, and facilitate pyroptosis by activating cisplatin-induced NLRP3/caspase-1/gasdermin D (GSDMD) pathway [134]. MEG3 can also enhance NSCLC sensitivity to cisplatin by sponging miR-21-5p and thereby upregulating SRY-box transcription factor 7 (SOX7) expression [135]; by sponging miR-141, MEG3 could overcome CRC cells chemoresistance to oxaliplatin and promote programmed cell death factor 4 (PDCD4) expression [129]. Subsequently, MEG3 could suppress cisplatin and cyclophosphamide resistance in T-cell lymphoblastic lymphoma cells through the PI3K/mTOR pathway [136].
MEG3 could also act as an agonist of other antitumor drugs. Through MEG3/miR-4513/phenazine biosynthesis-like domain-containing (PBLD) axis, MEG3 promoted breast cancer cells’ sensitivity to paclitaxel [137]. Furthermore, MEG3 suppresses the levels of drug-resistant transporters, including multidrug resistance-associated protein-1 (MRP1), multidrug resistance protein 1 (MDR1), and ATP binding cassette subfamily G member 2 (ABCG2), thus increasing CML cells’ sensitivity against imatinib; miR-21 mimics could reverse their levels [138]. Meanwhile, by sponging miR-155, MEG3 upregulated alpha-1,2-mannosyltransferase (ALG9) expression, thereby promoting AML cells’ sensitivity against adriamycin and vincristine [139]. Moreover, MEG3 could promote pancreatic cancer cells’ chemoresistance to gemcitabine [140].
Besides, MEG3 was closely related to 131I-sensitivity of thyroid carcinoma by sponging miR-182 [141]. Finally, very recent research showed that tumor-targeting therapy of osteosarcoma (OS) can be performed by a highly effective engineered and MEG3-loaded exosome, as a combination of MEG3 and exosome significantly increased MEG3 therapeutic effect [142]. Together, these findings suggest that MEG3 plays a significant role in enhancing chemotherapeutic drug sensitivity and radiosensitivity in a variety of human cancers, making it a potential therapeutic target for cancer treatment.
Table 3. Roles of MEG3 in therapeutic resistance of cancers.
Table 3. Roles of MEG3 in therapeutic resistance of cancers.
Cancer TypeExpressionTargetChemical-/RadioresistanceRefs
TNBCDownregulatedNLRP3/caspase-1/GSDMD pathwayCisplatin (DDP)[134]
NSCLCDownregulatedmiR-21-5p/SOX7Cisplatin[135]
T-cell lymphoblastic lymphomaDownregulatedPI3K/mTOR signalingCisplatin and Cyclophosphamide[136]
CRCDownregulatedmiR-141/PDCD4Oxaliplatin[129]
AMLDownregulatedmiR-21/MRP1, MDR1, and ABCG2Imatinib[138]
Breast cancerDownregulatedmiR-4513/PBLDPaclitaxel (PTX)[137]
ACLDownregulatedmiR-155/ALG9Adriamycin and Vincristine[139]
Thyroid carcinomaDownregulatedmiR-182131I[141]
Abbreviations: TNBC: triple-negative breast cancer; ESCC: esophageal squamous cell carcinoma; CRC: colorectal cancer; AML: chronic myeloid leukemia; ACL: acute myeloid leukemia.

5. Conclusions and Perspectives

MEG3 has emerged as a potential tumor suppressor that could regulate various hallmarks of cancer including cell proliferation, cell death, invasion and metastasis, metabolic reprogramming, angiogenesis, and drug resistance (Figure 3). MEG3 expression is downregulated in most malignant tumors, including glioma, HCC, CRC, and breast cancer. As shown in Figure 2, MEG3 regulation on tumor progression occurs through its function as a sponge that adsorbs miRNA, transcription, protein translation and post-translational modifications. However, in some cases, for example in angiogenesis, the role of MEG3 is still unclear, as current studies provide paradoxical results that require further detailed investigation. It is also noteworthy that a recent study showed that MEG3 could promote HCC cell senescence by sponging miR-16-5p, leading to the decrease in vestigial like family member 4 (VGLL4), which is a tumor suppressor and transcriptional cofactor, while increasing the levels of senescence-related markers p21 and p16 [143].
Hence, while more detailed studies are still needed to investigate whether MEG3 could regulate other hallmarks of cancer, such as avoiding immune destruction, genome instability and mutation, non-mutational epigenetic reprogramming, unlocking phenotypic plasticity and polymorphic microbiomes and whether there are exceptions for its tumor suppressive effects in certain hallmarks of cancer, present results demonstrate the tumor suppressive function of MEG3. Furthermore, although detailed investigations are still needed, MEG3 is a potential diagnostic biomarker and anti-tumor therapeutic target.

Author Contributions

Conceptualization, V.K. and S.W.; investigation, L.Z. and F.Z.; writing—original draft preparation, L.Z. and F.Z.; writing—review and editing, V.K. and S.W.; visualization, W.L. and G.S.; supervision, V.K. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Natural Science Foundation of China (No. 11832008, No. 81872273, No. 31871367, and No. 32070715).

Acknowledgments

Our intention is to summarize the state of art. However, due to space limitations, we would like to apologize to authors whose works are not cited here. Their contributions should not be consid-ered less important than those that are cited.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ABCG2: ATP binding cassette subfamily G member 2; AIFM2: apoptosis-inducing factor mitochondrion-associated 2; ALG9: alpha-1,2-mannosyltransferase; APAF-1: apoptotic protease activating factor-1; ATG3: autophagy related 3; ATG5: autophagy related 5; ATF6: activating transcription factor 6; ATG12: autophagy related 12; Bax: Bcl-2-associated X; Bcl-2: B-cell lymphoma 2; bFGF: basic fibroblast growth factor; Bmi1: B lymphoma Mo-MLV insertion region 1; BRI3: brain protein I3; BTG2: B-cell translocation gene; ceRNA: competitive endogenous RNA; CHOP: C/EBP-homologous protein; CLU: Clusterin; CNV: copy number variation; CRC: colorectal cancer; DDP: cisplatin; DNMT1: DNA methyltransferase 1; DRE: distal regulatory region; EGFR: epithelial growth factor receptor; EMT: epithelial-to-mesenchymal transition; EN-2: Engrailed-2; ER: endoplasmic reticulum; EZH2: enhancer of zeste homologue 2; FoxM1: forkhead box M1; FOXO1: forkhead box O1; GRP78: glucose-regulated protein 78; GSDMD: gasdermin D; GSK: glycogen synthase kinase; Gtl2: gene trap loci 2; HDAC1: histone deacetylase 1; HK2: hexokinase 2; HMGA2: high mobility group AT-hook 2; H3K27me3: trimethylation of histone H3 lysine 27; JARID2: Jumonji and AT-rich interaction domain containing 2; KPS: karnofsky performance score; LATS2: large tumor suppressor 2; LC3 II: microtubule-associated protein light chain 3 II; LDHA: lactate dehydrogenase A; LncRNAs: long non-coding RNAs; MCL-1: myeloid cell leukemia-1; MDM2: murine double minute 2; MDR1: multidrug resistance protein 1; MEG3: maternally expressed gene 3; miRNA: microRNA; MMP-2: matrix metalloproteinase 2; MMP-3: matrix metalloproteinase 3; MMP-9: matrix metalloproteinase 9; MREs: microRNA response elements; MRP1: multidrug resistance-associated protein-1; NF-κB: nuclear factor-κB; OTUD4: OTU deubiquitinase 4; OS: overall survival; PBLD: phenazine biosynthesis-like domain-containing; PCNA: proliferating cell nuclear antigen; PD-1: programmed death-1; PDCD4: programmed cell death 4; PERK: protein kinase R-like endoplasmic reticulum kinase; PGF: placental growth factor; PHLPP2: PH domain and leucine-rich repeat protein phosphatase 2; PKD1: protein kinase D1; PKM2: pyruvate kinase muscle isozyme M2; PRC2: polycomb repressive complex 2; PSAT1: phosphoserine aminotransferase 1; p-STAT3: phosphorylated signal transducer and activator of transcription 3; PTX: paclitaxel; PTEN: phosphatase and tensin homolog; Rac1: Rac family small GTPase 1; RASL11B: RAS like family 11 member B; RBMS3: RNA binding motif single stranded interacting protein 3; SDF-1: stromal cell-derived factor-1; SDH: succinate dehydrogenase; SOCS3: cytokine signaling 3; sORFs: short open reading frames; SOCS5: suppressor of cytokine signaling 5; SOCS6: suppressor of cytokine signaling 6; SOX7: SRY-box transcription factor 7; ST3Gal1: β-galactoside α-2,3-sialyltransferase 1; TGFBR1: transforming growth factor beta receptor 1; TGF-β1: transforming growth factor β1; RNF2: RING finger protein 2; TPM1: tropomyosin 1; TRAF3: TNF receptor-associated factor 3; VEGFA: vascular endothelial growth factor A; VEGFB: vascular endothelial growth factor B.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef] [PubMed]
  3. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  5. Pitolli, C.; Wang, Y.; Mancini, M.; Shi, Y.; Melino, G.; Amelio, I. Do Mutations Turn p53 into an Oncogene? Int. J. Mol. Sci. 2019, 20, 6241. [Google Scholar] [CrossRef] [Green Version]
  6. Han, M.; Jia, L.; Lv, W.; Wang, L.; Cui, W. Epigenetic Enzyme Mutations: Role in Tumorigenesis and Molecular Inhibitors. Front. Oncol. 2019, 9, 194. [Google Scholar] [CrossRef]
  7. Zhang, X.; Wang, W.; Zhu, W.; Dong, J.; Cheng, Y.; Yin, Z.; Shen, F. Mechanisms and Functions of Long Non-Coding RNAs at Multiple Regulatory Levels. Int. J. Mol. Sci. 2019, 20, 5573. [Google Scholar] [CrossRef] [Green Version]
  8. Lin, C.; Yang, L. Long Noncoding RNA in Cancer: Wiring Signaling Circuitry. Trends Cell Biol. 2018, 28, 287–301. [Google Scholar] [CrossRef]
  9. Chen, S.; Zhang, Y.; Ding, X.; Li, W. Identification of lncRNA/circRNA-miRNA-mRNA ceRNA Network as Biomarkers for Hepatocellular Carcinoma. Front. Genet. 2022, 13, 838869. [Google Scholar] [CrossRef]
  10. Yang, G.; Lu, X.; Yuan, L. LncRNA: A link between RNA and cancer. Biochim. Biophys. Acta 2014, 1839, 1097–1109. [Google Scholar] [CrossRef]
  11. Hartford, C.C.R.; Lal, A. When Long Noncoding Becomes Protein Coding. Mol. Cell. Biol. 2020, 40, e00528-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  13. Abdi, E.; Latifi-Navid, S.; Latifi-Navid, H. Long noncoding RNA polymorphisms and colorectal cancer risk: Progression and future perspectives. Environ. Mol. Mutagen. 2022, 63, 98–112. [Google Scholar] [CrossRef]
  14. Tian, J.; Luo, B. Identification of Three Prognosis-Related Differentially Expressed lncRNAs Driven by Copy Number Variation in Thyroid Cancer. J. Immunol. Res. 2022, 2022, 9203796. [Google Scholar] [CrossRef] [PubMed]
  15. Zamaraev, A.V.; Volik, P.I.; Sukhikh, G.T.; Kopeina, G.S.; Zhivotovsky, B. Long non-coding RNAs: A view to kill ovarian cancer. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188584. [Google Scholar] [CrossRef]
  16. Schmitt, A.M.; Chang, H.Y. Long Noncoding RNAs in Cancer Pathways. Cancer Cell 2016, 29, 452–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhou, Y.; Zhang, X.; Klibanski, A. MEG3 noncoding RNA: A tumor suppressor. J. Mol. Endocrinol. 2012, 48, R45–R53. [Google Scholar] [CrossRef]
  18. Schuster-Gossler, K.; Bilinski, P.; Sado, T.; Ferguson-Smith, A.; Gossler, A. The mouse Gtl2 gene is differentially expressed during embryonic development, encodes multiple alternatively spliced transcripts, and may act as an RNA. Dev. Dyn. 1998, 212, 214–228. [Google Scholar] [CrossRef]
  19. Miyoshi, N.; Wagatsuma, H.; Wakana, S.; Shiroishi, T.; Nomura, M.; Aisaka, K.; Kohda, T.; Surani, M.A.; Kaneko-Ishino, T.; Ishino, F. Identification of an imprinted gene, Meg3/Gtl2 and its human homologue MEG3, first mapped on mouse distal chromosome 12 and human chromosome 14q. Genes Cells 2000, 5, 211–220. [Google Scholar] [CrossRef]
  20. Kagami, M.; O’Sullivan, M.J.; Green, A.J.; Watabe, Y.; Arisaka, O.; Masawa, N.; Matsuoka, K.; Fukami, M.; Matsubara, K.; Kato, F.; et al. The IG-DMR and the MEG3-DMR at human chromosome 14q32.2: Hierarchical interaction and distinct functional properties as imprinting control centers. PLoS Genet. 2010, 6, e1000992. [Google Scholar] [CrossRef]
  21. Sherpa, C.; Rausch, J.W.; Le Grice, S.F. Structural characterization of maternally expressed gene 3 RNA reveals conserved motifs and potential sites of interaction with polycomb repressive complex 2. Nucleic Acids Res. 2018, 46, 10432–10447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Li, R.; Fang, L.; Pu, Q.; Bu, H.; Zhu, P.; Chen, Z.; Yu, M.; Li, X.; Weiland, T.; Bansal, A.; et al. MEG3-4 is a miRNA decoy that regulates IL-1beta abundance to initiate and then limit inflammation to prevent sepsis during lung infection. Sci. Signal. 2018, 11, eaao2387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Azam, S.; Hou, S.; Zhu, B.; Wang, W.; Hao, T.; Bu, X.; Khan, M.; Lei, H. Nuclear retention element recruits U1 snRNP components to restrain spliced lncRNAs in the nucleus. RNA Biol. 2019, 16, 1001–1009. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Liu, X.; Bai, X.; Lin, Y.; Li, Z.; Fu, J.; Li, M.; Zhao, T.; Yang, H.; Xu, R.; et al. Melatonin prevents endothelial cell pyroptosis via regulation of long noncoding RNA MEG3/miR-223/NLRP3 axis. J. Pineal. Res. 2018, 64, e12449. [Google Scholar] [CrossRef]
  25. Sathishkumar, C.; Prabu, P.; Mohan, V.; Balasubramanyam, M. Linking a role of lncRNAs (long non-coding RNAs) with insulin resistance, accelerated senescence, and inflammation in patients with type 2 diabetes. Hum. Genom. 2018, 12, 41. [Google Scholar] [CrossRef] [PubMed]
  26. Yan, H.; Rao, J.; Yuan, J.; Gao, L.; Huang, W.; Zhao, L.; Ren, J. Long non-coding RNA MEG3 functions as a competing endogenous RNA to regulate ischemic neuronal death by targeting miR-21/PDCD4 signaling pathway. Cell Death Dis. 2017, 8, 3211. [Google Scholar] [CrossRef]
  27. Ghafouri-Fard, S.; Taheri, M. Maternally expressed gene 3 (MEG3): A tumor suppressor long non coding RNA. Biomed. Pharmacother. 2019, 118, 109129. [Google Scholar] [CrossRef]
  28. Benetatos, L.; Vartholomatos, G.; Hatzimichael, E. MEG3 imprinted gene contribution in tumorigenesis. Int. J. Cancer 2011, 129, 773–779. [Google Scholar] [CrossRef]
  29. Zhang, X.; Zhou, Y.; Mehta, K.R.; Danila, D.C.; Scolavino, S.; Johnson, S.R.; Klibanski, A. A pituitary-derived MEG3 isoform functions as a growth suppressor in tumor cells. J. Clin. Endocrinol. Metab. 2003, 88, 5119–5126. [Google Scholar] [CrossRef] [Green Version]
  30. Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P.P. A ceRNA hypothesis: The Rosetta Stone of a hidden RNA language? Cell 2011, 146, 353–358. [Google Scholar] [CrossRef]
  31. Tay, Y.; Rinn, J.; Pandolfi, P.P. The multilayered complexity of ceRNA crosstalk and competition. Nature 2014, 505, 344–352. [Google Scholar] [CrossRef] [Green Version]
  32. Moradi, M.T.; Fallahi, H.; Rahimi, Z. Interaction of long noncoding RNA MEG3 with miRNAs: A reciprocal regulation. J. Cell. Biochem. 2019, 120, 3339–3352. [Google Scholar] [CrossRef] [PubMed]
  33. Karreth, F.A.; Pandolfi, P.P. ceRNA cross-talk in cancer: When ce-bling rivalries go awry. Cancer Discov. 2013, 3, 1113–1121. [Google Scholar] [CrossRef] [Green Version]
  34. Dan, J.; Wang, J.; Wang, Y.; Zhu, M.; Yang, X.; Peng, Z.; Jiang, H.; Chen, L. LncRNA-MEG3 inhibits proliferation and metastasis by regulating miRNA-21 in gastric cancer. Biomed. Pharmacother. 2018, 99, 931–938. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Z.; Yang, L.; Liu, X.; Nie, Z.; Luo, J. Long noncoding RNA MEG3 inhibits proliferation of chronic myeloid leukemia cells by sponging microRNA21. Biomed. Pharmacother. 2018, 104, 181–192. [Google Scholar] [CrossRef] [PubMed]
  36. Wu, L.; Zhu, L.; Li, Y.; Zheng, Z.; Lin, X.; Yang, C. LncRNA MEG3 promotes melanoma growth, metastasis and formation through modulating miR-21/E-cadherin axis. Cancer Cell Int. 2020, 20, 12. [Google Scholar] [CrossRef]
  37. Lin, L.; Liu, X.; Lv, B. Long non-coding RNA MEG3 promotes autophagy and apoptosis of nasopharyngeal carcinoma cells via PTEN up-regulation by binding to microRNA-21. J. Cell. Mol. Med. 2021, 25, 61–72. [Google Scholar] [CrossRef]
  38. Peng, W.; Si, S.; Zhang, Q.; Li, C.; Zhao, F.; Wang, F.; Yu, J.; Ma, R. Long non-coding RNA MEG3 functions as a competing endogenous RNA to regulate gastric cancer progression. J. Exp. Clin. Cancer Res. 2015, 34, 79. [Google Scholar] [CrossRef] [Green Version]
  39. Shen, X.; Bai, H.; Zhu, H.; Yan, Q.; Yang, Y.; Yu, W.; Shi, Q.; Wang, J.; Li, J.; Chen, L. Long Non-Coding RNA MEG3 Functions as a Competing Endogenous RNA to Regulate HOXA11 Expression by Sponging miR-181a in Multiple Myeloma. Cell. Physiol. Biochem. 2018, 49, 87–100. [Google Scholar] [CrossRef]
  40. Ji, Y.; Feng, G.; Hou, Y.; Yu, Y.; Wang, R.; Yuan, H. Long noncoding RNA MEG3 decreases the growth of head and neck squamous cell carcinoma by regulating the expression of miR-421 and E-cadherin. Cancer Med. 2020, 9, 3954–3963. [Google Scholar] [CrossRef]
  41. Zhu, J.; Han, S. Lidocaine inhibits cervical cancer cell proliferation and induces cell apoptosis by modulating the lncRNA-MEG3/miR-421/BTG1 pathway. Am. J. Transl. Res. 2019, 11, 5404–5416. [Google Scholar]
  42. Zhang, W.; Shi, S.; Jiang, J.; Li, X.; Lu, H.; Ren, F. LncRNA MEG3 inhibits cell epithelial-mesenchymal transition by sponging miR-421 targeting E-cadherin in breast cancer. Biomed. Pharmacother. 2017, 91, 312–319. [Google Scholar] [CrossRef] [PubMed]
  43. Chan, J.J.; Tay, Y. Noncoding RNA:RNA Regulatory Networks in Cancer. Int. J. Mol. Sci. 2018, 19, 1310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhang, X.; Rice, K.; Wang, Y.; Chen, W.; Zhong, Y.; Nakayama, Y.; Zhou, Y.; Klibanski, A. Maternally expressed gene 3 (MEG3) noncoding ribonucleic acid: Isoform structure, expression, and functions. Endocrinology 2010, 151, 939–947. [Google Scholar] [CrossRef] [Green Version]
  45. Uroda, T.; Anastasakou, E.; Rossi, A.; Teulon, J.M.; Pellequer, J.L.; Annibale, P.; Pessey, O.; Inga, A.; Chillon, I.; Marcia, M. Conserved Pseudoknots in lncRNA MEG3 Are Essential for Stimulation of the p53 Pathway. Mol. Cell. 2019, 75, 982–995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Huang, Z.F.; Tang, Y.L.; Shen, Z.L.; Yang, K.Y.; Gao, K. UXT, a novel DNMT3b-binding protein, promotes breast cancer progression via negatively modulating lncRNA MEG3/p53 axis. Mol. Ther. Oncolytics. 2022, 24, 497–506. [Google Scholar] [CrossRef]
  47. Zhu, Y.; Chen, P.; Gao, Y.; Ta, N.; Zhang, Y.; Cai, J.; Zhao, Y.; Liu, S.; Zheng, J. MEG3 Activated by Vitamin D Inhibits Colorectal Cancer Cells Proliferation and Migration via Regulating Clusterin. EBioMedicine 2018, 30, 148–157. [Google Scholar] [CrossRef] [Green Version]
  48. Zhang, J.; Gao, Y. Long non-coding RNA MEG3 inhibits cervical cancer cell growth by promoting degradation of P-STAT3 protein via ubiquitination. Cancer Cell Int. 2019, 19, 175. [Google Scholar] [CrossRef] [Green Version]
  49. Mirzaei, S.; Gholami, M.H.; Hushmandi, K.; Hashemi, F.; Zabolian, A.; Canadas, I.; Zarrabi, A.; Nabavi, N.; Aref, A.R.; Crea, F.; et al. The long and short non-coding RNAs modulating EZH2 signaling in cancer. J. Hematol. Oncol. 2022, 15, 18. [Google Scholar] [CrossRef]
  50. Trotman, J.B.; Braceros, K.C.A.; Cherney, R.E.; Murvin, M.M.; Calabrese, J.M. The control of polycomb repressive complexes by long noncoding RNAs. Wiley Interdiscip. Rev. RNA 2021, 12, e1657. [Google Scholar] [CrossRef]
  51. Terashima, M.; Tange, S.; Ishimura, A.; Suzuki, T. MEG3 Long Noncoding RNA Contributes to the Epigenetic Regulation of Epithelial-Mesenchymal Transition in Lung Cancer Cell Lines. J. Biol. Chem. 2017, 292, 82–99. [Google Scholar] [CrossRef]
  52. Mondal, T.; Subhash, S.; Vaid, R.; Enroth, S.; Uday, S.; Reinius, B.; Mitra, S.; Mohammed, A.; James, A.R.; Hoberg, E.; et al. MEG3 long noncoding RNA regulates the TGF-beta pathway genes through formation of RNA-DNA triplex structures. Nat. Commun. 2015, 6, 7743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Zhou, Y.; Yang, H.; Xia, W.; Cui, L.; Xu, R.; Lu, H.; Xue, D.; Tian, Z.; Ding, T.; Cao, Y.; et al. LncRNA MEG3 inhibits the progression of prostate cancer by facilitating H3K27 trimethylation of EN2 through binding to EZH2. J. Biochem. 2020, 167, 295–301. [Google Scholar] [CrossRef]
  54. Ye, M.; Gao, R.; Chen, S.; Wei, M.; Wang, J.; Zhang, B.; Wu, S.; Xu, Y.; Wu, P.; Chen, X.; et al. Downregulation of MEG3 and upregulation of EZH2 cooperatively promote neuroblastoma progression. J. Cell. Mol. Med. 2022, 26, 2377–2391. [Google Scholar] [CrossRef] [PubMed]
  55. Levine, A.J. p53: 800 million years of evolution and 40 years of discovery. Nat. Rev. Cancer 2020, 20, 471–480. [Google Scholar] [CrossRef] [PubMed]
  56. Zhu, J.; Liu, S.; Ye, F.; Shen, Y.; Tie, Y.; Zhu, J.; Wei, L.; Jin, Y.; Fu, H.; Wu, Y.; et al. Long Noncoding RNA MEG3 Interacts with p53 Protein and Regulates Partial p53 Target Genes in Hepatoma Cells. PLoS ONE 2015, 10, e0139790. [Google Scholar] [CrossRef] [PubMed]
  57. Zhou, Y.; Zhong, Y.; Wang, Y.; Zhang, X.; Batista, D.L.; Gejman, R.; Ansell, P.J.; Zhao, J.; Weng, C.; Klibanski, A. Activation of p53 by MEG3 non-coding RNA. J. Biol. Chem. 2007, 282, 24731–24742. [Google Scholar] [CrossRef] [Green Version]
  58. Lu, K.H.; Li, W.; Liu, X.H.; Sun, M.; Zhang, M.L.; Wu, W.Q.; Xie, W.P.; Hou, Y.Y. Long non-coding RNA MEG3 inhibits NSCLC cells proliferation and induces apoptosis by affecting p53 expression. BMC Cancer 2013, 13, 461. [Google Scholar] [CrossRef] [Green Version]
  59. Zhu, X.; Lv, L.; Wang, M.; Fan, C.; Lu, X.; Jin, M.; Li, S.; Wang, F. DNMT1 facilitates growth of breast cancer by inducing MEG3 hyper-methylation. Cancer Cell Int. 2022, 22, 56. [Google Scholar] [CrossRef]
  60. Gong, X.; Huang, M. Long non-coding RNA MEG3 promotes the proliferation of glioma cells through targeting Wnt/beta-catenin signal pathway. Cancer Gene Ther. 2017, 24, 381–385. [Google Scholar] [CrossRef]
  61. Fan, F.Y.; Deng, R.; Yi, H.; Sun, H.P.; Zeng, Y.; He, G.C.; Su, Y. The inhibitory effect of MEG3/miR-214/AIFM2 axis on the growth of T-cell lymphoblastic lymphoma. Int. J. Oncol. 2017, 51, 316–326. [Google Scholar] [CrossRef] [PubMed]
  62. He, H.; Dai, J.; Zhuo, R.; Zhao, J.; Wang, H.; Sun, F.; Zhu, Y.; Xu, D. Study on the mechanism behind lncRNA MEG3 affecting clear cell renal cell carcinoma by regulating miR-7/RASL11B signaling. J. Cell. Physiol. 2018, 233, 9503–9515. [Google Scholar] [CrossRef] [PubMed]
  63. Wu, X.; Li, J.; Ren, Y.; Zuo, Z.; Ni, S.; Cai, J. MEG3 can affect the proliferation and migration of colorectal cancer cells through regulating miR-376/PRKD1 axis. Am. J. Transl. Res. 2019, 11, 5740–5751. [Google Scholar]
  64. Zhang, Y.Y.; Feng, H.M. MEG3 Suppresses Human Pancreatic Neuroendocrine Tumor Cells Growth and Metastasis by Down-Regulation of Mir-183. Cell. Physiol. Biochem. 2017, 44, 345–356. [Google Scholar] [CrossRef] [Green Version]
  65. Tan, J.; Xiang, L.; Xu, G. LncRNA MEG3 suppresses migration and promotes apoptosis by sponging miR-548d-3p to modulate JAK-STAT pathway in oral squamous cell carcinoma. IUBMB Life 2019, 71, 882–890. [Google Scholar] [CrossRef]
  66. Li, Z.Y.; Yang, L.; Liu, X.J.; Wang, X.Z.; Pan, Y.X.; Luo, J.M. The Long Noncoding RNA MEG3 and its Target miR-147 Regulate JAK/STAT Pathway in Advanced Chronic Myeloid Leukemia. EBioMedicine 2018, 34, 61–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Zhang, J.; Yao, T.; Wang, Y.; Yu, J.; Liu, Y.; Lin, Z. Long noncoding RNA MEG3 is downregulated in cervical cancer and affects cell proliferation and apoptosis by regulating miR-21. Cancer Biol. Ther. 2016, 17, 104–113. [Google Scholar] [CrossRef] [Green Version]
  68. Wang, L.; Yu, M.; Zhao, S. lncRNA MEG3 modified epithelial-mesenchymal transition of ovarian cancer cells by sponging miR-219a-5p and regulating EGFR. J. Cell. Biochem. 2019, 120, 17709–17722. [Google Scholar] [CrossRef]
  69. Qin, N.; Tong, G.F.; Sun, L.W.; Xu, X.L. Long Noncoding RNA MEG3 Suppresses Glioma Cell Proliferation, Migration, and Invasion by Acting as a Competing Endogenous RNA of miR-19a. Oncol. Res. 2017, 25, 1471–1478. [Google Scholar] [CrossRef]
  70. Wu, J.; Pang, R.; Li, M.; Chen, B.; Huang, J.; Zhu, Y. m6A-Induced LncRNA MEG3 Suppresses the Proliferation, Migration and Invasion of Hepatocellular Carcinoma Cell Through miR-544b/BTG2 Signaling. Onco Targets Ther. 2021, 14, 3745–3755. [Google Scholar] [CrossRef]
  71. Huang, C.; Liao, X.; Jin, H.; Xie, F.; Zheng, F.; Li, J.; Zhou, C.; Jiang, G.; Wu, X.R.; Huang, C. MEG3, as a Competing Endogenous RNA, Binds with miR-27a to Promote PHLPP2 Protein Translation and Impairs Bladder Cancer Invasion. Mol. Ther. Nucleic Acids 2019, 16, 51–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zuo, S.; Wu, L.; Wang, Y.; Yuan, X. Long Non-coding RNA MEG3 Activated by Vitamin D Suppresses Glycolysis in Colorectal Cancer via Promoting c-Myc Degradation. Front. Oncol. 2020, 10, 274. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, Q.; Sun, H.; Wang, X.; Yu, X.; Zhang, J.; Guo, B.; Hexige, S. Metabolic changes during malignant transformation in primary cells of oral lichen planus: Succinate accumulation and tumour suppression. J. Cell. Mol. Med. 2020, 24, 1179–1188. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, C.Y.; Yu, M.S.; Li, X.; Zhang, Z.; Han, C.R.; Yan, B. Overexpression of long non-coding RNA MEG3 suppresses breast cancer cell proliferation, invasion, and angiogenesis through AKT pathway. Tumour Biol. 2017, 39, 1010428317701311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, H.; Wang, J.; Lv, S.; Zhang, Y.; Zhang, C.; Lige, B.; Dan, S.; Sun, Y. Long noncoding RNA MEG3 plays a promoting role in the proliferation, invasion, and angiogenesis of lung adenocarcinoma cells through the AKT pathway. J. Cell. Biochem. 2019, 120, 16143–16152. [Google Scholar] [CrossRef] [PubMed]
  76. Buccarelli, M.; Lulli, V.; Giuliani, A.; Signore, M.; Martini, M.; D’Alessandris, Q.G.; Giannetti, S.; Novelli, A.; Ilari, R.; Giurato, G.; et al. Deregulated expression of the imprinted DLK1-DIO3 region in glioblastoma stemlike cells: Tumor suppressor role of lncRNA MEG3. Neuro Oncol. 2020, 22, 1771–1784. [Google Scholar] [CrossRef]
  77. Dong, S.; Ma, M.; Li, M.; Guo, Y.; Zuo, X.; Gu, X.; Zhang, M.; Shi, Y. LncRNA MEG3 regulates breast cancer proliferation and apoptosis through miR-141-3p/RBMS3 axis. Genomics 2021, 113, 1689–1704. [Google Scholar] [CrossRef]
  78. Shan, G.; Tang, T.; Xia, Y.; Qian, H.J. MEG3 interacted with miR-494 to repress bladder cancer progression through targeting PTEN. J. Cell. Physiol. 2020, 235, 1120–1128. [Google Scholar] [CrossRef]
  79. Han, T.; Zhuo, M.; Yuan, C.; Xiao, X.; Cui, J.; Qin, G.; Wang, L.; Jiao, F. Coordinated silencing of the Sp1-mediated long noncoding RNA MEG3 by EZH2 and HDAC3 as a prognostic factor in pancreatic ductal adenocarcinoma. Cancer Biol. Med. 2020, 17, 953–969. [Google Scholar] [CrossRef]
  80. Li, J.; Jiang, X.; Li, C.; Liu, Y.; Kang, P.; Zhong, X.; Cui, Y. LncRNA-MEG3 inhibits cell proliferation and invasion by modulating Bmi1/RNF2 in cholangiocarcinoma. J. Cell. Physiol. 2019, 234, 22947–22959. [Google Scholar] [CrossRef]
  81. Wang, J.; Xu, W.; He, Y.; Xia, Q.; Liu, S. LncRNA MEG3 impacts proliferation, invasion, and migration of ovarian cancer cells through regulating PTEN. Inflamm. Res. 2018, 67, 927–936. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, Y.; Liu, J.; Lv, Y.; Zhang, C.; Guo, S. LncRNA meg3 suppresses hepatocellular carcinoma in vitro and vivo studies. Am. J. Transl. Res. 2019, 11, 4089–4099. [Google Scholar]
  83. Gong, A.; Zhao, X.; Pan, Y.; Qi, Y.; Li, S.; Huang, Y.; Guo, Y.; Qi, X.; Zheng, W.; Jia, L. The lncRNA MEG3 mediates renal cell cancer progression by regulating ST3Gal1 transcription and EGFR sialylation. J. Cell. Sci. 2020, 133, jcs244020. [Google Scholar] [CrossRef]
  84. Zhang, L.; Liang, X.; Li, Y. Long non-coding RNA MEG3 inhibits cell growth of gliomas by targeting miR-93 and inactivating PI3K/AKT pathway. Oncol. Rep. 2017, 38, 2408–2416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Dai, Y.; Wan, Y.; Qiu, M.; Wang, S.; Pan, C.; Wang, Y.; Ou, J. lncRNA MEG3 Suppresses the Tumorigenesis of Hemangioma by Sponging miR-494 and Regulating PTEN/ PI3K/AKT Pathway. Cell. Physiol. Biochem. 2018, 51, 2872–2886. [Google Scholar] [CrossRef] [PubMed]
  86. Peitzsch, C.; Tyutyunnykova, A.; Pantel, K.; Dubrovska, A. Cancer stem cells: The root of tumor recurrence and metastases. Semin. Cancer Biol. 2017, 44, 10–24. [Google Scholar] [CrossRef] [PubMed]
  87. Zhao, Y.; Dong, Q.; Li, J.; Zhang, K.; Qin, J.; Zhao, J.; Sun, Q.; Wang, Z.; Wartmann, T.; Jauch, K.W.; et al. Targeting cancer stem cells and their niche: Perspectives for future therapeutic targets and strategies. Semin. Cancer Biol. 2018, 53, 139–155. [Google Scholar] [CrossRef]
  88. Chen, P.Y.; Hsieh, P.L.; Peng, C.Y.; Liao, Y.W.; Yu, C.H.; Yu, C.C. LncRNA MEG3 inhibits self-renewal and invasion abilities of oral cancer stem cells by sponging miR-421. J. Formos. Med. Assoc. 2021, 120, 1137–1142. [Google Scholar] [CrossRef]
  89. Zhang, S.; Ji, W.W.; Wei, W.; Zhan, L.X.; Huang, X. Long noncoding RNA Meg3 sponges miR-708 to inhibit intestinal tumorigenesis via SOCS3-repressed cancer stem cells growth. Cell Death Dis. 2021, 13, 25. [Google Scholar] [CrossRef]
  90. Carneiro, B.A.; El-Deiry, W.S. Targeting apoptosis in cancer therapy. Nat. Rev. Clin. Oncol. 2020, 17, 395–417. [Google Scholar] [CrossRef]
  91. Roberts, J.Z.; Crawford, N.; Longley, D.B. The role of Ubiquitination in Apoptosis and Necroptosis. Cell Death Differ. 2022, 29, 272–284. [Google Scholar] [CrossRef] [PubMed]
  92. Budihardjo, I.; Oliver, H.; Lutter, M.; Luo, X.; Wang, X. Biochemical pathways of caspase activation during apoptosis. Annu. Rev. Cell. Dev. Biol. 1999, 15, 269–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Luo, G.; Wang, M.; Wu, X.; Tao, D.; Xiao, X.; Wang, L.; Min, F.; Zeng, F.; Jiang, G. Long Non-Coding RNA MEG3 Inhibits Cell Proliferation and Induces Apoptosis in Prostate Cancer. Cell. Physiol. Biochem. 2015, 37, 2209–2220. [Google Scholar] [CrossRef] [PubMed]
  94. Shen, B.; Zhou, N.; Hu, T.; Zhao, W.; Wu, D.; Wang, S. LncRNA MEG3 negatively modified osteosarcoma development through regulation of miR-361-5p and FoxM1. J. Cell. Physiol. 2019, 234, 13464–13480. [Google Scholar] [CrossRef]
  95. Liu, G.; Zhao, X.; Zhou, J.; Cheng, X.; Ye, Z.; Ji, Z. Long non-coding RNA MEG3 suppresses the development of bladder urothelial carcinoma by regulating miR-96 and TPM1. Cancer Biol. Ther. 2018, 19, 1039–1056. [Google Scholar] [CrossRef] [Green Version]
  96. Zhu, D.; Xiao, Z.; Wang, Z.; Hu, B.; Duan, C.; Zhu, Z.; Gao, N.; Zhu, Y.; Wang, H. MEG3/MIR-376B-3P/HMGA2 axis is involved in pituitary tumor invasiveness. J. Neurosurg. 2020, 134, 499–511. [Google Scholar] [CrossRef] [PubMed]
  97. Huang, Z.L.; Chen, R.P.; Zhou, X.T.; Zhan, H.L.; Hu, M.M.; Liu, B.; Wu, G.D.; Wu, L.F. Long non-coding RNA MEG3 induces cell apoptosis in esophageal cancer through endoplasmic reticulum stress. Oncol. Rep. 2017, 37, 3093–3099. [Google Scholar] [CrossRef] [PubMed]
  98. Huang, Y.; Xie, F.; Cheng, X.; Chi, J.; Lei, Y.; Guo, R.; Han, J. LncRNA MEG3 promotes endoplasmic reticulum stress and suppresses proliferation and invasion of colorectal carcinoma cells through the MEG3/miR-103a-3p/PDHB ceRNA pathway. Neoplasma 2021, 68, 362–374. [Google Scholar] [CrossRef]
  99. Zhang, X.; Wu, N.; Wang, J.; Li, Z. LncRNA MEG3 inhibits cell proliferation and induces apoptosis in laryngeal cancer via miR-23a/APAF-1 axis. J. Cell. Mol. Med. 2019, 23, 6708–6719. [Google Scholar] [CrossRef] [Green Version]
  100. Jin, L.; Cai, Q.; Wang, S.; Wang, S.; Mondal, T.; Wang, J.; Quan, Z. Long noncoding RNA MEG3 regulates LATS2 by promoting the ubiquitination of EZH2 and inhibits proliferation and invasion in gallbladder cancer. Cell Death Dis. 2018, 9, 1017. [Google Scholar] [CrossRef] [Green Version]
  101. Bao, D.; Yuan, R.X.; Zhang, Y. Effects of lncRNA MEG3 on proliferation and apoptosis of gallbladder cancer cells through regulating NF-kappaB signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 6632–6638. [Google Scholar] [CrossRef] [PubMed]
  102. Lu, W.X. Long non-coding RNA MEG3 represses cholangiocarcinoma by regulating miR-361-5p/TRAF3 axis. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7356–7368. [Google Scholar] [CrossRef] [PubMed]
  103. Linder, B.; Kogel, D. Autophagy in Cancer Cell Death. Biology 2019, 8, 82. [Google Scholar] [CrossRef] [PubMed]
  104. Lin, L.; Baehrecke, E.H. Autophagy, cell death, and cancer. Mol. Cell. Oncol. 2015, 2, e985913. [Google Scholar] [CrossRef] [Green Version]
  105. Das, G.; Shravage, B.V.; Baehrecke, E.H. Regulation and function of autophagy during cell survival and cell death. Cold Spring Harb. Perspect. Biol. 2012, 4, a008813. [Google Scholar] [CrossRef] [Green Version]
  106. Ye, M.; Lu, H.; Tang, W.; Jing, T.; Chen, S.; Wei, M.; Zhang, J.; Wang, J.; Ma, J.; Ma, D.; et al. Downregulation of MEG3 promotes neuroblastoma development through FOXO1-mediated autophagy and mTOR-mediated epithelial-mesenchymal transition. Int. J. Biol. Sci. 2020, 16, 3050–3061. [Google Scholar] [CrossRef]
  107. Meirson, T.; Gil-Henn, H.; Samson, A.O. Invasion and metastasis: The elusive hallmark of cancer. Oncogene 2020, 39, 2024–2026. [Google Scholar] [CrossRef]
  108. van Zijl, F.; Krupitza, G.; Mikulits, W. Initial steps of metastasis: Cell invasion and endothelial transmigration. Mutat. Res. 2011, 728, 23–34. [Google Scholar] [CrossRef]
  109. Lambert, A.W.; Pattabiraman, D.R.; Weinberg, R.A. Emerging Biological Principles of Metastasis. Cell 2017, 168, 670–691. [Google Scholar] [CrossRef] [Green Version]
  110. Majidpoor, J.; Mortezaee, K. Steps in metastasis: An updated review. Med. Oncol. 2021, 38, 3. [Google Scholar] [CrossRef]
  111. Xu, G.; Meng, L.; Yuan, D.; Li, K.; Zhang, Y.; Dang, C.; Zhu, K. MEG3/miR21 axis affects cell mobility by suppressing epithelialmesenchymal transition in gastric cancer. Oncol. Rep. 2018, 40, 39–48. [Google Scholar] [CrossRef] [Green Version]
  112. Tang, H.; Long, Q.; Zhuang, K.; Yan, Y.; Han, K.; Guo, H.; Lu, X. miR-665 promotes the progression of gastric adenocarcinoma via elevating FAK activation through targeting SOCS3 and is negatively regulated by lncRNA MEG3. J. Cell. Physiol. 2020, 235, 4709–4719. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, D.; Dong, P.; Xiong, Y.; Chen, R.; Konno, Y.; Ihira, K.; Yue, J.; Watari, H. PD-L1 Is a Tumor Suppressor in Aggressive Endometrial Cancer Cells and Its Expression Is Regulated by miR-216a and lncRNA MEG3. Front. Cell. Dev. Biol. 2020, 8, 598205. [Google Scholar] [CrossRef] [PubMed]
  114. Liu, Y.; Xu, Y.; Ding, L.; Yu, L.; Zhang, B.; Wei, D. LncRNA MEG3 suppressed the progression of ovarian cancer via sponging miR-30e-3p and regulating LAMA4 expression. Cancer Cell. Int. 2020, 20, 181. [Google Scholar] [CrossRef] [PubMed]
  115. Li, M.; Liao, H.; Wu, J.; Chen, B.; Pang, R.; Huang, J.; Zhu, Y. Long noncoding RNA matrilineal expression gene 3 inhibits hepatocellular carcinoma progression by targeting microRNA-5195-3p and regulating the expression of forkhead box O1. Bioengineered 2021, 12, 12880–12890. [Google Scholar] [CrossRef]
  116. Li, M.K.; Liu, L.X.; Zhang, W.Y.; Zhan, H.L.; Chen, R.P.; Feng, J.L.; Wu, L.F. Long noncoding RNA MEG3 suppresses epithelialtomesenchymal transition by inhibiting the PSAT1dependent GSK3beta/Snail signaling pathway in esophageal squamous cell carcinoma. Oncol. Rep. 2020, 44, 2130–2142. [Google Scholar] [CrossRef]
  117. Wang, C.; Yan, G.; Zhang, Y.; Jia, X.; Bu, P. Long non-coding RNA MEG3 suppresses migration and invasion of thyroid carcinoma by targeting of Rac1. Neoplasma 2015, 62, 541–549. [Google Scholar] [CrossRef] [Green Version]
  118. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [Green Version]
  119. Burk, D.; Schade, A.L. On respiratory impairment in cancer cells. Science 1956, 124, 270–272. [Google Scholar] [CrossRef]
  120. Pavlova, N.N.; Thompson, C.B. The Emerging Hallmarks of Cancer Metabolism. Cell Metab. 2016, 23, 27–47. [Google Scholar] [CrossRef] [Green Version]
  121. Carmeliet, P.; Jain, R.K. Angiogenesis in cancer and other diseases. Nature 2000, 407, 249–257. [Google Scholar] [CrossRef] [PubMed]
  122. Lugano, R.; Ramachandran, M.; Dimberg, A. Tumor angiogenesis: Causes, consequences, challenges and opportunities. Cell. Mol. Life Sci. 2020, 77, 1745–1770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Xu, X.; Zhong, Z.; Shao, Y.; Yi, Y. Prognostic Value of MEG3 and Its Correlation With Immune Infiltrates in Gliomas. Front. Genet. 2021, 12, 679097. [Google Scholar] [CrossRef]
  124. Wang, W.; Xie, Y.; Chen, F.; Liu, X.; Zhong, L.L.; Wang, H.Q.; Li, Q.C. LncRNA MEG3 acts a biomarker and regulates cell functions by targeting ADAR1 in colorectal cancer. World J. Gastroenterol. 2019, 25, 3972–3984. [Google Scholar] [CrossRef] [PubMed]
  125. Zhao, H.; Wang, X.; Feng, X.; Li, X.; Pan, L.; Liu, J.; Wang, F.; Yuan, Z.; Yang, L.; Yu, J.; et al. Long non-coding RNA MEG3 regulates proliferation, apoptosis, and autophagy and is associated with prognosis in glioma. J. Neurooncol. 2018, 140, 281–288. [Google Scholar] [CrossRef] [PubMed]
  126. Ma, J.; Li, T.F.; Han, X.W.; Yuan, H.F. Downregulated MEG3 contributes to tumour progression and poor prognosis in oesophagal squamous cell carcinoma by interacting with miR-4261, downregulating DKK2 and activating the Wnt/beta-catenin signalling. Artif. Cells Nanomed. Biotechnol. 2019, 47, 1513–1523. [Google Scholar] [CrossRef]
  127. Cui, X.; Yi, Q.; Jing, X.; Huang, Y.; Tian, J.; Long, C.; Xiang, Z.; Liu, J.; Zhang, C.; Tan, B.; et al. Mining Prognostic Significance of MEG3 in Human Breast Cancer Using Bioinformatics Analysis. Cell. Physiol. Biochem. 2018, 50, 41–51. [Google Scholar] [CrossRef]
  128. Gao, Y.; Huang, P.; Zhang, J. Hypermethylation of MEG3 promoter correlates with inactivation of MEG3 and poor prognosis in patients with retinoblastoma. J. Transl. Med. 2017, 15, 268. [Google Scholar] [CrossRef] [Green Version]
  129. Wang, H.; Li, H.; Zhang, L.; Yang, D. Overexpression of MEG3 sensitizes colorectal cancer cells to oxaliplatin through regulation of miR-141/PDCD4 axis. Biomed. Pharmacother. 2018, 106, 1607–1615. [Google Scholar] [CrossRef]
  130. Wang, C.; Tao, X.; Wei, J. Effects of LncRNA MEG3 on immunity and autophagy of non-small cell lung carcinoma through IDO signaling pathway. World J. Surg. Oncol. 2021, 19, 244. [Google Scholar] [CrossRef]
  131. Chen, B.; Dragomir, M.P.; Yang, C.; Li, Q.; Horst, D.; Calin, G.A. Targeting non-coding RNAs to overcome cancer therapy resistance. Signal Transduct. Target. Ther. 2022, 7, 121. [Google Scholar] [CrossRef] [PubMed]
  132. Gao, L.; Wu, Z.X.; Assaraf, Y.G.; Chen, Z.S.; Wang, L. Overcoming anti-cancer drug resistance via restoration of tumor suppressor gene function. Drug Resist. Updat. 2021, 57, 100770. [Google Scholar] [CrossRef]
  133. Buttarelli, M.; De Donato, M.; Raspaglio, G.; Babini, G.; Ciucci, A.; Martinelli, E.; Baccaro, P.; Pasciuto, T.; Fagotti, A.; Scambia, G.; et al. Clinical Value of lncRNA MEG3 in High-Grade Serous Ovarian Cancer. Cancers 2020, 12, 966. [Google Scholar] [CrossRef]
  134. Yan, H.; Luo, B.; Wu, X.; Guan, F.; Yu, X.; Zhao, L.; Ke, X.; Wu, J.; Yuan, J. Cisplatin Induces Pyroptosis via Activation of MEG3/NLRP3/caspase-1/GSDMD Pathway in Triple-Negative Breast Cancer. Int. J. Biol. Sci. 2021, 17, 2606–2621. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, P.; Chen, D.; Ma, H.; Li, Y. LncRNA MEG3 enhances cisplatin sensitivity in non-small cell lung cancer by regulating miR-21-5p/SOX7 axis. Onco. Targets Ther. 2017, 10, 5137–5149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Deng, R.; Fan, F.Y.; Yi, H.; Liu, F.; He, G.C.; Sun, H.P.; Su, Y. MEG3 affects the progression and chemoresistance of T-cell lymphoblastic lymphoma by suppressing epithelial-mesenchymal transition via the PI3K/mTOR pathway. J. Cell. Biochem. 2018, 120, 8144–8153. [Google Scholar] [CrossRef]
  137. Zhu, M.; Wang, F.; Mi, H.; Li, L.; Wang, J.; Han, M.; Gu, Y. Long noncoding RNA MEG3 suppresses cell proliferation, migration and invasion, induces apoptosis and paclitaxel-resistance via miR-4513/PBLD axis in breast cancer cells. Cell Cycle 2020, 19, 3277–3288. [Google Scholar] [CrossRef]
  138. Zhou, X.; Yuan, P.; Liu, Q.; Liu, Z. LncRNA MEG3 Regulates Imatinib Resistance in Chronic Myeloid Leukemia via Suppressing MicroRNA-21. Biomol. Ther. 2017, 25, 490–496. [Google Scholar] [CrossRef] [Green Version]
  139. Yu, Y.; Kou, D.; Liu, B.; Huang, Y.; Li, S.; Qi, Y.; Guo, Y.; Huang, T.; Qi, X.; Jia, L. LncRNA MEG3 contributes to drug resistance in acute myeloid leukemia by positively regulating ALG9 through sponging miR-155. Int. J. Lab. Hematol. 2020, 42, 464–472. [Google Scholar] [CrossRef]
  140. Ma, L.; Wang, F.; Du, C.; Zhang, Z.; Guo, H.; Xie, X.; Gao, H.; Zhuang, Y.; Kornmann, M.; Gao, H.; et al. Long non-coding RNA MEG3 functions as a tumour suppressor and has prognostic predictive value in human pancreatic cancer. Oncol. Rep. 2018, 39, 1132–1140. [Google Scholar] [CrossRef]
  141. Liu, Y.; Yue, P.; Zhou, T.; Zhang, F.; Wang, H.; Chen, X. LncRNA MEG3 enhances (131)I sensitivity in thyroid carcinoma via sponging miR-182. Biomed. Pharmacother. 2018, 105, 1232–1239. [Google Scholar] [CrossRef] [PubMed]
  142. Huang, X.; Wu, W.; Jing, D.; Yang, L.; Guo, H.; Wang, L.; Zhang, W.; Pu, F.; Shao, Z. Engineered exosome as targeted lncRNA MEG3 delivery vehicles for osteosarcoma therapy. J. Control. Release 2022, 343, 107–117. [Google Scholar] [CrossRef] [PubMed]
  143. Tao, Y.; Yue, P.; Miao, Y.; Gao, S.; Wang, B.; Leng, S.X.; Meng, X.; Zhang, H. The lncRNA MEG3/miR-16-5p/VGLL4 regulatory axis is involved in etoposide-induced senescence of tumor cells. J. Gene Med. 2021, 23, e3291. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of DLK1-MEG3 locus on human chromosome 14. The 837 kb-long DLK1-MEG3 locus contains the protein-coding genes DIO3, RTL1, and DLK1. The MEG3 gene has ten exons and is 35 kb long. The IG-DMR is 13 kb upstream of the MEG3 gene. The MEG3-DMR overlaps with the MEG3 promoter. IG-DMR: intergenic differentially methylated region.
Figure 1. Schematic diagram of DLK1-MEG3 locus on human chromosome 14. The 837 kb-long DLK1-MEG3 locus contains the protein-coding genes DIO3, RTL1, and DLK1. The MEG3 gene has ten exons and is 35 kb long. The IG-DMR is 13 kb upstream of the MEG3 gene. The MEG3-DMR overlaps with the MEG3 promoter. IG-DMR: intergenic differentially methylated region.
Cancers 14 06032 g001
Figure 2. MEG3 inhibits cancer progression through different mechanisms. MEG3 is involved in tumor progression in two ways, such as acting as a sponge for miRNA and regulating its targets through transcriptional as well as post-translational regulations.
Figure 2. MEG3 inhibits cancer progression through different mechanisms. MEG3 is involved in tumor progression in two ways, such as acting as a sponge for miRNA and regulating its targets through transcriptional as well as post-translational regulations.
Cancers 14 06032 g002
Figure 3. MEG3 and the hallmarks of cancer. In this review, we mainly focus on five hallmarks of cancer regulated by MEG3, including inhibiting tumor cell proliferation, inducing cell death, reducing invasion and metastasis, reprogramming energy metabolism, and regulating angiogenesis.
Figure 3. MEG3 and the hallmarks of cancer. In this review, we mainly focus on five hallmarks of cancer regulated by MEG3, including inhibiting tumor cell proliferation, inducing cell death, reducing invasion and metastasis, reprogramming energy metabolism, and regulating angiogenesis.
Cancers 14 06032 g003
Table 1. Biological implications of MEG3 on hallmarks of cancer.
Table 1. Biological implications of MEG3 on hallmarks of cancer.
Cancer TypemiRNARelated GenesHallmarksRefs
Breast cancermiR-494-3pOTUD4Growth inhibition[59]
Glioma/Wnt/β-cateninCell cycle regulation[60]
T-cell lymphoblastic lymphomamiR-214AIFM2, Ki-67, PCNAGrowth inhibition[61]
Clear cell renal cell carcinomamiR-7RASL11BGrowth inhibition[62]
CRCmiR-376PKD1Cell cycle regulation[63]
Pancreatic neuroendocrine tumormiR-183BRI3Growth inhibition[64]
OSCCmiR-548d-3pSOCS5, SOCS6Apoptosis induction[65]
CMLmiR-147JAK/STAT3Apoptosis induction[66]
Cervical cancermiR-21-5pp53, caspase3Apoptosis induction[67]
Breast cancermiR-421E-cadherinEMT inhibition[42]
Ovarian cancermiR-219a-5pEGFREMT inhibition[68]
GliomamiR-19aPTENMetastasis inhibition[69]
HCCmiR-544bBTG2Metastasis inhibition[70]
Bladder cancermiR-27aPHLPP2, c-MycMetastasis inhibition[71]
CRC/LDHA, PKM2, HK2Metabolic reprogramming[72]
OSCCmiR-361-5psuccinateMetabolic reprogramming[73]
Breast cancer/VEGFA, PGF, bFGF, TGF-β1, MMP-9, AKTAngiogenesis inhibition[74]
Lung cancer/VEGFA, VEGFB, bFGF, SDF-1, TGF-β, angiogenin, MMP-9Angiogenesis promotion[75]
Abbreviations: CRC: colorectal cancer; OSCC: oral squamous cell carcinoma; CML: chronic myeloid leukemia; HCC: hepatocellular carcinoma; EMT: epithelial-mesenchymal transition.
Table 2. MEG3 expression and relevant clinical characteristics in human cancers.
Table 2. MEG3 expression and relevant clinical characteristics in human cancers.
Cancer TypeExpressionRelevant Clinical CharacteristicsRefs
GliomaDownregulatedOverall survival rates, Advanced WHO grade, Karnofsky performance score, IDH wild-type, tumor recurrence, progression-free survival[125]
ESCCDownregulatedTumor size, lymph node metastasis, poor prognosis[126]
NSCLCDownregulatedSurvival rate[130]
GliomaDownregulatedTumor grade[60]
CRCDownregulatedLymph node metastasis, TNM staging, Overall survival[129]
GlioblastomaDownregulatedSurvival[76]
Breast cancerDownregulatedOverall survival, Relapse-free survival, Distant metastasis-free survival, Disease-specific survival[127]
RetinoblastomaDownregulatedSurvival[128]
GliomaDownregulatedOverall survival, Progression-free survival[123]
Abbreviations: ESCC: esophageal squamous cell carcinoma; NSCLC: non-small cell lung carcinoma; CRC: colorectal cancer.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, L.; Zhao, F.; Li, W.; Song, G.; Kasim, V.; Wu, S. The Biological Roles and Molecular Mechanisms of Long Non-Coding RNA MEG3 in the Hallmarks of Cancer. Cancers 2022, 14, 6032. https://doi.org/10.3390/cancers14246032

AMA Style

Zhang L, Zhao F, Li W, Song G, Kasim V, Wu S. The Biological Roles and Molecular Mechanisms of Long Non-Coding RNA MEG3 in the Hallmarks of Cancer. Cancers. 2022; 14(24):6032. https://doi.org/10.3390/cancers14246032

Chicago/Turabian Style

Zhang, Lei, Fuqiang Zhao, Wenfang Li, Guanbin Song, Vivi Kasim, and Shourong Wu. 2022. "The Biological Roles and Molecular Mechanisms of Long Non-Coding RNA MEG3 in the Hallmarks of Cancer" Cancers 14, no. 24: 6032. https://doi.org/10.3390/cancers14246032

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop