Next Article in Journal
Mitochondrial Phenotypes in Parkinson’s Diseases—A Focus on Human iPSC-Derived Dopaminergic Neurons
Previous Article in Journal
MXD3 Promotes Obesity and the Androgen Receptor Signaling Pathway in Gender-Disparity Hepatocarcinogenesis
Previous Article in Special Issue
Snail Upregulates Transcription of FN, LEF, COX2, and COL1A1 in Hepatocellular Carcinoma: A General Model Established for Snail to Transactivate Mesenchymal Genes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aspects of the Epigenetic Regulation of EMT Related to Cancer Metastasis

Department of Biotechnology and Genetic Engineering, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia, 40-055 Katowice, Poland
*
Author to whom correspondence should be addressed.
Cells 2021, 10(12), 3435; https://doi.org/10.3390/cells10123435
Submission received: 30 September 2021 / Revised: 28 November 2021 / Accepted: 3 December 2021 / Published: 6 December 2021
(This article belongs to the Special Issue Latest Research on Epithelial-Mesenchymal Transition (EMT))

Abstract

:
Epithelial to mesenchymal transition (EMT) occurs during the pathological process associated with tumor progression and is considered to influence and promote the metastatic cascade. Characterized by loss of cell adhesion and apex base polarity, EMT enhances cell motility and metastasis. The key markers of the epithelial to mesenchymal transition are proteins characteristic of the epithelial phenotype, e.g., E-cadherin, cytokeratins, occludin, or desmoplakin, the concentration and activity of which are reduced during this process. On the other hand, as a result of acquiring the characteristics of mesenchymal cells, an increased amount of N-cadherin, vimentin, fibronectin, or vitronectin is observed. Importantly, epithelial cells undergo partial EMT where some of the cells show both epithelial and mesenchymal characteristics. The significant influence of epigenetic regulatory mechanisms is observed in the gene expression involved in EMT. Among the epigenetic modifications accompanying incorrect genetic reprogramming in cancer are changes in the level of DNA methylation within the CpG islands and posttranslational covalent changes of histone proteins. All observed modifications, which are stable but reversible changes, affect the level of gene expression leading to the development and progression of the disease, and consequently affect the uncontrolled growth of the population of cancer cells.

1. Epithelial to Mesenchymal Transition

Epithelial to mesenchymal transition (EMT) is primarily involved in embryogenesis and is associated with adult tissue regeneration, wound healing, and fibrosis. During biological processes, EMT is a highly regulated process, while the other type of EMT occurring during tumor progression is deregulated and transient [1]. During transition of epithelial cells, loss of apical and basolateral polarity, disruption of cell adhesion, remodeling of the cytoskeleton, and changes in cell to matrix adhesion are observed, and cells become migratory mesenchymal cells [2]. All of these processes are associated with changes in the expression of genes encoding proteins characterizing a specific cellular phenotype. These changes relate, for example, to genes encoding extracellular matrix components; cytoskeletal genes mediating cell adhesion, migration, motility, and morphogenesis; genes controlling cell differentiation, development, growth, and proliferation; as well as signal transduction and transcription factor genes that cause EMT and all of its associated processes [3,4]. The main characteristics of epithelial cells are gradually lost, and partial EMT occurs where some cells show both epithelial and mesenchymal characteristics. Those partial EMT states are more relevant to cancer metastasis than the defined mesenchymal phenotype, and the number of potential intermediate EMT states within a tumor can be an indicator of malignancy [5]. There is evidence that not all cells in the primary tumor change morphology towards the mesenchymal state and express EMT-TFs. It is important that malignization, understood as ability of tumor cells to disseminate, is not universal for all cells in a tumor. Some studies suggest that EMT is required for metastasis, while other studies suggest that EMT may be dispensable or irrelevant for metastasis. However, the effect of EMT on the tumorigenesis of different cancers including prostate, lung, liver, pancreatic, and breast cancers has been demonstrated [5,6].
EMT is reversible by the mesenchymal to epithelial transition (MET), which is thought to affect circulating cancer cells when they reach a desirable metastatic niche and settle down to proliferate and develop secondary tumors [7].

2. Cancer Metastasis

Cancer metastasis requires multiple steps, including local invasion, intravasation, and colonization of distant organs. Cytoskeletal rearrangements within the cancer cells, combined with the action of adhesive interactions, secreted extracellular matrix metalloproteinases (MMPs), and cathepsins, drive cancer cell invasion and migration through the stroma. Cancer cells may migrate as single cells boring a path through the extracellular matrix, move along collagen fibers, or migrate collectively as ensembles that forge ahead from the tumor invasion front [8].
Cancer cells disseminate from tumors by invading blood vessels and the lymphatic system. Hematogenous dissemination is the main form of metastatic spread to organs. Circulating tumor cells (CTCs) in the blood circulation express progenitor and epithelial to mesenchymal transition markers, suggesting that these cells are the beginning of metastatic tumors [4,9].
Genomic instability is the cause of tumor initiation and progression and occurs in a metastatically competent subclone of a primary tumor. Genomic changes are reflected in many cellular phenotypes, any one of which may have all the necessary properties to complete the metastatic process. The most predominant genes that are changed in metastasis include tumor protein p53 (TP53), cyclin-dependent kinase inhibitor 2A (CDKN2A), phosphatase and tensin homolog (PTEN), phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA), and retinoblastoma (RB1) [4].

3. Epigenetic Regulation of EMT Biomarkers

Molecular mechanisms governing EMT leading to a downregulation of epithelial markers and an increase in mesenchymal markers and EMT-linked transcription factors can depend on epigenetic modifications, and in certain cases are tissue-specific. Detailed literature data are provided in the Table 1. DNA methylation is one of the best characterized mechanisms of epigenetic gene regulation and may be involved in regulating EMT [10].

3.1. Epigenetic Changes of Genes Mediating Cell Adhesion, Migration, and Motility

The E-cadherin (CDH1 gene encodes a classical cadherin of the cadherin superfamily, epithelial calcium-dependent adhesion protein, a component of the adherent junctions between epithelial cells with tumor suppressor activity) gene promoter, which possesses several regulatory sequences including three E-boxes that mediate CDH1 transcriptional repression in mesenchymal cells, is transcriptionally and epigenetically regulated during EMT [11]. Besides the promoter, there are seven regions in the CDH1 gene with regulatory functions, and six of them are located in intron 2, which has been previously identified as a cis-regulatory element with a pivotal role in controlling the gene’s expression [35]. In addition, the methylation of CpG sites located in the CDH1 enhancers correlates with the low gene expression [35].
CDH1 transcription may directly or indirectly depend on expression control of miRNAs and long noncoding RNAs, due to the number of pathways regulating CDH1 expression through the activity of Snail, Slug, ZEB1/2, EZH2, and Twist transcription factors [36]. miR-101 has been reported to act as a tumor suppressor by targeting CDH1 inhibitors, such as ZEB1/ZEB2 and EZH2 in different tumors. miR-200 family members are known as transcriptional repressors of E-cadherin by targeting ZEB1/ZEB2, meanwhile elevating expression of β-catenin and relocating mostly membrane-associated β-catenin to nucleus, thus activating the Wnt/β-catenin signaling pathway. Aberrant activation of the Wnt/β-catenin signaling pathway promotes cell proliferation, metastasis, and tumorigenesis [37].
N-cadherin (CDH2 gene encodes a classical cadherin and member of the cadherin superfamily, neuronal calcium-dependent adhesion protein) is the transmembrane protein that mediates cell adhesion and promotes cell migration and invasion, which makes it an important factor in the epithelial to mesenchymal transition (EMT) [38]. A defined cell–cell junction staining of N-cadherin was visible after transfection with miR-338-3p, which is epigenetically silenced in gastric cancer. The Ras related protein (Rab-14) and Hedgehog acyltransferase (Hhat) are direct targets of miR-338-3p and enforced expression of miR-338-3p, and small interfering RNA induced Rab14-mediated accumulation of N-cadherin in the cell junctions [39].
Fibronectin (FN1 gene encodes fibronectin, a glycoprotein present in a soluble dimeric form in plasma, and in a dimeric or multimeric form at the cell surface and in extracellular matrix) is widely expressed by several cell types and is involved in the migration, proliferation, and adhesion of cells, and in changes of the extracellular matrix through signaling with integrins in several types of cancers [40]. Upregulation of FN1 may depend on miR-9-3p downregulation in cancer cells. The antitumor role of miR-9-3p including suppressing cell proliferation, migration, and invasion via downregulating FN1 (among others) and inactivating the EMT process was confirmed in nasopharyngeal carcinoma [16].
Matrix Metallopeptidase 2 (MMP2) is a member of the matrix metalloproteinase gene family, the zinc dependent enzymes capable of cleaving components of the extracellular matrix and molecules involved in signal transduction. Different statuses of epigenetic changes within the MMP2 promoter were observed in various neoplasms. MMP2 promoter regions are hypermethylated in non-migratory breast cancer cells and hypomethylated in the highly migratory glioma cells [17,41].
For Matrix Metallopeptidase 9 (MMP9) the enzyme encoded by this gene degrades type IV and V collagens and is involved in the breakdown of the extracellular matrix during arthritis or metastasis. Changes in the degree of methylation, both within CpG DNA islands and histone proteins, have been observed in cancer with different expression levels of MMP9. The extensive CpG islands present in the coding region are epigenetically repressed in glioma cells. In breast carcinoma cells, the elevated expression of MMP9 correlated with the presence of H3K4me3 methylation [41]. DNA promoter methylation is important for the regulation of MMP9 expression, and H3K4me3 is high in the promoter DNA regions, showing the importance of these binding sites in MMP9 transcription in breast cancer cell lines [19]. Hypermethylation at the CpG-2 intragenic region is observed with higher MMP9 expression levels, and the MMP9 intragenic methylation hotspot may be responsible for the MMP9 overexpression in melanoma cell lines [42]. Additionally, miR-211 acted as a tumor suppressor and inhibited EMT by targeting MMP9 expression and may be a potential target of gastric cancer treatment [20]. Importantly, MMP9 is one of the downstream target genes of SMYD3 (SET and MYND domain containing protein 3), the histone H3-K4 di-/tri- and H4-K5-methyltransferase, and overexpression of SMYD3 can induce MMP9 expression in human tumor models, e.g., gastric cancer [43].
Desmoplakin (DSP) encodes a protein that anchors intermediate filaments to the desmosomal plaques and is involved in the organization of the desmosomal cadherin plakoglobin complexes. Loss of desmoplakin as a tumor suppressor was reportedly attributed to invasive tumorigenic potential. Decreased expression of DSP in a majority of lung cancer cell lines and primary lung tumors caused by methylation in the promoter region and in intron 1 indicated that epigenetic regulation is responsible for gene silencing [21].
Collagens (fibrillar forming collagen COL1A1 encodes the proalpha 1 chains of type I collagen whose triple helix comprises two alpha 1 chains and one alpha 2 chain), found in most connective tissues, are involved in tumor invasion and progression. Collagen type I is composed of three polypeptide chains transcribed from COL1A1 and COL1A2 gene. Epigenetic alterations of collagen genes have been reported in various neoplasms, and each gene is methylated in several human cancer cells with coordinately decreased collagen expression. Although COL1A1 mRNA is usually upregulated in tumor tissues, there is a small group of tumors that has downregulated mRNA expression, mainly due to promoter methylation. These downregulated cases were correlated with poor overall survival in patients with hepatocellular carcinoma [23,44].

3.2. Epigenetic Changes of Genes Controlling Cell Differentiation and Proliferation

Keratins (KRTs) are a family of intermediate filament proteins that are responsible for the structural integrity of epithelial cells and play significant roles in the occurrence, progression, and metastasis of multiple cancers [45]. KRT8 and KRT19 are two oncogenes important in development of human cancers, with high expression and hypomethylated promoters [29].
KRT5, KRT14, and KRT17 overexpression is related to poor survival across different types of cancers. H3K27me3 promotes KRT14 gene expression by altering the recruitment pattern of its transcription factor Sp1 [25]. Epigenetic silencing of KRT13 genes might be useful for the development of diagnostic markers and novel therapeutic approaches [24]. KRT15 is associated with tumorigenesis and has different expression characteristics across different types of cancer [46]. Decreased expression of KRT18 may be dependent on DNA hypermethylation and may serve as a biomarker for the diagnosis of metastasis [26]. The CpGs in the promoter of the KRT19 gene and the methylation levels of KRT19 are significantly inversely associated with gene expression, which suggests that DNA methylation is a potential epigenetic regulator in malignancy [28]. Decreased methylation correlates with increased KRT23 transcript expression, which is inducible with the demethylating agent (5-Aza-2′-deoxycytidine), suggesting a potential epigenetic regulation for KRT23 [30].
Occludin (the OCLN gene) encodes an integral membrane protein that is required for cytokine-induced regulation of the tight junctions (TJs) (paracellular permeability membrane). The hallmark of epithelial cells and their subsequent loss from cancer cells tends to be considered as a direct effect of epithelial to mesenchymal transition (EMT). OCLN can be partially methylated in the promoter sequence, and miR-122 has been reported to regulate gene expression [31].
Vimentin (the VIM gene encodes a type III intermediate filament protein, which is responsible for maintaining shape and integrity of the cytoplasm and stabilizing cytoskeletal interactions) plays a significant role in adhesion, migration, and signaling and is considered as a classic mesenchymal cell marker. VIM may be used for the early detection of several cancers and hypermethylation of the promoter is important for the transcriptional silencing of the gene in cancer cells [32]. On the other hand, increased levels of vimentin in the serum may be caused by hypomethylation of the vimentin gene and increased expression of the vimentin gene intracellularly, leading to secretion of high levels of soluble vimentin in the serum [34]. Additionally, varying degrees of methylation within a gene’s CpG islands leads to various levels of gene silencing, and DNA methylation is considered as an epigenetic biomarker for cancer detection [47]. VIM methylation seems to be a possible biomarker for the detection of tumor DNA in the serum of different colorectal carcinoma patients [33].

3.3. Transcription Factor Genes

EMT can be attained through cooperative functioning of different transcription factors such as SNAI1/2, TWIST, and ZEB1/2 with others (FOX, SOX, PRX, and HMGA2), leading to cell-forward consensus differentiations that make tumor cells capable of motility and metastasis. SNAIL1 (Snail) and SNAIL2 (Slug) activation downregulates the expression of genes involved in tight junctions (occludin and claudin 1), apical polarity (Crumbs Cell Polarity Complex Component 3, CRB3), and other cell adhesions such as E-cadherin, but upregulates the expression of N-cadherin, fibronectin, vimentin, and MMPs through different signaling pathways. Expression of E-cadherin, β-catenin, and γ-catenin is masked by TWIST1, but the expression of vimentin is increasing. Importantly, even a strong EMT-inducing transcription factor operates differentially in each cell type belonging to a different EMT status [48,49].
SNAI1 (Snail family transcriptional repressor 1) is the member of the SNAIL family that is the best studied EMT transcription factor and is predominantly expressed in cell lines with an epithelial phenotype. It is reported that SNAI1 expression at the onset of EMT and other EMT factors are induced at later time points to strengthen a mesenchymal state. Induction of SNAI1 leads to repression of SNAI2 and simultaneous induction of ZEB1/2 transcripts at an early stage, highlighting that execution of this functional circuit preceded other molecular and phenotypical changes that can lead to EMT. Epigenetic repression of SNAI1 on its target genes has been shown to be carried out through recruitment of HDAC1/2 and corepressor SIN3A on its SNAG domain [50]. Overexpression of SNAI1 is a pathologically relevant event in human acute myeloid leukemia (AML) that contributes to impaired differentiation and enhanced proliferation of immature myeloid cells, and ectopic expression of SNAI1 in hematopoietic cells predisposes to AML development, which is mediated by the interaction with the histone lysine-specific demethylase 1A KDM1A/LSD1 [51]. Suppression of SNAIL1 mediated gene reprogramming by SETDB1 occurs through increased H3K9 methylation at the SNAI1 gene promoter that represses its H3K9 acetylation imposed by activated Smad3/4 complexes. SETDB1 regulates the activity of Smad3 at the SNAI1 gene through a balance between histone methylation and acetylation [52].
ZEB1 (Zinc finger E-box binding homeobox 1) acts as a transcriptional repressor that inhibits interleukin-2 gene expression and represses the E-cadherin promoter, which induces an epithelial to mesenchymal transition (EMT) by recruiting SMARCA4/BRG1 [53]. ZEB1 recruits the histone deacetylase HDAC1 or the methyltransferase DNMT1 to the CDH1 promoter to maintain the hypermethylation status and inhibit the transcription. The chromatin modifier SETD1B, a histone methyltransferase of Lys4 at histone H3, is a direct transcriptional target of ZEB1, which regulates the expression pattern of the SETD1B in colorectal cancer cells. ZEB1 binds the promoter region of the SETD1B gene and SETD1B-dependent active H3K4me3 histone code seems to open the ZEB1 promoter, forming a positive feedback loop [54]. The characteristic epigenetic regulation of plasticity in lung cancer cells is the interaction between the miR-200/ZEB axis in non-small-cell lung cancer. The miR-200 DNA promoters were hypermethylated in the human lung adenocarcinoma PC9 cell line in the TGF-β1-induced EMT process. Treatment with 5-aza-2′-deoxycytidine reversed hypermethylation, and ZEB1/2 expression was restored, leading to reversal of EMT [55]. Further epigenetic regulation of ZEB1 depends on the protein arginine methyltransferases PRMT1, which induce the EMT process in breast cancer cells by mediating the asymmetric dimethylation of arginine 3 of histone H4 (H4R3me2as) at the ZEB1 promoter and activating its transcription [56]. UTX (Ubiquitously Transcribed Tetratricopeptide Repeat on chromosome X) is identified as a histone demethylase that specifically targets di- and tri-methyl groups on lysine 27 of histone H3 (H3K27me2/3), which was shown to epigenetically induce the expression of ZEB1 in brain metastases during the metastasis of lung adenocarcinoma to the brain [57].
TWIST1 (Twist family bHLH transcription factor 1) is a basic helix−loop−helix (bHLH) transcription factor that plays an important role in embryonic development.
TWIST has been long considered as a transcription repressor of the EMT, and increasing the expression of TWIST1 is directly associated with tumor invasion and metastasis and mediates the loss of E-cadherin, a key epithelial marker, by binding to the promoter and blocking its transcription [58]. TWIST is responsible for inducing the formation of actin-rich invadopodia, which recruit MMP7, MMP9, and MMP14 to the leading edge where they degrade ECM and basement membranes, facilitating tumor invasion and metastasis. TWIST1 expression can be activated in response to the methyltransferase MMSET (nuclear receptor SET domain, NSD protein), leading to an increase in H3K36me2 and H3K27me3 methylation and suggesting a direct role of MMSET in the regulation of TWIST1 expression, in turn leading to an increase in migration and invasion, and changes in cell morphology and gene expression consistent with an epithelial to mesenchymal transition (EMT) [59]. The lysine acetyltransferase TIP60/KAT5 mediates K73/76 diacetylation of the TWIST motif function as diacetylation in the histone H4 mimic GK-X-GK motif recruits BRD4 (a member of the BET, bromodomain and extra-terminal domain) protein and related transcriptional components to the super-enhancer of its targeted genes during progression of basal-like breast cancer [58]. TWIST1 is upregulated and FOXO3a (the forkhead box class O 3a, which belongs to the FOXO subclass of forkhead transcription factors) is downregulated in breast cancer tissues, and overexpression of TWIST1 can reverse the effect of FOXO3a on proliferation, invasion, migration, and EMT. Additionally, the overexpression of FOXO3a or knockdown of TWIST1 suppresses the proliferation, invasion, migration, and EMT, and FOXO3a suppresses the growth and metastasis of breast cancer by targeting TWIST1 in vivo [60]. EZH2 (Enhancer of zeste homolog 2), a histone methyltransferase and a catalytic component of PRC2, catalyzes tri-methylation of histone H3 at Lys 27 (H3K27me3), which was shown to epigenetically suppress the expression of TWIST in brain metastases during the metastasis of lung adenocarcinoma to the brain. However, the role of EZH2 has not been fully established in lung adenocarcinoma. The major role of EZH2 in oncogenesis is to engage in the proliferation and development rather than the EMT process of the cancer. EZH2 is an interesting target for anticancer therapy because of upregulation in multiple cancers [57].
FOX. Deregulation of FOX factors is correlated with carcinogenesis and tumor progression. The FOXA factors clustered together with epithelial genes that are involved in the formation of adherens and tight junctions and the loss of FOXA1, 2, 3 impair the structure and function of enhancers at the tumor suppressor genes CDH1, EPHB3, and CDX2, which are integral to the differentiation and tissue integrity of the intestinal epithelium. Conversely, expression of FOXA factors in cells with inactive CDX2 and EPHB3 enhancers led to chromatin opening and de novo deposition of the H3K4me1 and H3K27ac marks. In cellular EMT models, ectopically expressed SNAIL1 directly represses FOXA1 and triggers downregulation of all FOXA family members, suggesting that loss of FOXA expression promotes EMT [61]. FOXF2 is specifically expressed in most basal-like breast cells, and the deficiency enhances metastatic ability and induces epithelial to mesenchymal transition (EMT) of cells by upregulating the transcription of TWIST1, which is a transcriptional target of FOXF2 [62].

3.4. Hypoxia-Induced EMT

Due to the metabolic conditions of the cancer, hypoxia is a common phenomenon in most solid tumors. A hypoxic microenvironment triggers varied molecular responses, but importantly, hypoxia and EMT share a number of common signaling pathways. This fact prompted researchers to even distinguish the term hypoxia-induced EMT. Among other stress-regulated factors, hypoxia-inducible factors (HIFs) are known to be the master regulators of gene expression under oxygen-deprived conditions. Transcriptional regulation by HIF proteins is mediated by an HIF complex consisting of oxygen-dependent HIFα and is expressed constitutively as HIFβ subunits, which bind to the hypoxia-responsive element (HRE) present in the promoter region of target genes [63]. HIF1 provokes EMT through upregulating transcription factors or repressors, activating the EMT-associated signaling pathways, modulating EMT-associated inflammatory cytokines, and regulating other pathways such as epigenetic regulators. It has been shown that HIF1 induces EMT through the transcriptional control of E-cadherin, SNAIL, ZEB1, and TWIST [64]. Hypoxia affects EMT through regulating signaling pathways, modulating expression and signaling of transcription factors, and regulating miRNA and lncRNA networks associated with EMT [65]. miR-205 is the most down-modulated miRNA in prostate cancer cells upon cancer-associated fibroblast stimulation, due to direct transcriptional repression by HIF1, a known redox-sensitive transcription factor. Ectopic miR-205 overexpression in prostate cancer cells counteracts EMT, thus impairing enhancement of cell invasion, acquisition of stem cell traits, tumorigenicity, and metastatic dissemination. Additionally, miR-205 inhibits tumor-driven activation of surrounding fibroblasts by reducing pro inflammatory cytokine secretion [66]. During hypoxia, most lncRNAs (HALs) are upregulated. HIF could directly promote the expression of these hypoxia-inducible lncRNAs through binding to the HREs (hypoxia response elements) located in their promoter [67]. Tumor hypoxia seems to play an important role to consider as a driver of EMT, tumor heterogeneity, and tumor immune escape.

3.5. Immunological Aspects of EMT

Inflammation contributes significantly to tumor cell metastasis and has been established as a key inducer of EMT during the progression of cancer. Inflammatory cytokines such as TGFβ, TNFα, IL-1, IL-6, and IL-8 activate transcription factors such as SMAD, NF-κB, STAT3, SNAIL, TWIST, and ZEB, which are considered to be potent inducers of EMT in different cancer types [68]. It has been confirmed that IL-6 is involved in the epithelial to mesenchymal transition in head and neck tumor cells via the activation of the STAT3/SNAIL signaling pathway, and STAT3 knock down significantly reverses IL-6-mediated EMT changes [69,70]. There are also results indicating that TNFα and IL-1b (but to a lower extent) induce EMT properties in breast tumor cells [71]. Additionally, TNF-α promotes phenotypic transition and increases the expression of EMT markers compared with TGF-β1 alone, which leads to the assumption that a pro-inflammatory tumor microenvironment enriched in TNF-α may accentuate TGF-β1-induced EMT in colon cancer cells [72]. TGF-β1 in the tumor microenvironment secreted from tumor-associated macrophages (M2 subtype (M2-TAMs)) activates the SMAD2/3 pathway and influences the increase in the expression levels of SOX4 and SOX2, which results in elevation of the migration ability of the cells in vitro, by influencing the expression of genes associated with the EMT process in glioma progression [73]. EMT induced by SNAIL accelerates cancer metastasis due to enhanced invasive ability but also induction of multiple immunosuppression and immune resistance mechanisms including immunosuppressive cytokines, regulatory T cells, impaired dendritic cells, and cytotoxic T lymphocyte resistance. Murine and human melanoma cells with typical EMT features after SNAIL transduction induced regulatory T cells and impaired dendritic cells partly through thrombospondin-1 (TSP1) production [74]. Activated human T cells are involved in the synthesis of IL-6, TNFα, and TGFβ, which induce the expression of mesenchymal proteins such as fibronectin, vimentin, and ZEB1 in inflammatory breast cancer cells [75]. The inflammatory tumor microenvironment (TME) associated with EMT of ‘mesenchymal’ lung adenocarcinoma and lung squamous cell carcinoma show an increased expression of all cytokines and growth factor genes compared to ‘epithelial’ non-small cell lung cancer (NSCLCs). Expression of EMT-inducing cytokines IFN-γ and TNF-α are significantly increased in ‘mesenchymal’ cells, while expression of immunosuppressive cytokines IL-10 and TGF-β are also significantly increased in NSCLCs. TGF-β is not only immunosuppressive but also has EMT inducing and T-cell inhibiting functions [76]. It has an important role in the phenomenon of epithelial to mesenchymal transition (EMT), a phenotypic switch observed with carcinomas that promotes progression towards metastasis in relation to cancer inflammation [77].

4. Epidrugs in Cancer Therapy

DNA methylation is directly linked with histone deacetylation. DNA hypermethylation and hypoacetylation of histones H3 and H4 are linked to cancer progression. As a result, clinical studies have focused on inhibitors of DNA methyltransferases (DNMT) and HDAC inhibitors as a potential therapeutic approach to reverse cancer-promoting epigenetic changes [78]. There is evidence that some epidrugs play a significant role in the regulation of global DNA hypomethylation, which represses EMT [79]. There are data indicating that DNA methylation causally underlies EMT-driven resistance to cancer treatment, rather than being a mere consequence, e.g., several cell lines derived from various cancer types, including hepatocellular carcinoma, pancreatic, lung and ovarian cancer, acquired resistance to cancer therapies by activating an EMT. Importantly, after 5-aza-2′-deoxycytidine-mediated DNA demethylation, treatment-resistant cells lost their growth advantage under therapeutic pressure compared to control cells [80]. The first-generation of epidrugs included DNA methyltransferase inhibitors (the nucleoside analogs of DNMTi: 5-azacytidine, AZA [81], Vidaza®; 5-aza-2′-deoxycytidine, decytabine [82], Dacogen®), and histone deacetylase inhibitors (HDACi, the hydroxamic acids: suberoylanilide hydroxamic acid, SAHA, vorinostat [83], Zolinza®; trichostatin A, TSA [84]; cyclic peptides: trapoxin A, FK228, romidepsin [85], Istodax®) with a low degree of selectivity, and showed undesirable pharmacokinetic properties, poor target selectivity, and low bioavailability, were more active within non-physiological pH ranges, and are targets of cellular deaminases, which ultimately mean a short half-life for these compounds. The second-generation of epidrugs have improved physiological properties, including DNA methyltransferase inhibitors (the nucleoside analogs of DNMTi: zebularine; guadecitabine, S110 or SGI-110 [79], and non-nucleoside analogs: hydralazine, procainamide, RG108), and histone deacetylase inhibitors (HDACi: hydroxamic acid: belinostat, Beleodaq®; dacinostat; panobinostat [85], Farydak®; quisinostat [86]; benzamides: etinostat; mocetinostat [81]; chidamide, Tucidinostat, Epidaza®; carboxylic acids: valproic acid [87]). The mechanism of interaction of the third generation of epidrugs corresponds to the assumption that epigenetic factors could write, delete, or read epigenetic marks in the form of protein complexes, which is essential for the design of highly selective epidrugs. The third generation of epidrugs includes histone methyltransferase inhibitors (HMTi, e.g., EZH2i: GSK126 [81], GSK2816126A; tazemetostat, Tazverik®), histone demethylase inhibitors (HDMi), and bromodomain and extra-terminal domain inhibitors (BETi) [78,88]. Previous results have supported the assumption that co-treatment of epigenetic modifying drugs: 5-azacitidine (an inhibitor of DNA methyltransferases (DNMTs)), GSK126 (an inhibitor of Enhancer Of Zeste Homolog 2 (EZH2)), and/or mocetinostat (an inhibitor of class I histone deacetylases (HDACs)), with overexpression of transcription factor EMT suppressor Grainyhead-like 2 (GRHL2), could induce MET to different extents in ovarian cancer cell lines with an intermediate EMT or a full EMT state [81]. A series of epigenetic drugs for EMT are being tested in trials. Decitabine and 5-azacytidine have been approved by the FDA for acute myeloid leukemia and myelodysplasia treatment. HDAC inhibitors and HDM inhibitors are still under testing. The effectiveness of the use of this group of drugs is especially carefully observed, especially taking into account some study results showing contradictory evidence about the stimulatory effect on EMT in tumor and metastasis [89,90]. This problem must be taken into account in the conducted research. The development of the most effective epidrugs remains an open question.

5. Conclusions

Epithelial to mesenchymal transition is a highly dynamic process involved in the metastatic cascade, as well as in specific hallmarks of cancer, such as cancer cell stemness, therapy resistance, and immune evasion. There is still some controversy about the involvement of EMT in metastasis. Due to the mutated state in cancer cells, chromatin regulators very often have pro-metastatic effects and activate EMT. For example, previous studies have confirmed that PRC2 methyltransferase Enhancer of Zeste Homolog 2 (EZH2) positively regulates breast cancer cell EMT through interaction with SOX4. The stimulatory effect results from transcriptional repression of epithelial genes [91]. Disturbances in maintaining the balance of functional antagonism of chromatin modulators, often resulting from mutations, may have a different effect on EMT in cancer [92]. Due to the nature of epigenetic modifications, the possibility of modulating these changes and understanding their functions in EMT and metastasis will provide new insight into our understanding of tumor progression and metastasis. Owing to the reversible nature of epigenetic modifications, a thorough understanding of their functions in EMT provides new insights into our knowledge of cancer progression and metastasis. Moreover, it can further facilitate the development of diagnostic and therapeutic strategies for the treatment of human cancer. Epigenetic changes associated with EMT are clinically relevant as a potential biomarker. Given the complexity of epigenetic transformations and, importantly, the need to balance the processes involved in epigenetic modifications within target DNA or histones, it remains open how to develop epigenetic treatment strategies aimed at restoring the balance of cancer cells/tissues with minimal impact on normal cells. So called context-dependent behaviors of epigenetic-regulated networks in cancer is crucial to understanding and proposing appropriate therapeutic solutions. Peripheral therapy does not seem to be an accurate solution, and the approach analogous to other therapeutic strategies, including the use of targeted therapy, has its justification. We believe that a detailed understanding of epigenetic aspects in EMT regulation and metastasis will undoubtedly be a good way to develop prognostic biomarkers in cancer, whereas the enzymatic basis of epigenetic changes could become a good research model for testing new epidrugs.

Author Contributions

E.N. and I.B. designed the article. E.N. performed the literature search. E.N. drafted the work. I.B. revised the work. Both authors contributed to the article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Medical University of Silesia, Katowice, Poland, grant number PCN-2-115/K/0/B, KNW-2-B13/N/9/K, PCN-1-174/N/0/1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, D.; Xing, T.; Yang, Z.; Dudek, R.; Lu, Q.; Chen, Y.-H. Epithelial Mesenchymal Transition in Embryonic Development, Tissue Repair and Cancer: A Comprehensive Overview. J. Clin. Med. 2017, 7, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Roche, J. The epithelial-to-mesenchymal transition in cancer. Cancers 2018, 10, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Vergara, D.; Simeone, P.; Franck, J.; Trerotola, M.; Giudetti, A.; Capobianco, L.; Tinelli, A.; Bellomo, C.; Fournier, I.; Gaballo, A.; et al. Translating epithelial mesenchymal transition markers into the clinic: Novel insights from proteomics. EuPA Open Proteom. 2016, 10, 31–41. [Google Scholar] [CrossRef] [Green Version]
  4. Fares, J.; Fares, M.Y.; Khachfe, H.H.; Salhab, H.A.; Fares, Y. Molecular principles of metastasis: A hallmark of cancer revisited. Signal Transduct. Target. Ther. 2020, 5, 28. [Google Scholar] [CrossRef] [PubMed]
  5. Ribatti, D.; Tamma, R.; Annese, T. Epithelial-Mesenchymal Transition in Cancer: A Historical Overview. Transl. Oncol. 2020, 13, 100773. [Google Scholar] [CrossRef] [PubMed]
  6. Brabletz, T.; Kalluri, R.; Nieto, M.A.; Weinberg, R.A. EMT in cancer. Nat. Rev. Cancer 2018, 18, 128–134. [Google Scholar] [CrossRef]
  7. Bakir, B.; Chiarella, A.M.; Pitarresi, J.R.; Rustgi, A.K. EMT, MET, Plasticity, and Tumor Metastasis. Trends Cell Biol. 2020, 30, 764–776. [Google Scholar] [CrossRef]
  8. Steeg, P.S. Targeting metastasis. Nat. Rev. Cancer 2016, 16, 201–218. [Google Scholar] [CrossRef]
  9. Ganesh, K.; Massagué, J. Targeting metastatic cancer. Nat. Med. 2021, 27, 34–44. [Google Scholar] [CrossRef]
  10. Peixoto, P.; Etcheverry, A.; Aubry, M.; Missey, A.; Lachat, C.; Perrard, J.; Hendrick, E.; Delage-Mourroux, R.; Mosser, J.; Borg, C.; et al. EMT is associated with an epigenetic signature of ECM remodeling genes. Cell Death Dis. 2019, 10, 205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Gao, L.M.; Xu, S.F.; Zheng, Y.; Wang, P.; Zhang, L.; Shi, S.S.; Wu, T.; Li, Y.; Zhao, J.; Tian, Q.; et al. Long non-coding RNA H19 is responsible for the progression of lung adenocarcinoma by mediating methylation-dependent repression of CDH1 promoter. J. Cell. Mol. Med. 2019, 23, 6411–6428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Liu, J.; Sun, X.; Qin, S.; Wang, H.; Du, N.; Li, Y.; Pang, Y.; Wang, C.; Xu, C.; Ren, H. CDH1 promoter methylation correlates with decreased gene expression and poor prognosis in patients with breast cancer. Oncol. Lett. 2016, 11, 2635–2643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Shokraii, F.; Moharrami, M.; Motamed, N.; Shahhoseini, M.; Totonchi, M.; Ezzatizadeh, V.; Firouzi, J.; Khosravani, P.; Ebrahimi, M. Histone modification marks strongly regulate Cdh1 promoter in prostospheres as a model of prostate cancer stem like cells. Cell J. 2018, 21, 124–134. [Google Scholar] [CrossRef]
  14. Cui, H.; Hu, Y.; Guo, D.; Zhang, A.; Gu, Y.; Zhang, S.; Zhao, C.; Gong, P.; Shen, X.; Li, Y.; et al. DNA methyltransferase 3A isoform b contributes to repressing E-cadherin through cooperation of DNA methylation and H3K27/H3K9 methylation in EMT-related metastasis of gastric cancer. Oncogene 2018, 37, 4358–4371. [Google Scholar] [CrossRef] [PubMed]
  15. Guvakova, M.A.; Prabakaran, I.; Wu, Z.; Hoffman, D.I.; Huang, Y.; Tchou, J.; Zhang, P.J. CDH2/N-cadherin and early diagnosis of invasion in patients with ductal carcinoma in situ. Breast Cancer Res. Treat. 2020, 183, 333–346. [Google Scholar] [CrossRef]
  16. Ding, Y.; Pan, Y.; Liu, S.; Jiang, F.; Jiao, J. Elevation of MiR-9–3p suppresses the epithelial-mesenchymal transition of nasopharyngeal carcinoma cells via down-regulating FN1, ITGB1 and ITGAV. Cancer Biol. Ther. 2017, 18, 414–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Pereira, I.T.; Ramos, E.A.S.; Costa, E.T.; Camargo, A.A.; Manica, G.C.M.; Klassen, L.M.B.; Chequin, A.; Braun-Prado, K.; Pedrosa, F.D.O.; Souza, E.M.; et al. Fibronectin affects transient MMP2 gene expression through DNA demethylation changes in non-invasive breast cancer cell lines. PLoS ONE 2014, 9, e105806. [Google Scholar] [CrossRef] [Green Version]
  18. Song, C.; Zhu, S.; Wu, C.; Kang, J. Histone deacetylase (HDAC) 10 suppresses cervical cancer metastasis through inhibition of matrix metalloproteinase (MMP) 2 and 9 expression. J. Biol. Chem. 2013, 288, 28021–28033. [Google Scholar] [CrossRef] [Green Version]
  19. Klassen, L.M.B.; Chequin, A.; Manica, G.C.M.; Biembengut, I.V.; Toledo, M.B.; Baura, V.A.; Pedrosa, F.D.O.; Ramos, E.A.S.; Costa, F.F.; de Souza, E.M.; et al. MMP9 gene expression regulation by intragenic epigenetic modifications in breast cancer. Gene 2018, 642, 461–466. [Google Scholar] [CrossRef]
  20. Wang, X.D.; Wen, F.X.; Liu, B.C.; Song, Y. MiR-211 inhibits cell epithelial-mesenchymal transition by targeting MMP9 in gastric cancer. Int. J. Clin. Exp. Pathol. 2017, 10, 7551–7558. [Google Scholar]
  21. Yang, L.; Chen, Y.; Cui, T.; Knösel, T.; Zhang, Q.; Albring, K.F.; Huber, O.; Petersen, I. Desmoplakin acts as a tumor suppressor by inhibition of the Wnt/β-catenin signaling pathway in human lung cancer. Carcinogenesis 2012, 33, 1863–1870. [Google Scholar] [CrossRef] [Green Version]
  22. Galoian, K.; Qureshi, A.; Wideroff, G.; Temple, H.T. Restoration of desmosomal junction protein expression and inhibition of H3K9-specific histone demethylase activity by cytostatic proline-rich polypeptide-1 leads to suppression of tumorigenic potential in human chondrosarcoma cells. Mol. Clin. Oncol. 2015, 3, 171–178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hayashi, M.; Nomoto, S.; Hishida, M.; Inokawa, Y.; Kanda, M.; Okamura, Y.; Nishikawa, Y.; Tanaka, C.; Kobayashi, D.; Yamada, S.; et al. Identification of the collagen type 1 alpha 1 gene (COL1A1) as a candidate survival-related factor associated with hepatocellular carcinoma. BMC Cancer 2014, 14, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Naganuma, K.; Hatta, M.; Ikebe, T.; Yamazaki, J. Epigenetic alterations of the keratin 13 gene in oral squamous cell carcinoma. BMC Cancer 2014, 14, 988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Verma, A.; Singh, A.; Nengroo, M.A.; Saini, K.K.; Sinha, A.; Singh, A.K.; Datta, D. H3K27me3 mediated KRT14 upregulation promotes TNBC peritoneal metastasis. bioRxiv 2021. [Google Scholar] [CrossRef]
  26. Butler, C.; Sprowls, S.; Szalai, G.; Arsiwala, T.; Saralkar, P.; Straight, B.; Hatcher, S.; Tyree, E.; Yost, M.; Kohler, W.J.; et al. Translational Oncology Hypomethylating Agent Azacitidine Is Effective in Treating Brain Metastasis Triple-Negative Breast Cancer Through Regulation of DNA Methylation of Keratin 18 Gene. Transl. Oncol. 2020, 13, 100775. [Google Scholar] [CrossRef] [PubMed]
  27. Lai, Y.C.C.; Cheng, C.C.; Lai, Y.S.; Liu, Y.H. Cytokeratin 18-associated histone 3 modulation in hepatocellular carcinoma: A mini review. Cancer Genom. Proteom. 2017, 14, 219–223. [Google Scholar] [CrossRef] [Green Version]
  28. Yokomichi, N.; Nishida, N.; Umeda, Y.; Taniguchi, F.; Yasui, K.; Toshima, T.; Mori, Y.; Nyuya, A.; Tanaka, T.; Yamada, T.; et al. Heterogeneity of Epigenetic and Epithelial Mesenchymal Transition Marks in Hepatocellular Carcinoma with Keratin 19 Proficiency. Liver Cancer 2019, 8, 239–254. [Google Scholar] [CrossRef]
  29. Wang, W.; He, J.; Lu, H.; Kong, Q.; Lin, S. KRT8 and KRT19, associated with EMT, are hypomethylated and overexpressed in lung adenocarcinoma and link to unfavorable prognosis. Biosci. Rep. 2020, 40, BSR20193468. [Google Scholar] [CrossRef]
  30. Birkenkamp-Demtröder, K.; Hahn, S.A.; Mansilla, F.; Thorsen, K.; Maghnouj, A.; Christensen, R.; Øster, B.; Ørntoft, T.F. Keratin23 (KRT23) Knockdown Decreases Proliferation and Affects the DNA Damage Response of Colon Cancer Cells. PLoS ONE 2013, 8, e73593. [Google Scholar] [CrossRef]
  31. Conceição, A.L.G.; Da Silva, C.T.; Badial, R.M.; Valsechi, M.C.; Stuqui, B.; Gonçalves, J.D.; Jasiulionis, M.G.; De Freitas Calmon, M.; Rahal, P. Downregulation of OCLN and GAS1 in clear cell renal cell carcinoma. Oncol. Rep. 2017, 37, 1487–1496. [Google Scholar] [CrossRef]
  32. Jung, S.; Yi, L.; Kim, J.; Jeong, D.; Oh, T.; Kim, C.H.; Kim, C.J.; Shin, J.; An, S.; Lee, M.S. The role of vimentin as a methylation biomarker for early diagnosis of cervical cancer. Mol. Cells 2011, 31, 405–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Shirahata, A.; Hibi, K. Serum Vimentin Methylation as a Potential Marker for Colorectal Cancer. Anticancer Res. 2014, 34, 4121–4125. [Google Scholar] [PubMed]
  34. Dwedar, F.I.; Khalil, G.I.; Nayer, S.A.; Farouk, A. Aberrant Vimentin Methylation Is Characteristic of Breast Cancer. Adv. Breast Cancer Res. 2016, 684, 150–162. [Google Scholar] [CrossRef] [Green Version]
  35. Tedaldi, G.; Molinari, C.; Celina, S.; Barbosa-Matos, R.; André, A.; Danesi, R.; Arcangeli, V.; Ravegnani, M.; Saragoni, L.; Morgagni, P.; et al. Genetic and Epigenetic Alterations of CDH1 Regulatory Regions in Hereditary and Sporadic Gastric Cancer. Pharmaceuticals 2021, 14, 457. [Google Scholar] [CrossRef]
  36. Rossi, T.; Tedaldi, G.; Petracci, E.; Khouzam, R.A.; Ranzani, G.N.; Morgagni, P.; Saragoni, L.; Monti, M.; Calistri, D.; Ulivi, P.; et al. E-cadherin Downregulation and microRNAs in Sporadic Intestinal-Type Gastric Cancer. Int. J. Mol. Sci. 2019, 20, 4452. [Google Scholar] [CrossRef] [Green Version]
  37. Cong, N.; Du, P.; Zhang, A.; Shen, F.; Su, J.; Pu, P.; Wang, T.A.O.; Zjang, J.I.E.; Kang, C.; Zhang, Q. Downregulated microRNA-200a promotes EMT and tumor growth through the Wnt/β -catenin pathway by targeting the E-cadherin repressors ZEB1/ZEB2 in gastric adenocarcinoma. Oncol. Rep. 2013, 29, 1579–1587. [Google Scholar] [CrossRef] [Green Version]
  38. Da, C.; Wu, K.; Yue, C.; Bai, P.; Wang, R.; Wang, G.; Zhao, M.; Lv, Y.; Hou, P. N-cadherin promotes thyroid tumorigenesis through modulating major signaling pathways. Oncotarget 2017, 8, 8131–8142. [Google Scholar] [CrossRef] [Green Version]
  39. Guo, B.; Zhang, J.; Li, Q.; Zhao, Z.; Wang, W.; Zhou, K.; Wang, X.; Tong, D.; Zhao, L.; Yang, J.; et al. Hypermethylation of miR-338-3p and Impact of its Suppression on Cell Metastasis Through N-Cadherin Accumulation at the Cell-Cell Junction and Degradation of MMP in Gastric Cancer. Cell. Physiol. Biochem. 2018, 50, 411–425. [Google Scholar] [CrossRef]
  40. Deraya, I.E.; Hestiantoro, A.; Muharam, R.; Marwali, M.L.S.; Darmawi; Asmarinah. The mRNA expression and DNA methylation level of fibronectin 1 (FN1) gene encoding focal adhesion molecule in endometrial endometriosis. IOP Conf. Ser. Earth Environ. Sci. 2020, 457, 012079. [Google Scholar] [CrossRef]
  41. Chernov, A.V.; Strongin, A.Y. Epigenetic regulation of matrix metalloproteinases and their collagen substrates in cancer. Biomol. Concepts 2011, 2, 135–147. [Google Scholar] [CrossRef] [Green Version]
  42. Falzone, L.; Salemi, R.; Travali, S.; Scalisi, A.; McCubrey, J.A.; Candido, S.; Libra, M. MMP-9 overexpression is associated with intragenic hypermethylation of MMP9 gene in melanoma. Aging 2016, 8, 933–944. [Google Scholar] [CrossRef] [Green Version]
  43. Liu, Y.; Liu, H.; Luo, X.; Deng, J.; Pan, Y.; Liang, H. Overexpression of SMYD3 and matrix metalloproteinase-9 are associated with poor prognosis of patients with gastric cancer. Tumor Biol. 2015, 36, 4377–4386. [Google Scholar] [CrossRef]
  44. Li, J.; Ding, Y.; Li, A. Identification of COL1A1 and COL1A2 as candidate prognostic factors in gastric cancer. World J. Surg. Oncol. 2016, 14, 1–5. [Google Scholar] [CrossRef] [Green Version]
  45. Han, W.; Hu, C.; Fan, Z.J.; Shen, G.L. Transcript levels of keratin prognostic indicators in melanoma patients. Sci. Rep. 2021, 11, 1023. [Google Scholar] [CrossRef] [PubMed]
  46. Zhong, P.; Shu, R.; Wu, H.; Liu, Z.; Shen, X.; Hu, Y. Low KRT15 expression is associated with poor prognosis in patients with breast invasive carcinoma. Exp. Ther. Med. 2021, 21, 305. [Google Scholar] [CrossRef] [PubMed]
  47. Leygo, C.; Williams, M.; Jin, H.C.; Chan, M.W.Y.; Chu, W.K.; Grusch, M.; Cheng, Y.Y. Review Article DNA Methylation as a Noninvasive Epigenetic Biomarker for the Detection of Cancer. Dis. Markers 2017, 2017, 3726595. [Google Scholar] [CrossRef] [PubMed]
  48. Ghaderi, M.; Niknejad, A. Tumor Microenvironment: Involved Factors and Signaling Pathways in Epithelial-Mesenchymal Transition. Int. J. Cancer Manag. 2021, 14, e113121. [Google Scholar] [CrossRef]
  49. Georgakopoulos-Soares, I.; Chartoumpekis, D.V.; Kyriazopoulou, V.; Zaravinos, A. EMT Factors and Metabolic Pathways in Cancer. Front. Oncol. 2020, 10, 499. [Google Scholar] [CrossRef]
  50. Sundararajan, V.; Tan, M.; Tan, T.Z.; Ye, J.; Thiery, J.P.; Huang, R.Y.J. SNAI1 recruits HDAC1 to suppress SNAI2 transcription during epithelial to mesenchymal transition. Sci. Rep. 2019, 9, 8295. [Google Scholar] [CrossRef] [Green Version]
  51. Carmichael, C.L.; Wang, J.; Nguyen, T.; Kolawole, O.; Benyoucef, A.; de Mazière, C.; Milne, A.R.; Samuel, S.; Gillinder, K.; Hediyeh-Zadeh, S.; et al. The EMT modulator SNAI1 contributes to AML pathogenesis via its interaction with LSD1. Blood 2020, 136, 957–973. [Google Scholar] [CrossRef]
  52. Du, D.; Katsuno, Y.; Meyer, D.; Budi, E.H.; Chen, S.; Koeppen, H.; Wang, H.; Akhurst, R.J.; Derynck, R. Smad3-mediated recruitment of the methyltransferase SETDB1/ESET controls Snail1 expression and epithelial–mesenchymal transition. EMBO Rep. 2018, 19, 135–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Romero, S.; Musleh, M.; Bustamante, M.; Stambuk, J.; Pisano, R.; Lanzarini, E.; Chiong, H.; Rojas, J.; Castro, V.G.; Jara, L.; et al. Polymorphisms in TWIST1 and ZEB1 are associated with prognosis of gastric cancer patients. Anticancer Res. 2018, 38, 3871–3877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lindner, P.; Paul, S.; Eckstein, M.; Hampel, C.; Muenzner, J.K.; Erlenbach-Wuensch, K.; Ahmed, H.P.; Mahadevan, V.; Brabletz, T.; Hartmann, A.; et al. EMT transcription factor ZEB1 alters the epigenetic landscape of colorectal cancer cells. Cell Death Dis. 2020, 11, 147. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, N.; Wang, Y. Decitabine reverses TGF-β 1-induced epithelial–mesenchymal transition in non-small-cell lung cancer by regulating miR-200/ZEB axis. Drug Des. Devel. Ther. 2017, 11, 969–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Gao, Y.; Zhao, Y.; Zhang, J.; Lu, Y.; Liu, X.; Geng, P.; Huang, B.; Zhang, Y.; Lu, J. The dual function of PRMT1 in modulating epithelial-mesenchymal transition and cellular senescence in breast cancer cells through regulation of ZEB1. Sci. Rep. 2016, 6, 19874. [Google Scholar] [CrossRef]
  57. Lee, Y.M.; Kim, S.H.; Kim, M.S.; Kim, D.C.; Lee, E.H.; Lee, J.S.; Lee, S.; Kim, Y.Z. Epigenetic Role of Histone Lysine Methyltransferase and Demethylase on the Expression of Transcription Factors Associated with the Epithelial-to-Mesenchymal Transition of Lung Adenocarcinoma Metastasis to the Brain. Cancers 2020, 12, 3632. [Google Scholar] [CrossRef]
  58. Shi, J.; Cao, J.; Zhou, B.P. Twist-BRD4 Complex: Potential Drug Target for Basal-like Breast Cancer. Physiol. Behav. 2017, 176, 139–148. [Google Scholar] [CrossRef] [Green Version]
  59. Ezponda, T.; Popovic, R.; Shah, M.Y.; Martinez-Garcia, E.; Zheng, Y.; Min, D.-J.; Will, C.; Neri, A.; Kelleher, N.L.; Yu, J.; et al. The Histone Methyltransferase MMSET/WHSC1 Activates TWIST1 to Promote an Epithelial-Mesenchymal Transition and Invasive Properties of Prostate Cancer. Oncogene 2013, 32, 2882–2890. [Google Scholar] [CrossRef] [Green Version]
  60. Jin, L.; Zhang, J.; Fu, H.Q.; Zhang, X.; Pan, Y.L. FOXO3a inhibits the EMT and metastasis of breast cancer by regulating TWIST-1 mediated miR-10b/CADM2 axis. Transl. Oncol. 2021, 14, 101096. [Google Scholar] [CrossRef]
  61. Jägle, S.; Busch, H.; Freihen, V.; Beyes, S.; Schrempp, M.; Boerries, M.; Hecht, A. SNAIL1-mediated downregulation of FOXA proteins facilitates the inactivation of transcriptional enhancer elements at key epithelial genes in colorectal cancer cells. PLoS Genet. 2017, 13, e1007109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wang, Q.S.; Kong, P.Z.; Li, X.Q.; Yang, F.; Feng, Y.M. FOXF2 deficiency promotes epithelial-mesenchymal transition and metastasis of basal-like breast cancer. Breast Cancer Res. 2015, 17, 30. [Google Scholar] [CrossRef] [Green Version]
  63. Saxena, K.; Jolly, M.K.; Balamurugan, K. Hypoxia, partial EMT and collective migration: Emerging culprits in metastasis. Transl. Oncol. 2020, 13, 100845. [Google Scholar] [CrossRef]
  64. Gao, T.; Li, J.-Z.; Lu, Y.; Zhang, C.-Y.; Li, Q.; Mao, J.; Li, L.H. The mechanism between epithelial mesenchymal transition in breast cancer and hypoxia microenvironment. Biomed. Pharmacother. 2016, 80, 393–405. [Google Scholar] [CrossRef]
  65. Hapke, R.Y.; Haake, S.M. Hypoxia-induced epithelial to mesenchymal transition in cancer. Cancer Lett. 2020, 487, 10–20. [Google Scholar] [CrossRef]
  66. Giannoni, E.; Casamichele, A.; Gandellini, P.; Taddei, M.L.; Callari, M.; Piovan, C.; Valdagni, R.; Pierotti, M.A.; Zaffaroni, N.; Chiarugi, P. miR-205 Hinders the Malignant Interplay Between Prostate Cancer Cells and Associated Fibroblasts. Antioxidants Redox Signal. 2014, 20, 1045–1059. [Google Scholar] [CrossRef] [Green Version]
  67. Kuo, T.; Kung, H.; Shih, J. Signaling in and out: Long-noncoding RNAs in tumor hypoxia. J. Biomed. Sci. 2020, 27, 59. [Google Scholar] [CrossRef]
  68. Zhao, Y.; Tian, B.; Sadygov, R.G.; Zhang, Y.; Brasier, A.R. Integrative proteomic analysis reveals reprograming tumor necrosis factor signaling in epithelial mesenchymal transition. Physiol. Behav. 2017, 176, 139–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Hirano, T. IL-6 in inflammation, autoimmunity and cancer. Int. Immunol. 2021, 33, 127–148. [Google Scholar] [CrossRef] [PubMed]
  70. Yadav, A.; Kumar, B.; Datta, J.; Teknos, T.N.; Kumar, P. IL-6 promotes head and neck tumor metastasis by inducing epithelial-mesenchymal transition via the JAK-STAT3-SNAIL signaling pathway. Mol. Cancer Res. 2011, 9, 1658–1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Soria, G.; Ofri-Shahak, M.; Haas, I.; Yaal-Hahoshen, N.; Leider-Trejo, L.; Leibovich-Rivkin, T.; Weitzenfeld, P.; Meshel, T.; Shabtai, E.; Gutman, M.; et al. Inflammatory mediators in breast cancer: Coordinated expression of TNFα & IL-1β with CCL2 & CCL5 and effects on epithelial-to-mesenchymal transition. BMC Cancer 2011, 11, 130. [Google Scholar] [CrossRef] [Green Version]
  72. Li, Y.; Zhu, G.; Zhai, H.; Jia, J.; Yang, W.; Li, X.; Liu, L. Simultaneous stimulation with tumor necrosis factor-α and transforming growth factor-β1 induces epithelial-mesenchymal transition in colon cancer cells via the nf-κb pathway. Oncol. Lett. 2018, 15, 6873–6880. [Google Scholar] [CrossRef]
  73. Liu, Z.; Kuang, W.; Zhou, Q.; Zhang, Y. TGF-β1 secreted by M2 phenotype macrophages enhances the stemness and migration of glioma cells via the SMAD2/3 signalling pathway. Int. J. Mol. Med. 2018, 42, 3395–3403. [Google Scholar] [CrossRef] [Green Version]
  74. Kudo-Saito, C.; Shirako, H.; Takeuchi, T.; Kawakami, Y. Cancer Metastasis Is Accelerated through Immunosuppression during Snail-Induced EMT of Cancer Cells. Cancer Cell 2009, 15, 195–206. [Google Scholar] [CrossRef] [Green Version]
  75. Chattopadhyay, I.; Ambati, R.; Gundamaraju, R. Exploring the Crosstalk between Inflammation and Epithelial-Mesenchymal Transition in Cancer. Mediat. Inflamm. 2021, 2021, 9918379. [Google Scholar] [CrossRef]
  76. Chae, Y.K.; Chang, S.; Ko, T.; Anker, J.; Agte, S.; Iams, W.; Choi, W.M.; Lee, K.; Cruz, M. Epithelial-mesenchymal transition (EMT) signature is inversely associated with T-cell infiltration in non-small cell lung cancer (NSCLC). Sci. Rep. 2018, 8, 2–9. [Google Scholar] [CrossRef] [Green Version]
  77. Dominguez, C.; David, J.M.; Palena, C. Epithelial-mesenchymal transition and inflammation at the site of the primary tumor. Semin. Cancer Biol. 2017, 47, 177–184. [Google Scholar] [CrossRef] [PubMed]
  78. Buocikova, V.; Rios-Mondragon, I.; Pilalis, E.; Chatziioannou, A.; Miklikova, S.; Mego, M.; Pajuste, K.; Rucins, M.; El Yamani, N.; Longhin, E.M.; et al. Epigenetics in breast cancer therapy—New strategies and future nanomedicine perspectives. Cancers 2020, 12, 3622. [Google Scholar] [CrossRef] [PubMed]
  79. Cardenas, H.; Vieth, E.; Lee, J.; Segar, M.; Liu, Y.; Nephew, K.P.; Matei, D. TGF-beta induces global changes in DNA methylation during the epithelial-to-mesenchymal transition in ovarian cancer cells. Epigenetics 2014, 9, 1461–1472. [Google Scholar] [CrossRef] [Green Version]
  80. Galle, E.; Thienpont, B.; Cappuyns, S.; Venken, T.; Busschaert, P.; Van Haele, M.; Van Cutsem, E.; Roskams, T.; Van Pelt, J.; Verslype, C.; et al. DNA methylation-driven EMT is a common mechanism of resistance to various therapeutic agents in cancer. Clin. Epigenetics 2020, 12, 27. [Google Scholar] [CrossRef] [PubMed]
  81. Chung, V.Y.; Tan, T.Z.; Ye, J.; Huang, R.L.; Lai, H.C.; Kappei, D.; Wollmann, H.; Guccione, E.; Huang, R.Y.J. The role of GRHL2 and epigenetic remodeling in epithelial–mesenchymal plasticity in ovarian cancer cells. Commun. Biol. 2019, 2, 272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wang, X.; Wang, J.; Jia, Y.; Wang, Y.; Han, X.; Duan, Y.; Lv, W.; Ma, M.; Liu, L. Methylation decreases the Bin1 tumor suppressor in ESCC and restoration by decitabine inhibits the epithelial mesenchymal transition. Oncotarget 2017, 8, 19661–19673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Sakamoto, T.; Kobayashi, S.; Yamada, D.; Nagano, H.; Tomokuni, A.; Tomimaru, Y.; Noda, T.; Gotoh, K.; Asaoka, T.; Wada, H.; et al. A Histone deacetylase inhibitor suppresses epithelial-mesenchymal transition and attenuates chemoresistance in biliary tract cancer. PLoS ONE 2016, 11, e0145985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Wang, X.; Chen, S.; Shen, T.; Lu, H.; Xiao, D.; Zhao, M.; Yao, Y.; Li, X.; Zhang, G.; Zhou, X.; et al. Trichostatin A reverses epithelial-mesenchymal transition and attenuates invasion and migration in MCF-7 breast cancer cells. Exp. Ther. Med. 2020, 19, 1687–1694. [Google Scholar] [CrossRef]
  85. Chang, T.C.; Matossian, M.D.; Elliott, S.; Burks, H.E.; Sabol, R.A.; Ucar, D.A.; Wathieu, H.; Zabaleta, J.; De Valle, L.; Gill, S.; et al. Evaluation of deacetylase inhibition in metaplastic breast carcinoma using multiple derivations of preclinical models of a new patient-derived tumor. PLoS ONE 2020, 15, e0226464. [Google Scholar] [CrossRef]
  86. Fukuda, K.; Takeuchi, S.; Arai, S.; Kita, K.; Tanimoto, A.; Nishiyama, A.; Yano, S. Glycogen synthase kinase-3 inhibition overcomes epithelial-mesenchymal transition-associated resistance to osimertinib in EGFR-mutant lung cancer. Cancer Sci. 2020, 111, 2374–2384. [Google Scholar] [CrossRef]
  87. Kanamoto, A.; Ninomiya, I.; Harada, S.; Tsukada, T.; Okamoto, K.; Nakanuma, S.; Sakai, S.; Makino, I.; Kinoshita, J.; Hayashi, H.; et al. Valproic acid inhibits irradiation-induced epithelial-mesenchymal transition and stem cell-like characteristics in esophageal squamous cell carcinoma. Int. J. Oncol. 2016, 49, 1859–1869. [Google Scholar] [CrossRef]
  88. Montalvo-Casimiro, M.; González-Barrios, R.; Meraz-Rodriguez, M.A.; Juárez-González, V.T.; Arriaga-Canon, C.; Herrera, L.A. Epidrug Repurposing: Discovering New Faces of Old Acquaintances in Cancer Therapy. Front. Oncol. 2020, 10, 674. [Google Scholar] [CrossRef]
  89. Ji, M.; Lee, E.J.; Kim, K.B.; Kim, Y.; Sung, R.; Lee, S.J.; Kim, D.S.; Park, S.M. HDAC inhibitors induce epithelial-mesenchymal transition in colon carcinoma cells. Oncol. Rep. 2015, 33, 2299–2308. [Google Scholar] [CrossRef] [Green Version]
  90. Cardenas, H.; Zhao, J.; Vieth, E.; Nephew, K.P.; Matei, D. EZH2 inhibition promotes epithelial-to-mesenchymal transition in ovarian cancer cells. Oncotarget 2016, 7, 84453–84467. [Google Scholar] [CrossRef] [Green Version]
  91. Tiwari, N.; Tiwari, V.K.; Waldmeier, L.; Balwierz, P.J.; Arnold, P.; Pachkov, M.; Meyer-Schaller, N.; Schübeler, D.; van Nimwegen, E.; Christofori, G. Sox4 Is a Master Regulator of Epithelial-Mesenchymal Transition by Controlling Ezh2 Expression and Epigenetic Reprogramming. Cancer Cell 2013, 23, 768–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Serresi, M.; Kertalli, S.; Li, L.; Schmitt, M.J.; Dramaretska, Y.; Wierikx, J.; Hulsman, D.; Gargiulo, G. Functional antagonism of chromatin modulators regulates epithelial-mesenchymal transition. Sci. Adv. 2021, 7, eabd7974. [Google Scholar] [CrossRef] [PubMed]
Table 1. Examples of epigenetic regulatory mechanisms of epithelial and mesenchymal markers.
Table 1. Examples of epigenetic regulatory mechanisms of epithelial and mesenchymal markers.
GeneMechanism of Action Types of Cancer Ref.
CDH1DNA metylationlung adenocarcinoma[11]
promoter DNA methylationbreast cancer[12]
enhanced H3K27me3,
reduced H3K9ac, H3K4me3, H3K9me2
prostate cancer stem-like cells[13]
Overexpression of DNMT3Ab
DNA hypermethylation,
H3K9me2 and H3K27me3
gastric cancer[14]
CDH2DNA hypermethylation at CpG site
near the AP-2 binding sequence
ductal carcinoma in situ[15]
FN1miR-9-3pnasopharyngeal carcinoma[16]
MMP2DNA promoter methylationbreast cancer[17]
HDAC10 suppresses expression by interacting with the promoter regions and deacetylating histones H3 and H4cervical cancer[18]
MMP9DNA promoter methylation
H3K4me3
breast cancer[19]
miR-211gastric cancer[20]
HDAC10 suppresses expression by interacting with the promoter regions and deacetylating histones H3 and H4cervical cancer[18]
DSPDNA promoter methylationlung cancer[21]
H3K9 demethylasechondrosarcoma[22]
COL1DNA promoter methylationhepatocellular carcinoma[23]
KRT13promoter abberant hypermethylation
PRC2-mediated H3K27me3
oral squamous cel carcinoma
(cell lines)
[24]
KRT14H3K27me3 enhances transcription by attenuating binding of Sp1 to promotertriple negative breast cancer[25]
KRT18DNA hypermethylationbrain metastasis of breast cancer[26]
histone H3 hypoacetylationhepatocellular carcinoma[27]
KRT19promoter hypermethylationhepatocellular carcinoma[28]
KRT8
KRT19
promoter hypomethylationlung adenocarcinoma[29]
KRT23promoter hypomethylationcolon adenocarcinoma[30]
OCLNdownregulation corelated with increase
in miR-122 and miR-34a expression
clear cell renal carcinoma[31]
VIMpromoter hypermethylationcervical cancer cells[32]
DNA methylation (detected in the serum)colorectal cancer at every disease stage[33]
low methylation associated with
elevated vimentin protein level in the serum
breast cancer[34]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nowak, E.; Bednarek, I. Aspects of the Epigenetic Regulation of EMT Related to Cancer Metastasis. Cells 2021, 10, 3435. https://doi.org/10.3390/cells10123435

AMA Style

Nowak E, Bednarek I. Aspects of the Epigenetic Regulation of EMT Related to Cancer Metastasis. Cells. 2021; 10(12):3435. https://doi.org/10.3390/cells10123435

Chicago/Turabian Style

Nowak, Ewa, and Ilona Bednarek. 2021. "Aspects of the Epigenetic Regulation of EMT Related to Cancer Metastasis" Cells 10, no. 12: 3435. https://doi.org/10.3390/cells10123435

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