Next Article in Journal
Plumbagin, a Biomolecule with (Anti)Osteoclastic Properties
Next Article in Special Issue
High Mobility Group Box 1 Promotes Lung Cancer Cell Migration and Motility via Regulation of Dynamin-Related Protein 1
Previous Article in Journal
Bcl-xL: A Focus on Melanoma Pathobiology
Previous Article in Special Issue
Clinical Perspectives of ERCC1 in Bladder Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Prominent Role of Histone Modifications in the Regulation of Tumor Metastasis

Department of Biological Chemistry, Medical School, National and Kapodistrian University of Athens, 11527 Athens, Greece
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Joint last and corresponding authors.
Int. J. Mol. Sci. 2021, 22(5), 2778; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052778
Submission received: 12 February 2021 / Accepted: 8 March 2021 / Published: 9 March 2021

Abstract

:
Tumor aggressiveness and progression is highly dependent on the process of metastasis, regulated by the coordinated interplay of genetic and epigenetic mechanisms. Metastasis involves several steps of epithelial to mesenchymal transition (EMT), anoikis resistance, intra- and extravasation, and new tissue colonization. EMT is considered as the most critical process allowing cancer cells to switch their epithelial characteristics and acquire mesenchymal properties. Emerging evidence demonstrates that epigenetics mechanisms, DNA methylation, histone modifications, and non-coding RNAs participate in the widespread changes of gene expression that characterize the metastatic phenotype. At the chromatin level, active and repressive histone post-translational modifications (PTM) in association with pleiotropic transcription factors regulate pivotal genes involved in the initiation of the EMT process as well as in intravasation and anoikis resistance, playing a central role in the progression of tumors. Herein, we discuss the main epigenetic mechanisms associated with the different steps of metastatic process, focusing in particular on the prominent role of histone modifications and the modifying enzymes that mediate transcriptional regulation of genes associated with tumor progression. We further discuss the development of novel treatment strategies targeting the reversibility of histone modifications and highlight their importance in the future of cancer therapy.

1. Introduction

Cancer is associated with increased incidence and mortality worldwide due to complex reasons reflecting population age and growth as well as the prevalence of main cancer risk factors. Cancer prognosis is challenging to generalize and depends on host and tumor heterogeneity, including cancer subtype, tumor staging and grading that often associate with infiltration, aggressiveness, and dismal prognosis. Tumor stage is generally more informative than grade in determining prognosis and is based on the assessment of tumor size with direct invasion into nearby structures, lymph node involvement, as well as metastasis which commonly correlates with worst patient outcome.
Metastasis involves the growth of secondary tumors at distant sites from the primary location and is tightly connected to a well-coordinated gene regulation machinery. For metastasis to occur, cancer cells need to undergo epithelial to mesenchymal transition (EMT) characterized by loss of cell adhesion proteins, such as vimentins and cadherins, adapting a nonpolarized, spindle-shaped, and fibroblast-like appearance [1]. This further leads to loss of cell-to-cell and cell-to-matrix interactions, allowing neoplastic cells to escape from the primary site into the circulation. Intravasation requires resistance to anoikis, a form of apoptosis induced by loss of cell-to-cell and cell-to-matrix interaction, thus promoting the survival of neoplastic cells inside the blood vessels [2]. Under physiological conditions, cells anchored to extracellular matrix or to adjacent cells are equipped with a pro-survival machinery that can overwhelm apoptotic signals and promote normal cell function. Loss of anchorage allows cancer cell to overcome apoptosis and travel to a distal location [3]. Extravasation and colonization of the new tissue represent the final steps of the metastatic process. The ‘homing’ of cancer cells at target tissues is a highly specific process depending on adhesion and extravasation as well as the properties of tissue microenvironment [4].
An interplay of genetic and epigenetic changes has been associated with the hallmarks of metastasis, representing critical regulators of genes required for cancer spread in a specific and reversible manner [5] (Figure 1). Epigenetic gene regulation can be achieved by chemically modifying DNA or histones, or by non-coding RNAs, such as microRNAs (miRNAs) and small interfering RNAs (siRNAs) [6]. Proper DNA and chromatin folding relies on histones integrity, stabilisation, and condensation. Increased condensation is associated with suppression of gene expression, whereas reduced condensation enhances gene activation. This process is highly dependent on epigenetic post-translational modifications (PTMs) taking place on the N- and C-terminals of histone proteins or on DNA. Histone modifications are broadly categorized into active or repressive, based on their effects on gene expression. Activating histone modifications include mainly acetylation, demethylation, and, in some cases, methylation of certain histones [7]. On the other hand, repressive modifications include histone methylation, deacetylation, and, less often, demethylation [7]. On top of these modifications, DNA methylation taking place on cytosines, mainly in CpG dinucleotides, induces further repression of gene expression. In parallel, non-coding RNAs act to alter gene transcription and further affect the post-transcriptional fate of the gene product [8]. In the following sections, we discuss main epigenetic mechanisms utilized by cancer cells to regulate gene expression, alter their transcriptional phenotype, and promote metastasis, focusing particularly on histone PTMs and their targeting options.

2. Epigenetic Modifications Regulate Epithelial to Mesenchymal Transition (EMT)

The EMT process can be activated by extracellular signals, such as extracellular matrix components, soluble growth factors such as the Transforming growth factor beta (TGF-β) and fibroblast growth factor (FGF), the signaling Wnt and Notch proteins, or by several intracellular cues which enhance signaling cascades, leading eventually to alterations in cytoskeletal organization and disassembly of cell–cell junction complexes [1,2,3]. A hallmark of EMT is the functional loss of the adhesion molecule E-cadherin present in most epithelial cells that leads to disaggregation of adjacent cancer cells and contributes to dissemination (Figure 1). The promoter of E-cadherin gene (CDH1) harbors E-box elements which are bound to transcription factors such as the Snail family of zinc finger proteins (SNAI1/2/3), the two double zinc finger and homeodomain factors ZEB1/2, and the family of bHLH factors (TWIST1/2, E12/E47)) that are ultimately involved in gene regulation [2,9]. Upon activation, these transcription factors act as molecular switches, induce transcriptional reprogramming, and mediate epithelial-mesenchymal plasticity, providing a rapid regulatory mechanism for cancer progression. Additional transcription factors such as the Forkhead Box C2 (FOXC2) and Transcription Factor 4 (TCF4) are involved in EMT without binding to CDH1 but by suppressing the expression of claudins and desmosomes present in cell junctions, or through interaction with other transcription factors.
During transcriptional reprogramming, alterations in histone modifications play a pivotal regulatory role, orchestrating the simultaneous repression of epithelial genes and activation of mesenchymal genes. Based on their effects on gene transcription and the associated chromatin environment, they are broadly characterized either as permissive or repressive. Permissive PTMs may induce the upregulation of mesenchymal genes, such as N-cadherin (CDH2), VE-cadherin (CDH5), fibronectin (FN1), and vimentin, whereas repressive histone modifications mainly participate in the suppression of epithelial genes, such as CDH1, claudin 1 (CLDN1), claudin 10 (CLDN10), and junction plakoglobin epithelial gene (JUP) [10].

2.1. Activating Histone Modifications Involved in EMT

Activating histone PTMs include mainly acetylation and demethylation. It is well-established that actively transcribed euchromatin is mainly characterized by high lysine acetylation levels catalysed by the acetylases Lysine Acetyltransferase 2B (PCAF), Histone Acetyltransferase P300/CREB Binding Protein (p300/CBP), and Tat Interacting Protein, 60 kDa (TIP60) (Figure 2).
During EMT, β-catenin, which is a core component of the cadherin protein complex, has been shown to interact with the T cell factor, in order to translocate to the nucleus and recruit the p300/CBP complex [11]. This further induces acetylation of the transcription factor ZEB1, enhancing its binding to the microRNA-200c/141 promoter, ultimately leading to its expression [12].
Histone 3 lysine 4 acetylation (H3K4Ac) marks mediated by TIP60, are often detected in the promoters of EMT marker genes, including CDH1, glioma-associated oncogene homolog 1 (GLI1) and smoothened homolog precursor (SMO) of the Hedgehog pathway which are involved in cancer cell migration and invasion [13]. H3K4Ac marks are also present in the transcription regulators Forkhead Box F1 (FOXF1) and BMI1 Proto-Oncogene, Polycomb Ring Finger (BMI1), that promote tumor progression and stemness [13,14].
Increased H3K9 acetylation was observed on the promoter of Transforming Growth Factor Beta Receptor 2 (TGFBR2) after the activation of SNAI1/2, leading to transcription of TGF-β, a major player in the regulation of cancer metastasis [15]. In addition, acetylation of H4K16, mediated by histone acetyltransferase Males-absent-on-the-first (hMOF/KAT8/MYST1) has been demonstrated to maintain the expression of EMT-related tumor suppressor genes, such as target of methylation-induced silencing (TMS1), CDH1, and Estrogen Receptor 1 (ESR1) [16]. hMOF is therefore often downregulated in breast cancer and medulloblastoma, resulting in genomic instability and impaired DNA damage response [16].
Another activating histone modification, demethylation is mainly mediated by the enzymes Jumonji Domain-Containing Protein 1A/B (JMJD1A/B), Jumonji Domain Containing 2 (JMJD2) and the PHD Finger Protein 2/8 (PHF2/8) [17]. JMJD1A has been shown to upregulate Metastasis Associated Lung Adenocarcinoma Transcript 1 (MALAT1), a long non-coding RNA that is involved in dysregulation of cell signaling and EMT induction, cancer cell migration and invasion [18]. Jumonji Domain Containing 2 B (JMJD2B) can interact with β-catenin to mediate H3K9 demethylation of the vimentin (VIM) gene promoter, and further promote EMT and metastasis [19]. Similarly, PHF8 causes H3K9 demethylation on the promoter of integrin genes including Integrin Subunit B2 (ITGB2), Integrin Subunit Alpha M (ITGAM), and Integrin Subunit Alpha 9 (ITGA9) to induce their expression [20]. Lastly, Jumonji Domain-Containing Protein 3 (JMJ3), a histone demethylase that removes methyl marks from H3K27, is induced by TGF-β and activates SNAI1 expression to facilitate EMT [21].

2.2. Repressive Histone Modifications Involved in EMT

Repressive histone modifications mainly involve methylation and deacetylation. Histone methylation occurs on arginine and lysine residues by histone methyltransferases (HMTs) which use S-adenosyl methionine (SAM) as a donor for methyl groups [8] (Figure 3). The Enhancer of zeste homolog 2 (EZH2) methyltransferase, part of the repressive Polycomb Repressive Complex 2 (PRC2) complex has been overexpressed in cancer and especially in metastatic tumors [22]. EZH2 mediates the H3K27me3 marker which has been associated to silencing of the CDH1 gene, Wnt antagonist genes such as Axis Inhibition Protein 2 (AXIN2), Naked1 (NKD1), Protein Phosphatase 2 Regulatory Subunit 2 beta (PPP2R2B), Prickle Planar Cell Polarity Protein 1 (PRICKLE1), Secreted Frizzled Related Protein 5 (SFRP5) and of the metastasis suppressor gene Raf Kinase Inhibitory Protein (RKIP) [22], promoting EMT and cancer metastasis [23] (Figure 3). The expression of PRC2 complex on the other hand has been inversely linked to metastasis and PRC2 loss has been involved in the enhancement of tumor metastatic potential [24].
The methyltransferase G9a that mediates H3K9me2 can interact with SNAI1 and promote EMT by silencing a wide array of genes [25]. Furthermore, it can recruit DNA methyltransferases (DNMT) on the promoter of CDH1 through SNAIL interaction, further suppressing E-cadherin expression and promoting EMT in vitro and in vivo [26]. Hypoxia promotes G9a stability and promotes indirectly tumor progression [27]. The effects of G9a can be partly attributed to its interaction with SNAI1, which induces EMT by silencing a wide array of genes [25] (Figure 3).
Another histone methyltransferase SET Domain Bifurcated Histone Lysine Methyltransferase 1 (SETDB1) mediates the repressive H3K9me3 mark and inhibits gene transcription. It has been shown to potentiate EMT by up-regulation of Signal Transducer And Activator Of Transcription 3 (STAT3) and induction of TWIST and c-myc [28]. Additionally, SETDB1 cooperates with ΔNp63α to induce its stabilization, and ΔNp63α recruits SETDB1 to target genes that are going to be repressed [29]. However, SETDB1 can also interact with Mothers Against Decapentaplegic Homolog 2/3 (SMAD2/3) to suppress EMT and metastasis [30].
Moreover, the histone methyltransferase SET Domain Containing Lysine Methyltransferase 8 (SET8) can monomethylate H4K20 and cooperate with TWIST to reduce CDH1 and upregulate CDH2 in a mouse model in vitro [31]. Di- and tri-methylation of H4K20 is mediated by the methyltransferase Suppressor Of Variegation 4-20 Homolog 1/2 (SUV420H1/2), commonly lost in many tumors where it affects chromatin integrity [32,33] and regulates tensin 3 (TNS3) adhesion protein [34].
A subgroup of ATP-dependent chromatin remodeling complexes, the SWItch/Sucrose Non-Fermentable (SWI/SNF) comprised of two catalytic subunits, BRM/SWI2-Related Gene 1 (BRG1) and BRG1-Associated Factor 60C (BAF60c) has been involved in EMT and tumor progression. BRG1 is recruited by ZEB1 on the CDH1 promoter inducing gene silencing [35], while BAF60c upregulates WNT5a, a member of WNT signaling, activated during EMT [36].
In addition, the Snail family of transcription factors can recruit the lysine-specific demethylase 1 (LSD1) which catalyzes H3K4me2 demethylation, acting downstream to silence Snail-regulated genes, such as CDH1. Transcription factors that activate Snail also interact with Histone Deacetylase 1 (HDAC1), HDAC2 [37], and PRC2 [1]. Snail has been also implicated in the formation of bivalent chromatin state harboring both repressive H3K27me3 marks as well as active H3K4me3 marks. Bivalent genes exist in a poised state, which makes them readily available for activation upon stimulation [38]. The property of bivalency is of major importance in cancer metastasis, contributing to EMT reversibility.
The other repressive histone PTM, deacetylation involves the removal of permissive marks from histones, inhibiting gene expression. Histone deacetylases (HDACs) can form large complexes that are recruited to specific areas, decreasing gene expression. SNAI1, for instance, can recruit the SIN3 Transcription Regulator Family Member A (SIN3A) repressor complex on the CDH1 promoter [39]. This complex contains HDAC 1/2 capable to remove the acetyl marks in gene promoter and decrease its expression. Moreover, SNAI2 can form a complex with the evolutionarily conserved SNAG (Snail/Gfi) domain and bind on the E-box of the BRCA2 promoter, where it recruits the oncogenic transcriptional corepressor C-terminal-binding protein (CtBP) complex that contains HDAC1, further inducing gene repression [40]. ZEB1/2 has been shown to recruit this complex on the promoter of CDH1 [41]. Furthermore, HDAC6 and Sirtuin 1 (SIRT1) have been shown to counteract the p300-mediated acetylation on cortactin and thus enhance its ability to bind F-actin, leading to induction of EMT and tumor progression [42].

2.3. DNA Methylation Involved in EMT

Aberrant DNA methylation is often observed in cancer, mediated by DNA methyltransferases (DNMTs) which are responsible for the formation of methylome, a dynamic chromatin state where genes can be silenced or activated in a quick, specific way, without altering their DNA sequence [43]. Upon induction of EMT, the methylome undergoes specific changes which alter the expression profile of cancer cells and promote the transcription of EMT-associated genes [44]. Silencing of CDH1 during EMT has been associated with promoter hypermethylation, induced by the transcription factors SNA1 [26] and ZEB1 [45] that recruit DNMTs. The DNMT1 cross-talks with the TGF-β/SMAD2 pathway [46] and along with DNMT3A play a significant role in silencing the miR-200 family members (miR-200b, miR-200a, miR-429, miR-200c, and miR-141) during EMT [47,48,49]. DNMTs also recruit methyl-DNA-binding domain proteins such as Methyl-CpG Binding Protein 2 (MeCP2) and Methyl-CpG Binding Domain Protein 1-4 (MBD1-4) [50] that mediate chromatin compaction upon DNA methylation [51]. Methylation is also a reversible event with DNA demethylases such as the three Ten-Eleven Translocation (TET) proteins, creating the dynamic setting mentioned above [52].

2.4. Non-Coding RNAs (Nc-RNAs) Involved in EMT

Non-coding RNAs (ncRNAs) are also involved in the regulation of gene transcription and EMT. MicroRNAs (miRs), a subtype of ncRNAs, target mRNAs post-transcriptionally and promote degradation, inhibition of translation, or both [53]. During EMT, miR-200s, miR-205, and ZEB1/2 act to suppress each other, forming an interactive loop [54]. In this loop, miRs were shown to suppress EMT while ZEB1/2 promotes EMT and cancer progression [55]. TGF-β2 and β-catenin are also repressed by miR-200s, thus inhibiting the EMT process [56]. Moreover, miRs have been shown to target other pro-EMT genes, including EZH2, SNAI1/2, TWIST1/2 [57,58]. Additional miRNAs, such as the Myc-activated miR-9, can promote EMT through targeting of the CDH1 gene [59] (Figure 1).
Long non-coding RNAs (lncRNAs) have also been shown to contribute to EMT and tumor progression through several pathways [60]. LncRNA homeobox (HOX) Transcript Antisense RNA (HOTAIR) that has been associated to metastasis and poor prognosis in several cancer types [61], can interact with both PRC2 and LSD1 in order to repress HOX genes such as HOXD8, HOXD9, HOXD10, HOXD11 along with other metastasis suppressor genes, such as P53-Responsive Gene 1 (PRG1), Junctional Adhesion Molecule 2 (JAM2), Protocadherin 10 (PCDH10), and Protocadherin Beta 5 (PCDHB5) [62]. Furthermore, TGF-β signaling may induce lncRNA-HIT (HOXA transcript induced by transforming growth factor (TGF)-β), which in turn inhibits the expression of E-cadherin, thus promoting EMT and the metastatic process [63]. Collectively, these data depict the complexity of epigenetic regulation that uses several different repressive methods to achieve quick control of gene expression levels.

3. Epigenetic Modifications Involved in Intravasation and Anoikis Resistance

Intravasation and anoikis resistance are the next crucial steps in the process of metastasis. Cancer cells that have undergone EMT and have lost their anchorage, need to suppress apoptotic signals. These signals act by promoting the death receptor activated pathway which involves activation of the initiator caspase-8 along with several other caspases [64]. Cancer cells are able to block this pathway by overexpressing the FLICE-like inhibitory protein (FLIP) which inhibits caspase-8, and is associated with worse prognosis [65] (Figure 1). Moreover, anoikis resistance is also regulated by the Frizzled Class Receptor 7 (FZD7)-TWIST1 interaction where FZD7 regulates the expression of TWIST1 via histone PTMs, such as H3K4me3 and H3K27Ac. These activating modifications increase the expression of transcription factor TWIST1, as validated by the reduction of TWIST1 expression after FZD7 knockdown [66]. BCL2 levels were shown to increase in parallel to TWIST1 levels. Both FZD7 and BCL2 are correlated with the Wnt pathway, and the knockdown of WNT5A, a ligand of FZD7, was shown to decrease BCL2 expression [66].

4. Epigenetic Modifications Involved in Extravasation and Colonization

The last steps in metastasis, extravasation, and colonization require circulating cancer cells to alter again their phenotype in order to ‘home’ new tissues. In this process, cancer cells are required to attain an epithelial-like phenotype which is achieved by reversing EMT in a process called mesenchymal to epithelial transition (MET). This allows the cancer cells to express adhesion molecules and cytokines which crosstalk with the cells of the local tissue and induce recruitment of immune cells [67]. In this way, cancer cells orchestrate their extravasation [68] by upregulating the expression of integrins and selectins [69], leading to elevated cadherin expression [70]. Epigenetic mechanisms are also implicated in this final step of metastasis mainly through histone methylation. Regulation of the cell differentiation N-Myc-Downstream Regulated gene 1 (NDRG1) that induces MET change and is involved in cancer cell metastasis, can be achieved through histone methylation marks [71] (Figure 1). Notably, in cell lines with increased expression of NDRG1 and low metastatic potential, increased methylation of H3K4 has been detected in NDRG1 gene. Importantly, NDRG1 is often silenced in most metastatic cancers [72].

5. Cancer-Type Specific Histone Modifications Regulating Tumor Metastasis

Specific epigenetic changes have been associated with the regulation of metastasis and the acquisition of invasive phenotype in several tumors, as described in this section.

5.1. Head and Neck Cancer

Acetylation of H3K4 has been detected on the promoters of EMT marker genes, including CDH1, GLI1, and SMO which are involved in cell migration and invasion in head and neck squamous cell carcinoma, serving as prognostic markers [13]. Of interest, HDAC3 has also been detected on the promoters of activated mesenchymal genes. It has been shown to remove H3K4Ac from specific gene repressing transcription factors such as Ikaros, allowing the upregulation of EMT genes through activation of Notch signaling, and enhancing metastasis [13].

5.2. Lung Cancer

In lung cancer, H3K4Ac marks have been detected in the transcription regulators FOXF1, and Bmi1 that promote tumor progression and stemness [13,14]. FOXF1 induces EMT through upregulation of lysyl oxidase (LOX) that is critical for the crosslinking of extracellular matrix proteins as well as through suppression of Smad2/3 signaling or transcriptional activation of SNAI1. Moreover, the actin binding protein profilin-2 interacts with HDAC1 to inhibit its binding to the promoters of Smad2 and Smad3, leading to Smad protein activation and subsequently enhancing the TGF-β-induced EMT and angiogenesis in lung cancer cells [73].
Apart of acetylation, histone methylation has also been observed in lung cancer cells. In non-invasive tumor cells, TGF-β favors the association of methyltransferase SETDB1 with Smad3 which mediates H3K9me in the Snail promoter, repressing its expression [74]. On the other hand, in invasive tumor cells undergoing TGF-β-induced EMT, repression of SETDB1 causes de-repression of the Snail promoter [74]. The association of SETDB1 with Smad3 has been also shown to suppress metastasis in lung cancer by repressing IL-2 and the Ca2+-dependent RNA-binding protein annexin A2 (ANXA2) that interacts with the mRNA of the nuclear oncogene, c-myc [30,75].
Additionally, H3K9me3 in the Tripartite Motif Containing 33 (TRIM33)-Smad2/3 complex displaces the heterochromatin protein 1 (HP1), allowing the subsequent activation of mesendoderm differentiation genes, such as Goosecoid Homeobox (GSC) and Mix Paired-Like Homeobox (MIXL1) [76,77].
TGF-β can also activate the Jumonji and AT-Rich Interaction Domain Containing 2 (JARID2), a component of the PRC2 complex that downregulates E-cadherin expression in lung cancer cells. JARID2 occupies the promoters of CDH1 and miR-200 family members and controls the recruitment of PRC and G9a methyltransferase, promoting methylation of H3K27, H3K9, and gene repression [78]. The pro-metastatic effect of G9a observed in lung cancer is attributed to gene silencing of epithelial cell adhesion molecule (ep-CAM), enhancing invasion [79].
Moreover, the protein arginine methyltransferase 5 (PRMT5) forms complex with the methylosome protein 50 (MEP50) to catalyze histone mono- and dimethylation of important EMT genes in lung cancer cells [80]. Specifically, the PRMT5-MEP50 complex mediates H3R2me through recruitment of the nuclear scaffolding protein WD Repeat Domain 5 (WDR5), causing the activation of EMT-promoting genes, such as VIM, Snail and Slug (SNAI1/2). They also mediate H3K4 and H4R3 methylation to repress metastasis suppressing genes, such as CDH1 and the Growth arrest-specific 1 (GAS1) [80]. Additionally, overexpression of a H3K4 histone demethylase, Jumonji, AT Rich Interactive Domain 1A (JARID1A) leads to the activation of cyclin D1/E1 and Integrin Subunit Beta 1 (ITGB1) expression, promoting lung cancer cell growth and metastasis [81].
In parallel with histone modifications, DNA methylation is also implicated in lung cancer metastasis. Overexpression of the programmed death-1 (PD-1) ligand 1 (PD-L1) that regulates tumor microenvironment during EMT has been attributed to decreased DNMT1 levels in lung cancer cells which lead to demethylation of the PD-L1 promoter, changing its expression [82].

5.3. Breast Cancer

Acetylation marks (H3K4Ac) are present in estrogen receptor (ER) signaling-responsive genes in breast cancer cells and their detection has proved useful in predicting early stages of tumor progression [83]. The acetyltransferase TIP60 that establishes H3K4, is involved in oncogene-induced DNA damage response and is often downregulated during hypoxia-induced EMT in ER-negative tumors. Reduced TIP60 expression was shown to block anti-tumor responses, including the DNA Damage Response (DDR) and p53 pathway, suggesting that H3K4Ac reduction favors ER-negative breast cancer progression [84]. On the other hand, TIP60 depletion in ER-positive breast cancer inhibits tumor development. This is possibly attributed to TIP60 recruitment by estrogens, promoting gene transcription. TIP60 interaction with ER-α results in the recruitment of histone methyltransferase Mixed Lineage Leukemia 1 (MLL1), which increases the H3K4me and H2AK5Ac marks, activating target genes that participate in the development of breast cancer [84].
The histone acetyltransferase General Control Of Amino Acid Synthesis Protein 5-Like 2 (GCN5) plays a critical role in the TGF-β/Smad signaling pathway in breast cancer cells [85]. GCN5 inhibition prevents EMT, migration and invasion of breast cancer cells. Another acetyltransferase family member, PCAF promotes EMT and cancer metastasis [86]. The p300 acetylase has been shown to interact with a complex comprised of the Disruptor of telomeric silencing 1-like (DOT1L) histone lysine methyltransferase and c-Myc, activating EMT regulators in breast cancer, and serving as a potential oncogene responsible for the formation of an aggressive phenotype and transformation of cancer stem cells (CSCs) [87].
Histone deacetylases are also highly involved in breast cancer metastasis. HDAC1 induces the growth and migration of breast cancer cells by upregulating Snail/IL-8 signals [88]. HDAC2 enables the motility of breast cancer cells by upregulating Matrix Metalloproteinase 2 (MMP2) and N-cadherin and HDAC3 associates with Epidermal Growth Factor Receptor (EGFR) and c-Src to promote breast cancer cell invasion [89,90]. In contrast, HDAC8 forms a complex with SMAD3/4, in order to transcriptionally suppress SIRT7 and thus inhibit metastasis, increasing the efficacy of chemotherapy in breast cancer [91]. Finally, SIRT7 reduces breast cancer metastasis by degrading SMAD4, the key factor in TGF-β pathway, thus inhibiting this pathway [91].
Moving towards methylation, the histone methyltransferase PRMT1 was shown to bind ZEB1 promoter and induce EMT in breast cancer cells [92]. Similarly, PRMT7 regulates E-cadherin expression and promotes EMT [93] whereas PRMT6 attenuates p21 signaling [94]. Additionally, PRMT9 activates the Phosphoinositide 3-kinase/AKT Serine/Threonine Kinase/Glycogen Synthase Kinase 3 Beta/Snail (PI3K/Akt/GSK-3b/Snail) pathway, enhancing cell migration and invasion [95].
In contrast, the methyltransferase SUV420H2 attenuates EMT and tumor progression by establishing H4K20me3, and decreasing the expression of the focal adhesion protein tensin-3 (TNS3) [34].
The lysine-specific demethylase JARID1A, upregulates the expression of the extracellular matrix protein Tenascin C (TN-C), and favors breast cancer cell invasion [96]. The demethylase JARID1B, promotes EMT by suppressing miR-200s [97] or Phosphatase And Tensin Homolog (PTEN) [98] through H3K4 demethylation of its promoter. However, JARID1B can act in concert with LSD1, to remove three H3K4 methylation marks from the C-C Motif Chemokine Ligand 14 (CCL14) promoter, resulting in inhibition of chemokine-mediated migration, angiogenesis and breast cancer metastasis [99].
The Ubiquitously transcribed tetratricopeptide repeat, X chromosome (UTX/KDM6A) demethylase activates several oncogenes and pro-metastatic genes, such as metalloproteinases MMP-9/11 and Homeobox protein Sineoculis homeobox homolog 1, Sine Oculis Homeobox Homolog 1 (SIX1), upon interaction with the H3K4 methyltransferase, MLL4 complex, enhancing EMT and breast cancer metastasis [100]. However, UTX has been also shown to cooperate with LSD1, HDAC1 and DNMT1 to compete with the MLL histone methyltransferase complex and disrupt the recruitment of c-Myc and p300. This induces the inhibition of H3K4 methylation on the promoters of EMT regulators, including Snail, ZEB1, and ZEB2 [101], as well as on the promoter of CDH1.
At this point, it is important to note that different histone PTMs are associated with specific breast cancer subtypes. H3K9ac and H3K36me3 marks are often detected in HER2-positive breast cancer cells, while H3K9Ac, H3K4me3 and H3K79me2 are predominant in triple negative breast cancer (TNBC) [102]. H3K27me3 is present in lower grade tumors, Luminal A and B1 subtypes [103,104] where it associates with repression of the Forkhead Box C1 (FOXC1) gene in Luminal B breast cancer, resulting in metastatic behavior [105]. SIRT1 is upregulated in luminal and HER2-positive subtypes but is significantly downregulated in TNBC, whereas H3K4ac, H3K9ac, and H4K16ac are relatively upregulated in TNBC but greatly reduced in luminal and HER2-positive subtypes [106].

5.4. Gastrointestinal Cancers

Several deacetylases have a predominant role in gastrointestinal cancers. SIRT2 is involved in the regulation of Akt/GSK-3β/β-catenin pathway to mediate EMT and associates with prognosis in hepatocellular carcinoma and esophageal squamous cell carcinoma. In hepatoma cells, HDACs catalyze H3K4/56 deacetylation at the CDH1 promoter, inducing E-cadherin repression by Snail2, in favor of EMT [107]. HDAC1/2 are recruited to ZEB1 on CDH1 promoter, inducing its repression in pancreatic cancer cells [41]. Moreover, overexpression of FLIP has been shown to mediate a concomitant overexpression of the anti-apoptotic factor Bcl-XL, and inhibit the anoikis apoptotic pathway in a pancreatic cancer cell model [108]. This result was counteracted by HDAC inhibition, which reduced FLIP levels and induced apoptosis of cancer cells, suggesting that HDACs participate in anoikis resistance of cancer cells [108].
Furthermore, amplification of the SETDB1 methyltransferase is responsible for the outgrowth of liver cancer cells by methylating and stabilizing the oncogenic p53 mutants [109].
The PRC2 complex that establishes repressive H3K27me3 marks has been shown recruited on the Kruppel Like Factor 2 (KLF2) and CDH1 promoters in gastric cancer cells [110]. Α switch of active H3K4me3 to repressive H3K27me3 has been detected on 102 EMT marker gene promoters, whereas the opposite switch of H3K27me3 to the activating H3K4me3 was observed on a series of upregulated mesenchymal gene promoters, such as ZEB2, CDH2, Platelet Derived Growth Factor Receptor α (PDGFRα) and ESRP1 in a TWIST1-induced EMT cell model [111]. Additionally, a third set of promoters containing both H3K4me3 and H3K27me3 marks were detected to regulate several bivalent genes in the same model, indicating an increased level of plasticity of the mesenchymal cell state [111].
Finally, DNA methylation has been also shown to regulate TWIST1/2 promoter methylation and correlate inversely with the TWIST1/2 expression levels in colorectal cancer. TWIST1/2-positive high-grade tumors exhibit both lymphatic vessel invasion and lymph node metastasis suggesting that TWIST promoter methylation may serve as a prognostic marker for patients with colorectal cancer [112]. Similarly, ZEB2 expression is also regulated by promoter methylation in pancreatic and hepatocellular carcinoma [113,114].

5.5. Prostate Cancer

Several histone modifications have been detected in prostate cancer cells with NAD-dependent deacetylase SIRT1 being recruited in ZEB1 to induce silencing of CDH1 and promoting EMT [115]. ZEB1-induced EMT is accompanied by repression of other epithelial genes, such as Epithelial cell adhesion molecule (EPCAM), Suppressor of tumorigenicity 14 (ST14), Epithelial Splicing Regulatory Protein 1 (ESRP1), and RAB25, reduced H3K9Ac and H3K27Ac on their promoters and a global H3K27 deacetylation [116]. Additionally, TGF-β promotes H3K4me3 and Retinoblastoma-Binding Protein 5 (RbBP5) binding to the promoter of Snail by recruiting Smad2/3 and CBP, leading to enhanced Snail expression in prostate cancer cells [78]. Furthermore, PHF8 demethylase was demonstrated to remove H3K9 methylation on the promoter of integrin genes and Rho-Associated Protein Kinase 1 (ROCK) kinase in prostate cancer cells and inducing their expression and promoting migration and invasion, thus being correlated with poor prognosis [20].

6. Targeting Options

The high predominance of epigenetic changes in the metastatic process of several tumors, along with their reversibility, has pointed research efforts towards their therapeutic targeting for the development of new epigenetic drugs that will be patient- and cancer-subtype specific (Table 1).

6.1. Low Molecular Weight EMT Inhibitors

An interesting EMT therapeutic approach is the use of small molecular weight inhibitors of EMT promoting genes, which target specific binding proteins that recognize different histone modifications with promising results in vitro [117]. These include small molecule compounds that target the acetylated-histone binding bromodomain (BD), specifically blocking the binding to acetylated lysine by the BD-containing transcriptional co-activator, Bromodomain Containing 4 (BRD4) [118,119]. Since BRD4 is found in the super-enhancer region of the Myc oncogene, its inhibition prevents Myc function in tumor cells [120]. More specifically, treatment with the BRD4 inhibitor MS417 was shown to reduce metastasis of colorectal cancer in mouse models by altering the expression of key EMT genes, causing a sharp increase in the epithelial marker E-cadherin, as well as a decrease in the mesenchymal marker vimentin [121]. Another BRD4 inhibitor JQ1 along with MS417 were demonstrated to interfere with BRD4 binding to Keratin 73 (K73)/K76Ac2 on TWIST transcription factor and repress WNT5A expression, thus reversing EMT and metastasis of breast cancer in vitro and in vivo [122]. Currently BRD4 inhibitors are tested in multiple clinical trials of different cancer types. Other small molecules, such as UNC1215 and UNC3866 that target specific methylated histone binders have been also used, even though their function in reversing EMT and tumor progression has yet to be clarified. They were shown to block the binding of methyl-lysine by the Malignant Brain Tumor (MBT) domain-containing protein Lethal(3)Malignant Brain Tumor-Like Protein 3 (L3MBTL3) [123] and CD-containing protein Chromobox 4/7 (CBX4/7) [124], respectively. Furthermore, in the context of EMT, TGF-β1 treatment in gastric cancer cells induced expression of JARID1A demethylase, which is recruited by p-SMAD3 to CDH1 promoter, inducing its silencing, and promoting malignant progression [125]. Finally, drugs, such as salinomycin that target CSCs, which play a great role in EMT have been successfully used in preclinical settings of various tumors [126,127,128,129,130]. Experimental data support that these epigenetic cancer treatments may influence the Twist-Snail/ZEB-E-cadherin axis and EMT inducers, such as Wnt-TGFβ-Bone Morphogenetic Protein (BMP). Finally, RNA interference techniques also seem to rise as a more promising approach for repressing these TFs on mRNA level [131].

6.2. Histone-Modifying Drugs

Several epigenetic drugs have already been approved for clinical use, including the combination regimen of DNA methyltransferase inhibitors (DNMTi) with histone deacetylase inhibitors (HDACi) in the treatment of myelodysplastic syndromes [132]. Clinical trials (up to phase IIb) with the same regimen have also been conducted for non-Hodgkin’s lymphomas, such as T-cell and diffuse large B-cell lymphoma, as well as acute myeloid leukemia [133]. In other solid tumors, for example non-small cell lung cancer, these drugs have yet to be established as a preferred treatment option and are only being used as experimental therapies in advanced, recurrent, or refractory states [133,134]. Moreover, panobinostat, a novel potent inhibitor of all HDAC enzymes has been effective towards cancer proliferation and apoptosis [135], as well as towards the expression of differentiation and EMT markers in in vivo hepatoma models [136]. Panobinostat was shown to upregulate the epithelial cytokeratin marker and downregulate the mesenchymal vimentin marker. The combination of the histone deacetylase inhibitor, Suberoylanilide hydroxamic acid (SAHA), and the methyltransferase inhibitor, Zebularine affects the differentiation of pancreatic cancer models [137]. Peroxiredoxin-2 (TSA) has also been used to target class I HDACs both in vivo and in vitro and resulted in reduced cancer cell growth along with EMT and metastasis suppression [138]. Mocetinonstat was also the only HDACi that demonstrated specific antagonism of ZEB1-mediated miR-203 repression in pancreatic cancer cells. miR-203 silencing by ZEB1 is strongly associated with tumor recurrence after treatment with gemcitabine [139]. A number of HDACi have also been investigated in the treatment of solid tumors, including breast cancer [140,141]. They have yet to be approved for clinical use, but they comprise a promising target for treating breast cancer and especially the refractory hormone-positive subtypes. HDACi prevent breast tumor progression via transcriptional inhibition of EMT-related genes, modulation of human epidermal growth factor receptor 2 (HER2) expression or induction of estrogen receptor (ER) in hormone receptor-negative tumors, as well as by increasing the sensitivity of hormonal therapy in ER positive tumors [142]. Only one phase III clinical trial exists that demonstrates the anti-tumor effects of HDACi in breast cancer [143]. HDACi have also been shown to target metastatic triple-negative breast cancer [144].
HMT inhibitors targeting PRMT3-6 in breast cancer have also been used preclinically [145], with PRMT5 inhibition displaying the greatest effects among the other family members. Its inhibition seems to reduce the metastatic potential and proliferation rate of breast cancer cells [146] and influence breast cancer sensitivity to other drugs [80]. On the other hand, PRMT2 methyltransferase itself was shown to improve the sensitivity of tamoxifen in ER positive breast cancer cells by transcriptionally suppressing the 36 kDa variant of estrogen receptor a, ER-a36 [147]. Similarly, PRMT4 activation was inhibited in endocrine resistant breast cancer cells [148].
Furthermore, EZH2 methyltransferase inhibitors have been used in clinical trials [149] and can be combined with Extracellular Signal-Regulated Kinase (Erk) inhibitors to suppress TGF-β-induced EMT [150]. The EZH2 inhibitor Tazemetostat, is currently under clinical trial in lymphomas and other tumors [151]. However, a recent clinical study of the EZH2 inhibitor GSK2816126 on lymphomas, solid tumors and multiple myelomas showed little effect in treatment efficiency and was terminated [152].
The non-selective Suppressor Of Variegation 3-9 Homolog 1 (SUV39H1) methyltransferase inhibitor, chaetocin [153] has been demonstrated to restore E-cadherin expression and the tumor suppressor p15INK4B, by reducing the levels of H3K9me3 on the promoters of these genes [154]. BIX01294, another histone methyltransferase inhibitor was shown to induce E-cadherin expression through inhibition of G9a and G9a-Like Protein (GLP) and reduction of H3K9me2 on its promoter [155,156]. UNC0638, another specific inhibitor of G9a and GLP methyltransferases with lower toxicity profile compared to BIX01294, was shown to exert similar effects and reduce the global H3K9me2 levels in in vivo [157].
The demethylase LSD1 implicated in the promotion of EMT can be also targeted to suppress metastasis. In vivo and in vitro studies showed that LSD1 inhibition by parnate or by an inhibitor targeting LSD1-SNAI1 interaction, leads to decreased cancer motility and invasiveness while also increasing E-cadherin [158]. Similar results were observed with the LSD1 inhibitor pargyline [159,160]. On the contrary, another clinical study on the LSD1 inhibitor GSK2879552 in patients with relapsed/refractory small cell lung carcinoma showed that the risk benefit was not significant to justify further studies on this drug [161]. However, more LSD1 inhibitors are currently tested in clinical trials for evaluation of their efficacy [162].

6.3. DNA Methylation Inhibitors

DNA methylation has also been explored as a possible therapeutic target for metastatic cancers. The DNA methylation status of cancer cells can predict their responsiveness to treatment with cisplatin, a standard chemotherapeutic agent used in gastric cancer. Furthermore, the DNA methylation inhibitor, decitabine applied in myelodysplastic syndromes (MDS) and acute myeloid leukemia (AML) has been shown to induce changes in cancer cell morphology, colony formation, and cell differentiation markers’ expression [163]. The hypomethylating agent 5-aza-2′-deoxycytidine can influence the differentiation status of cultured cancer cells and inhibit the High Mobility Group AT-Hook Protein 2 (HMGA2)-induced EMT [164,165,166]. 5-aza-2′-deoxycytidine was also able to restore epithelial phenotypes and inhibit EMT by promoting E-cadherin re-expression [167]. Moreover, 5-aza-2′-deoxycytidine-mediated targeting of BMP-4 that is epigenetically upregulated in cisplatin-resistant gastric cancer cell lines, was shown to improve their sensitivity to chemotherapy [168]. However, this effect of DNA demethylating agents, although capable of restoring genes silenced in cancer is not specific, and may result in the aberrant expression of many non-epithelial genes, affecting positively EMT progression and tumor metastasis [169].

6.4. NcRNAs

Lastly, the use of non-coding RNAs has been proposed as a treatment modality to reduce cancer metastasis. miRNAs and lncRNAs can decrease the expression of EMT-promoting transcription factors as well as the activity of epigenetic modifying enzymes. Therefore, increasing the expression of these non-coding RNAs or reintroducing them in the organism are under study. Despite the lack of in-depth knowledge, preclinical studies demonstrate promising results for future implementation of these epigenetic modulators in the treatment of cancer [170].
Table 1. Drug targeting of histone modifications.
Table 1. Drug targeting of histone modifications.
Drug CategoryDrug NameCancer TypeEffectReference
LMW EMT inhibitorsBRD4 InhibitorsMS417ColorectalInhibits metastasis via induction of E-cadherin and inhibition of Vimentin[121]
JQ1BreastBlocks BRD4 binding to K73/K76Ac2 on the TWIST transcription factor, resulting in decreased WNT5A expression and inhibition of EMT and metastasis[122]
Histone binder inhibitorUNC3866ProstateBlocks the binding of CBX4/7 to methyl-lysine; unclarified role in EMT reversal[124]
CSC targeting drugsSalinomycinBreast, GI, Leukemia, Lung, ProstateInfluences the Twist-Snail/ZEB-E-cadherin axis and Wnt-TGF-β-BMP to inhibit EMT[126,127,128,129,130]
Histone-modifying drugsDNMT InhibitorsChaetocinAMLRestores E-cadherin and p15INK4B expression by reducing SUV39H1- mediated H3K9me3 on their promoters[154]
BIX01294HeLa cellsInduces E-cadherin expression by inhibiting G9a and GLP-mediated H3K9me2 deposition on its promoter.[155,156]
UNC0638PancreaticInduces E-cadherin expression by inhibiting G9a and GLP-mediated H3K9me2 deposition on its promoter. [157]
ZebularinePancreaticUsed in combination with SAHA to promote the differentiation of cancer cells[137]
HDAC InhibitorsSAHAPancreaticCombined with Zebularine in order to promote differentiation of cancer cells[137]
PanobinostatHepatocellularInhibits cancer proliferation and induces apoptosis of cancer cells. Also increases differentiation markers in vivo[135]
MocetinostatPancreaticSpecific antagonism of ZEB1-mediated silencing of miR-203, which is associated with tumor recurrence[139]
EntinostatBreastIn phase III clinical trial[143]
PRMT5 InhibitorsshRNA- mediated inhibitionBreast, LungReduces metastatic potential, EMT and proliferation rate of cancer cells[80,146]
EZH2 InhibitorsTazemetostatAdvanced solid tumors and lymphomasUnder clinical trial investigation[152]
GSK2816126Lymphomas, solid tumors and MM Clinical study terminated due to little effect in treatment efficiency[152]
Erk InhibitorsAZD6244BreastCombination with EZH2 inhibitors suppresses TGF-β-induced EMT[150]
LSD1 InhibitorsParnateMyelogenous leukemiaDecreases motility and invasiveness, increases E-cadherin[158]
PargylineLung Suppresses proliferation, migration and invasion of cancer cells[159,160]
GSK2879552LungClinical study discontinued [161]
DNA methylation inhibitorsDecitabineMDS, AMLChanges cancer cell morphology, differentiation markers and inhibits proliferation[163]
5-aza-2′-deoxycytidineGastricImproves sensitivity to chemotherapy, restores epithelial phenotypes by promoting E-cadherin re-expression and inhibits EMT[167,168]
LMW; Low molecular weight, EMT; Epithelial to Mesenchymal Transition, BRD4; Bromodomain-containing protein 4, CSC; Cancer stem cell, GI; Gastrointestinal, WNT5A; Wnt Family Member 5A, CBX4/7; Chromobox 4/7, DNMT; DNA Methyltransferase, HDAC; Histone Deacetylase, SUV39H1; Suppressor Of Variegation 3-9 Homolog 1, G9a; Euchromatic histone-lysine N-methyltransferase 2 (EHMT2), GLP; G9a-like protein, ZEB1; Zinc finger E-box-binding homeobox 1, PRMT5; Protein Arginine Methyltransferase 5, EZH2; Enhancer of zeste homolog 2, MM; Multiple Myeloma, MDS; Myelodysplastic Syndromes, AML; Acute Myeloid Leukemia.

7. Conclusions—Future Perspectives

Collectively, metastasis is a critical step in cancer progression and worsens drastically patients’ survival. Its complexity stems from the wide variety of crosstalk pathways, as well as the heterogeneity observed in different cancer types and subtypes that either promote or inhibit the process. The subdivision of metastasis in three major steps has helped research to focus more on the specific checkpoints that regulate the process and elucidate previously unknown regulatory mechanisms for potential targeting.
Epigenetic modifications have recently emerged as critical regulators and drivers of the metastatic process. Targeting of these pathways holds a great therapeutic potential, with scientists being able not only to determine the epigenetic alterations, but also use newly discovered epigenetic modulators, to specifically target and reverse them. In this way, an effort is made to reset the process of metastasis and reconstitute the loss of epithelial-like phenotype in cancer cells. Until recently, most drugs targeting the epigenetic state of cells were only evaluated in vitro or in in vivo models. However, the constant progress in the field of epigenetics, has allowed their future implementation in clinical trials and in the clinical setting as individualized therapies that specifically target each cancer’s unique “epigenome”.
Despite these encouraging results, epigenetic therapies need extensive investigation to overcome the current lack of specificity of some agents, to elucidate the exact role of the different epigenetic alterations, and to determine the therapeutic effect of epigenetic regulation and the way to implement them in treatment schemes as a monotherapy or in conjunction with other agents. Future research should focus on unraveling the complex regulatory network induced by epigenetic mechanisms along with tumor microenvironment, aiming to define the best therapeutic targets and biomarkers for evaluating tumor aggressiveness and potential response to therapy. In these efforts, new technologies such as ATAC-seq, single-cell RNA-seq, and ChIP-seq will also prove invaluable and help scientists shed light in the entirety of the epigenome, its alterations in cancer, and the ways it can be targeted to increase patient prognosis, survival, and quality of life.

Author Contributions

Conceptualization, C.P.; methodology, M.M. and D.S.; software, M.M. and D.S.; validation, A.G.P. and C.P.; formal analysis, M.M.; investigation, M.M. and D.S.; resources, E.K.B. and C.P.; data curation, M.M.; writing—original draft preparation, M.M. and D.S.; writing—review and editing, A.G.P. and C.P.; visualization, M.M.; supervision, C.P.; project administration, C.P.; funding acquisition, E.K.B. and A.G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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.

Abbreviations

EMTEpithelial to Mesenchymal Transition
P300/CBPE1A Binding Protein P300/CREB Binding Protein
ZEB1Zinc Finger E-Box Binding Homeobox 1
BRG1BRM/SWI2-Related Gene 1
CtBPC-terminal-binding protein
HDACHistone deacetylase
SNAI1Snail Family of zinc finger proteins 1
SNAI2Slug
JMJ3Jumonji Domain Containing 3 Histone Lysine Demethylase
G9aEuchromatic Histone Lysine Methyltransferase 2
SIN3ASIN3 Transcription Regulator Family Member A
SET8Lysine Methyltransferase 5A
HHistone
KLysine
memethylation
acacetylation
RArginine
TWIST1Twist Family BHLH Transcription Factor 1
EZH2Enhancer of Zeste 2 Polycomb Repressive Complex 2 Subunit
PRC2Polycomb Repressive Complex 2
FZD7Frizzled Class Receptor 7
FLIPCASP8 and FADD Like Apoptosis Regulator
JMJD2BJumonji Domain-Containing Protein 2B
PHF8PHD Finger Protein 8
NDRG1N-Myc Downstream Regulated 1
GLI1Glioma-Associated Oncogene Homolog 1
SMOSmoothened, Frizzled Class Receptor
FOXF1Forkhead Box F1
Bmi1Polycomb Complex Protein BMI-1
SIRT2Sirtuin 2
TGFBR2TGF-β Receptor type 2
GCN5General Control of Amino Acid Synthesis Protein 5-Like 2
PCAFP300/CBP-Associated Factor
TIP60Tat Interacting Protein, 60 kDa
DDRDNA Damage Response
hMOFMales-absent-on-the-first histone acetyltransferase
TMS1Target of Methylation-induced Silencing
ESR1Estrogen Receptor 1
Smad2Mothers Against Decapentaplegic Homolog 2
EPCAMEpithelial Cell Adhesion Molecule
ST14Suppression of Tumorigenicity 14
ESRP1Epithelial Splicing Regulatory Protein 1
RAB25RAB25, Member RAS Oncogene Family
FLIPCASP8 And FADD Like Apoptosis Regulator
PRMTProtein Arginine Methyltransferase
MEP50Methylosome Protein 50
GAS1Growth Arrest Specific 1
PRCPolycomb Repressive Complex
KLF2Kruppel Like Factor 2
JARIDJumonji, AT Rich Interactive Domain
TRIM33Tripartite Motif Containing 33
Smad2/3Mothers Against Decapentaplegic Homolog 2/3
HP1Heterochromatin Protein 1-Alpha
GSCGoosecoid Homeobox
MIXL1Mix Paired-Like Homeobox
WDR5WD Repeat-Containing Protein 5
KDM6BLysine (K)-Specific Demethylase 6B
ITGIntegrin
UTXUbiquitously- Transcribed X Chromosome Tetratricopeptide Repeat Protein
MMP-9/11Metalloproteinase-9/11
SIX1Sine Oculis Homeobox Homolog 1
MLL4Mixed-Lineage Leukemia Protein 4
ROCK kinaseRho-Associated Protein Kinase 1
FOXC1Forkhead Box C1
SETDB1SET Domain Bifurcated Histone Lysine Methyltransferase 1
CDH1Cadherin 1
AXIN2Axis Inhibition Protein 2
NKD1Naked1
PPP2R2BProtein Phosphatase 2 Regulatory Subunit 2 beta
PRICKLE1Prickle Planar Cell Polarity Protein
SFRP5Secreted Frizzled Related Protein 5
RKIPRaf Kinase Inhibitory Protein

References

  1. Lin, T.F.; Ponn, A.; Hu, X.; Law, B.K.; Lu, J.L.D.P. Requirement of the histone demethylase LSD1 in Snai1-mediated transcriptional repression during epithelial-mesenchymal transition. Oncogene 2010, 29, 4896–4904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Thiery, J.P. Epithelial-mesenchymal transitions in tumour progression. Nat. Rev. Cancer 2002, 2, 442–454. [Google Scholar] [CrossRef]
  3. Paoli, P.; Giannoni, E.; Chiarugi, P. Anoikis molecular pathways and its role in cancer progression. Biochim. Biophys. Acta 2013, 1833, 3481–3498. [Google Scholar] [CrossRef] [Green Version]
  4. Sadanandam, A.; Varney, M.L.; Kinarsky, L.; Ali, H.; Mosley, R.L.; Singh, R.K. Identification of Functional Cell Adhesion Molecules with a Potential Role in Metastasis by a Combination of in vivo Phage Display and in silico Analysis. OMICS A J. Integr. Biol. 2007, 11, 41–57. [Google Scholar] [CrossRef]
  5. Wang, H.; Zhang, Y.; Kriska, A.; Chen, H. Chapter 28—Epigenetic Regulation in Cancer Metastasis; Tollefsbol, T.O.B.T.-M.E., Ed.; Academic Press: Boston, MA, USA, 2016; pp. 499–514. ISBN 978-0-12-803239-8. [Google Scholar]
  6. Costa, F.F. Epigenomics in cancer management. Cancer Manag. Res. 2010, 2, 255–265. [Google Scholar] [CrossRef]
  7. Hake, S.B.; Xiao, A.; Allis, C.D. Linking the epigenetic “language” of covalent histone modifications to cancer. Br. J. Cancer 2004, 90, 761–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Strepkos, D.; Markouli, M.; Klonou, A.; Papavassiliou, A.G.; Piperi, C. Histone methyltransferase SETDB1: A common denominator of tumorigenesis with therapeutic potential. Cancer Res. 2020. [Google Scholar] [CrossRef]
  9. Zhou, B.; Liu, Y.; Kahn, M.; Ann, D.K.; Han, A.; Wang, H.; Nguyen, C.; Flodby, P.; Zhong, Q.; Krishnaveni, M.S.; et al. Interactions between β-catenin and transforming growth factor-β signaling pathways mediate epithelial-mesenchymal transition and are dependent on the transcriptional co-activator cAMP-response element-binding protein (CREB)-binding protein (CBP). J. Biol. Chem. 2012, 287, 7026–7038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Wu, M.-Z.; Tsai, Y.-P.; Yang, M.-H.; Huang, C.-H.; Chang, S.-Y.; Chang, C.-C.; Teng, S.-C.; Wu, K.-J. Interplay between HDAC3 and WDR5 is essential for hypoxia-induced epithelial-mesenchymal transition. Mol. Cell 2011, 43, 811–822. [Google Scholar] [CrossRef]
  11. Hao, S.; He, W.; Li, Y.; Ding, H.; Hou, Y.; Nie, J.; Hou, F.F.; Kahn, M.; Liu, Y. Targeted inhibition of β-catenin/CBP signaling ameliorates renal interstitial fibrosis. J. Am. Soc. Nephrol. 2011, 22, 1642–1653. [Google Scholar] [CrossRef] [PubMed]
  12. Mizuguchi, Y.; Specht, S.; Lunz, J.G.; Isse, K.; Corbitt, N.; Takizawa, T.; Demetris, A.J. Cooperation of p300 and PCAF in the control of microRNA 200c/141 transcription and epithelial characteristics. PLoS ONE 2012, 7, e32449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, J.-Q.; Yan, F.-Q.; Wang, L.-H.; Yin, W.-J.; Chang, T.-Y.; Liu, J.-P.; Wu, K.-J. Identification of new hypoxia-regulated epithelial-mesenchymal transition marker genes labeled by H3K4 acetylation. Genes. Chromosomes Cancer 2020, 59, 73–83. [Google Scholar] [CrossRef]
  14. Yang, M.-H.; Hsu, D.S.-S.; Wang, H.-W.; Wang, H.-J.; Lan, H.-Y.; Yang, W.-H.; Huang, C.-H.; Kao, S.-Y.; Tzeng, C.-H.; Tai, S.-K.; et al. Bmi1 is essential in Twist1-induced epithelial-mesenchymal transition. Nat. Cell Biol. 2010, 12, 982–992. [Google Scholar] [CrossRef]
  15. Dhasarathy, A.; Phadke, D.; Mav, D.; Shah, R.R.; Wade, P.A. The transcription factors Snail and Slug activate the transforming growth factor-beta signaling pathway in breast cancer. PLoS ONE 2011, 6, e26514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kapoor-Vazirani, P.; Kagey, J.D.; Powell, D.R.; Vertino, P.M. Role of hMOF-dependent histone H4 lysine 16 acetylation in the maintenance of TMS1/ASC gene activity. Cancer Res. 2008, 68, 6810–6821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Højfeldt, J.W.; Agger, K.; Helin, K. Histone lysine demethylases as targets for anticancer therapy. Nat. Rev. Drug Discov. 2013, 12, 917–930. [Google Scholar] [CrossRef] [PubMed]
  18. Tee, A.E.; Ling, D.; Nelson, C.; Atmadibrata, B.; Dinger, M.E.; Xu, N.; Mizukami, T.; Liu, P.Y.; Liu, B.; Cheung, B.; et al. The histone demethylase JMJD1A induces cell migration and invasion by up-regulating the expression of the long noncoding RNA MALAT1. Oncotarget 2014, 5, 1793–1804. [Google Scholar] [CrossRef] [Green Version]
  19. Zhao, L.; Li, W.; Zang, W.; Liu, Z.; Xu, X.; Yu, H.; Yang, Q.; Jia, J. JMJD2B Promotes Epithelial—Mesenchymal Transition by Cooperating with β-Catenin and Enhances Gastric Cancer Metastasis. Clin. Cancer Res. 2013, 19, 6419–6429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Björkman, M.; Östling, P.; Härmä, V.; Virtanen, J.; Mpindi, J.-P.; Rantala, J.; Mirtti, T.; Vesterinen, T.; Lundin, M.; Sankila, A.; et al. Systematic knockdown of epigenetic enzymes identifies a novel histone demethylase PHF8 overexpressed in prostate cancer with an impact on cell proliferation, migration and invasion. Oncogene 2012, 31, 3444–3456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ramadoss, S.; Chen, X.; Wang, C.-Y. Histone demethylase KDM6B promotes epithelial-mesenchymal transition. J. Biol. Chem. 2012, 287, 44508–44517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ferraro, A.; Mourtzoukou, D.; Kosmidou, V.; Avlonitis, S.; Kontogeorgos, G.; Zografos, G.; Pintzas, A. EZH2 is regulated by ERK/AKT and targets integrin alpha2 gene to control Epithelial—Mesenchymal Transition and anoikis in colon cancer cells. Int. J. Biochem. Cell Biol. 2013, 45, 243–254. [Google Scholar] [CrossRef] [PubMed]
  23. Cao, Q.; Yu, J.; Dhanasekaran, S.M.; Kim, J.H.; Mani, R.-S.; Tomlins, S.A.; Mehra, R.; Laxman, B.; Cao, X.; Yu, J.; et al. Repression of E-cadherin by the polycomb group protein EZH2 in cancer. Oncogene 2008, 27, 7274–7284. [Google Scholar] [CrossRef] [Green Version]
  24. Vanharanta, S.; Shu, W.; Brenet, F.; Hakimi, A.A.; Heguy, A.; Viale, A.; Reuter, V.E.; Hsieh, J.J.-D.; Scandura, J.M.; Massagué, J. Epigenetic expansion of VHL-HIF signal output drives multiorgan metastasis in renal cancer. Nat. Med. 2013, 19, 50–56. [Google Scholar] [CrossRef] [Green Version]
  25. Dong, C.; Wu, Y.; Wang, Y.; Wang, C.; Kang, T.; Rychahou, P.G.; Chi, Y.-I.; Evers, B.M.; Zhou, B.P. Interaction with Suv39H1 is critical for Snail-mediated E-cadherin repression in breast cancer. Oncogene 2013, 32, 1351–1362. [Google Scholar] [CrossRef] [Green Version]
  26. Dong, C.; Wu, Y.; Yao, J.; Wang, Y.; Yu, Y.; Rychahou, P.G.; Evers, B.M.; Zhou, B.P. G9a interacts with Snail and is critical for Snail-mediated E-cadherin repression in human breast cancer. J. Clin. Investig. 2012, 122, 1469–1486. [Google Scholar] [CrossRef] [Green Version]
  27. Lee, J.S.; Kim, Y.; Bhin, J.; Shin, H.-J.R.; Nam, H.J.; Lee, S.H.; Yoon, J.-B.; Binda, O.; Gozani, O.; Hwang, D.; et al. Hypoxia-induced methylation of a pontin chromatin remodeling factor. Proc. Natl. Acad. Sci. USA 2011, 108, 13510–13515. [Google Scholar] [CrossRef] [Green Version]
  28. Zhang, H.; Cai, K.; Wang, J.; Wang, X.; Cheng, K.; Shi, F.; Jiang, L.; Zhang, Y.; Dou, J. MiR-7, Inhibited Indirectly by LincRNA HOTAIR, Directly Inhibits SETDB1 and Reverses the EMT of Breast Cancer Stem Cells by Downregulating the STAT3 Pathway. Stem Cells 2014, 32, 2858–2868. [Google Scholar] [CrossRef] [PubMed]
  29. Regina, C.; Compagnone, M.; Peschiaroli, A.; Lena, A.; Annicchiarico-Petruzzelli, M.; Piro, M.C.; Melino, G.; Candi, E. Setdb1, a novel interactor of ΔNp63, is involved in breast tumorigenesis. Oncotarget 2016, 7, 28836–28848. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, P.-C.; Lu, J.-W.; Yang, J.-Y.; Lin, I.-H.; Ou, D.-L.; Lin, Y.-H.; Chou, K.-H.; Huang, W.-F.; Wang, W.-P.; Huang, Y.-L.; et al. H3K9 Histone Methyltransferase, KMT1E/SETDB1, Cooperates with the SMAD2/3 Pathway to Suppress Lung Cancer Metastasis. Cancer Res. 2014, 74, 7333–7343. [Google Scholar] [CrossRef] [Green Version]
  31. Yang, F.; Sun, L.; Li, Q.; Han, X.; Lei, L.; Zhang, H.; Shang, Y. SET8 promotes epithelial-mesenchymal transition and confers TWIST dual transcriptional activities. EMBO J. 2012, 31, 110–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Yang, H.; Mizzen, C.A. The multiple facets of histone H4-lysine 20 methylation. Biochem. Cell Biol. 2009, 87, 151–161. [Google Scholar] [CrossRef]
  33. Fraga, M.F.; Ballestar, E.; Villar-Garea, A.; Boix-Chornet, M.; Espada, J.; Schotta, G.; Bonaldi, T.; Haydon, C.; Ropero, S.; Petrie, K.; et al. Loss of acetylation at Lys16 and trimethylation at Lys20 of histone H4 is a common hallmark of human cancer. Nat. Genet. 2005, 37, 391–400. [Google Scholar] [CrossRef] [PubMed]
  34. Shinchi, Y.; Hieda, M.; Nishioka, Y.; Matsumoto, A.; Yokoyama, Y.; Kimura, H.; Matsuura, S.; Matsuura, N. SUV420H2 suppresses breast cancer cell invasion through down regulation of the SH2 domain-containing focal adhesion protein tensin-3. Exp. Cell Res. 2015, 334, 90–99. [Google Scholar] [CrossRef] [PubMed]
  35. Sánchez-Tilló, E.; Lázaro, A.; Torrent, R.; Cuatrecasas, M.; Vaquero, E.C.; Castells, A.; Engel, P.; Postigo, A. ZEB1 represses E-cadherin and induces an EMT by recruiting the SWI/SNF chromatin-remodeling protein BRG1. Oncogene 2010, 29, 3490–3500. [Google Scholar] [CrossRef] [Green Version]
  36. Jordan, N.V.; Prat, A.; Abell, A.N.; Zawistowski, J.S.; Sciaky, N.; Karginova, O.A.; Zhou, B.; Golitz, B.T.; Perou, C.M.; Johnson, G.L. SWI/SNF chromatin-remodeling factor Smarcd3/Baf60c controls epithelial-mesenchymal transition by inducing Wnt5a signaling. Mol. Cell. Biol. 2013, 33, 3011–3025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Maeda, G.; Chiba, T.; Aoba, T.; Imai, K. Epigenetic inactivation of E-cadherin by promoter hypermethylation in oral carcinoma cells. Odontology 2007, 95, 24–29. [Google Scholar] [CrossRef] [PubMed]
  38. Bernstein, B.E.; Mikkelsen, T.S.; Xie, X.; Kamal, M.; Huebert, D.J.; Cuff, J.; Fry, B.; Meissner, A.; Wernig, M.; Plath, K.; et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 2006, 125, 315–326. [Google Scholar] [CrossRef] [Green Version]
  39. Peinado, H.; Ballestar, E.; Esteller, M.; Cano, A. Snail mediates E-cadherin repression by the recruitment of the Sin3A/histone deacetylase 1 (HDAC1)/HDAC2 complex. Mol. Cell. Biol. 2004, 24, 306–319. [Google Scholar] [CrossRef] [Green Version]
  40. Tripathi, M.K.; Misra, S.; Khedkar, S.V.; Hamilton, N.; Irvin-Wilson, C.; Sharan, C.; Sealy, L.; Chaudhuri, G. Regulation of BRCA2 Gene Expression by the SLUG Repressor Protein in Human Breast Cells. J. Biol. Chem. 2005, 280, 17163–17171. [Google Scholar] [CrossRef] [Green Version]
  41. Aghdassi, A.; Sendler, M.; Guenther, A.; Mayerle, J.; Behn, C.-O.; Heidecke, C.-D.; Friess, H.; Büchler, M.; Evert, M.; Lerch, M.M.; et al. Recruitment of histone deacetylases HDAC1 and HDAC2 by the transcriptional repressor ZEB1 downregulates E-cadherin expression in pancreatic cancer. Gut 2012, 61, 439–448. [Google Scholar] [CrossRef]
  42. Zhang, X.; Yuan, Z.; Zhang, Y.; Yong, S.; Salas-Burgos, A.; Koomen, J.; Olashaw, N.; Parsons, J.T.; Yang, X.-J.; Dent, S.R.; et al. HDAC6 modulates cell motility by altering the acetylation level of cortactin. Mol. Cell 2007, 27, 197–213. [Google Scholar] [CrossRef] [Green Version]
  43. Sandoval, J.; Esteller, M. Cancer epigenomics: Beyond genomics. Curr. Opin. Genet. Dev. 2012, 22, 50–55. [Google Scholar] [CrossRef]
  44. Carmona, F.J.; Davalos, V.; Vidal, E.; Gomez, A.; Heyn, H.; Hashimoto, Y.; Vizoso, M.; Martinez-Cardus, A.; Sayols, S.; Ferreira, H.J.; et al. A comprehensive DNA methylation profile of epithelial-to-mesenchymal transition. Cancer Res. 2014, 74, 5608–5619. [Google Scholar] [CrossRef] [Green Version]
  45. Fukagawa, A.; Ishii, H.; Miyazawa, K.; Saitoh, M. δEF1 associates with DNMT1 and maintains DNA methylation of the E-cadherin promoter in breast cancer cells. Cancer Med. 2015, 4, 125–135. [Google Scholar] [CrossRef]
  46. Papageorgis, P.; Lambert, A.W.; Ozturk, S.; Gao, F.; Pan, H.; Manne, U.; Alekseyev, Y.O.; Thiagalingam, A.; Abdolmaleky, H.M.; Lenburg, M.; et al. Smad signaling is required to maintain epigenetic silencing during breast cancer progression. Cancer Res. 2010, 70, 968–978. [Google Scholar] [CrossRef] [Green Version]
  47. Davalos, V.; Moutinho, C.; Villanueva, A.; Boque, R.; Silva, P.; Carneiro, F.; Esteller, M. Dynamic epigenetic regulation of the microRNA-200 family mediates epithelial and mesenchymal transitions in human tumorigenesis. Oncogene 2012, 31, 2062–2074. [Google Scholar] [CrossRef] [PubMed]
  48. Vrba, L.; Jensen, T.J.; Garbe, J.C.; Heimark, R.L.; Cress, A.E.; Dickinson, S.; Stampfer, M.R.; Futscher, B.W. Role for DNA methylation in the regulation of miR-200c and miR-141 expression in normal and cancer cells. PLoS ONE 2010, 5, e8697. [Google Scholar] [CrossRef] [PubMed]
  49. Wiklund, E.D.; Bramsen, J.B.; Hulf, T.; Dyrskjøt, L.; Ramanathan, R.; Hansen, T.B.; Villadsen, S.B.; Gao, S.; Ostenfeld, M.S.; Borre, M.; et al. Coordinated epigenetic repression of the miR-200 family and miR-205 in invasive bladder cancer. Int. J. Cancer 2011, 128, 1327–1334. [Google Scholar] [CrossRef]
  50. Sansom, O.J.; Maddison, K.; Clarke, A.R. Mechanisms of disease: Methyl-binding domain proteins as potential therapeutic targets in cancer. Nat. Clin. Pract. Oncol. 2007, 4, 305–315. [Google Scholar] [CrossRef]
  51. McCabe, M.T.; Brandes, J.C.; Vertino, P.M. Cancer DNA methylation: Molecular mechanisms and clinical implications. Clin. Cancer Res. 2009, 15, 3927–3937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kohli, R.M.; Zhang, Y. TET enzymes, TDG and the dynamics of DNA demethylation. Nature 2013, 502, 472–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kim, V.N.; Han, J.; Siomi, M.C. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell Biol. 2009, 10, 126–139. [Google Scholar] [CrossRef]
  54. Park, S.-M.; Gaur, A.B.; Lengyel, E.; Peter, M.E. The miR-200 family determines the epithelial phenotype of cancer cells by targeting the E-cadherin repressors ZEB1 and ZEB2. Genes Dev. 2008, 22, 894–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Khew-Goodall, Y.; Goodall, G.J. The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef]
  56. Xia, H.; Ng, S.S.; Jiang, S.; Cheung, W.K.C.; Sze, J.; Bian, X.-W.; Kung, H.-F.; Lin, M.C. miR-200a-mediated downregulation of ZEB2 and CTNNB1 differentially inhibits nasopharyngeal carcinoma cell growth, migration and invasion. Biochem. Biophys. Res. Commun. 2010, 391, 535–541. [Google Scholar] [CrossRef]
  57. Varambally, S.; Cao, Q.; Mani, R.-S.; Shankar, S.; Wang, X.; Ateeq, B.; Laxman, B.; Cao, X.; Jing, X.; Ramnarayanan, K.; et al. Genomic loss of microRNA-101 leads to overexpression of histone methyltransferase EZH2 in cancer. Science 2008, 322, 1695–1699. [Google Scholar] [CrossRef] [Green Version]
  58. Abba, M.L.; Patil, N.; Leupold, J.H.; Allgayer, H. MicroRNA Regulation of Epithelial to Mesenchymal Transition. J. Clin. Med. 2016, 5, 8. [Google Scholar] [CrossRef] [PubMed]
  59. Ma, L.; Young, J.; Prabhala, H.; Pan, E.; Mestdagh, P.; Muth, D.; Teruya-Feldstein, J.; Reinhardt, F.; Onder, T.T.; Valastyan, S.; et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nat. Cell Biol. 2010, 12, 247–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Schmitz, S.U.; Grote, P.; Herrmann, B.G. Mechanisms of long noncoding RNA function in development and disease. Cell. Mol. Life Sci. 2016, 73, 2491–2509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Gupta, R.A.; Shah, N.; Wang, K.C.; Kim, J.; Horlings, H.M.; Wong, D.J.; Tsai, M.-C.; Hung, T.; Argani, P.; Rinn, J.L.; et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010, 464, 1071–1076. [Google Scholar] [CrossRef]
  62. Tsai, M.-C.; Manor, O.; Wan, Y.; Mosammaparast, N.; Wang, J.K.; Lan, F.; Shi, Y.; Segal, E.; Chang, H.Y. Long noncoding RNA as modular scaffold of histone modification complexes. Science 2010, 329, 689–693. [Google Scholar] [CrossRef] [Green Version]
  63. Richards, E.J.; Zhang, G.; Li, Z.-P.; Permuth-Wey, J.; Challa, S.; Li, Y.; Kong, W.; Dan, S.; Bui, M.M.; Coppola, D.; et al. Long non-coding RNAs (LncRNA) regulated by transforming growth factor (TGF) β: LncRNA-hit-mediated TGFβ-induced epithelial to mesenchymal transition in mammary epithelia. J. Biol. Chem. 2015, 290, 6857–6867. [Google Scholar] [CrossRef] [Green Version]
  64. Marconi, A.; Atzei, P.; Panza, C.; Fila, C.; Tiberio, R.; Truzzi, F.; Wachter, T.; Leverkus, M.; Pincelli, C. FLICE/caspase-8 activation triggers anoikis induced by beta1-integrin blockade in human keratinocytes. J. Cell Sci. 2004, 117, 5815–5823. [Google Scholar] [CrossRef] [Green Version]
  65. Mawji, I.A.; Simpson, C.D.; Hurren, R.; Gronda, M.; Williams, M.A.; Filmus, J.; Jonkman, J.; Da Costa, R.S.; Wilson, B.C.; Thomas, M.P.; et al. Critical role for Fas-associated death domain-like interleukin-1-converting enzyme-like inhibitory protein in anoikis resistance and distant tumor formation. J. Natl. Cancer Inst. 2007, 99, 811–822. [Google Scholar] [CrossRef] [Green Version]
  66. Tan, M.; Asad, M.; Heong, V.; Wong, M.K.; Tan, T.Z.; Ye, J.; Kuay, K.T.; Thiery, J.P.; Scott, C.; Huang, R.Y.-J. The FZD7-TWIST1 axis is responsible for anoikis resistance and tumorigenesis in ovarian carcinoma. Mol. Oncol. 2019, 13, 757–780. [Google Scholar] [CrossRef] [Green Version]
  67. Smith, H.A.; Kang, Y. The metastasis-promoting roles of tumor-associated immune cells. J. Mol. Med. 2013, 91, 411–429. [Google Scholar] [CrossRef]
  68. McDonald, B.; Spicer, J.; Giannais, B.; Fallavollita, L.; Brodt, P.; Ferri, L.E. Systemic inflammation increases cancer cell adhesion to hepatic sinusoids by neutrophil mediated mechanisms. Int. J. Cancer 2009, 125, 1298–1305. [Google Scholar] [CrossRef]
  69. Khatib, A.M.; Kontogiannea, M.; Fallavollita, L.; Jamison, B.; Meterissian, S.; Brodt, P. Rapid induction of cytokine and E-selectin expression in the liver in response to metastatic tumor cells. Cancer Res. 1999, 59, 1356–1361. [Google Scholar]
  70. Läubli, H.; Stevenson, J.L.; Varki, A.; Varki, N.M.; Borsig, L. L-Selectin Facilitation of Metastasis Involves Temporal Induction of Fut7-Dependent Ligands at Sites of Tumor Cell Arrest. Cancer Res. 2006, 66, 1536–1542. [Google Scholar] [CrossRef] [Green Version]
  71. Guan, R.J.; Ford, H.L.; Fu, Y.; Li, Y.; Shaw, L.M.; Pardee, A.B. Drg-1 as a differentiation-related, putative metastatic suppressor gene in human colon cancer. Cancer Res. 2000, 60, 749–755. [Google Scholar]
  72. Bandyopadhyay, S.; Pai, S.K.; Hirota, S.; Hosobe, S.; Takano, Y.; Saito, K.; Piquemal, D.; Commes, T.; Watabe, M.; Gross, S.C.; et al. Role of the putative tumor metastasis suppressor gene Drg-1 in breast cancer progression. Oncogene 2004, 23, 5675–5681. [Google Scholar] [CrossRef] [Green Version]
  73. Tang, Y.-N.; Ding, W.-Q.; Guo, X.-J.; Yuan, X.-W.; Wang, D.-M.; Song, J.-G. Epigenetic regulation of Smad2 and Smad3 by profilin-2 promotes lung cancer growth and metastasis. Nat. Commun. 2015, 6, 8230. [Google Scholar] [CrossRef] [PubMed]
  74. Du, D.; Katsuno, Y.; Meyer, D.; Budi, E.H.; Chen, S.-H.; 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] [Green Version]
  75. Wakabayashi, Y.; Tamiya, T.; Takada, I.; Fukaya, T.; Sugiyama, Y.; Inoue, N.; Kimura, A.; Morita, R.; Kashiwagi, I.; Takimoto, T.; et al. Histone 3 lysine 9 (H3K9) methyltransferase recruitment to the interleukin-2 (IL-2) promoter is a mechanism of suppression of IL-2 transcription by the transforming growth factor-β-smad pathway. J. Biol. Chem. 2011, 286, 35456–35465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Massagué, J. TGFβ signalling in context. Nat. Rev. Mol. Cell Biol. 2012, 13, 616–630. [Google Scholar] [CrossRef]
  77. David, C.J.; Massagué, J. Contextual determinants of TGFβ action in development, immunity and cancer. Nat. Rev. Mol. Cell Biol. 2018, 19, 419–435. [Google Scholar] [CrossRef] [PubMed]
  78. Li, D.; Sun, H.; Sun, W.-J.; Bao, H.-B.; Si, S.-H.; Fan, J.-L.; Lin, P.; Cui, R.-J.; Pan, Y.-J.; Wen, S.-M.; et al. Role of RbBP5 and H3K4me3 in the vicinity of Snail transcription start site during epithelial-mesenchymal transition in prostate cancer cell. Oncotarget 2016, 7, 65553–65567. [Google Scholar] [CrossRef] [Green Version]
  79. Chen, M.-W.; Hua, K.-T.; Kao, H.-J.; Chi, C.-C.; Wei, L.-H.; Johansson, G.; Shiah, S.-G.; Chen, P.-S.; Jeng, Y.-M.; Cheng, T.-Y.; et al. H3K9 histone methyltransferase G9a promotes lung cancer invasion and metastasis by silencing the cell adhesion molecule Ep-CAM. Cancer Res. 2010, 70, 7830–7840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Chen, H.; Lorton, B.; Gupta, V.; Shechter, D. A TGFβ-PRMT5-MEP50 axis regulates cancer cell invasion through histone H3 and H4 arginine methylation coupled transcriptional activation and repression. Oncogene 2017, 36, 373–386. [Google Scholar] [CrossRef] [Green Version]
  81. Teng, Y.-C.; Lee, C.-F.; Li, Y.-S.; Chen, Y.-R.; Hsiao, P.-W.; Chan, M.-Y.; Lin, F.-M.; Huang, H.-D.; Chen, Y.-T.; Jeng, Y.-M.; et al. Histone demethylase RBP2 promotes lung tumorigenesis and cancer metastasis. Cancer Res. 2013, 73, 4711–4721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Asgarova, A.; Asgarov, K.; Godet, Y.; Peixoto, P.; Nadaradjane, A.; Boyer-Guittaut, M.; Galaine, J.; Guenat, D.; Mougey, V.; Perrard, J.; et al. PD-L1 expression is regulated by both DNA methylation and NF-kB during EMT signaling in non-small cell lung carcinoma. Oncoimmunology 2018, 7, e1423170. [Google Scholar] [CrossRef] [Green Version]
  83. Messier, T.L.; Gordon, J.A.R.; Boyd, J.R.; Tye, C.E.; Browne, G.; Stein, J.L.; Lian, J.B.; Stein, G.S. Histone H3 lysine 4 acetylation and methylation dynamics define breast cancer subtypes. Oncotarget 2016, 7, 5094–5109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Judes, G.; Dubois, L.; Rifaï, K.; Idrissou, M.; Mishellany, F.; Pajon, A.; Besse, S.; Daures, M.; Degoul, F.; Bignon, Y.-J.; et al. TIP60: An actor in acetylation of H3K4 and tumor development in breast cancer. Epigenomics 2018, 10, 1415–1430. [Google Scholar] [CrossRef]
  85. Zhao, M.; Mishra, L.; Deng, C.-X. The role of TGF-β/SMAD4 signaling in cancer. Int. J. Biol. Sci. 2018, 14, 111–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wang, L.-T.; Liu, K.-Y.; Jeng, W.-Y.; Chiang, C.-M.; Chai, C.-Y.; Chiou, S.-S.; Huang, M.-S.; Yokoyama, K.K.; Wang, S.-N.; Huang, S.-K.; et al. PCAF-mediated acetylation of ISX recruits BRD4 to promote epithelial-mesenchymal transition. EMBO Rep. 2020, 21, e48795. [Google Scholar] [CrossRef]
  87. Cho, M.-H.; Park, J.-H.; Choi, H.-J.; Park, M.-K.; Won, H.-Y.; Park, Y.-J.; Lee, C.H.; Oh, S.-H.; Song, Y.-S.; Kim, H.S.; et al. DOT1L cooperates with the c-Myc-p300 complex to epigenetically derepress CDH1 transcription factors in breast cancer progression. Nat. Commun. 2015, 6, 7821. [Google Scholar] [CrossRef] [Green Version]
  88. Tang, Z.; Ding, S.; Huang, H.; Luo, P.; Qing, B.; Zhang, S.; Tang, R. HDAC1 triggers the proliferation and migration of breast cancer cells via upregulation of interleukin-8. Biol. Chem. 2017, 398, 1347–1356. [Google Scholar] [CrossRef]
  89. Karamouzis, M.V.; Badra, F.A.; Papavassiliou, A.G. Breast cancer: The upgraded role of HER-3 and HER-4. Int. J. Biochem. Cell Biol. 2007, 39, 851–856. [Google Scholar] [CrossRef]
  90. Kwak, S.-M.; Seo, J.; Hwang, J.-T.; Sung, G.-J.; Song, J.-H.; Jeong, J.-H.; Lee, S.-H.; Yoon, H.-G.; Choi, H.-K.; Choi, K.-C. EGFR-c-Src-Mediated HDAC3 Phosphorylation Exacerbates Invasion of Breast Cancer Cells. Cells 2019, 8, 930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Tang, X.; Li, G.; Su, F.; Cai, Y.; Shi, L.; Meng, Y.; Liu, Z.; Sun, J.; Wang, M.; Qian, M.; et al. HDAC8 cooperates with SMAD3/4 complex to suppress SIRT7 and promote cell survival and migration. Nucleic Acids Res. 2020, 48, 2912–2923. [Google Scholar] [CrossRef]
  92. Li, Z.; Hou, P.; Fan, D.; Dong, M.; Ma, M.; Li, H.; Yao, R.; Li, Y.; Wang, G.; Geng, P.; et al. The degradation of EZH2 mediated by lncRNA ANCR attenuated the invasion and metastasis of breast cancer. Cell Death Differ. 2017, 24, 59–71. [Google Scholar] [CrossRef] [Green Version]
  93. Geng, P.; Zhang, Y.; Liu, X.; Zhang, N.; Liu, Y.; Liu, X.; Lin, C.; Yan, X.; Li, Z.; Wang, G.; et al. Automethylation of protein arginine methyltransferase 7 and its impact on breast cancer progression. FASEB J. 2017, 31, 2287–2300. [Google Scholar] [CrossRef] [Green Version]
  94. Nakakido, M.; Deng, Z.; Suzuki, T.; Dohmae, N.; Nakamura, Y.; Hamamoto, R. PRMT6 increases cytoplasmic localization of p21CDKN1A in cancer cells through arginine methylation and makes more resistant to cytotoxic agents. Oncotarget 2015, 6, 30957–30967. [Google Scholar] [CrossRef] [Green Version]
  95. Jiang, H.; Zhou, Z.; Jin, S.; Xu, K.; Zhang, H.; Xu, J.; Sun, Q.; Wang, J.; Xu, J. PRMT9 promotes hepatocellular carcinoma invasion and metastasis via activating PI3K/Akt/GSK-3β/Snail signaling. Cancer Sci. 2018, 109, 1414–1427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Cao, J.; Liu, Z.; Cheung, W.K.C.; Zhao, M.; Chen, S.Y.; Chan, S.W.; Booth, C.J.; Nguyen, D.X.; Yan, Q. Histone demethylase RBP2 is critical for breast cancer progression and metastasis. Cell Rep. 2014, 6, 868–877. [Google Scholar] [CrossRef] [Green Version]
  97. Enkhbaatar, Z.; Terashima, M.; Oktyabri, D.; Tange, S.; Ishimura, A.; Yano, S.; Suzuki, T. KDM5B histone demethylase controls epithelial-mesenchymal transition of cancer cells by regulating the expression of the microRNA-200 family. Cell Cycle 2013, 12, 2100–2112. [Google Scholar] [CrossRef]
  98. Tang, B.; Qi, G.; Tang, F.; Yuan, S.; Wang, Z.; Liang, X.; Li, B.; Yu, S.; Liu, J.; Huang, Q.; et al. JARID1B promotes metastasis and epithelial-mesenchymal transition via PTEN/AKT signaling in hepatocellular carcinoma cells. Oncotarget 2015, 6, 12723–12739. [Google Scholar] [CrossRef] [Green Version]
  99. Li, Q.; Shi, L.; Gui, B.; Yu, W.; Wang, J.; Zhang, D.; Han, X.; Yao, Z.; Shang, Y. Binding of the JmjC demethylase JARID1B to LSD1/NuRD suppresses angiogenesis and metastasis in breast cancer cells by repressing chemokine CCL14. Cancer Res. 2011, 71, 6899–6908. [Google Scholar] [CrossRef] [Green Version]
  100. Kim, J.-H.; Sharma, A.; Dhar, S.S.; Lee, S.-H.; Gu, B.; Chan, C.-H.; Lin, H.-K.; Lee, M.G. UTX and MLL4 coordinately regulate transcriptional programs for cell proliferation and invasiveness in breast cancer cells. Cancer Res. 2014, 74, 1705–1717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Choi, H.-J.; Park, J.-H.; Park, M.; Won, H.-Y.; Joo, H.-S.; Lee, C.H.; Lee, J.-Y.; Kong, G. UTX inhibits EMT-induced breast CSC properties by epigenetic repression of EMT genes in cooperation with LSD1 and HDAC1. EMBO Rep. 2015, 16, 1288–1298. [Google Scholar] [CrossRef] [Green Version]
  102. Xi, Y.; Shi, J.; Li, W.; Tanaka, K.; Allton, K.L.; Richardson, D.; Li, J.; Franco, H.L.; Nagari, A.; Malladi, V.S.; et al. Histone modification profiling in breast cancer cell lines highlights commonalities and differences among subtypes. BMC Genom. 2018, 19, 150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Karsli-Ceppioglu, S.; Dagdemir, A.; Judes, G.; Lebert, A.; Penault-Llorca, F.; Bignon, Y.-J.; Bernard-Gallon, D. The Epigenetic Landscape of Promoter Genome-wide Analysis in Breast Cancer. Sci. Rep. 2017, 7, 6597. [Google Scholar] [CrossRef]
  104. Healey, M.A.; Hu, R.; Beck, A.H.; Collins, L.C.; Schnitt, S.J.; Tamimi, R.M.; Hazra, A. Association of H3K9me3 and H3K27me3 repressive histone marks with breast cancer subtypes in the Nurses’ Health Study. Breast Cancer Res. Treat. 2014, 147, 639–651. [Google Scholar] [CrossRef]
  105. Hirukawa, A.; Smith, H.W.; Zuo, D.; Dufour, C.R.; Savage, P.; Bertos, N.; Johnson, R.M.; Bui, T.; Bourque, G.; Basik, M.; et al. Targeting EZH2 reactivates a breast cancer subtype-specific anti-metastatic transcriptional program. Nat. Commun. 2018, 9, 2547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Rifaï, K.; Judes, G.; Idrissou, M.; Daures, M.; Bignon, Y.-J.; Penault-Llorca, F.; Bernard-Gallon, D. SIRT1-dependent epigenetic regulation of H3 and H4 histone acetylation in human breast cancer. Oncotarget 2018, 9, 30661–30678. [Google Scholar] [CrossRef]
  107. Hu, Y.; Zheng, Y.; Dai, M.; Wang, X.; Wu, J.; Yu, B.; Zhang, H.; Cui, Y.; Kong, W.; Wu, H.; et al. G9a and histone deacetylases are crucial for Snail2-mediated E-cadherin repression and metastasis in hepatocellular carcinoma. Cancer Sci. 2019, 110, 3442–3452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Natoni, F.; Diolordi, L.; Santoni, C.; Montani, M.S.G. Sodium butyrate sensitises human pancreatic cancer cells to both the intrinsic and the extrinsic apoptotic pathways. Biochim. Biophys. Acta Mol. Cell Res. 2005, 1745, 318–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Fei, Q.; Shang, K.; Zhang, J.; Chuai, S.; Kong, D.; Zhou, T.; Fu, S.; Liang, Y.; Li, C.; Chen, Z.; et al. Histone methyltransferase SETDB1 regulates liver cancer cell growth through methylation of p53. Nat. Commun. 2015, 6, 8651. [Google Scholar] [CrossRef] [Green Version]
  110. Xia, R.; Jin, F.; Lu, K.; Wan, L.; Xie, M.; Xu, T.; De, W.; Wang, Z. SUZ12 promotes gastric cancer cell proliferation and metastasis by regulating KLF2 and E-cadherin. Tumor Biol. 2015, 36, 5341–5351. [Google Scholar] [CrossRef]
  111. Malouf, G.G.; Taube, J.H.; Lu, Y.; Roysarkar, T.; Panjarian, S.; Estecio, M.R.; Jelinek, J.; Yamazaki, J.; Raynal, N.J.-M.; Long, H.; et al. Architecture of epigenetic reprogramming following Twist1-mediated epithelial-mesenchymal transition. Genome Biol. 2013, 14, R144. [Google Scholar] [CrossRef]
  112. Chaffer, C.L.; Thompson, E.W.; Williams, E.D. Mesenchymal to epithelial transition in development and disease. Cells Tissues Organs 2007, 185, 7–19. [Google Scholar] [CrossRef]
  113. Yang, J.; Weinberg, R.A. Epithelial-mesenchymal transition: At the crossroads of development and tumor metastasis. Dev. Cell 2008, 14, 818–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Neureiter, D.; Herold, C.; Ocker, M. Gastrointestinal Cancer—Only a Deregulation of Stem Cell Differentiation? Int. J. Mol. Med. 2006, 17, 483–489. [Google Scholar] [CrossRef] [Green Version]
  115. Byles, V.; Zhu, L.; Lovaas, J.D.; Chmilewski, L.K.; Wang, J.; Faller, D.V.; Dai, Y. SIRT1 induces EMT by cooperating with EMT transcription factors and enhances prostate cancer cell migration and metastasis. Oncogene 2012, 31, 4619–4629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Roche, J.; Nasarre, P.; Gemmill, R.; Baldys, A.; Pontis, J.; Korch, C.; Guilhot, J.; Ait-Si-Ali, S.; Drabkin, H. Global Decrease of Histone H3K27 Acetylation in ZEB1-Induced Epithelial to Mesenchymal Transition in Lung Cancer Cells. Cancers 2013, 5, 334–356. [Google Scholar] [CrossRef] [PubMed]
  117. Thiery, J.-P.; Chua, K.; Sim, W.J.; Huang, R. Epithelial mesenchymal transition during development in fibrosis and in the progression of carcinoma. Bull. Cancer 2010, 97, 1285–1295. [Google Scholar] [CrossRef] [PubMed]
  118. Delmore, J.E.; Issa, G.C.; Lemieux, M.E.; Rahl, P.B.; Shi, J.; Jacobs, H.M.; Kastritis, E.; Gilpatrick, T.; Paranal, R.M.; Qi, J.; et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 2011, 146, 904–917. [Google Scholar] [CrossRef] [Green Version]
  119. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I.; et al. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef] [Green Version]
  120. Lovén, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 2013, 153, 320–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Hu, Y.; Zhou, J.; Ye, F.; Xiong, H.; Peng, L.; Zheng, Z.; Xu, F.; Cui, M.; Wei, C.; Wang, X.; et al. BRD4 inhibitor inhibits colorectal cancer growth and metastasis. Int. J. Mol. Sci. 2015, 16, 1928–1948. [Google Scholar] [CrossRef] [Green Version]
  122. Shi, J.; Wang, Y.; Zeng, L.; Wu, Y.; Deng, J.; Zhang, Q.; Lin, Y.; Li, J.; Kang, T.; Tao, M.; et al. Disrupting the interaction of BRD4 with diacetylated Twist suppresses tumorigenesis in basal-like breast cancer. Cancer Cell 2014, 25, 210–225. [Google Scholar] [CrossRef] [Green Version]
  123. James, L.I.; Barsyte-Lovejoy, D.; Zhong, N.; Krichevsky, L.; Korboukh, V.K.; Herold, J.M.; MacNevin, C.J.; Norris, J.L.; Sagum, C.A.; Tempel, W.; et al. Discovery of a chemical probe for the L3MBTL3 methyllysine reader domain. Nat. Chem. Biol. 2013, 9, 184–191. [Google Scholar] [CrossRef]
  124. Stuckey, J.I.; Dickson, B.M.; Cheng, N.; Liu, Y.; Norris, J.L.; Cholensky, S.H.; Tempel, W.; Qin, S.; Huber, K.G.; Sagum, C.; et al. A cellular chemical probe targeting the chromodomains of Polycomb repressive complex 1. Nat. Chem. Biol. 2016, 12, 180–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Liang, X.; Zeng, J.; Wang, L.; Shen, L.; Ma, X.; Li, S.; Wu, Y.; Ma, L.; Ci, X.; Guo, Q.; et al. Histone demethylase RBP2 promotes malignant progression of gastric cancer through TGF-β1-(p-Smad3)-RBP2-E-cadherin-Smad3 feedback circuit. Oncotarget 2015, 6, 17661–17674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Ketola, K.; Hilvo, M.; Hyötyläinen, T.; Vuoristo, A.; Ruskeepää, A.-L.; Orešič, M.; Kallioniemi, O.; Iljin, K. Salinomycin inhibits prostate cancer growth and migration via induction of oxidative stress. Br. J. Cancer 2012, 106, 99–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wang, Y. Effects of salinomycin on cancer stem cell in human lung adenocarcinoma A549 cells. Med. Chem. 2011, 7, 106–111. [Google Scholar] [CrossRef] [PubMed]
  128. Gong, C.; Yao, H.; Liu, Q.; Chen, J.; Shi, J.; Su, F.; Song, E. Markers of tumor-initiating cells predict chemoresistance in breast cancer. PLoS ONE 2010, 5, e15630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Bardsley, M.R.; Horváth, V.J.; Asuzu, D.T.; Lorincz, A.; Redelman, D.; Hayashi, Y.; Popko, L.N.; Young, D.L.; Lomberk, G.A.; Urrutia, R.A.; et al. Kitlow stem cells cause resistance to Kit/platelet-derived growth factor alpha inhibitors in murine gastrointestinal stromal tumors. Gastroenterology 2010, 139, 942–952. [Google Scholar] [CrossRef] [Green Version]
  130. Fuchs, D.; Daniel, V.; Sadeghi, M.; Opelz, G.; Naujokat, C. Salinomycin overcomes ABC transporter-mediated multidrug and apoptosis resistance in human leukemia stem cell-like KG-1a cells. Biochem. Biophys. Res. Commun. 2010, 394, 1098–1104. [Google Scholar] [CrossRef] [PubMed]
  131. Kiesslich, T.; Pichler, M.; Neureiter, D. Epigenetic control of epithelial-mesenchymal-transition in human cancer. Mol. Clin. Oncol. 2013, 1, 3–11. [Google Scholar] [CrossRef] [Green Version]
  132. Stintzing, S.; Kemmerling, R.; Kiesslich, T.; Alinger, B.; Ocker, M.; Neureiter, D. Myelodysplastic syndrome and histone deacetylase inhibitors: “To be or not to be acetylated”? J. Biomed. Biotechnol. 2011, 2011, 214143. [Google Scholar] [CrossRef]
  133. Lane, A.A.; Chabner, B.A. Histone deacetylase inhibitors in cancer therapy. J. Clin. Oncol. 2009, 27, 5459–5468. [Google Scholar] [CrossRef] [Green Version]
  134. Batty, N.; Malouf, G.G.; Issa, J.P.J. Histone deacetylase inhibitors as anti-neoplastic agents. Cancer Lett. 2009, 280, 192–200. [Google Scholar] [CrossRef]
  135. Di Fazio, P.; Schneider-Stock, R.; Neureiter, D.; Okamoto, K.; Wissniowski, T.; Gahr, S.; Quint, K.; Meissnitzer, M.; Alinger, B.; Montalbano, R.; et al. The pan-deacetylase inhibitor panobinostat inhibits growth of hepatocellular carcinoma models by alternative pathways of apoptosis. Cell. Oncol. 2010, 32, 285–300. [Google Scholar] [CrossRef]
  136. Di Fazio, P.; Montalbano, R.; Quint, K.; Alinger, B.; Kemmerling, R.; Kiesslich, T.; Ocker, M.; Neureiter, D. The pan-deacetylase inhibitor panobinostat modulates the expression of epithelial-mesenchymal transition markers in hepatocellular carcinoma models. Oncol. Lett. 2013, 5, 127–134. [Google Scholar] [CrossRef] [Green Version]
  137. Neureiter, D.; Zopf, S.; Leu, T.; Dietze, O.; Hauser-Kronberger, C.; Hahn, E.G.; Herold, C.; Ocker, M. Apoptosis, proliferation and differentiation patterns are influenced by Zebularine and SAHA in pancreatic cancer models. Scand. J. Gastroenterol. 2007, 42, 103–116. [Google Scholar] [CrossRef]
  138. Takai, N.; Desmond, J.C.; Kumagai, T.; Gui, D.; Said, J.W.; Whittaker, S.; Miyakawa, I.; Koeffler, H.P. Histone Deacetylase Inhibitors Have a Profound Antigrowth Activity in Endometrial Cancer Cells. Clin. Cancer Res. 2004, 10, 1141–1149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Meidhof, S.; Brabletz, S.; Lehmann, W.; Preca, B.-T.; Mock, K.; Ruh, M.; Schüler, J.; Berthold, M.; Weber, A.; Burk, U.; et al. ZEB1-associated drug resistance in cancer cells is reversed by the class I HDAC inhibitor mocetinostat. EMBO Mol. Med. 2015, 7, 831–847. [Google Scholar] [CrossRef] [PubMed]
  140. Zhuang, J.; Zhang, J.; Wu, M.; Zhang, Y. A Dynamic 3D Tumor Spheroid Chip Enables More Accurate Nanomedicine Uptake Evaluation. Adv. Sci. 2019, 6, 1901462. [Google Scholar] [CrossRef]
  141. Ediriweera, M.K.; Cho, S.K. Targeting miRNAs by histone deacetylase inhibitors (HDACi): Rationalizing epigenetics-based therapies for breast cancer. Pharmacol. Ther. 2020, 206, 107437. [Google Scholar] [CrossRef] [PubMed]
  142. Connolly, R.M.; Rudek, M.A.; Piekarz, R. Entinostat: A promising treatment option for patients with advanced breast cancer. Future Oncol. 2017, 13, 1137–1148. [Google Scholar] [CrossRef]
  143. Yeruva, S.L.H.; Zhao, F.; Miller, K.D.; Tevaarwerk, A.J.; Wagner, L.I.; Gray, R.J.; Sparano, J.A.; Connolly, R.M. E2112: Randomized phase III trial of endocrine therapy plus entinostat/placebo in patients with hormone receptor-positive advanced breast cancer. NPJ Breast Cancer 2018, 4, 1. [Google Scholar] [CrossRef] [Green Version]
  144. Terranova-Barberio, M.; Thomas, S.; Ali, N.; Pawlowska, N.; Park, J.; Krings, G.; Rosenblum, M.D.; Budillon, A.; Munster, P.N. HDAC inhibition potentiates immunotherapy in triple negative breast cancer. Oncotarget 2017, 8, 114156–114172. [Google Scholar] [CrossRef] [Green Version]
  145. Wang, S.-C.M.; Dowhan, D.H.; Muscat, G.E.O. Epigenetic arginine methylation in breast cancer: Emerging therapeutic strategies. J. Mol. Endocrinol. 2019, 62, R223–R237. [Google Scholar] [CrossRef] [PubMed]
  146. Chiang, K.; Davies, C.C. Linking PRMT5 to breast cancer stem cells: New therapeutic opportunities? Mol. Cell. Oncol. 2018, 5, e1441628. [Google Scholar] [CrossRef] [Green Version]
  147. Shen, Y.; Zhong, J.; Liu, J.; Liu, K.; Zhao, J.; Xu, T.; Zeng, T.; Li, Z.; Chen, Y.; Ding, W.; et al. Protein arginine N-methyltransferase 2 reverses tamoxifen resistance in breast cancer cells through suppression of ER-α36. Oncol. Rep. 2018, 39, 2604–2612. [Google Scholar] [CrossRef]
  148. Hiken, J.F.; McDonald, J.I.; Decker, K.F.; Sanchez, C.; Hoog, J.; VanderKraats, N.D.; Jung, K.L.; Akinhanmi, M.; Rois, L.E.; Ellis, M.J.; et al. Epigenetic activation of the prostaglandin receptor EP4 promotes resistance to endocrine therapy for breast cancer. Oncogene 2017, 36, 2319–2327. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, J.-Q.; Wu, K.-J. Epigenetic regulation of epithelial-mesenchymal transition by hypoxia in cancer: Targets and therapy. Curr. Pharm. Des. 2015, 21, 1272–1278. [Google Scholar] [CrossRef] [PubMed]
  150. Lu, C.; Sidoli, S.; Kulej, K.; Ross, K.; Wu, C.H.; Garcia, B.A. Coordination between TGF-β cellular signaling and epigenetic regulation during epithelial to mesenchymal transition. Epigenet. Chromatin 2019, 12, 11. [Google Scholar] [CrossRef]
  151. Open-Label, Multicenter, Phase 1/2 Study of Tazemetostat (EZH2 Histone Methyl Transferase [HMT] Inhibitor) as a Single Agent in Subjects with Adv. Solid Tumors or with B-cell Lymphomas and Tazemetostat in Combination with Prednisolone in Subjects with DLBCL. Available online: https://clinicaltrials.gov/ct2/show/NCT01897571 (accessed on 21 January 2021).
  152. A Study to Investigate the Safety, Pharmacokinetics, Pharmacodynamics and Clinical Activity of GSK2816126 in Subjects with Relapsed/Refractory Diffuse Large B Cell Lymphoma, Transformed Follicular Lymphoma, Other Non-Hodgkin’s Lymphomas, Solid Tumors and Multiple Myeloma. Available online: https://clinicaltrials.gov/ct2/show/NCT02082977 (accessed on 21 January 2021).
  153. Greiner, D.; Bonaldi, T.; Eskeland, R.; Roemer, E.; Imhof, A. Identification of a specific inhibitor of the histone methyltransferase SU(VAR)3-9. Nat. Chem. Biol. 2005, 1, 143–145. [Google Scholar] [CrossRef]
  154. Lakshmikuttyamma, A.; Scott, S.A.; DeCoteau, J.F.; Geyer, C.R. Reexpression of epigenetically silenced AML tumor suppressor genes by SUV39H1 inhibition. Oncogene 2010, 29, 576–588. [Google Scholar] [CrossRef] [Green Version]
  155. Kubicek, S.; O’Sullivan, R.J.; August, E.M.; Hickey, E.R.; Zhang, Q.; Teodoro, M.L.; Rea, S.; Mechtler, K.; Kowalski, J.A.; Homon, C.A.; et al. Reversal of H3K9me2 by a small-molecule inhibitor for the G9a histone methyltransferase. Mol. Cell 2007, 25, 473–481. [Google Scholar] [CrossRef] [PubMed]
  156. Chang, Y.; Zhang, X.; Horton, J.R.; Upadhyay, A.K.; Spannhoff, A.; Liu, J.; Snyder, J.P.; Bedford, M.T.; Cheng, X. Structural basis for G9a-like protein lysine methyltransferase inhibition by BIX-01294. Nat. Struct. Mol. Biol. 2009, 16, 312–317. [Google Scholar] [CrossRef]
  157. Pan, M.-R.; Hsu, M.-C.; Chen, L.-T.; Hung, W.-C. G9a orchestrates PCL3 and KDM7A to promote histone H3K27 methylation. Sci. Rep. 2015, 5, 18709. [Google Scholar] [CrossRef] [Green Version]
  158. Ferrari-Amorotti, G.; Fragliasso, V.; Esteki, R.; Prudente, Z.; Soliera, A.R.; Cattelani, S.; Manzotti, G.; Grisendi, G.; Dominici, M.; Pieraccioli, M.; et al. Inhibiting interactions of lysine demethylase LSD1 with snail/slug blocks cancer cell invasion. Cancer Res. 2013, 73, 235–245. [Google Scholar] [CrossRef] [Green Version]
  159. Lin, Y.; Wu, Y.; Li, J.; Dong, C.; Ye, X.; Chi, Y.-I.; Evers, B.M.; Zhou, B.P. The SNAG domain of Snail1 functions as a molecular hook for recruiting lysine-specific demethylase 1. EMBO J. 2010, 29, 1803–1816. [Google Scholar] [CrossRef] [PubMed]
  160. Lv, T.; Yuan, D.; Miao, X.; Lv, Y.; Zhan, P.; Shen, X.; Song, Y. Over-expression of LSD1 promotes proliferation, migration and invasion in non-small cell lung cancer. PLoS ONE 2012, 7, e35065. [Google Scholar] [CrossRef] [Green Version]
  161. Investigation of GSK2879552 in Subjects with Relapsed/Refractory Small Cell Lung Carcinoma. Available online: https://clinicaltrials.gov/ct2/show/NCT02034123 (accessed on 21 January 2021).
  162. Maes, T.; Tirapu, I.; Mascaró, C.; Ortega, A.; Estiarte, A.; Valls, N.; Castro-Palomino, J.; Buesa Arjol, C.; Kurz, G. Preclinical characterization of a potent and selective inhibitor of the histone demethylase KDM1A for MLL leukemia. J. Clin. Oncol. 2013, 31, e13543. [Google Scholar] [CrossRef]
  163. Ryningen, A.; Stapnes, C.; Bruserud, Ø. Clonogenic acute myelogenous leukemia cells are heterogeneous with regard to regulation of differentiation and effect of epigenetic pharmacological targeting. Leuk. Res. 2007, 31, 1303–1313. [Google Scholar] [CrossRef]
  164. Jones, P.A.; Taylor, S.M. Cellular differentiation, cytidine analogs and DNA methylation. Cell 1980, 20, 85–93. [Google Scholar] [CrossRef]
  165. Jones, P.A.; Taylor, S.M.; Wilson, V. DNA modification, differentiation, and transformation. J. Exp. Zool. 1983, 228, 287–295. [Google Scholar] [CrossRef] [PubMed]
  166. Tan, E.-J.; Kahata, K.; Idås, O.; Thuault, S.; Heldin, C.-H.; Moustakas, A. The high mobility group A2 protein epigenetically silences the Cdh1 gene during epithelial-to-mesenchymal transition. Nucleic Acids Res. 2015, 43, 162–178. [Google Scholar] [CrossRef] [Green Version]
  167. Nam, J.-S.; Ino, Y.; Kanai, Y.; Sakamoto, M.; Hirohashi, S. 5-aza-2′-deoxycytidine restores the E-cadherin system in E-cadherin-silenced cancer cells and reduces cancer metastasis. Clin. Exp. Metastasis 2004, 21, 49–56. [Google Scholar] [CrossRef] [PubMed]
  168. Ivanova, T.; Zouridis, H.; Wu, Y.; Cheng, L.L.; Tan, I.B.; Gopalakrishnan, V.; Ooi, C.H.; Lee, J.; Qin, L.; Wu, J.; et al. Integrated epigenomics identifies BMP4 as a modulator of cisplatin sensitivity in gastric cancer. Gut 2013, 62, 22–33. [Google Scholar] [CrossRef]
  169. Lin, Y.-T.; Wu, K.-J. Epigenetic regulation of epithelial-mesenchymal transition: Focusing on hypoxia and TGF-β signaling. J. Biomed. Sci. 2020, 27, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Berman, M.; Mattheolabakis, G.; Suresh, M.; Amiji, M. Reversing epigenetic mechanisms of drug resistance in solid tumors using targeted microRNA delivery. Expert Opin. Drug Deliv. 2016, 13, 987–998. [Google Scholar] [CrossRef]
Figure 1. Epigenetic mechanisms regulate critical checkpoints of metastasis. Metastasis checkpoints are regulated by highly coordinated epigenetic mechanisms. EMT is mainly characterized by loss of E-cadherin expression. This can be achieved through regulation of transcription factors ZEB1, 2, SNAIL, and TWIST1. The p300/CBP complex acetylates the promoter of ZEB1, which further activates BRG1, in order to suppress E-cadherin via histone methylation. Additionally, ZEB1/ZEB2 have the ability tο recruit the CtBP complex, which enables HDAC1 to suppress the E-cadherin gene. At the same time, TGF-β, which is activated by TGFBR2-mediated induction of SNAI1/2, can activate both JMJ3 demethylase and lncRNA-HIT. JMJ3 is part of a loop along with TGF-β and SNAI1, regulating each other’s induction. Histone methyltransferase G9 targets SNAI1 which recruits SIN3A to suppress E-cadherin via deacetylation. Additionally, SET8 monomethylates H4K20 and cooperates with TWIST1 to increase N-cadherin and suppress E-cadherin. The methyltransferase EZH2, part of the PRC2 complex and the MYC-activated miR-9 can also inhibit E-cadherin expression by histone methylation and targeting of the E-cadherin mRNA. During the second step of intravasation and anoikis resistance, cancer cells express anti-apoptotic factors, such as BCL2 as well as mesenchymal markers. FZD7 induces activation of TWIST1 which increases BCL2 and FLIP inhibits anoikis by suppressing caspase-8 production. Moreover, JMJD2B interacts with β-catenin to upregulate the mesenchymal marker vimentin and PHF8 is involved in elevation of integrins expression. In the final step of extravasation and colonization, the cell differentiation NDRG1 gene plays a crucial role in reversing EMT and promoting metastasis.
Figure 1. Epigenetic mechanisms regulate critical checkpoints of metastasis. Metastasis checkpoints are regulated by highly coordinated epigenetic mechanisms. EMT is mainly characterized by loss of E-cadherin expression. This can be achieved through regulation of transcription factors ZEB1, 2, SNAIL, and TWIST1. The p300/CBP complex acetylates the promoter of ZEB1, which further activates BRG1, in order to suppress E-cadherin via histone methylation. Additionally, ZEB1/ZEB2 have the ability tο recruit the CtBP complex, which enables HDAC1 to suppress the E-cadherin gene. At the same time, TGF-β, which is activated by TGFBR2-mediated induction of SNAI1/2, can activate both JMJ3 demethylase and lncRNA-HIT. JMJ3 is part of a loop along with TGF-β and SNAI1, regulating each other’s induction. Histone methyltransferase G9 targets SNAI1 which recruits SIN3A to suppress E-cadherin via deacetylation. Additionally, SET8 monomethylates H4K20 and cooperates with TWIST1 to increase N-cadherin and suppress E-cadherin. The methyltransferase EZH2, part of the PRC2 complex and the MYC-activated miR-9 can also inhibit E-cadherin expression by histone methylation and targeting of the E-cadherin mRNA. During the second step of intravasation and anoikis resistance, cancer cells express anti-apoptotic factors, such as BCL2 as well as mesenchymal markers. FZD7 induces activation of TWIST1 which increases BCL2 and FLIP inhibits anoikis by suppressing caspase-8 production. Moreover, JMJD2B interacts with β-catenin to upregulate the mesenchymal marker vimentin and PHF8 is involved in elevation of integrins expression. In the final step of extravasation and colonization, the cell differentiation NDRG1 gene plays a crucial role in reversing EMT and promoting metastasis.
Ijms 22 02778 g001
Figure 2. Histone acetylation regulates genes involved in cancer progression. Histone acetylation is an activating histone modification that favors gene transcription. H3K4Ac marks are present on the promoters of several genes, GLI1, SMO, FOXF1, Bmi1, and SIRT2, which are upregulated in cancer, favoring tumor progression. Upon SNAI1/2 activation, increased H3K9 acetylation has been observed on the promoter of TGFBR2, causing its upregulation. Critical acetyltransferases TIP60, GCN5, PCAF, and p300 catalyze histone acetylation of metastasis-promoting genes. hMOF, a H4K16 histone acetyltransferase normally maintains the expression of EMT-related tumor suppressor genes, such as TMS1, CDH1, and ESR1, but is often downregulated in cancer. TIP60 is also commonly downregulated in various tumors, where it prevents the activity of anti-tumor DDR and p53 pathways, indicating how reduction in histone acetylation promotes cancer progression. HDAC3 can also be detected on the promoters of activated mesenchymal genes, even though most HDACs are associated with gene repression. Upon Notch signaling activation, H3K4Ac is removed by HDAC3, allowing for Notch-mediated expression of EMT-related genes. Profilin-2 interacts with HDAC1 and inhibits its binding to the promoters of Smad2 and Smad3, causing Smad protein activation and subsequently enhancing TGF-β-induced EMT and angiogenesis. HDACs also catalyze H3K4/56 deacetylation at the CDH1 promoter, thus repressing E-cadherin. Similarly, HDAC1 and 2 are recruited by ZEB1 to CDH1 promoter, inducing its repression. In the same context, ZEB1 recruits SIRT1, a nicotinamide adenine dinucleotide (NAD)-dependent deacetylase, associated with a global H3K27 deacetylation and reduced H3K9Ac and H3K27Ac levels on the promoters of CDH1 and other epithelial genes, such as EPCAM, ST14, ESRP1, and RAB25, promoting EMT and metastasis. Finally, overexpression of FLIP leads to the upregulation of the anti-apoptotic Bcl-XL, inhibiting the anoikis apoptotic pathway. HDAC inhibition prevents this outcome by reducing FLIP levels and inducing apoptosis of cancer cells, suggesting that HDACs participate in anoikis resistance of cancer cells through yet unknown mechanisms.
Figure 2. Histone acetylation regulates genes involved in cancer progression. Histone acetylation is an activating histone modification that favors gene transcription. H3K4Ac marks are present on the promoters of several genes, GLI1, SMO, FOXF1, Bmi1, and SIRT2, which are upregulated in cancer, favoring tumor progression. Upon SNAI1/2 activation, increased H3K9 acetylation has been observed on the promoter of TGFBR2, causing its upregulation. Critical acetyltransferases TIP60, GCN5, PCAF, and p300 catalyze histone acetylation of metastasis-promoting genes. hMOF, a H4K16 histone acetyltransferase normally maintains the expression of EMT-related tumor suppressor genes, such as TMS1, CDH1, and ESR1, but is often downregulated in cancer. TIP60 is also commonly downregulated in various tumors, where it prevents the activity of anti-tumor DDR and p53 pathways, indicating how reduction in histone acetylation promotes cancer progression. HDAC3 can also be detected on the promoters of activated mesenchymal genes, even though most HDACs are associated with gene repression. Upon Notch signaling activation, H3K4Ac is removed by HDAC3, allowing for Notch-mediated expression of EMT-related genes. Profilin-2 interacts with HDAC1 and inhibits its binding to the promoters of Smad2 and Smad3, causing Smad protein activation and subsequently enhancing TGF-β-induced EMT and angiogenesis. HDACs also catalyze H3K4/56 deacetylation at the CDH1 promoter, thus repressing E-cadherin. Similarly, HDAC1 and 2 are recruited by ZEB1 to CDH1 promoter, inducing its repression. In the same context, ZEB1 recruits SIRT1, a nicotinamide adenine dinucleotide (NAD)-dependent deacetylase, associated with a global H3K27 deacetylation and reduced H3K9Ac and H3K27Ac levels on the promoters of CDH1 and other epithelial genes, such as EPCAM, ST14, ESRP1, and RAB25, promoting EMT and metastasis. Finally, overexpression of FLIP leads to the upregulation of the anti-apoptotic Bcl-XL, inhibiting the anoikis apoptotic pathway. HDAC inhibition prevents this outcome by reducing FLIP levels and inducing apoptosis of cancer cells, suggesting that HDACs participate in anoikis resistance of cancer cells through yet unknown mechanisms.
Ijms 22 02778 g002
Figure 3. Role of histone methylation in gene regulation during tumor progression. Trimethylation of H3K27 and H3K4 has been associated with suppression of epithelial genes or activation of mesenchymal ones, respectively. In this process, the PRMT5-MEP50 complex suppresses CDH1 and GAS1 by mediating H3K4 and H4R3 methylation. E-cadherin is suppressed by SNAI1-mediated PRC2 trimethylation of H3K27. PRC2 also suppresses KLF2 expression. JARID2 recruits both PRC2 and G9a to promote H3K27 and H3K9 methylation at the promoters of CDH1 and miR-200. G9a also silences the ep-CAM gene, promoting metastasis. The TRIM33/Smad2/3 complex inhibits the binding of HP1 on the DNA, leading to expression of GSC and MIXL1 mesenchymal genes. The PRMT5-MEP50 complex recruits WDR5 and causes H3R2 methylation, activating the EMT-promoting genes VIM and SNAI1/2. Over-expression of the histone demethylase, JARID1A leads to the activation of cyclin D1/E1 and ITGB1 expression, promoting tumor progression. The H3K27 demethylase, UTX, activates several pro-metastatic genes, such as MMP-9/11 and SIX1, after interacting with the MLL4 complex, which includes a H3K4 methyltransferase, thus causing enhancement of EMT and metastasis. The PHF8 demethylase upregulates the expression of Vimentin, Integrin, and Rho-associated protein kinase (ROCK) kinase by removing H3K9 methylation marks, thus favoring metastasis. JMJD2B, a H3K9 demethylase, induces the expression of Vimentin, after interacting with β-catenin and also demethylates the Integrin (ITGB2, ITGAM, ITGA9, ITGAB2) gene promoters, further promoting EMT and metastasis. JMJ3, a H3K27 histone demethylase is induced by TGF-β and activates SNAI1 expression to facilitate EMT. EZH2 establishes repressive H3K27me3 marks and downregulates FOXC, inducing metastasis. It also silences the CDH1, AXIN2, NKD1, PPP2R2B, PRICKLE1, SFRP5, RKIP, promoting EMT and cancer metastasis. SETDB1 methyltransferase amplification is responsible for cancer cell growth by methylating and stabilizing the oncogenic p53 mutants. In this context, PRMT6 methyltransferase reduces the expression of p21, promoting tumor cell migration.
Figure 3. Role of histone methylation in gene regulation during tumor progression. Trimethylation of H3K27 and H3K4 has been associated with suppression of epithelial genes or activation of mesenchymal ones, respectively. In this process, the PRMT5-MEP50 complex suppresses CDH1 and GAS1 by mediating H3K4 and H4R3 methylation. E-cadherin is suppressed by SNAI1-mediated PRC2 trimethylation of H3K27. PRC2 also suppresses KLF2 expression. JARID2 recruits both PRC2 and G9a to promote H3K27 and H3K9 methylation at the promoters of CDH1 and miR-200. G9a also silences the ep-CAM gene, promoting metastasis. The TRIM33/Smad2/3 complex inhibits the binding of HP1 on the DNA, leading to expression of GSC and MIXL1 mesenchymal genes. The PRMT5-MEP50 complex recruits WDR5 and causes H3R2 methylation, activating the EMT-promoting genes VIM and SNAI1/2. Over-expression of the histone demethylase, JARID1A leads to the activation of cyclin D1/E1 and ITGB1 expression, promoting tumor progression. The H3K27 demethylase, UTX, activates several pro-metastatic genes, such as MMP-9/11 and SIX1, after interacting with the MLL4 complex, which includes a H3K4 methyltransferase, thus causing enhancement of EMT and metastasis. The PHF8 demethylase upregulates the expression of Vimentin, Integrin, and Rho-associated protein kinase (ROCK) kinase by removing H3K9 methylation marks, thus favoring metastasis. JMJD2B, a H3K9 demethylase, induces the expression of Vimentin, after interacting with β-catenin and also demethylates the Integrin (ITGB2, ITGAM, ITGA9, ITGAB2) gene promoters, further promoting EMT and metastasis. JMJ3, a H3K27 histone demethylase is induced by TGF-β and activates SNAI1 expression to facilitate EMT. EZH2 establishes repressive H3K27me3 marks and downregulates FOXC, inducing metastasis. It also silences the CDH1, AXIN2, NKD1, PPP2R2B, PRICKLE1, SFRP5, RKIP, promoting EMT and cancer metastasis. SETDB1 methyltransferase amplification is responsible for cancer cell growth by methylating and stabilizing the oncogenic p53 mutants. In this context, PRMT6 methyltransferase reduces the expression of p21, promoting tumor cell migration.
Ijms 22 02778 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Markouli, M.; Strepkos, D.; Basdra, E.K.; Papavassiliou, A.G.; Piperi, C. Prominent Role of Histone Modifications in the Regulation of Tumor Metastasis. Int. J. Mol. Sci. 2021, 22, 2778. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052778

AMA Style

Markouli M, Strepkos D, Basdra EK, Papavassiliou AG, Piperi C. Prominent Role of Histone Modifications in the Regulation of Tumor Metastasis. International Journal of Molecular Sciences. 2021; 22(5):2778. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052778

Chicago/Turabian Style

Markouli, Mariam, Dimitrios Strepkos, Efthimia K. Basdra, Athanasios G. Papavassiliou, and Christina Piperi. 2021. "Prominent Role of Histone Modifications in the Regulation of Tumor Metastasis" International Journal of Molecular Sciences 22, no. 5: 2778. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052778

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