Skip to main content

The molecular mechanisms and therapeutic strategies of EMT in tumor progression and metastasis

Abstract

Epithelial–mesenchymal transition (EMT) is an essential process in normal embryonic development and tissue regeneration. However, aberrant reactivation of EMT is associated with malignant properties of tumor cells during cancer progression and metastasis, including promoted migration and invasiveness, increased tumor stemness, and enhanced resistance to chemotherapy and immunotherapy. EMT is tightly regulated by a complex network which is orchestrated with several intrinsic and extrinsic factors, including multiple transcription factors, post-translational control, epigenetic modifications, and noncoding RNA-mediated regulation. In this review, we described the molecular mechanisms, signaling pathways, and the stages of tumorigenesis involved in the EMT process and discussed the dynamic non-binary process of EMT and its role in tumor metastasis. Finally, we summarized the challenges of chemotherapy and immunotherapy in EMT and proposed strategies for tumor therapy targeting EMT.

Introduction

The epithelial–mesenchymal transition (EMT) is a reversible process in which epithelial cells lose their properties and become mesenchymal cells, with altered expression of cell adhesion molecules and cytoskeleton. As a result, the cells develop motility-invasive properties, allowing them to transition between epithelial and mesenchymal states in a highly dynamic and plastic manner. The reverse process, which is known as mesenchymal–epithelial transformation (MET), occurs frequently during development (such as heart development, kidney morphogenesis, and somite formation) and cancer [1]. In living organisms, EMT is involved in embryogenesis, inflammation, fibrosis, wound healing, cancer development, and other physiological and pathological processes [2,3,4,5]. EMT is a continuous process of transition along the EMT spectrum, through which cells undergo loss of apical polarity, increase in anterior–posterior polarity, decrease in cell adhesion, shift from an epithelial to a mesenchymal phenotype, and gain of mesenchymal properties [6]. Biological processes of EMT, which can be classified as EMT-type 1, 2, and 3, are associated with embryogenesis, tissue regeneration, and cancer progression, respectively [7]. During embryonic development, EMT promotes pro-intestinal formation, neural crest stratification, mesodermal development, endocardial morphogenesis, and the generation of new cell and tissue types [6]. In wound healing and inflammation sites, EMT plays a central role in restoring epithelial and endothelial integrity. However, reactivation of EMT during pathological processes has an important role in cancer progression, as EMT can confer metastatic properties to tumor cells, enhance invasion, invade surrounding tissues, and colonize distant organs [8]. From the perspective of cancer cell dynamics, EMT is a dynamic and continuous spectrum process along the transition from the epithelial to the mesenchymal cell state. Tumor cells in the intermediate state of the epithelial to mesenchymal spectrum have both epithelial and mesenchymal properties and can better survive, metastasize and colonize distal organs [9].

The progression of EMT is regulated by the expression of EMT-translational factors (TFs) (such as SNAIL, ZEB, TWIST, and others) and miRNAs, as well as epigenetic and posttranslational regulators [10]. The three TFs of the SNAIL family (SNAIL1, SNAIL2, and SNAIL3) and those of the basic helix–loop–helix (BHLH) family (TWIST1 and TWIST2) can downregulate the expression of epithelial genes and upregulate the expression of mesenchymal genes. The ZEB family of zinc finger TFs (ZEB1 and ZEB2) can activate or repress transcription by binding E-box regulatory gene sequences. Noncoding miRNAs can also selectively bind mRNA to promote its degradation or inhibit its translation. Because a variety of miRNAs can act directly on the SNAIL family to regulate EMT, changes in miRNA expression affect the course and metastasis of EMT. Multiple signaling pathways such as TGF-β, Wnt, Notch, and PI3K-AKT are also involved in the regulatory network of EMT. Furthermore, post-translational regulation can induce EMT and promote metastasis of tumor cells. Epigenetic modifications and regulation can control the expression of related EMT-TFs, which are critical in regulating the molecular pathways of metabolism, transcription, differentiation, and apoptosis in the EMT process.

During cancer progression, cancer cells in EMT states are highly plastic when they transition to epithelial/mesenchymal states (i.e., partial EMT). Such a process is similar to cells during embryonic development, making them the drivers of tumorigenesis [11, 12]. Epithelial cells acquire some distinct mesenchymal features during cancer development that can be isolated in the primary tumor, enabling them to invade adjacent tissues before spreading distally. The phenotypic status of tumor cells undergoing this process can be graded by a combination of epithelial and mesenchymal markers. Individual cells that progress to different states along the E to M spectrum can generate extensive phenotypic heterogeneity within the tumor, and this phenotypic plasticity and heterogeneity can provide cancer cells with greater adaptability and resistance [13]. This is confirmed by the observation that skin and breast primary tumors have multiple E/M cell subpopulations with distinct chromatin landscapes and gene expression profiles, which are spatially localized to specific sites in the tumor [14]. Currently, several clinical trial studies applied novel personalized therapies based on molecular levels, which directly or indirectly inhibit EMT, using the expression of specific EMT markers as the selection criterion. Examples include the effect of aspirin on CTC subtypes (epithelial/mesenchymal/mixed) in metastatic breast and colorectal cancers [15]. Activation of EMT can facilitate the progression of this phenotype toward increased invasiveness and inhibit the sensitivity of tumor cells to chemotherapy by altering the microenvironment, where quasi-mesenchymal cells exhibit higher resistance to therapeutic regimens such as chemotherapy and immunotherapy [16, 17]. Notably, both EMT and MET are required for the process of metastasis. While EMT mobilizes the cells in the primary tumor, MET terminates the migration process and thereby resulting in the distal colonizing of cancer cells [6, 18, 19]. Furthermore, EMT is aberrantly activated during organ fibrosis and is required for the fibrotic response [17, 20]. EMT induces the formation of mesenchymal myofibroblasts near epithelial cells, which then aggregate and secrete fibrosis-promoting factors that promote tissue degeneration. Tissue regeneration is associated with organ fibrosis and disease progression and ultimately causes organ failure [21].

In this review article, we described the cytoskeletal and compositional changes that occur during the EMT process. The regulatory network of numerous regulatory factors in the EMT process was summarized, and the controversial topic of whether EMT is required for cancer development was discussed. The significance of partial EMT in promoting tumor cell plasticity and metastasis was also discussed. Furthermore, the role of EMT in cancer stem cells (CSCs) and circulating tumor cells (CTCs) was described. Finally, a summary of current EMT therapeutic modalities was presented, as well as the various salient issues and challenges associated with the current therapeutic processes.

Cytoskeletal changes between epithelial and mesenchymal cells

Cells differentiate into different states throughout the development of an organism, and they can be broadly classified into epithelial or mesenchymal phenotypes [7, 22]. Epithelial cells have a flat and polygonal shape that is maintained synergistically by the actin cytoskeleton and intermediate filaments [23]. Epithelial cells have a strong apical and basal polarity, with plasma membranes oriented toward and away from the lumen, respectively [24]. Each membrane is composed of different proteins that allow for the targeted transport of molecules and the localization of various activities to specific cellular regions [25]. Therefore, apical and basal polarization is imperative for many biological functions such as endocytosis, exocytosis, and vesicle transport [26]. Additionally, adjacent individual epithelial cells develop tight junctions and coalesce into a single cell layer to form the basic epithelial tissue [27, 28]. Epithelial cell mobility is limited by shape and attachment, allowing them to migrate only in patches and whole blocks [29]. Epithelial cells usually communicate with each other through tight junctions, adherence junctions, desmosomes, and gap junctions [30] (Fig. 1). When epithelial junctions are dissolutive, epithelial cells lose their apical–basal polarity [30].

Fig. 1
figure 1

Changes in the cytoskeleton and properties between epithelial and mesenchymal during the EMT process. Epithelial cells exhibit apical–basal polarity with cell–cell and cell–matrix attachment. Three multi-protein complexes (Scribble complex, Crumbs complex, and PAR complex) interact to regulate the spatial separation of apical and basal structural domains together to establish cell polarity. Intercellular adhesion and communication are provided by intercellular junctions and maintain tissue stability and integrity. Tight junctions (TJs) form strips around cells that help separate apical and basal regions and form sealed spaces between adjacent cells, preventing the flow of material. Adherens junctions (AJs) are located below TJs, surround cells, and provide intercellular adhesion, but they are relatively permeable. Gap junctions are gaps located on the outer surface of cells and are hydrophilic ion transport channels between adjacent cells. Bridging granules provide sites of cell adhesion and intermediate filament binding to disklike structures located on the outer surface of the cell. The occurrence of EMT leads to the dissolution of intercellular junctions and loss of cell polarity allowing cytoskeletal rearrangements that alter the shape of the cell, transforming the cell into a mesenchymal phenotype and promoting cell motility and invasion. Based on a synthesis of the literature, we conclude that as the EMT progresses, the cell characteristics are changed, including reduces in drug sensitivity, proliferation, and response to apoptosis signals and increases in drug efflux, invasion, and immune evasion. The partial EMT with intermediate state has properties of enhanced stemness and tumor initiation capacity, and stronger ability to adapt to the changes in immune microenvironment and metabolism

Mesenchymal cells are distinguished from epithelial cells by their irregular morphology, with an elongated, spindle-shaped form and less rigid topography [26]. They are also characterized chiefly by the acquisition of anterior–posterior polarity while losing cellular adhesion or uniform composition. In mesenchymal cells, keratin is downregulated, whereas vimentins are upregulated, which increases the strength of the cytoskeleton and flexibility, and the ability to migrate and invade tissues [31]. In parallel, bundles of stress fibers are structured by actin filaments to generate new actin-rich membrane protrusions, allowing for various types of movement and sensory reception [32].

EMT is also a process that routinely promotes the migration of cells toward the extracellular matrix, which is known as invasion. The process of invasion differs from the controlled and seemingly stable interactions between epithelial and endothelial cells in intact tissue, as well as from the underlying basement membrane [33]. To undergo directed cell migration, as shown for high-grade cancer cells, epithelial cells must divert their apical–basal cell polarity to frontal polarity [20]. In vertebrate cells, polarization is usually regulated by the apical compartment (partitioning‑defective (PAR) and Crumbs complexes) and the basolateral compartment (Scribble complexes), and disruption of these complexes confers loss of apical–basal polarization during EMT [20, 34]. The Scribble complex is considered a tumor suppressor that is found in the basal structural domain, which maintains basolateral polarity [35]. The PAR complex defines the boundary between the basal and apical structural domains, while the Crumb complex is associated with the LIN-1 (PALS1) related protein (PALS1) and the PALS1-associated tight junction (PATJ) complex, which controls apical structural domain formation [20, 36]. These two complexes are located in tight junctions and work together to maintain the apical structural domain.

The PAR and Scribble complexes are antagonistic to each other, and the PAR complex helps in enhancing the activity of the Crumb complex. Apical–basal polarity is associated with the integrity of epithelial or endothelial junctions, whereas redirected cell polarity is associated with the deconstruction of lateral cell–cell junctions of the epithelium or endothelium during EMT [20]. Tight junction instability is accompanied by decreased expression of claudin and occludin, and the diffusion of zonula occludens 1 (ZO1) from cell contacts [20, 30]. During transformation, the expression or function of epithelial genes such as E-cadherin-specific cytokeratin and zone of occlusion 1 (ZO-1) is lost, whereas the expression of genes defining the mesenchymal phenotype (such as vimentin, fibronectin, N-cadherin, β1 and β3 integrins) is enhanced [18, 37, 38]. Initiation of EMT also disrupts bridging granules, allowing low levels of connexin to compromise the integrity of gap junctions [39]. When adhesion junctions are destabilized, epithelial calmodulin (E-cadherin) on the cell membrane is cleaved and degraded [29]. Changes in the aforementioned molecular readouts can characterize the specific cellular features of the process [40].

Epithelial–mesenchymal transition cells can develop a spindle-shaped mesenchymal morphology and motility by forming protrusions due to the reorganization of the actin cytoskeletal structure and the aforementioned changes [14]. The cortical organization of the actin cytoskeleton is repositioned as bundles of stress fibers and clustered near the ventral surface of cultured cells [24]. The stress fiber bundles facilitate various types of movement and sensory reception by forming new actin-rich membrane protrusions [31]. Furthermore, keratin downregulation and vimentins upregulation strengthen the cytoskeleton, making it less susceptible to damage during migration and more flexible. In addition, cells with EMT characteristics can degrade and invade their extracellular matrix at the invasion front of individual cells or populations of cells by activating proteases such as matrix metalloproteinase (MMPs) [41].

In addition, tumor cells in epithelial–mesenchymal hybrid state or partial EMT during the transition have the greater migratory ability, allowing them to detach from their original tissue and wander throughout the body [14]. Besides, changes in the intracellular and intravascular environment make EMT complex and diverse [24]. According to ongoing research, EMT represents a spectrum of states with intrinsic plasticity, progressively forming different intermediates for transformation rather than being determined by only two cellular states [6]. Certain EMT states exhibit both epithelial and mesenchymal cell characteristics in both in vitro and in vivo models of many developmental and disease processes [14]. These EMT cells may co-express epithelial and mesenchymal markers or may lose epithelial markers without obtaining mesenchymal markers [42]. Therefore, the cellular and molecular characteristics of EMT should be evaluated in a context-dependent manner. We described “partial EMT” in greater detail in the subsequent section.

Regulatory networks of EMT

Transcriptional regulation of EMT

The transition of cells from epithelial to mesenchymal states is mediated by key transcription factors, which primarily regulate intercellular adhesion, cell polarity, and motility [18]. Transcription factors induce mesenchymal gene expression by suppressing genes associated with the epithelial phenotype, resulting in the EMT cellular signature [43]. The major EMT-inducible transcription factors are Zinc finger binding transcription factors SNAIL1 and SNAIL2, zinc finger E box binding homology frame factors ZEB1 and ZEB2, and basic helix–loop–helix (BHLH) factors TWIST1 and TWIST2 [20]. SNAIL1/2, ZEB1/2, and TWIST1/2 are thought to be major regulators of the transcriptional pathway that drives EMT, and they converge to activate the expression of transcription factors (Fig. 2) [44].

Fig. 2
figure 2

Regulatory network of EMT. EMT is tightly regulated by a complex network which is composed of several factors, including transcription regulation, posttranslational control, epigenetic modifications, and noncoding RNA-mediated regulation. Snail, Twist, Zeb, and other EMT-related transcription factors are regulated by multiple signaling pathways at post-transcriptional and posttranslational level and orchestrate with other epigenetic factors to regulate downstream transcriptional networks, further mediating the biological effects of EMT

SNAIL1 and SNAIL2 have critical roles in the activation of the EMT program during embryonic development, fibrosis, and cancer [45]. In a physiological context, they enhance neural crest development and migration [46]. SNAIL1 and SNAIL2 bind to the E box sequence of the CDH1 promoter region and directly repress transcription by recruiting the polycomb repressor complex (PRC), which contains components such as G9a, methyltransferases enhancer of zeste homolog 2 (EZH2), histone deacetylases 1, 2, and the Lys‑specific demethylase 1 (LSD1) [47]. All of these factors coordinate histone hypermethylation and deacetylation to repress epithelial gene expression [47,48,49,50]. SNAIL1 could directly inhibit TJ formation and epithelial markers such as E-cadherin and claudins while upregulating mesenchymal phenotypic markers including vimentin and fibronectin [51,52,53]. SNAIL2 promotes the loss of cell adhesion and polarity and confers migration and invasion capabilities [54]. Moreover, they both cooperate with other transcription regulators to control gene expression. For example, SNAIL1 cooperates with ETS1 to activate MMP expression [55]. SNAIL expression in specific physiological settings could be activated via multiple signaling pathways, including TGF-β, Wnt, Notch, and growth factors acting on receptor tyrosine kinases (RTKs) [56], indicating that the SNAIL-associated EMT program is driven by multiple mechanisms. SNAIL1 and SNAIL2 play key roles in EMT-induced tumor progression. For example, the expression levels of SNAIL1 could be used as a prognostic indicator. E-cadherin expression is positively correlated with patient survival, whereas overexpression of MMPs is associated with tumor cell aggressiveness [57]. Furthermore, it has been shown that stable SNAIL2 expression decreases e-calmodulin levels and thus enhances metastasis in breast and ovarian cancers [58, 59].

ZEB1 and ZEB2, which belong to the human ZEB family, are zinc finger TFs that bind to regulatory gene sequences at E-box and activate or repress transcription [60]. ZEB-mediated transcriptional repression usually involves the recruitment of C-terminal-binding protein (CTBP) [61, 62]. In certain cancer cells, ZEB interacts with the switch/sucrose nonfermentable (SWI/SNF) chromatin-remodeling protein BRG1 (Brahma-related gene 1) to suppress gene transcription [63]. Furthermore, ZEB1 can switch from a transcriptional suppressor to an activator by interacting with co-activators PCAF and P300 [64]. ZEB proteins suppress the expression of cell polarity complex components and downregulate TJ genes, thereby driving EMT [65]. However, ZEB proteins enhance the expression of the mesenchymal protein vimentin and N-cadherin [66]. The TGF-β and Wnt signaling pathways, as well as other growth factors that activate the Ras-MAPK signaling pathway can all induce ZEB protein expression [20]. TWIST1 can cooperate with SNAIL1 to induce ZEB1 expression, which is often followed by SNAIL activation [67]. According to genetic evidence from related studies, ZEB1 expression is essential for effective invasion and metastasis in mouse models of pancreatic cancer [68]. The expression of ZEB1/2 in epithelial cells causes an EMT and mesenchymal phenotype in tumor stem cells, which promotes invasion, metastatic dissemination, and a dedifferentiated state [69]. Moreover, ZEB1 expression is associated with poor clinical outcomes in solid tumors (including breast, colorectal, and pancreatic tumors) [70,71,72,73], and poor prognosis and survival in various tumor types [74]. However, it is worth noting that ZEB1 and ZEB2 have opposite effects on different cell lines, implying that tissue context is critical for the function of EMT-TF [64, 75, 76]. For instance, in a mouse model of melanoma, ZEB2 inhibited tumor metastasis, while ZEB1 drove tumorigenesis and progression. In clinical studies, ZEB2 deficiency was associated with reduced survival in melanoma patients, while ZEB1 expression was associated with poorer clinical outcomes. Additionally, ZEB1 and ZEB2 can have different functions due to the regulation of the transforming growth factor-β (TGF-β) and bone morphogenetic protein (BMP) signaling pathways. ZEB1 induces osteoblast differentiation and associated growth arrest, whereas ZEB2 has opposite functions.

TWIST1 and TWIST2 belong to the basic helix–loop–helix (BHLH) family of transcription factors. Both proteins form homodimers or heterodimers with E12 and E47 to regulate E-box DNA response elements in order to suppress or activate transcription [77]. TWIST1, similar to SNIAL, can suppress the expression of E‑cadherin and promote the expression of N‑cadherin, resulting in decreased cell adhesion and increased cell motility [78, 79]. The transcriptional regulation of TWIST proteins is essential for the recruitment of methyltransferase SET8 or activation of B lymphoma Mo‑MLV insertion region 1 homolog (BMI1) [80, 81]. A related study found that SNAIL2 knockdown blocked the ability of TWIST to activate EMT in mammary cells, suggesting that TWIST can indirectly induce transcriptional repression of E- cadherin. [82]. Multiple signaling pathways activate TWIST during the EMT program. Importantly, the hypoxia-inducible factor-1α (HIF-1α) transcription factor activates TWIST expression and promotes EMT and tumor cell dissemination under hypoxic conditions [78]. Related studies, using a mouse model of spontaneous squamous cell carcinoma, demonstrated that cancer cells undergo EMT and spread to the circulation, which is facilitated by activation of the EMT-inducible transcription factor TWIST1 [79]. Furthermore, TWIST has been shown to play a crucial role in the development of benign skin tumors in mice. Conditional ablation of TWIST expression in the skin prevented DMBA/TPA-induced skin cancer and significantly decreased tumorigenesis [83]. Moreover, TWIST overexpression is associated with tumor invasion and metastasis [84, 85].

In addition to SNAIL, ZEB, and TWIST, the EMT program is regulated by various transcription factors in tissue development and cancer. For example, several forkhead box (FOX) proteins, such as FOXC1, FOXC2, and FOXQ1, can promote mesenchymal differentiation and decrease the expression of proteins involved in polarity complexes and cell–cell junctional [86,87,88]. SRY box (SOX) transcription factors cooperate with SNAIL proteins to promote EMT [89]. According to research, paired-related homeobox 1 (PRRX1) protein is a novel regulator of EMT [90]. However, the interaction between various EMT-TFs needs to be clarified. In addition, the precise regulation and functions of the EMT-TFs in different EMT contexts should be further investigated.

Translational and post-translational control

Translational control has a significant effect on EMT (Fig. 2). The forced expression of Y-box-binding protein-1 (YB-1) induces EMT and promotes metastasis by directly activating the cap-independent translation of SNAIL1 mRNA and other mesenchymal factors in RAS-transformed mammary epithelial cells [91]. Embryonic lethal abnormal vision-like RNA promotes the EMT process by enhancing the stability of SNAIL1 mRNA [92]. In breast cancer, cytoplasmic polyadenylation element-binding protein 1 (CPEB1) mediates EMT and metastasis by enhancing the shortening of the polyA tail of MMP9, which lowers MMP9 translation [93].

Recent studies have underlined the importance of post-translational modification at the proteome level for the EMT program. Protein phosphorylation, the most common post-translational modification, is required for the regulation of multiple molecular pathways in metabolism, transcription, differentiation, and apoptosis [94]. As a post-translational modification, phosphorylation can control the expression of SNAIL [95]. Glycogen synthase kinase-3β (GSK-3β), a classical kinase involved in many signaling pathways, phosphorylates SNAIL through two consecutive motifs, thereby controlling its ubiquitination and subcellular localization. First, GSK-3β binds to and phosphorylates Ser97 and Ser101 in the SNAIL motif 1 to induce nuclear export of SNAIL. Subsequently, GSK-3β phosphorylates Ser108, Ser112, Ser116, and Ser120 in motif 2 to promote β-Trcp-mediated degradation of the SNAIL proteasome [96]. Several signaling pathways, including Wnt, NF-κB, Notch, and PI3K-AKT, inhibit GSK-3β‑mediated phosphorylation or disrupt the GSK-3β-SNAIL interaction to boost SNAIL stability and downregulate E-cadherin, resulting in EMT program activation [96,97,98,99]. Protein kinase D1 (PKD1) phosphorylates SNAIL1 and promotes its export from the nuclear [100]. On the contrary, small C-terminal domain phosphatase 1 (SCP1) could dephosphorylate SNAIL1 to retain it in the nucleus and enhance its activity [101]. Several other kinases, including p21-activated kinase 1 (PAK1) and large tumor suppressor 2 (LATS2), phosphorylate SNAILS, affect its activity positively or negatively [102, 103]. Furthermore, phosphorylation regulates the stability of TWIST. TWIST phosphorylation at Ser68 by MAPKs prevents ubiquitin-mediated degradation and enhances TWIST activity [104].

Other post-translational modifications can modulate the course of EMT by regulating the activity of key EMT-TFs. Sumoylation of ZEB2 by PRC2 promotes its export from the nucleus, abolishing ZEB-mediated gene regulation [105]. The stability and activity of EMT-TFs are regulated by ubiquitination, which activates or suppresses the EMT program. An atypical ubiquitin E3 ligase complex, Skp1-PamFbxo45, controls the EMT program by regulating the degradation of different EMT-TFs [106]. In breast cancer, ZEB1 is regulated by the E3 ubiquitin ligase SIAH, which marks its degradation [107]. However, deubiquitination of ZEB1 by the ubiquitin-specific protease-51 (USP51) promotes its stabilization [108]. The same post-translational regulation of epithelial proteins is possible during EMT. Hakai, an E3 ubiquitin ligase, ubiquitinates E-cadherin, inducing its endocytosis and destruction [109]. Acetylation of SNAIL1 protein by CBP inhibits the formation of the suppressor complex and converts SNAIL1 from a gene repressor to an activator [110]. Moreover, p300 can acetylate SNAIL1 and TWIST1, regulating their stability, location, and interactions with other proteins [111]. Members of the miR-200 family act as oncogenic miRNAs, enhancing E-cadherin expression while suppressing the expression of ZEB1 and ZEB2. The sumo modification of FoxM1 at lysine 463, which is a posttranslational modification, is required for complete suppression of miR-200b/c in breast cancer cells [112].

Epigenetic modifications

Epithelial–mesenchymal transition is associated with important epigenetic alterations, which are often required to mediate the function of EMT-TFs [113] (Fig. 2). Over the past few decades, specific modifications, including DNA methylation and histone modifications, as well as multiple epigenetic regulators, have been identified as key regulators of the EMT process. In a variety of human tumors including breast, bladder, lung, liver, gastric, and prostate cancers, CDH1 promoter methylation has been implicated as a key factor in EMT [114,115,116]. Transcription factors such as SNAIL, SLUG, ZEB1, and ZEB2/SIP1 bind to the E-box on the CDH1 promoter and are considered to be direct inhibitors of E-cadherin [117,118,119]. Research indicates that ZEB1 mediates CDH1 downregulation in basal cell-like breast cancer and recruits DNMT1 to the CDH1 promoter to maintain the methylation status of the promoter [120]. The histone demethylase LSD1, also known as KDM1A and AOF2, plays an essential role in EMT [121]. It was discovered that SNAIL interacts with LSD1 through its SNAG (SNAIL1/GF) structural domain and recruits LSD1 to the CDH1 promoter. Therefore, the methyl group on lysine 4 of histone H3 will be removed (H3K4m2) [48]. Other demethylases such as KDM6B and PHF8, as well as the methyltransferases PRMT5, EZH2/SUZ12, SUV39H1, and G9a have been reported to regulate EMT by regulating the expression of EMT-TF (mainly SNAIL1 and ZEB) or interacting with these factors to affect the expression of downstream genes [122,123,124,125,126].

During the transition of trophoblast stem cells from an epithelial to a mesenchymal state, the histone deacetylase HDAC6 directly deacetylates the promoter of the TJ gene, resulting in decreased cell–cell adhesion, which is one of the earliest EMT events [127]. The ZEB and TWIST families of transcription factors also bind and recruit the nucleosome remodeling deacetylase NuRD complex to their target promoters [128]. Altogether, epigenetic modifications are the basis for determining the expression of key proteins in the EMT. Notably, these modifications are often reversible and can play a key role in defining EMT plasticity.

Noncoding RNA-mediated regulation

Several microRNAs (miRNAs) can directly regulate the expression of EMT transcription factors (Fig. 2). Noncoding miRNAs inhibit the translation or promote degradation of mRNAs by selectively binding to them. The miR-200 family, which has five miRNAs, and miR-205 suppress ZEB1/2 expression [129, 130]. Although they target different ZEB sequences, they cooperate to enhance ZEB suppression [131]. In liver carcinoma cell, p53 inhibits the EMT program by increasing the levels of miR‑200 and miR‑192/miR-215, resulting in low expression of ZEB1/2 [132]. In colorectal, breast, prostate, and hepatocytes cancer cells, miR-34, miR-203, miR-29b, and miR-30 can similarly suppress the expression of SNAIL1 [133,134,135,136]. Furthermore, miR-1 and miR-200 can suppress the expression of SNAIL2 in prostate adenocarcinoma cells [137]. The expression of miRNAs and EMT-TFs are regulated in a double-negative feedback loop. It is widely known that the miR-200 family suppresses the production and activation of ZEB, which in turn suppresses the expression of miR-200 family [129]. In addition, similar to the activities of miR-200 family members, miR-1199-5p acts as a guardian of epithelial cell phenotype in a reciprocal double-negative feedback loop with ZEB1 [138]. Other double-negative feedback loops operate between miRNAs and EMT-TFs were reported in multiple processes of tumor epithelial–mesenchymal transition, including miR-34/miR-203 and SNAIL1, miR-1/miR-200 and SNAIL2, miR-33a-5p and ZEB1, miR-145 and ZEB2, miR-200 and Foxf2, miR-30a and SOX4, or miR-15a/16-1 and AP4 [133, 134, 137, 139,140,141,142,143]. These feedback loops may explain how imbalanced expression between microRNAs and EMT-TFs causes reinforced activation of EMT and steady mesenchymal specification once EMT is completed, as well as the reversibility of EMT and MET [144].

Furthermore, miRNAs also regulate the expression of EMT-TFs indirectly, thereby controlling EMT progression. miRNA let-7 and miR-365 control the expression of SNAIL1 and TWIST by deregulating high-mobility group A2 (HMGA2), a chromatin-binding protein that activates SNIAL and TWIST [145, 146]. In addition to regulating major transcription factors, miRNAs could regulate EMT by directly interacting with epithelial or mesenchymal genes [20]. For example, the miR-9 increased the motility and invasiveness of tumor cells by directly suppressing the expression of the E-cadherin-encoding messenger RNA [147]. In addition, miR-194 directly interacted with several 3′ untranslated regions (3′-UTR) of multiple mRNAs such as N-cadherin mRNA and lowers its expression in advanced-stage gastric cells [148]. Overexpression of miR-194 inhibited migration, invasion, and metastasis of hepatic cancer cells [149]. Furthermore, several miRNAs, including miR‑491‑5p, miR‑155, miR‑24, and miR‑124, regulate EMT progression by targeting cell architectural components [150,151,152,153].

Long noncoding RNAs (lncRNAs) are also involved in EMT regulation. For example, lncRNA-activated by TGF-β (lncRNA-ATB) and lncRNA-PNUTS are thought to act as sponges of the miR-200 family and miR-205, respectively, and isolation of these miRNAs prevents them from suppressing EMT-TF transcription [154, 155]. In addition, translational regulator lncRNA negatively regulates the translation of CDH1 mRNA to promote EMT [156]. H19 can mediate EMT by differentially binding to miR-200b, miR-200c, Let-7, SNAIL2, and Ezh2 [157]. In EMT cases of murine and human breast cancer, lncRNA EMT-associated transcript 20 (ET-20) was found to be transcribed in an antisense manner through the Sox4 EMT master transcription factor and co-regulates with the Tnc gene to bind to bridging proteins on the cell membrane. This process thus results in the impairment of intercellular junctions and enhancement of EMT. [158]. In conclusion, the regulatory activities of miRNAs and lncRNAs as post-transcriptional regulators form a complex regulatory network that controls EMT. In particular, while miRNAs control EMT by regulating EMT transcription factors or regulators, overexpression of lncRNAs in various cancers can induce EMT and promote tumor metastasis.

The relationship between EMT and tumors

Unlike normal tumor cells, cancer stem cells (CSCs) have the function of initiating and maintaining tumor growth, self-renewal, and proliferation. As a subgroup of CSCs, metastatic cancer stem cells (MCSCs) can receive matrix signals from the distal organ environment and escape from the boundary of the primary tumor [24]. Activation of EMT promotes the invasive phenotype of MCSCs, and makes MCSCs participate in the cascade process which is composed of three steps: first, the tumor cells invade to surrounding tissues; second, the trans-endothelial migration of tumor cells into the circulation system; third, the tumor cells colonize in distal tissue which results in the formation of metastatic foci (Fig. 3). EMT-TFs act as key regulators of CSC; thus, EMT is closely related to the acquisition of tumor cell stemness.

Fig. 3
figure 3

Role of EMT in the tumor metastasis. Tumor cells in situ are induced by EMT to initiate the metastatic cascade process. Intermediate state and mesenchymal stem cells (MCSCs) lose intercellular junctions, detach from tumor tissue, and invade surrounding tissues. Metastatic tumor cells enter the circulation through the endothelial barrier by active or passive trans-endothelial migration (TEM), invading the surrounding mesenchyme and disrupting endothelial junctions. After entering the circulation, single or clustered circulating tumor cells (CTCs) can bind to coagulation factors on platelets by expressing thrombin, forming a unique immune mechanism that protects metastatic cells from immune escape. Neutrophils are also recruited and bind to CTCs to promote tumor cells survival and proliferation. CTCs move slowly, roll along the endothelium, and then arrest. MCSCs anchor with endothelial cells for extravasation and then colonize distal organs via the MET process

The roles of EMT and MET in tumor metastasis

During the process of cell detachment from in situ, the downregulation or degradation of E-cadherin and the upregulation of N-cadherin destabilize AJs and disrupt the intercellular junctions of MCSCs, promoting cell migration and invasion [29, 159]. The activation of the “cadherin switch” can be used as a marker for the initiation of EMT. By regulating pro-apoptotic and anti-apoptotic genes, the “integrin switch” enables cells to make use of survival signals through overexpression or structural activation of integrins, and then changes the cell metabolism to prevent MCSCs from losing their nest due to apoptosis [160, 161]. During tumor cell invasion, the response to internal and external signals of the cell promotes the formation of stress fibers that inhibit cofilin and actin, allowing the cytoskeleton to reorganize dynamically [162]. The attachment sites formed by protrusions of actin between the cell and the ECM promote cell elongation and mesenchymal migration enabling cell motility [163]. The master cell guides cells that maintain intercellular junctions to migrate in narrow lines, clusters, or broadsheets, and the phenotype of the cells is influenced differently by the following cells [164, 165]. In contrast, cells that have lost their intercellular junctions can migrate in both mesenchymal and amoeboid forms [162].

After the early steps of local invasion of the metastatic tissue by EMT cells, MCSCs located at the tumor front begin to invade the surrounding tissues into the blood vessels or lymphatic vessels. EMT is a process in which cells transition from epithelium to mesenchyme and leading to the emergence of various hybrid phenotypes [13, 17, 41, 166]. The E/M hybrid phenotype, for example, is crucial in the metastatic spread of tumor cells because it has a high degree of epithelial–mesenchymal plasticity (EMP), a property that makes it heterogeneous during metastasis and contributes to better tumor cell aggregation and dissemination [167]. In related studies, a mixed E/M phenotype was found in heterogeneous circulating tumor cells (CTCs) in human lung and breast cancer patients [168,169,170]. During the intravasation process, MCSCs invade the surrounding stroma and disrupt endothelial junctions in order to transcend the endothelium barrier into the circulation by active or passive migration for trans-endothelial migration [171].

The expression of Notch, VEGF, and TGF-β signaling pathways have significant effects on endothelial function. Histone proteases can degrade BM and ECM components and activate urokinase-type plasminogen activator (uPA), which can then mediate ECM remodeling [172]. Matrix metalloproteinases can cleave E-cadherin, induce EMT via EGFR signaling, and promote neoangiogenesis by releasing growth factors from degraded ECM [173]. By expressing thrombin, MCSCs can bind to coagulation factors on platelets upon entry into the circulation, forming a unique immune mechanism to protect metastatic cells from immune escape and maintaining mesenchymal properties by activating the SMAD and Notch pathways [174, 175]. Collectively, migrating cells or aggregated individual cells in the circulation system can form swarms of mobile metastatic cells, which have better survival and metastatic potential [176, 177]. Next, the metastatic cells can find a suitable microenvironment to protect themselves from apoptosis and retain at the new distant metastasis sites. Then they migrate and exudate across endothelial cells and invade the tissue around the blood vessels. This homing mechanism and circulatory pathway enable MCSCs to migrate to and colonize distant organs [178].

The motility of circulating metastatic cells is slowed in capillaries that are similar in size to the cells themselves, which roll along the endothelium before arresting [179]. The binding of intercellular adhesion molecules (ICAM1), galactose lectin 3, and selectin expressed on endothelial cells to integrins, CD44, and MUC1 expressed on metastatic cells mediates the anchoring of MCSCs to the endothelium. Tumor cells eventually colonize through the MET process, and metastatic cells reacquire epithelial features during EMT reversal, generating distal secondary tumors that are histopathologically similar to the primary tumor [171]. A study examined mammary tumor cells in mice using in vivo microscopy techniques and found that they can undergo the EMT process spontaneously and revert to an epithelial state after mobility, migration, invasion, and colonization, corroborating the aforementioned process [180].

EMT is the key driver of tumor metastasis

EMT is a multidimensional and nonlinear process, it is difficult for in vitro experiments to accurately reflect the dynamic EMT process that cancer cells undergo in vivo [181]. Consequently, most of the available studies are based on cultured cell lines or xenograft models. In vivo studies focus on clarifying the characteristics of EMT by injecting or xenografting parental or manipulated cancer cells, whereas in vitro experiments mainly investigate the functional role of EMT-TFs through their acquisition or loss. EMT is hypothesized to be a driver of cancer progression. In related studies, researchers traced the process of SNAIL1 endogenous expression in tumors and found SNAIL1 activation and EMT in primary tumor cells, which eventually spread [182]. Furthermore, a decrease in SNAIL was found to inhibit the development of metastasis in the PyMT breast cancer model [183]. In another breast cancer model, the absence of TWIST resulted in a decrease in tumor metastasis [184]. Deletion of ZEB1 prevents tumor cells from invading and migrating [68]. In melanoma patients, ZEB2 deletion reduced their survival rate, which was verified in a mouse model of the disease [75, 76]. Therefore, these studies suggest that EMT is important for tumor metastasis.

Context-dependent EMT program

Although the EMT-TFs are the key drivers of tumor initiation, progression, and metastasis, several controversial results and conflicting data about the crucial role of EMT and EMT-TFs in cancer metastasis continue to be extensively debated [185,186,187,188,189]. It was observed that deletion of SNAIL or TWIST did not significantly inhibit metastasis of tumor cells in genetically engineered mouse models of pancreatic ductal adenocarcinoma [186]. In addition, a study conducted by Fisher et al. developed and used a mesenchymal-specific Cre-mediated fluorescent labeling switch system to track EMT in spontaneous breast-to-lung metastasis and found that a small proportion of primary epithelial tumor cells manifested EMT phenotype, but lung metastasis was achieved using non-EMT cells [185]. In a relevant mouse breast cancer model, by tracking the endogenous SNAIL activation, it was found that only primary tumors had SNAIL1 activation and endogenous EMT production [182, 190]. In the MMTV-PyMT mouse model of metastatic breast cancer, E-cadherin gene-negative, as well as N-cadherin gene-positive tumor cells undergoing EMT, were found to be the cells that migrated and possibly initiated the metastatic cascade. But vimentin gene activity was not detected during metastasis. These studies confirmed the presence of EMT but did not demonstrate its necessity for metastasis, suggesting that tumor cells may use the unique EMT gene program to metastasize and colonize distant organs [180, 191]. Therefore, the unnecessary association between EMT and metastasis may be influenced by context-dependent, which can be varied in different tumor types with different EMT requirements.

An excellent review has recently discussed the non-redundant functions of each particular EMT-TF and demonstrated that the tissue context is critical for the precise functions of EMT-TFs due to their distinguishing expression in development, tissue homeostasis, and different tumor types [43]. Even with the similar morphology of migratory tumor cells, carcinoma cells that have undergone EMT program in different tumor types may be regulated by diverse gene expression profiles [42]. A good example is that the CTCs from patients with lobular breast cancer display epithelial phenotypes, whereas those from HER2+ and triple-negative subtypes exhibit mesenchymal phenotypes [170]. The differences in attribute, expression pattern, regulation, target gene signature, and function among EMT-TFs may determine the differential hierarchical role of EMT-TFs in cancer biology [43]. In addition, the EMT-TFs have the ability to modulate each other in a complex, dynamic and interdependent manner. The ablation of one EMT driver could be compensated by an alternative EMT program [187]. Investigating the specific EMT-TFs activation in a given tumor type is critical to defining the precise role of EMT in tumor progression and metastasis.

The role of partial EMT

As a continuous flux between the extreme states of epithelial and mesenchymal, the process of EMT could be transited and reversible [43]. The binary transition between epithelial and mesenchymal phenotypes has been considered as the basis for EMT and MET. Recent studies have demonstrated that EMT is not a linear process but a spectrum and cancer cells are usually in a state between epithelium and mesenchyme [6, 14, 192]. Based on the spectral characteristics of EMT, related studies have successively introduced scoring algorithms for EMT. One study quantitatively estimated and scored the degree of EMT of tumor cell lines by the generic characteristics of EMT and correlated it with the efficacy of patient survival and drug response to assess patient prognosis [193]. Another study scored the partial EMT gene expression indexes based on genomics and proteomics, characterized and quantified the EMT state of the patient's tumor cells, and analyzed the survival of the patients [194].

Epithelial and mesenchymal markers are co-expressed or epithelial markers are absent without acquiring mesenchymal markers as a common manifestation of partial EMT, and this can place cancer cells in a dynamic window that may endow them with higher epithelial–mesenchymal plasticity for tumor progression and metastasis [43]. A recent study using a genealogical labeling approach and sequencing found that CDH1 epithelial protein was internalized in most tumors in the RAB11+ cycle, whereas at the protein level CDH1 mRNA remained unchanged in most cells [195]. Additionally, it has been found that tumor cell lines with partial EMT can produce clusters of circulating tumors in vivo and collectively migrate in vitro, whereas tumor cell lines with full EMT spread in form of single cells both in vivo and in vitro. In the KPCY model, it was found that cells in complete EMT often invaded and spread as single cells, while cells in the partial EMT stage often migrated more aggressively as a collective [195]. In the MMTV-PyMT mouse model of primary tumors, tumor cells rarely undergo a complete EMT program. Instead, they undergo partial EMT and can migrate, invade surrounding tissues at high rates, initiate metastatic growth, and return to an epithelial state to localize to distal organs via the MET program [196]. The application of the single-RNA sequencing technique can help us to delineate a clearer picture of EMT programs in tumor progression. Through single-cell sequencing technology, it was found that partial EMTs under-express SNAIL1/2, ZEB1/2, and TWIST1/2 during mouse organogenesis which is consistent with epithelial gene expression and they may be a transient population during development [197]. Puram et al. identified a subset of malignant cells with a gene signature associated with partial EMT in primary and metastatic head and neck squamous cell carcinomas using the single-RNA sequencing technique [198]. Interestingly, although these partial EMT cells exhibited several classical EMT features such as the expression of VIM, TGFβ-induced (TGFBI), and extracellular matrix genes, the EMT-TFs expression was significantly low, and the expression of the epithelial gene was still maintained [198]. Therefore, this confirmed the aggressive and highly plastic characteristics of partial EMT cells [198]. In a study conducted by Pastushenko et al. different combinations of the markers CD106, CD51, and CD61 were used to identify different states of tumor transition that occur during the EMT of cancer progression, distributed throughout the EMT spectrum, and introduction of the term “hybrid” EMT [14] (Fig. 4). Computational modeling studies have also revealed nonlinear multistable EMT dynamics. Besides, it was indicated that the intermediate hybrid EMT state is regulated by the feedback loops at the core of the EMT regulatory network, particularly the mutual inhibitory loops between several miRNAs and EMT-TFs [6, 199,200,201,202,203,204]. Future studies should focus on the dynamics of EMT spectrum, which will further clarify the role of EMT program in tumor metastasis.

Fig. 4
figure 4

Features of partial EMT. The partial EMT as an intermediate state is not only phenotypically different from the epithelial and mesenchymal states, but also has greatly altered cellular properties. There is a synthesis of the literature; it has intermediate polarity and loose intercellular junctions compared to the two and has metastatic potential, but the response to treatment is not clear. In addition, partial EMT has plasticity and can differentiate into different cells and move in clusters within the body

EMT and circulating tumor cells

Circulating tumor cells (CTCs) are cancer cells that have spread from the primary tumor to various organs through the vascular or lymphatic system [205]. Many studies have confirmed that certain CTC subpopulations of EMT metastases contain prognostic information. Circulating tumor cells represent a heterogeneous group, and the various hybrid phenotypes of EMT from the epithelium (E) to mesenchyme (M) play significant roles in tumor heterogeneity [13, 199]. Therefore, many EMT-related target genes can be used as the key indicators for CTC detection, among which vimentin and several epithelial adhesion molecules (E-cadherin, N-cadherin, and EpCAM among others) are frequently analyzed in CTC [167, 206, 207]. The ZEB (ZEB1 and ZEB2), SNAIL (SNAIL and Slug) family, and the TWIST family which are the core transcription factors regulating EMT target genes, are also frequently used as indicators for CTC assessment [6, 43, 208].

Because of the dynamic and reversible nature of EMT, it can improve the restrictive environment for tumor cell metastasis and EMT also affects the metastatic ability of the CTCs. First, EMT can promote CTC release; on the one hand, EMT stimulates angiogenesis and hence it has also been found that powerful angiogenic factor VEGF-A, which is an EMT target gene, is expressed in CTCs from patients with breast cancer [209]. In addition, EMT can promote the migration of tumor by inducing protein hydrolases such as matrix metalloproteinases (MMPs) [210]. High MMP activity has been reported in CTCs isolated from patients with prostate cancer [211]. A study found that platelets and EMT process jointly regulate intravasation/extravasation of CTCs [212]. During the process of intravasation, platelets bind to CTCs to release related cytokines, which promote the separation of tumor cells from the primary site and infiltrate into the blood and enter the circulation. Platelet-derived TGF-β and platelet-derived growth factors can drive the EMT process, thus endowing CTCs with metastasis and survival. The invasive ability enables CTCs to penetrate the vascular barrier to achieve extravasation.

The CTCs undergoing EMT are better able to resist apoptosis or anoikis because the EMT activates many survival pathways [213]. Typical EMT markers such as EMT transcription factors or stem cell markers which are related to molecules of survival pathways (such as EGFR, Akt, PI3K Bcl-2, and p53) were detected in CTCs isolated from colorectal, ovarian, breast, and other cancers [214,215,216,217]. Besides, clusters of CTCs that enter the circulation by the way of EMT may have different cellular phenotypes. A different study found that among the isolated CTCs, there were both cells with CTC clusters, cells with mesenchymal markers (including fibronectin, N-calmodulin, or PAI-1), and epithelial cell markers (such as EpCAM or cytokeratin), and the most of CTCs are EpCAM positive types [170].

Hordes of emergent EMT hybrid CTCs may establish the ability of cell–cell interactions to contribute to better resistance to anoikis/apoptosis and pro-transfer EMT-driven properties [218]. Therefore, CTCs travel in the form of a cluster with hybrid phenotypes of EMT may exhibit better survival and higher seeding efficiency in bloodstream and secondary sites. By expressing tissue factor (TF), the CTCs undergoing EMT will also be more effective in activating coagulation and building a protective cocoon. Tissue factor (TF) can be a target gene for EMTs and the EMT core transcription factors ZEB1 as well as SNAIL can regulate TF expression, with vimentin stabilizing TF mRNA [219, 220]. The TF/EMT relationship has also been confirmed through the correlation between vimentin and TF expression in studies of triple-negative breast cancer (TNBC).

Furthermore, the EMT induces multiple receptors that mediate the interaction of neutrophils with CTCs or CTCs with platelets/fibronectin (including CD44, ICAM1, αvβ3, or VCAM1), thereby driving cell cycle progression in the circulation system and expanding the metastatic potential of CTC [221,222,223,224,225,226]. Neutrophils have also been shown to induce EMT in several cell systems through the release of soluble factors, including CXCL-1, IL-17, or neutrophil elastase [227,228,229]. In addition to the above mechanisms, EMT can also induce immune escape, including increased expression of immune checkpoint proteins, altered autophagy, immunoproteasome defects, and immune synaptic dysfunction [230, 231].

A large number of clinical studies such as those in breast, prostate, and rectal cancer support the use of CTC counts as a valid prognostic biomarker before or during cancer treatment (chemotherapy or targeted therapy) [232,233,234,235,236]. Previous studies have shown that certain CTC subpopulations of EMT metastases have prognostic information and hence prognostic-related information can be found by detecting other typical EMT markers [237, 238]. A correlation between PD-L1 expression and EMT markers has also been demonstrated in tumors and CTCs, particularly in NSCLC and TNBC [239]. Therefore, combined detection of known therapeutic targets such as EGFR, PD-L1, or HER2 as well as EMT markers with poorer prognosis could point to potential combination therapies and improve patient care [240]. Furthermore, the emergence of CTCs phenotypes undergoing EMT correlates with drug resistance and the use of EMT as a concomitant marker in the treatment process may help predict the development of resistance and guide clinical treatment strategies [241,242,243]. There is still a discussion about EMT signatures in CTCs. Therefore, the selection of appropriate CTC isolation techniques, specific EMT markers, and transfer to determine clear parameters for clinical routine are important for future clinical translational application as well as the targeted therapy of CTC for EMT.

Relationship of EMT and tumor stemness

To explain how the propagation and dissemination of completely heterogeneous tumors at secondary sites is achieved by CSCs, the induction of EMT and stemness at the front of tumor invasion was firstly proposed [244]. Many epithelial tissues are maintained by stem cells that exhibit two remarkable phenotypes: one is the ability to differentiate and generate daughter cells with specific functions associated with epithelial phenotypes, whereas the other is the ability to self-renew, thus maintaining the stem cell pool [245]. In this section, we discussed the relationship between EMT and tumor cell stemness.

The CD44, CD24, CD133, and aldehyde dehydrogenase 1 (ALDH1) are the commonly used CSC surface markers. In the breast, EMT-derived stem cells are phenotypically similar to CSCs after induction, expressing CD44High, CD24low, and forming mammospheres [221]. Since then, this feature has been found in many CSC subpopulations and other studies have reported that high intracellular ALDH1 activity is a marker of stemness [246, 247]. Recent studies have shown that the activity of CD44high and ALDH1 do not usually coexist in CSC, which reveals the presence of different types of CSC [248]. A separate previous study focused on lung cancer with tobacco carcinogens exposure found a remarkable phenomenon. It was evident that CD44high/ CD24low cell population acquired a dramatic increase under brief carcinogen exposure. Further, the expression of CD133 and ALDH1 also improved but not significantly [249]. However, the expression of stem cell markers is not always consistent in the same subtype of breast cancer [250]. Therefore, the combination of CD44 subtype binding ALDH1 activity and integrins is promising in the study of primary tumors and their metastatic proliferation.

In addition, the EMT process plays a crucial role in enriching the CSCs pool in breast cancer. Specifically, PGC-1α and miR-200c are progressively inhibited by the EMT process and this results in mitofusin 1 aggregation. The NUMB gains phosphorylated and then dissociates from the cortical membrane with mitochondrial fusion, and the stem cells undergo asymmetric division to ensure a sustained CSC pool [251].

Signals of EMT-induced stemness

CSC are tumor cells with key characteristics of self-renewal, tumor initiation potential, and clonal long-term repopulation potential and are a small subset of malignant cells in tumors [252]. Further, the acquisition of stem cell properties is associated with the activation of EMT. Several EMT-TFs are involved in the stemness regulation of tumor stem cells as interference signals and are key regulators of CSC [253].

Among the different EMT-TFs, ZEB1 was the first one that was investigated in pancreatic cancer as a link from EMT to stemness. The experiment found that lowering ZEB1 causes pancreatic cell lines to lose the ability to form tumor spheres in vitro, and tumor initiation in vivo, hence confirming the requirement of stemness for Zeb1 [254].

Bmi1 is an epigenetic regulator that plays an important role in the maintenance of adult stem cell function and can also induce EMT [255]. Generally, Bmi1 is modulated using miR-200/-205 by binding 3’UTR sequence. Therefore, miRNA families repress CSC stemness through Bmi1 suppression-resulted cell apoptosis, senescence, and differentiation [256, 257]. Bmi1 can also be indirectly repressed by ZEB1. Furthermore, previous studies have shown that ZEB1 directly represses the miR-200 family, which in turn represses the expression of Bmi1 [258, 259]. The results explain how EMT regulates stem cells in this context and reveal a ZEB1/miR-200/Bmi1 pathway in pancreatic CSCs.

Upon inhibition by H3K27me3, the promoter of miR-200/-205 gets methylation, which relieves ZEB1 and increases EMT [249]. In addition, Bmi1 binds to PTEN (phosphatase and tensin homolog) at human nasopharyngeal epithelial cells and initiates PI3K/Akt/GSK-3β signaling to promote EMT in the nasopharyngeal carcinomas process [256]. Elsewhere, the pathway by which the ZEB1/miR-200/Notch signaling axis regulates CSCs was also elucidated [260].

The ability of pancreatic cancer cells to form tumor spheres was reduced through inhibition of the Notch pathway (achieved using γ-secretase inhibitors or knockdown of Jag1/Maml2/Maml3) suggesting that cell stemness of CSCs can be induced through the Notch pathway [261]. Notch signaling components Jag1, Maml2, and Maml3 are direct targets of miR-200 inhibition in breast and pancreatic cancer cell lines. Moreover, it has been found that Zeb1 inhibits miR-200 family members, resulting in de-repression of the expression of the three signaling components [262].

In the context of head and neck squamous cell carcinoma (HNSCC), it was found that EMT-TF, Twist can control CSC stemness by signaling to Bmi1. It directly binds the Bmi1 promoter and activates Bmi1, and hence TWIST1 can upregulate BMI1. Both the tumor spheroid and tumor initiation properties of HNSCC cells can be induced by Twist and Bmi1, and thus Bmi1 is essential for the stem induction ability of TWIST in these cells [81]. In epithelial ovarian cancer (EOC), TWIST1 is regulating the miRNAs miR-199a and miR-214 to control stem cell differentiation, and these miRNAs can control cell proliferation, apoptosis, and inflammation [263].

Of the SNAIL family, EMT-TF is not only associated with the stemness characteristics of the multilayered epithelium, but may also be involved in its signaling regulation [264]. It was found that the ability of SNAIL to significantly induce tumor spheres and tumor initiation in rectal cancer cell lines was the highest expressed EMT-TF. Expression of Numb which is an inhibitor of the Wnt signaling effector protein β-catenin can be inhibited by miR-146a. Further, it was evident that the SNAIL bound and activated the promoter of miR-146a, which in turn promoted the activation of the Wnt pathway, fully validating it as a stem regulator. Therefore, they did not only confirm the role of the SNAIL/miR-146/Numb/β-catenin pathway in promoting SC properties, but also found a poor prognostic and treatment resistance correlation of SNAIL High NUMB Low in a group of patients with colorectal cancer in a clinical study [265]. Elsewhere, it has been found that Slug which is a member of the SNAIL family, can induce mammary stem cell (MASC) through synergistic activation of different autoregulatory gene expression programs with sox9 [266]. Therefore, stemness and mesenchymal properties of tumor cells were found to be maintained through the KLF4/TGF-β1/Smad/SNAIL pathway in a human colorectal cancer model [267].

Although many reports suggest that cells undergoing full EMT may acquire the stemness, the relationship between complete EMT coupled with stemness has been challenged by some studies [268,269,270,271,272]. To reconcile these conflicting results, existing studies suggest that cells in a mixed E/M or partial EMT state are more likely to acquire stemness than cells in a pure epithelial or mesenchymal state [273,274,275,276]. This model is supported by several lines of evidence from studies of sphere formation and tumorigenicity in prostate and breast cancer models [6, 221, 268, 274]. Furthermore, in an ovarian cancer model, it has been found that the partial EMT phenotype increased tumor stemness, whereas loss of stem cell markers and tumorigenicity was associated with a fully epithelial or fully mesenchymal phenotype [275]. This implies that there is a “stemness window” between fully differentiated epithelial cells and fully differentiated mesenchymal cells [199, 277]. In other words, cells with a partial/hybrid EMT phenotype, rather than cells locked in a full EMT phenotype, have higher plasticity in tumor invasion and proliferation and can complete the invasive metastatic cascade.

In a clinical study, the subpopulations of CTC were analyzed according to stem and EMT markers after chemotherapy. It was found that only CSC+/partial EMT+ CTCs (co-expressing stem and partial EMT phenotypes) were highly enriched after chemotherapy which was associated with pulmonary metastases and lack of treatment response. Therefore, CSC+/partial EMT+ CTCs can be used as a prognostic marker for metastatic breast cancer patients receiving first-line chemotherapy [242]. In addition, effectively targeting a subpopulation of CTCs with stem cell properties and high metastatic potential has the potential to improve patient survival and may be a promising avenue for cancer treatment.

Therapeutic strategies and challenges for EMT

The EMT may be resistant to a variety of treatments, including chemotherapy, radiotherapy, and activation of EMT controls resistance to treatment at multiple levels [278]. Increased drug efflux or avoidance of apoptosis and necrosis are the common pathways of drug resistance [19]. As an important example, SLUG and SNAIL can avoid treatment-induced apoptosis by interfering with p53 function or inhibiting the tumor suppressor PTEN (phosphatase and tensin homolog) [213]. Two separate studies have found that in genetically engineered mouse models of breast and pancreatic cancer, the primary and metastatic tumor cells became resistant to chemotherapeutic agents with an EMT-dependent presentation [185, 186]. Further, the activation of EMT also confers carcinoma cell ability to induce local immunosuppression, hence compromising immunosurveillance and contributing resistance to immunotherapy. Mesenchymal carcinoma expressing EMT markers exert immunosuppressive effects in multiple ways, including secreting chemokines and cytokines, thus promoting the formation of regulatory T (Treg) cells, recruiting M2 macrophages, blocking cytotoxic activities of T lymphocytes (CTL) and natural killer (NK) cells, and inhibiting antigen presentation of dendritic cells (DCs) [110, 279,280,281,282].

Notably, EMT in carcinoma cells promotes the expression of PD-L1 and a higher EMT score correlates with tumors that respond best to CTL-A4, PD1, and PD-L1 antibodies and with tumors that express other increased immune checkpoint markers [193, 283,284,285]. In addition to inducing resistance to various treatments, EMT can also be induced after treatment, including activation of EMT-promoting pathways via TGF-b, NF-κB, WNT, FGF, and EGF/HER2. Therefore, induction of EMT may exhibit adaptations in response to treatment-induced cellular stress [230, 286]. Further, the activation of EMT after treatment leads to further acceleration of the disease process through mechanisms including increased proliferation, decreased apoptosis, immunosuppression, increased stemness, and metastasis [286].

Prevention or reversing the lethal effects of EMT is of great importance for cancer treatment. Currently, there are three main strategies for targeted EMT therapy (Fig. 5) (Table 1). First, it can inhibit tumorigenesis by blocking upstream signaling pathways. This includes ligand-neutralizing antibodies, decoy receptors, or inhibitors blocking TGFβ, NF-κB, EGFR, cMET, WNT, and Notch signaling [286,287,288]. In addition, effective inducers of EMT include a variety of pro-inflammatory signals such as TNF-α [289]. Another therapeutic strategy is to target the molecular drivers of EMT. Although EMT- TFs are the main drivers/regulators of the EMT process, direct targeting of the transcription factors (EMT-TFs) is challenging [68]. Further, several EMT-TFs have a complementary and redundant function because they tightly connect via feedback mechanisms. Therefore, targeting their interactions with important cofactors may be a more beneficial strategy while also targeting multiple EMT-TFs [68].

Fig. 5
figure 5

Therapeutic strategies for targeting EMT. EMT may be resistant to various treatments such as chemotherapy and radiotherapy, and the activation of EMT after treatment leads to further acceleration of the disease process by mechanisms including increased proliferation, decreased apoptosis, immunosuppression, stemness, and increased metastasis. There are three main strategies for targeted EMT treatment: I) inhibiting tumorigenesis by blocking upstream signaling pathways, II) targeting the molecular drivers of EMT, and III) targeting mesenchymal cells and outer stroma or inhibiting the MET process

Targeted regulation of epigenetics is a reversible and stable way of inducing the reprogramming required for phenotypic switching in the EMT process [290]. The application of modified synthetic miRNAs that can interfere with EMT-TF at the post-transcriptional level, such as liposomal miR-34 (MRX34) on tumor growth and metastasis has been previously used in clinical trials [291].

Another therapeutic strategy is to target mesenchymal cells and ECM or inhibition of the MET. Inducing re-differentiation or trans-differentiation by inhibiting the function of mesenchymal-specific proteins, blocking cell plasticity, and re-sensitizing tumor cells to standard treatments. For instance, using biologically active compounds or monoclonal antibodies against vimentin, fibronectin, and N-cadherin [292]. A study has found that there is a group of mixed epithelial/mesenchymal phenotypic cells called quasi-mesenchymal (qM) cells in the EMT process, which will metastasize and become resistant to immune checkpoint blocking therapy. In this study, knockout-derived factors (CD73, CSF1, or SPP1) targeting qM cancer cells can prevent the assembly of immunosuppressive tumor microenvironment, promote their transition into epithelial phenotypic, and improve the sensitivity to checkpoint immunotherapy [293].

Another therapeutic strategy is to target the interaction between cancer cells and ECM. Integrins, a family of ubiquitous cell membrane adhesion receptors, play an imperative role in physiological processes via their attachment to the ECM [294]. The interaction between cancer cells and the ECM are mostly mediated by integrins, which further activates the signals involved in the initiation, progression, and metastasis of solid tumor [288, 294]. Thus, integrins present themselves as an attractive target for cancer treatment due to their crucial role in tumor progress and metastasis. Indeed, inhibiting the function of integrin was shown to reduce tumor metastasis in animal models and maintain a stable disease state in clinical trials. [288, 294,295,296,297,298]. In parallel with the launch of EMT program in cancer cells, the degradation of ECM is initialed by the synthesis of proteases of the MMPs family, which further allows cancer cells to enter the circulatory system and implant in distant tissue [299]. In summary, several potential anti-cancer drugs MMPs and their inhibitors have been extensively studied [300, 301].

However, research on targeted EMT treatment strategies is challenging and the serious adverse effects caused by EMT-targeted therapy cannot be neglected. Targeted EMT therapy can inhibit the migration and invasive behavior of tumor cells in the primary tumor, but this approach only works if the early spread of cancer cells has not occurred [302, 303]. In two studies of genetically engineered mouse models of breast and pancreatic cancer, primary and metastatic tumor cells, EMT was shown to confer stronger resistance to chemotherapeutic drugs [185, 186]. Therefore, in the early stage of cancer development, EMT-targeted therapy combined with conventional chemotherapy can improve the sensitivity of tumor cells to drugs. EMT-targeted therapy can also prevent tumor cells from further spreading, and lock them in the primary site with a clear boundary, which is beneficial to the resection of the primary tumor. In the later stages of the disease, EMT inhibitors can reduce the proportion of CSCs in tumors, prevent CTC colonization of the primary tumor and reduce the ability of CTCs to generate secondary tumors when seeded at distant sites [278, 304]. In a study of triple-negative breast cancer, a small molecule inhibitor of EMT, GSK-3β inhibitor BIO, could effectively and selectively inhibit the EMT and CSC, and migration characteristics of cells with mesenchymal and stem cell phenotypes [305]. When tumor cells lose their CSC properties, the expression of ABC proteins will decrease, reducing drug efflux and increasing the efficacy of chemotherapy [288]. It is still not clear whether patients, with early-stage disease or with advanced disease (with CTC detection and evidence of distant metastases) will benefit most from anti-EMT therapy. Therefore, this makes the selection of the appropriate timing of targeted EMT to be critical. Anti-EMT treatment may induce a contrary result than originally expected. The reverse process may be driven during anti-EMT treatment, leading to the development of MET, promoting the colonization, and metastasis of circulating tumor cells, and increasing tumor proliferation will make cancer cells more susceptible to chemotherapeutic agents [32]. As EMT plays very imperative physiological roles, the targeting process of EMT not only has an effect on tumor cell subpopulations but may also have a negative impact on normal cells [24]. In addition, inhibition of the EMT process may affect the repair function of the body because EMT plays an important role in the physiological response to trauma and wound healing [164]. Therefore, the targeted therapy of EMT has two sides and the research process is concerned about its effect, whereas the consequences of side effects are also worthy of attention.

Conclusion and perspectives

EMT is a highly regulated dynamic process that is angelic and demonic to an organism. During embryonic development and tissue repair, EMT is an essential presence, but the promotion of this process by EMT during tumor development often has undesirable consequences [18, 42]. It confers motility, stem cell properties, and therapeutic resistance to epithelial cells in a variety of normal and cancerous tissues. This process can be activated by a variety of signals and a regulatory network of multiple transcription factors and processes such as post-transcriptional and posttranslational modifications as well as epigenetic modifications govern the execution of EMT.

Regardless of our growing familiarity with the EMT program, the precise requirement of EMT and EMT-TFs in tumor metastasis is still debatable [185,186,187,188,189]. These contradictory results collectively demonstrated that the different attributes and expression patterns in different tumor types determine their precise role and function in cancer biology. A clear illustration that EMT effectors are in diverse cancer types should be further delineated. It is also worth noting that EMT in tumors is a spectrum of intermediate states rather than a binary process. Tumor cells often show a partial EMT phenotype that exhibits various degrees of epithelial and mesenchymal markers expressions [187]. Thus, it might be defective to fully capture all the ongoing EMT events (such as partial EMT) by tracing cells based on the single gene expression as described in previous studies [185,186,187]. Future studies should combine some omics approaches such as the single-cell RNA sequencing technique to dissect the dynamics and involvement of partial EMT precisely and comprehensively in tumor metastasis.

CSCs in the EMT process place tumor cells in an intermediate state of the E to M spectrum, a state in which cells are highly resistant to chemotherapy and can survive and generate new tumor cells that eventually leads to clinical relapse [17, 306]. In addition, increased drug efflux and avoidance of apoptotic signaling pathways have also been suggested as relevant resistance mechanisms for EMT [307]. EMT contributes to immunosuppression within the immune microenvironment, and activation of EMT-TFs leads to the accumulation of immunosuppressive cells in the tumor microenvironment. For instance, a related study found that activation of SNAIL in ovarian cancer upregulates CXCL-1 and CXCL-2, and activation of ZEB1 in breast cancer upregulates IL-6 and IL-8, both of which recruit MDSCs to the tumor microenvironment, leading to immunotherapy resistance and promoting cancer progression [308, 309]. Moreover, immune cells can secrete cytokines and chemokines to regulate EMT, and tumor cells that receive EMT can in turn produce immunosuppressive cytokines or chemokines. These two processes can complement each other to further promote cancer progression [310]. Therefore, immunotherapy targeting immunosuppressive cells in combination with immune checkpoint inhibitors represents a promising anti-cancer therapy .

Table 1 Inhibitors of EMT in clinical phase trials

Although the involvement of EMT in invasion, dissemination, and extravasation is essential for the propagation of primary tumors, EMT alone cannot complete the metastatic colonization of distal organs using tumor cells but requires the subsequent process of MET to restore the epithelial phenotype to complete the growth of metastatic lesions [14, 311]. Therefore, EMT-TF is variable and tissue-specific in promoting metastasis as well as essential in some cases, but not necessary in others. However, this evidence refutes the view that EMT is non-essential for tumor metastasis. The SNAIL has been shown to have such properties in breast cancer and TWIST1 in squamous cell carcinoma [182, 271]. The dispensability of EMT is dependent on the dominant role of EMT-TF and there may be a hierarchical relationship between EMT-TF. In addition, different tumors have very different effects on EMT-TF and EMT, and it would be crucial to evaluate the step of EMT works in which tumor, which would be the key to blocking EMT or clearing EMT.

Neither is EMT a single-cell state nor a binary process but a mixture of different types of cell states, which provides a “stem cell window” and makes EMT plastic, especially partial EMT [312]. Related studies have shown that different EMTs have different characteristics of infiltration, metastasis, and differentiation, when the cells of mixed epithelium and mesenchyme play a greater role in reaching circulation, colonization, and metastasis [14]. The mixtures of epithelial and mesenchymal subpopulations can be distinguished from exclusively hybrid E/M cells based on gene expression [313]. Subpopulations of EMT are spatially distributed in specific tissues, non-randomly in various parts of the tumor, exist in different microenvironments, and are associated with different stromal cells. It is now widely believed that some of the cells produced by partial EMTs contain mixed epithelium and mesenchyme which is a key factor in the invasion and spread of cancer cells. Further, this promiscuous feature also plays a crucial role in the successful metastatic colonization of disseminated cells [33]. Therefore, deciphering the full morphology of EMT and the transformation of cells between different states will help in understanding tumor heterogeneity, growth, invasion, metastasis, and drug resistance.

For cancer, although conventionally targeted approaches can be used or the tumor microenvironment can be used for treatment, plasticity of EMT can also lead to drug resistance. Therefore, it is a great challenge to target and destroy the cells that receive EMT in tumor tissue. Inhibition of EMT may inadvertently promote secondary tumor formation because of the plasticity of the procedure and the need for tumor cells to implant in distant organs requiring a return to an epithelial state. Alternatively, a trans-differentiation strategy can be used to induce EMT cells into a harmless cell and push them to an extreme EMT state, leading to their eventual differentiation or apoptosis and this has also been shown to be effective in breast cancer-related studies [312]. In a breast cancer model, it was observed that the infiltrating breast cancer cells can give rise to adipocytes by trans-differentiation, thereby inhibiting cancer metastasis. Moreover, targeting downstream effectors of EMT/MET is a more accurate therapy. During MET, the miR-200s downregulates Tinagl1 (a secreted metastasis inhibitory protein) and thus recombinant Tinagl1 reduces tumorigenesis and metastasis, whereas EMT induction is prevented by directly targeting miR-200 s [314]. Whatever, researchers should continually investigate the role of EMT in tumor development, and develop better treatments for targeting the EMT process.

Availability of data and materials

Not applicable.

Abbreviations

EMT:

Epithelial–mesenchymal transition

MET:

Mesenchymal–epithelial transformation

TFs:

Translational factors

BHLH:

Basic helix–loop–helix

CSCs:

Cancer stem cells

CTCs:

Circulating tumor cells

PAR:

Partitioning‑defective

PATJ:

PALS1-associated tight junction

ZO1:

Zonula occludens 1

MMPs:

Matrix metalloproteinase

PRC:

Polycomb repressor complex

EZH2:

Enhancer of zeste homolog 2

LSD1:

Lys-specific demethylase 1

RTKs:

Receptor tyrosine kinases

CTBP:

C-terminal-binding protein

SWI/SNF:

Switch/sucrose nonfermentable

TGF-β:

Transforming growth factor-β

BMP:

Bone morphogenetic protein

BMI1:

B lymphoma Mo‑MLV insertion region 1 homolog

HIF-1α:

Hypoxia-inducible factor-1α

FOX:

Forkhead box

SOX:

SRY box

PRRX1:

Paired-related homeobox 1

YB-1:

Y-box-binding protein-1

CPEB1:

Cytoplasmic polyadenylation element-binding protein 1

GSK-3β:

Glycogen synthase kinase-3β

PKD1:

Protein kinase D1

SCP1:

Small C-terminal domain phosphatase 1

PAK1:

P21-activated kinase 1

LATS2:

Large tumor suppressor 2

USP51:

Ubiquitin-specific protease-51

HMGA2:

High-mobility group A2

3′-UTR:

3′ Untranslated regions

lncRNAs:

Long noncoding RNAs

ET-20:

EMT-associated transcript 20

MCSCs:

Metastatic cancer stem cells

EMP:

Epithelial–mesenchymal plasticity

uPA:

Urokinase-type plasminogen activator

ICAM1:

Intercellular adhesion molecules

TGFBI:

Transformation of growth factor beta-induced

TNBC:

Triple-negative breast cancer

ALDH1:

Aldehyde dehydrogenase 1

HNSCC:

Head and neck squamous cell carcinoma

EOC:

Epithelial ovarian cancer

MASC:

Mammary stem cell

HF:

Hair follicle

CTL:

Cytotoxic activities of T lymphocytes

NK:

Natural killer

DCs:

Dendritic cells

qM:

Quasi-mesenchymal

References

  1. Polyak K, Weinberg RA. Transitions between epithelial and mesenchymal states: acquisition of malignant and stem cell traits. Nat Rev Cancer. 2009;9:265–73.

    Article  CAS  PubMed  Google Scholar 

  2. Gros J, Tabin CJ. Vertebrate limb bud formation is initiated by localized epithelial-to-mesenchymal transition. Science. 2014;343:1253–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Correa-Costa M, Andrade-Oliveira V, Braga TT, Castoldi A, Aguiar CF, Origassa CS, Rodas AC, Hiyane MI, Malheiros DM, Rios FJ, et al. Activation of platelet-activating factor receptor exacerbates renal inflammation and promotes fibrosis. Lab Invest. 2014;94:455–66.

    Article  CAS  PubMed  Google Scholar 

  4. Ruscetti M, Quach B, Dadashian EL, Mulholland DJ, Wu H. Tracking and functional characterization of epithelial-mesenchymal transition and mesenchymal tumor cells during prostate cancer metastasis. Can Res. 2015;75:2749–59.

    Article  CAS  Google Scholar 

  5. Chen T, You Y, Jiang H, Wang ZZ. Epithelial-mesenchymal transition (EMT): a biological process in the development, stem cell differentiation, and tumorigenesis. J Cell Physiol. 2017;232:3261–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Nieto MA, Huang RY, Jackson RA, Thiery JP. EMT: 2016. Cell. 2016;2016(166):21–45.

    Article  CAS  Google Scholar 

  7. Kalluri R, Weinberg RA. The basics of epithelial-mesenchymal transition. J Clin Invest. 2009;119:1420–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Mittal V. Epithelial mesenchymal transition in tumor metastasis. Annu Rev Pathol. 2018;13:395–412.

    Article  CAS  PubMed  Google Scholar 

  9. Pal A, Barrett TF, Paolini R, Parikh A, Puram SV. Partial EMT in head and neck cancer biology: a spectrum instead of a switch. Oncogene. 2021;40:5049–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Cano A. The transcription factor snail controls epithelial-mesenchymal transitions by repressing E-cadherin expression. Nat Cell Biol. 2000;2:76–83.

    Article  CAS  PubMed  Google Scholar 

  11. Diepenbruck M, Christofori G. Epithelial-mesenchymal transition (EMT) and metastasis: yes, no, maybe? Curr Opin Cell Biol. 2016;43:7–13.

    Article  CAS  PubMed  Google Scholar 

  12. Nieto MA, Huang RY-J, Jackson RA, Thiery JP. EMT: 2016. Cell. 2016;166:21–45.

    Article  CAS  PubMed  Google Scholar 

  13. Yang J, Antin P, Berx G, Blanpain C, Brabletz T, Bronner M, Campbell K, Cano A, Casanova J, Christofori G, et al. Guidelines and definitions for research on epithelial-mesenchymal transition. Nat Rev Mol Cell Biol. 2020;21:341–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Pastushenko I, Brisebarre A, Sifrim A, Fioramonti M, Revenco T, Boumahdi S, Van Keymeulen A, Brown D, Moers V, Lemaire S, et al. Identification of the tumour transition states occurring during EMT. Nature. 2018;556:463–8.

    Article  CAS  PubMed  Google Scholar 

  15. Santamaria PG, Moreno-Bueno G, Portillo F, Cano A. EMT: present and future in clinical oncology. Mol Oncol. 2017;11:718–38.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Cai Y, Wang B, Li B, Huang X, Guo H, Liu Y, Chen B, Zhao S, Wu S, Li W, et al. Collection on reports of molecules linked to epithelial-mesenchymal transition in the process of treating metastasizing cancer: a narrative review. Ann Transl Med. 2021;9:946.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Dongre A, Weinberg RA. New insights into the mechanisms of epithelial-mesenchymal transition and implications for cancer. Nat Rev Mol Cell Biol. 2019;20:69–84.

    Article  CAS  PubMed  Google Scholar 

  18. Lu W, Kang Y. Epithelial-mesenchymal plasticity in cancer progression and metastasis. Dev Cell. 2019;49:361–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Brabletz S, Schuhwerk H, Brabletz T, Stemmler MP. Dynamic EMT: a multi-tool for tumor progression. EMBO J. 2021;40: e108647.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lamouille S, Xu J, Derynck R. Molecular mechanisms of epithelial-mesenchymal transition. Nat Rev Mol Cell Biol. 2014;15:178–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Thiery JP, Acloque H, Huang RY, Nieto MA. Epithelial-mesenchymal transitions in development and disease. Cell. 2009;139:871–90.

    Article  CAS  PubMed  Google Scholar 

  22. Nakaya Y, Sheng G. Epithelial to mesenchymal transition during gastrulation: an embryological view. Dev Growth Differ. 2008;50:755–66.

    Article  CAS  PubMed  Google Scholar 

  23. Axelrod JD. Cell shape in proliferating epithelia: a multifaceted problem. Cell. 2006;126:643–5.

    Article  CAS  PubMed  Google Scholar 

  24. Singh M, Yelle N, Venugopal C, Singh SKEMT. Mechanisms and therapeutic implications. Pharmacol Ther. 2018;182:80–94.

    Article  CAS  PubMed  Google Scholar 

  25. Lee JM, Dedhar S, Kalluri R, Thompson EW. The epithelial-mesenchymal transition: new insights in signaling, development, and disease. J Cell Biol. 2006;172:973–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Voulgari A, Pintzas A. Epithelial-mesenchymal transition in cancer metastasis: mechanisms, markers and strategies to overcome drug resistance in the clinic. Biochim Biophys Acta. 2009;1796:75–90.

    CAS  PubMed  Google Scholar 

  27. Lim J, Thiery JP. Epithelial-mesenchymal transitions: insights from development. Development. 2012;139:3471–86.

    Article  CAS  PubMed  Google Scholar 

  28. Fletcher DA, Mullins RD. Cell mechanics and the cytoskeleton. Nature. 2010;463:485–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Yilmaz M, Christofori G. EMT, the cytoskeleton, and cancer cell invasion. Cancer Metastasis Rev. 2009;28:15–33.

    Article  PubMed  Google Scholar 

  30. Huang RY, Guilford P, Thiery JP. Early events in cell adhesion and polarity during epithelial-mesenchymal transition. J Cell Sci. 2012;125:4417–22.

    Article  CAS  PubMed  Google Scholar 

  31. Williams CM, Engler AJ, Slone RD, Galante LL, Schwarzbauer JE. Fibronectin expression modulates mammary epithelial cell proliferation during acinar differentiation. Cancer Res. 2008;68:3185–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Pasquier J, Abu-Kaoud N, Al Thani H, Rafii A. Epithelial to mesenchymal transition in a clinical perspective. J Oncol. 2015;2015: 792182.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Derynck R, Weinberg RA. EMT and cancer: more than meets the eye. Dev Cell. 2019;49:313–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. St Johnston D, Ahringer J. Cell polarity in eggs and epithelia: parallels and diversity. Cell. 2010;141:757–74.

    Article  CAS  PubMed  Google Scholar 

  35. Lee M, Vasioukhin V. Cell polarity and cancer–cell and tissue polarity as a non-canonical tumor suppressor. J Cell Sci. 2008;121:1141–50.

    Article  CAS  PubMed  Google Scholar 

  36. Nakaya Y, Sheng G. EMT in developmental morphogenesis. Cancer Lett. 2013;341:9–15.

    Article  CAS  PubMed  Google Scholar 

  37. Weston JA, Thiery JP. Pentimento: neural crest and the origin of mesectoderm. Dev Biol. 2015;401:37–61.

    Article  CAS  PubMed  Google Scholar 

  38. Bolós V, Peinado H, Pérez-Moreno MA, Fraga MF, Esteller M, Cano A. The transcription factor Slug represses E-cadherin expression and induces epithelial to mesenchymal transitions: a comparison with Snail and E47 repressors. J Cell Sci. 2003;116:499–511.

    Article  PubMed  CAS  Google Scholar 

  39. Bax NA, Pijnappels DA, van Oorschot AA, Winter EM, de Vries AA, van Tuyn J, Braun J, Maas S, Schalij MJ, Atsma DE, et al. Epithelial-to-mesenchymal transformation alters electrical conductivity of human epicardial cells. J Cell Mol Med. 2011;15:2675–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Barsan V, Cristofanilli M, Mani SA. Circulating breast tumor cells exhibit dynamic changes in epithelial and mesenchymal composition. Breast Dis A Year Book Q. 2013;24:225–6.

    Article  Google Scholar 

  41. Pastushenko I, Blanpain C. EMT transition states during tumor progression and metastasis. Trends Cell Biol. 2019;29:212–26.

    Article  CAS  PubMed  Google Scholar 

  42. Aiello NM, Kang Y. Context-dependent EMT programs in cancer metastasis. J Exp Med. 2019;216:1016–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Stemmler MP, Eccles RL, Brabletz S, Brabletz T. Non-redundant functions of EMT transcription factors. Nat Cell Biol. 2019;21:102–12.

    Article  CAS  PubMed  Google Scholar 

  44. Mathsyaraja H, Ostrowski MC. Setting Snail2’s pace during EMT. Nat Cell Biol. 2012;14:1122–3.

    Article  CAS  PubMed  Google Scholar 

  45. Barrallo-Gimeno A, Nieto MA. The Snail genes as inducers of cell movement and survival: implications in development and cancer. Development. 2005;132:3151–61.

    Article  CAS  PubMed  Google Scholar 

  46. Aybar MJ, Nieto MA, Mayor R. Snail precedes slug in the genetic cascade required for the specification and migration of the Xenopus neural crest. Development. 2003;130:483–94.

    Article  CAS  PubMed  Google Scholar 

  47. Herranz N, Pasini D, Diaz VM, Franci C, Gutierrez A, Dave N, Escriva M, Hernandez-Munoz I, Di Croce L, Helin K, et al. Polycomb complex 2 is required for E-cadherin repression by the Snail1 transcription factor. Mol Cell Biol. 2008;28:4772–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lin T, Ponn A, Hu X, Law BK, Lu J. Requirement of the histone demethylase LSD1 in Snai1-mediated transcriptional repression during epithelial-mesenchymal transition. Oncogene. 2010;29:4896–904.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Tong ZT, Cai MY, Wang XG, Kong LL, Mai SJ, Liu YH, Zhang HB, Liao YJ, Zheng F, Zhu W, et al. EZH2 supports nasopharyngeal carcinoma cell aggressiveness by forming a co-repressor complex with HDAC1/HDAC2 and Snail to inhibit E-cadherin. Oncogene. 2012;31:583–94.

    Article  CAS  PubMed  Google Scholar 

  50. Dong C, Wu Y, Yao J, Wang Y, Yu Y, Rychahou PG, Evers BM, Zhou BP. G9a interacts with Snail and is critical for Snail-mediated E-cadherin repression in human breast cancer. J Clin Invest. 2012;122:1469–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Manfredi AA, Ramirez GA, Rovere-Querini P, Maugeri N. The neutrophil’s choice: phagocytose vs make neutrophil extracellular traps. Front Immunol. 2018;9:288.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Thiam HR, Wong SL, Qiu R, Kittisopikul M, Vahabikashi A, Goldman AE, Goldman RD, Wagner DD, Waterman CM. NETosis proceeds by cytoskeleton and endomembrane disassembly and PAD4-mediated chromatin decondensation and nuclear envelope rupture. Proc Natl Acad Sci USA. 2020;117:7326–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kajita M, McClinic KN, Wade PA. Aberrant expression of the transcription factors snail and slug alters the response to genotoxic stress. Mol Cell Biol. 2004;24:7559–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhou W, Gross KM, Kuperwasser C. Molecular regulation of Snai2 in development and disease. J Cell Sci. 2019;132: jcs235127.

    Article  CAS  PubMed  Google Scholar 

  55. Jordà M, Olmeda D, Vinyals A, Valero E, Cubillo E, Llorens A, Cano A, Fabra A. Upregulation of MMP-9 in MDCK epithelial cell line in response to expression of the Snail transcription factor. J Cell Sci. 2005;118:3371–85.

    Article  PubMed  CAS  Google Scholar 

  56. Peinado H, Olmeda D, Cano A. Snail, Zeb and bHLH factors in tumour progression: an alliance against the epithelial phenotype? Nat Rev Cancer. 2007;7:415–28.

    Article  CAS  PubMed  Google Scholar 

  57. Qiao B, Johnson NW, Gao J. Epithelial-mesenchymal transition in oral squamous cell carcinoma triggered by transforming growth factor-beta1 is Snail family-dependent and correlates with matrix metalloproteinase-2 and -9 expressions. Int J Oncol. 2010;37:663–8.

    CAS  PubMed  Google Scholar 

  58. Li W, Shen M, Jiang YZ, Zhang R, Zheng H, Wei Y, Shao ZM, Kang Y. Deubiquitinase USP20 promotes breast cancer metastasis by stabilizing SNAI2. Genes Dev. 2020;34:1310–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Fan L, Lei H, Zhang S, Peng Y, Fu C, Shu G, Yin G. Non-canonical signaling pathway of SNAI2 induces EMT in ovarian cancer cells by suppressing miR-222-3p transcription and upregulating PDCD10. Theranostics. 2020;10:5895–913.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Gheldof A, Hulpiau P, van Roy F, De Craene B, Berx G. Evolutionary functional analysis and molecular regulation of the ZEB transcription factors. Cell Mol Life Sci. 2012;69:2527–41.

    Article  CAS  PubMed  Google Scholar 

  61. Postigo AA, Dean DC. ZEB represses transcription through interaction with the corepressor CtBP. Proc Natl Acad Sci U S A. 1999;96:6683–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Grooteclaes ML, Frisch SM. Evidence for a function of CtBP in epithelial gene regulation and anoikis. Oncogene. 2000;19:3823–8.

    Article  CAS  PubMed  Google Scholar 

  63. Sánchez-Tilló E, Lázaro A, Torrent R, Cuatrecasas M, Vaquero EC, 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;29:3490–500.

    Article  PubMed  CAS  Google Scholar 

  64. Postigo AA. Opposing functions of ZEB proteins in the regulation of the TGFbeta/BMP signaling pathway. Embo j. 2003;22:2443–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Aigner K, Dampier B, Descovich L, Mikula M, Sultan A, Schreiber M, Mikulits W, Brabletz T, Strand D, Obrist P, et al. The transcription factor ZEB1 (deltaEF1) promotes tumour cell dedifferentiation by repressing master regulators of epithelial polarity. Oncogene. 2007;26:6979–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Bindels S, Mestdagt M, Vandewalle C, Jacobs N, Volders L, Noel A, van Roy F, Berx G, Foidart JM, Gilles C. Regulation of vimentin by SIP1 in human epithelial breast tumor cells. Oncogene. 2006;25:4975–85.

    Article  CAS  PubMed  Google Scholar 

  67. Dave N, Guaita-Esteruelas S, Gutarra S, Frias À, Beltran M, Peiró S, de Herreros AG. Functional cooperation between Snail1 and twist in the regulation of ZEB1 expression during epithelial to mesenchymal transition. J Biol Chem. 2011;286:12024–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Krebs AM, Mitschke J, Lasierra Losada M, Schmalhofer O, Boerries M, Busch H, Boettcher M, Mougiakakos D, Reichardt W, Bronsert P, et al. The EMT-activator Zeb1 is a key factor for cell plasticity and promotes metastasis in pancreatic cancer. Nat Cell Biol. 2017;19:518–29.

    Article  CAS  PubMed  Google Scholar 

  69. Chaffer CL, Marjanovic ND, Lee T, Bell G, Kleer CG, Reinhardt F, D’Alessio AC, Young RA, Weinberg RA. Poised chromatin at the ZEB1 promoter enables breast cancer cell plasticity and enhances tumorigenicity. Cell. 2013;154:61–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Chen B, Chen B, Zhu Z, Ye W, Zeng J, Liu G, Wang S, Gao J, Xu G, Huang Z. Prognostic value of ZEB-1 in solid tumors: a meta-analysis. BMC Cancer. 2019;19:635.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Zhang GJ, Zhou T, Tian HP, Liu ZL, Xia SS. High expression of ZEB1 correlates with liver metastasis and poor prognosis in colorectal cancer. Oncol Lett. 2013;5:564–8.

    Article  CAS  PubMed  Google Scholar 

  72. Jang MH, Kim HJ, Kim EJ, Chung YR, Park SY. Expression of epithelial-mesenchymal transition-related markers in triple-negative breast cancer: ZEB1 as a potential biomarker for poor clinical outcome. Hum Pathol. 2015;46:1267–74.

    Article  CAS  PubMed  Google Scholar 

  73. Kurahara H, Takao S, Maemura K, Mataki Y, Kuwahata T, Maeda K, Ding Q, Sakoda M, Iino S, Ishigami S, et al. Epithelial-mesenchymal transition and mesenchymal-epithelial transition via regulation of ZEB-1 and ZEB-2 expression in pancreatic cancer. J Surg Oncol. 2012;105:655–61.

    Article  CAS  PubMed  Google Scholar 

  74. Li MZ, Wang JJ, Yang SB, Li WF, Song XM. ZEB2 promotes tumor metastasis and correlates with poor prognosis of human colorectal cancer. Am J Transl Res. 2017;9:2838.

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Caramel J, Papadogeorgakis E, Hill L, Browne GJ, Richard G, Wierinckx A, Saldanha G, Osborne J, Hutchinson P, Tse G, et al. A switch in the expression of embryonic EMT-inducers drives the development of malignant melanoma. Cancer Cell. 2013;24:466–80.

    Article  CAS  PubMed  Google Scholar 

  76. Denecker G, Vandamme N, Akay O, Koludrovic D, Taminau J, Lemeire K, Gheldof A, De Craene B, Van Gele M, Brochez L, et al. Identification of a ZEB2-MITF-ZEB1 transcriptional network that controls melanogenesis and melanoma progression. Cell Death Differ. 2014;21:1250–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Franco HL, Casasnovas J, Rodriguez-Medina JR, Cadilla CL. Redundant or separate entities?–roles of Twist1 and Twist2 as molecular switches during gene transcription. Nucleic Acids Res. 2011;39:1177–86.

    Article  CAS  PubMed  Google Scholar 

  78. Yang MH, Wu MZ, Chiou SH, Chen PM, Chang SY, Liu CJ, Teng SC, Wu KJ. Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat Cell Biol. 2008;10:295–305.

    Article  CAS  PubMed  Google Scholar 

  79. Meng J, Chen S, Han JX, Qian B, Wang XR, Zhong WL, Qin Y, Zhang H, Gao WF, Lei YY, et al. Twist1 regulates Vimentin through Cul2 circular RNA to promote EMT in hepatocellular carcinoma. Cancer Res. 2018;78:4150–62.

    Article  CAS  PubMed  Google Scholar 

  80. 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;31:110–23.

    Article  CAS  PubMed  Google Scholar 

  81. Yang MH, Hsu DS, Wang HW, Wang HJ, Lan HY, Yang WH, Huang CH, Kao SY, Tzeng CH, Tai SK, et al. Bmi1 is essential in Twist1-induced epithelial-mesenchymal transition. Nat Cell Biol. 2010;12:982–92.

    Article  PubMed  CAS  Google Scholar 

  82. Casas E, Kim J, Bendesky A, Ohno-Machado L, Wolfe CJ, Yang J. Snail2 is an essential mediator of Twist1-induced epithelial mesenchymal transition and metastasis. Cancer Res. 2011;71:245–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Beck B, Lapouge G, Rorive S, Drogat B, Desaedelaere K, Delafaille S, Dubois C, Salmon I, Willekens K, Marine JC, Blanpain C. Different levels of Twist1 regulate skin tumor initiation, stemness, and progression. Cell Stem Cell. 2015;16:67–79.

    Article  CAS  PubMed  Google Scholar 

  84. Yang J, Mani SA, Donaher JL, Ramaswamy S, Itzykson RA, Come C, Savagner P, Gitelman I, Richardson A, Weinberg RA. Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell. 2004;117:927–39.

    Article  CAS  PubMed  Google Scholar 

  85. Lee TK, Poon RT, Yuen AP, Ling MT, Kwok WK, Wang XH, Wong YC, Guan XY, Man K, Chau KL, Fan ST. Twist overexpression correlates with hepatocellular carcinoma metastasis through induction of epithelial-mesenchymal transition. Clin Cancer Res. 2006;12:5369–76.

    Article  CAS  PubMed  Google Scholar 

  86. Han B, Qu Y, Jin Y, Yu Y, Deng N, Wawrowsky K, Zhang X, Li N, Bose S, Wang Q, et al. FOXC1 activates smoothened-independent hedgehog signaling in basal-like breast cancer. Cell Rep. 2015;13:1046–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Myatt SS, Lam EW. The emerging roles of forkhead box (Fox) proteins in cancer. Nat Rev Cancer. 2007;7:847–59.

    Article  CAS  PubMed  Google Scholar 

  88. Ou-Yang L, Xiao SJ, Liu P, Yi SJ, Zhang XL, Ou-Yang S, Tan SK, Lei X. Forkhead box C1 induces epithelial-mesenchymal transition and is a potential therapeutic target in nasopharyngeal carcinoma. Mol Med Rep. 2015;12:8003–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Sakai D, Suzuki T, Osumi N, Wakamatsu Y. Cooperative action of Sox9, Snail2 and PKA signaling in early neural crest development. Development. 2006;133:1323–33.

    Article  CAS  PubMed  Google Scholar 

  90. Guo J, Fu Z, Wei J, Lu W, Feng J, Zhang S. PRRX1 promotes epithelial-mesenchymal transition through the Wnt/β-catenin pathway in gastric cancer. Med Oncol. 2015;32:393.

    Article  PubMed  CAS  Google Scholar 

  91. Evdokimova V, Tognon C, Ng T, Ruzanov P, Melnyk N, Fink D, Sorokin A, Ovchinnikov LP, Davicioni E, Triche TJ, Sorensen PH. Translational activation of snail1 and other developmentally regulated transcription factors by YB-1 promotes an epithelial-mesenchymal transition. Cancer Cell. 2009;15:402–15.

    Article  CAS  PubMed  Google Scholar 

  92. Zhou Y, Chang R, Ji W, Wang N, Qi M, Xu Y, Guo J, Zhan L. Loss of scribble promotes snail translation through translocation of HuR and enhances cancer drug resistance. J Biol Chem. 2016;291:291–302.

    Article  CAS  PubMed  Google Scholar 

  93. Gumireddy K, Li A, Yan J, Setoyama T, Johannes GJ, Orom UA, Tchou J, Liu Q, Zhang L, Speicher DW, et al. Identification of a long non-coding RNA-associated RNP complex regulating metastasis at the translational step. Embo j. 2013;32:2672–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Ubersax JA, Ferrell JE Jr. Mechanisms of specificity in protein phosphorylation. Nat Rev Mol Cell Biol. 2007;8:530–41.

    Article  CAS  PubMed  Google Scholar 

  95. Saitoh M. Epithelial-mesenchymal transition is regulated at post-transcriptional levels by transforming growth factor-beta signaling during tumor progression. Cancer Sci. 2015;106:481–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Zhou BP, Deng J, Xia W, Xu J, Li YM, Gunduz M, Hung MC. Dual regulation of Snail by GSK-3beta-mediated phosphorylation in control of epithelial-mesenchymal transition. Nat Cell Biol. 2004;6:931–40.

    Article  CAS  PubMed  Google Scholar 

  97. Yook JI, Li XY, Ota I, Hu C, Kim HS, Kim NH, Cha SY, Ryu JK, Choi YJ, Kim J, et al. A Wnt-Axin2-GSK3beta cascade regulates Snail1 activity in breast cancer cells. Nat Cell Biol. 2006;8:1398–406.

    Article  CAS  PubMed  Google Scholar 

  98. Sahlgren C, Gustafsson MV, Jin S, Poellinger L, Lendahl U. Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc Natl Acad Sci USA. 2008;105:6392–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Wu Y, Deng J, Rychahou PG, Qiu S, Evers BM, Zhou BP. Stabilization of snail by NF-kappaB is required for inflammation-induced cell migration and invasion. Cancer Cell. 2009;15:416–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Du C, Zhang C, Hassan S, Biswas MH, Balaji KC. Protein kinase D1 suppresses epithelial-to-mesenchymal transition through phosphorylation of snail. Cancer Res. 2010;70:7810–9.

    Article  CAS  PubMed  Google Scholar 

  101. Wu Y, Evers BM, Zhou BP. Small C-terminal domain phosphatase enhances snail activity through dephosphorylation. J Biol Chem. 2009;284:640–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Yang Z, Rayala S, Nguyen D, Vadlamudi RK, Chen S, Kumar R. Pak1 phosphorylation of snail, a master regulator of epithelial-to-mesenchyme transition, modulates snail’s subcellular localization and functions. Cancer Res. 2005;65:3179–84.

    Article  CAS  PubMed  Google Scholar 

  103. Zhang K, Rodriguez-Aznar E, Yabuta N, Owen RJ, Mingot JM, Nojima H, Nieto MA, Longmore GD. Lats2 kinase potentiates Snail1 activity by promoting nuclear retention upon phosphorylation. Embo j. 2012;31:29–43.

    Article  PubMed  CAS  Google Scholar 

  104. Hong J, Zhou J, Fu J, He T, Qin J, Wang L, Liao L, Xu J. Phosphorylation of serine 68 of Twist1 by MAPKs stabilizes Twist1 protein and promotes breast cancer cell invasiveness. Cancer Res. 2011;71:3980–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Long J, Zuo D, Park M. Pc2-mediated sumoylation of Smad-interacting protein 1 attenuates transcriptional repression of E-cadherin. J Biol Chem. 2005;280:35477–89.

    Article  CAS  PubMed  Google Scholar 

  106. Xu M, Zhu C, Zhao X, Chen C, Zhang H, Yuan H, Deng R, Dou J, Wang Y, Huang J, et al. Atypical ubiquitin E3 ligase complex Skp1-Pam-Fbxo45 controls the core epithelial-to-mesenchymal transition-inducing transcription factors. Oncotarget. 2015;6:979–94.

    Article  PubMed  Google Scholar 

  107. Chen A, Wong C, Liu M, House CM, Möller A. The ubiquitin ligase Siah is a novel regulator of Zeb1 in breast cancer. Oncotarget. 2015;6:862–73.

    Article  CAS  PubMed  Google Scholar 

  108. Zhou Z, Zhang P, Hu X, Kim J, Yao F, Xiao Z, Zeng L, Chang L, Sun Y, Ma L. USP51 promotes deubiquitination and stabilization of ZEB1. Am J Cancer Res. 2017;7:2020–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Fujita Y, Krause G, Scheffner M, Zechner D, Leddy HE, Behrens J, Sommer T, Birchmeier W. Hakai, a c-Cbl-like protein, ubiquitinates and induces endocytosis of the E-cadherin complex. Nat Cell Biol. 2002;4:222–31.

    Article  CAS  PubMed  Google Scholar 

  110. Hsu DS, Wang HJ, Tai SK, Chou CH, Hsieh CH, Chiu PH, Chen NJ, Yang MH. Acetylation of snail modulates the cytokinome of cancer cells to enhance the recruitment of macrophages. Cancer Cell. 2014;26:534–48.

    Article  CAS  PubMed  Google Scholar 

  111. Chang R, Zhang Y, Zhang P, Zhou Q. Snail acetylation by histone acetyltransferase p300 in lung cancer. Thorac Cancer. 2017;8:131–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Wang CM, Liu R, Wang L, Nascimento L, Brennan VC, Yang WH. SUMOylation of FOXM1B alters its transcriptional activity on regulation of MiR-200 family and JNK1 in MCF7 human breast cancer cells. Int J Mol Sci. 2014;15:10233–51.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Hawkins RD, Hon GC, Lee LK, Ngo Q, Lister R, Pelizzola M, Edsall LE, Kuan S, Luu Y, Klugman S, et al. Distinct epigenomic landscapes of pluripotent and lineage-committed human cells. Cell Stem Cell. 2010;6:479–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Bornman DM, Mathew S, Alsruhe J, Herman JG, Gabrielson E. Methylation of the E-cadherin gene in bladder neoplasia and in normal urothelial epithelium from elderly individuals. Am J Pathol. 2001;159:831–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Graff JR, Herman JG, Lapidus RG, Chopra H, Xu R, Jarrard DF, Isaacs WB, Pitha PM, Davidson NE, Baylin SB. E-cadherin expression is silenced by DNA hypermethylation in human breast and prostate carcinomas. Cancer Res. 1995;55:5195–9.

    CAS  PubMed  Google Scholar 

  116. Tamura G, Yin J, Wang S, Fleisher AS, Zou T, Abraham JM, Kong D, Smolinski KN, Wilson KT, James SP, et al. E-Cadherin gene promoter hypermethylation in primary human gastric carcinomas. J Natl Cancer Inst. 2000;92:569–73.

    Article  CAS  PubMed  Google Scholar 

  117. Puisieux A, Brabletz T, Caramel J. Oncogenic roles of EMT-inducing transcription factors. Nat Cell Biol. 2014;16:488–94.

    Article  CAS  PubMed  Google Scholar 

  118. Bolós V, Peinado H, Pérez-Moreno MA, Fraga MF, Esteller M, Cano A. The transcription factor Slug represses E-cadherin expression and induces epithelial to mesenchymal transitions: a comparison with Snail and E47 repressors. J Cell Sci. 2016;129:1283–1283.

    Article  CAS  Google Scholar 

  119. Canel M, Serrels A, Frame MC, Brunton VG. E-cadherin-integrin crosstalk in cancer invasion and metastasis. J Cell Sci. 2013;126:393–401.

    Article  CAS  PubMed  Google Scholar 

  120. Fukagawa A, Ishii H, Miyazawa K, Saitoh M. deltaEF1 associates with DNMT1 and maintains DNA methylation of the E-cadherin promoter in breast cancer cells. Cancer Med. 2015;4:125–35.

    Article  CAS  PubMed  Google Scholar 

  121. Lin Y, Wu Y, Li J, Dong C, Ye X, Chi YI, Evers BM, Zhou BP. The SNAG domain of Snail1 functions as a molecular hook for recruiting lysine-specific demethylase 1. EMBO J. 2010;29:1803–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Ramadoss S, Chen X, Wang CY. Histone demethylase KDM6B promotes epithelial-mesenchymal transition. J Biol Chem. 2012;287:44508–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Shao P, Liu Q, Kariuki MP, Cui J, Bair TB, Li T, Shaikamjad U, Zhang W, Qi HH. Histone demethylase PHF8 promotes epithelial to mesenchymal transition and breast tumorigenesis. Nuclc Acids Res. 2017;45:1687.

    Article  CAS  Google Scholar 

  124. Hou Z, Peng H, Ayyanathan K, Yan KP, Langer EM, Longmore GD, Rauscher FJ 3rd. The LIM protein AJUBA recruits protein arginine methyltransferase 5 to mediate SNAIL-dependent transcriptional repression. Mol Cell Biol. 2008;28:3198–207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Battistelli C, Cicchini C, Santangelo L, Tramontano A, Grassi L, Gonzalez FJ, de Nonno V, Grassi G, Amicone L, Tripodi M. The Snail repressor recruits EZH2 to specific genomic sites through the enrollment of the lncRNA HOTAIR in epithelial-to-mesenchymal transition. Oncogene. 2017;36:942–55.

    Article  CAS  PubMed  Google Scholar 

  126. Dong C, Wu Y, Yao J, Wang Y, Zhou BP. G9a interacts with Snail and is critical for Snail-mediated E-cadherin repression in human breast cancer. J Clin Investig. 2012;122:1469–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Mobley RJ, Raghu D, Duke LD, Abell-Hart K, Zawistowski JS, Lutz K, Gomez SM, Roy S, Homayouni R, Johnson GL. MAP3K4 controls the chromatin modifier HDAC6 during trophoblast stem cell epithelial-to-mesenchymal transition. Cell Rep. 2017;18:2387–400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Fu J, Qin L, He T, Qin J, Hong J, Wong J, Liao L, Xu J. The TWIST/Mi2/NuRD protein complex and its essential role in cancer metastasis. Cell Res. 2011;21:275–89.

    Article  CAS  PubMed  Google Scholar 

  129. Bracken CP, Gregory PA, Kolesnikoff N, Bert AG, Wang J, Shannon MF, Goodall GJ. A double-negative feedback loop between ZEB1-SIP1 and the microRNA-200 family regulates epithelial-mesenchymal transition. Cancer Res. 2008;68:7846–54.

    Article  CAS  PubMed  Google Scholar 

  130. Burk U, Schubert J, Wellner U, Schmalhofer O, Vincan E, Spaderna S, Brabletz T. A reciprocal repression between ZEB1 and members of the miR-200 family promotes EMT and invasion in cancer cells. EMBO Rep. 2008;9:582–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Gregory PA, Bert AG, Paterson EL, Barry SC, Tsykin A, Farshid G, Vadas MA, Khew-Goodall Y, Goodall GJ. The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nat Cell Biol. 2008;10:593–601.

    Article  CAS  PubMed  Google Scholar 

  132. Kim T, Veronese A, Pichiorri F, Lee TJ, Jeon YJ, Volinia S, Pineau P, Marchio A, Palatini J, Suh SS, et al. p53 regulates epithelial-mesenchymal transition through microRNAs targeting ZEB1 and ZEB2. J Exp Med. 2011;208:875–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Siemens H, Jackstadt R, Hunten S, Kaller M, Menssen A, Gotz U, Hermeking H. miR-34 and SNAIL form a double-negative feedback loop to regulate epithelial-mesenchymal transitions. Cell Cycle. 2011;10:4256–71.

    Article  CAS  PubMed  Google Scholar 

  134. Moes M, Le Bechec A, Crespo I, Laurini C, Halavatyi A, Vetter G, Del Sol A, Friederich E. A novel network integrating a miRNA-203/SNAI1 feedback loop which regulates epithelial to mesenchymal transition. PLoS ONE. 2012;7: e35440.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Ru P, Steele R, Newhall P, Phillips NJ, Toth K, Ray RB. miRNA-29b suppresses prostate cancer metastasis by regulating epithelial-mesenchymal transition signaling. Mol Cancer Ther. 2012;11:1166–73.

    Article  CAS  PubMed  Google Scholar 

  136. Zhang J, Zhang H, Liu J, Tu X, Zang Y, Zhu J, Chen J, Dong L, Zhang J. miR-30 inhibits TGF-β1-induced epithelial-to-mesenchymal transition in hepatocyte by targeting Snail1. Biochem Biophys Res Commun. 2012;417:1100–5.

    Article  CAS  PubMed  Google Scholar 

  137. Liu YN, Yin JJ, Abou-Kheir W, Hynes PG, Casey OM, Fang L, Yi M, Stephens RM, Seng V, Sheppard-Tillman H, et al. MiR-1 and miR-200 inhibit EMT via Slug-dependent and tumorigenesis via Slug-independent mechanisms. Oncogene. 2013;32:296–306.

    Article  CAS  PubMed  Google Scholar 

  138. Diepenbruck M, Tiede S, Saxena M, Ivanek R, Kalathur RKR, Lüönd F, Meyer-Schaller N, Christofori G. miR-1199-5p and Zeb1 function in a double-negative feedback loop potentially coordinating EMT and tumour metastasis. Nat Commun. 2017;8:1168.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Ren D, Wang M, Guo W, Huang S, Wang Z, Zhao X, Du H, Song L, Peng X. Double-negative feedback loop between ZEB2 and miR-145 regulates epithelial-mesenchymal transition and stem cell properties in prostate cancer cells. Cell Tissue Res. 2014;358:763–78.

    Article  CAS  PubMed  Google Scholar 

  140. Dai Y, Wu Z, Lang C, Zhang X, He S, Yang Q, Guo W, Lai Y, Du H, Peng X, Ren D. Copy number gain of ZEB1 mediates a double-negative feedback loop with miR-33a-5p that regulates EMT and bone metastasis of prostate cancer dependent on TGF-β signaling. Theranostics. 2019;9:6063–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Kundu ST, Byers LA, Peng DH, Roybal JD, Diao L, Wang J, Tong P, Creighton CJ, Gibbons DL. The miR-200 family and the miR-183~96~182 cluster target Foxf2 to inhibit invasion and metastasis in lung cancers. Oncogene. 2016;35:173–86.

    Article  CAS  PubMed  Google Scholar 

  142. Liu Z, Mi M, Zheng X, Zhang C, Zhu F, Liu T, Wu G, Zhang L. miR-30a/SOX4 double negative feedback loop is modulated by Disulfiram and regulates EMT and stem cell-like properties in breast cancer. J Cancer. 2021;12:5053–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Shi L, Jackstadt R, Siemens H, Li H, Kirchner T, Hermeking H. p53-induced miR-15a/16-1 and AP4 form a double-negative feedback loop to regulate epithelial-mesenchymal transition and metastasis in colorectal cancer. Cancer Res. 2014;74:532–42.

    Article  CAS  PubMed  Google Scholar 

  144. Lamouille S, Subramanyam D, Blelloch R, Derynck R. Regulation of epithelial-mesenchymal and mesenchymal-epithelial transitions by microRNAs. Curr Opin Cell Biol. 2013;25:200–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Watanabe S, Ueda Y, Akaboshi S, Hino Y, Sekita Y, Nakao M. HMGA2 maintains oncogenic RAS-induced epithelial-mesenchymal transition in human pancreatic cancer cells. Am J Pathol. 2009;174:854–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Qi J, Rice SJ, Salzberg AC, Runkle EA, Liao J, Zander DS, Mu D. MiR-365 regulates lung cancer and developmental gene thyroid transcription factor 1. Cell Cycle. 2012;11:177–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Ma L, Young J, Prabhala H, Pan E, Mestdagh P, Muth D, Teruya-Feldstein J, Reinhardt F, Onder TT, Valastyan S, et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nat Cell Biol. 2010;12:247–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Song Y, Zhao F, Wang Z, Liu Z, Chiang Y, Xu Y, Gao P, Xu H. Inverse association between miR-194 expression and tumor invasion in gastric cancer. Ann Surg Oncol. 2012;19(Suppl 3):S509-517.

    Article  PubMed  Google Scholar 

  149. Meng Z, Fu X, Chen X, Zeng S, Tian Y, Jove R, Xu R, Huang W. miR-194 is a marker of hepatic epithelial cells and suppresses metastasis of liver cancer cells in mice. Hepatology. 2010;52:2148–57.

    Article  CAS  PubMed  Google Scholar 

  150. Zhou Q, Fan J, Ding X, Peng W, Yu X, Chen Y, Nie J. TGF-{beta}-induced MiR-491-5p expression promotes Par-3 degradation in rat proximal tubular epithelial cells. J Biol Chem. 2010;285:40019–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Kong W, Yang H, He L, Zhao JJ, Coppola D, Dalton WS, Cheng JQ. MicroRNA-155 is regulated by the transforming growth factor beta/Smad pathway and contributes to epithelial cell plasticity by targeting RhoA. Mol Cell Biol. 2008;28:6773–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Papadimitriou E, Vasilaki E, Vorvis C, Iliopoulos D, Moustakas A, Kardassis D, Stournaras C. Differential regulation of the two RhoA-specific GEF isoforms Net1/Net1A by TGF-β and miR-24: role in epithelial-to-mesenchymal transition. Oncogene. 2012;31:2862–75.

    Article  CAS  PubMed  Google Scholar 

  153. Zheng F, Liao YJ, Cai MY, Liu YH, Liu TH, Chen SP, Bian XW, Guan XY, Lin MC, Zeng YX, et al. The putative tumour suppressor microRNA-124 modulates hepatocellular carcinoma cell aggressiveness by repressing ROCK2 and EZH2. Gut. 2012;61:278–89.

    Article  CAS  PubMed  Google Scholar 

  154. Yuan JH, Yang F, Wang F, Ma JZ, Guo YJ, Tao QF, Liu F, Pan W, Wang TT, Zhou CC, et al. A long noncoding RNA activated by TGF-beta promotes the invasion-metastasis cascade in hepatocellular carcinoma. Cancer Cell. 2014;25:666–81.

    Article  CAS  PubMed  Google Scholar 

  155. Grelet S, Link LA, Howley B, Obellianne C, Palanisamy V, Gangaraju VK, Diehl JA, Howe PH. A regulated PNUTS mRNA to lncRNA splice switch mediates EMT and tumour progression. Nat Cell Biol. 2017;19:1105–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Dhamija S, Diederichs S. From junk to master regulators of invasion: lncRNA functions in migration, EMT and metastasis. Int J Cancer. 2016;139:269–80.

    Article  CAS  PubMed  Google Scholar 

  157. Kallen AN, Zhou XB, Xu J, Qiao C, Ma J, Yan L, Lu L, Liu C, Yi JS, Zhang H, et al. The imprinted H19 lncRNA antagonizes let-7 microRNAs. Mol Cell. 2013;52:101–12.

    Article  CAS  PubMed  Google Scholar 

  158. Saxena M, Hisano M, Neutzner M, Diepenbruck M, Ivanek R, Sharma K, Kalathur RKR, Burglin TR, Risoli S, Christofori G. The long non-coding RNA ET-20 mediates EMT by impairing desmosomes in breast cancer cells. J Cell Sci. 2021;134: jcs258418.

    Article  CAS  PubMed  Google Scholar 

  159. Cavallaro U, Christofori G. Cell adhesion and signalling by cadherins and Ig-CAMs in cancer. Nat Rev Cancer. 2004;4:118–32.

    Article  CAS  PubMed  Google Scholar 

  160. Bandyopadhyay A, Raghavan S. Defining the role of integrin alphavbeta6 in cancer. Curr Drug Targets. 2009;10:645–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Paoli P, Giannoni E, Chiarugi P. Anoikis molecular pathways and its role in cancer progression. Biochim Biophys Acta. 2013;1833:3481–98.

    Article  CAS  PubMed  Google Scholar 

  162. Rodriguez-Hernandez I, Cantelli G, Bruce F, Sanz-Moreno V. Rho, ROCK and actomyosin contractility in metastasis as drug targets. F1000Res. 2016;5:783.

    Article  CAS  Google Scholar 

  163. Machesky LM. Lamellipodia and filopodia in metastasis and invasion. FEBS Lett. 2008;582:2102–11.

    Article  CAS  PubMed  Google Scholar 

  164. Friedl P, Alexander S. Cancer invasion and the microenvironment: plasticity and reciprocity. Cell. 2011;147:992–1009.

    Article  CAS  PubMed  Google Scholar 

  165. Clark AG, Vignjevic DM. Modes of cancer cell invasion and the role of the microenvironment. Curr Opin Cell Biol. 2015;36:13–22.

    Article  CAS  PubMed  Google Scholar 

  166. Jolly MK, Somarelli JA, Sheth M, Biddle A, Tripathi SC, Armstrong AJ, Hanash SM, Bapat SA, Rangarajan A, Levine H. Hybrid epithelial/mesenchymal phenotypes promote metastasis and therapy resistance across carcinomas. Pharmacol Ther. 2019;194:161–84.

    Article  CAS  PubMed  Google Scholar 

  167. Francart ME, Lambert J, Vanwynsberghe AM, Thompson EW, Bourcy M, Polette M, Gilles C. Epithelial-mesenchymal plasticity and circulating tumor cells: Travel companions to metastases. Dev Dyn. 2018;247:432–50.

    Article  PubMed  Google Scholar 

  168. Zeinali M, Lee M, Nadhan A, Mathur A, Hedman C, Lin E, Harouaka R, Wicha MS, Zhao L, Palanisamy N, et al. High-throughput label-free isolation of heterogeneous circulating tumor cells and CTC clusters from non-small-cell lung cancer patients. Cancers (Basel). 2020;12:127.

    Article  CAS  Google Scholar 

  169. Hou JM, Krebs M, Ward T, Sloane R, Priest L, Hughes A, Clack G, Ranson M, Blackhall F, Dive C. Circulating tumor cells as a window on metastasis biology in lung cancer. Am J Pathol. 2011;178:989–96.

    Article  PubMed  PubMed Central  Google Scholar 

  170. Yu M, Bardia A, Wittner BS, Stott SL, Smas ME, Ting DT, Isakoff SJ, Ciciliano JC, Wells MN, Shah AM, et al. Circulating breast tumor cells exhibit dynamic changes in epithelial and mesenchymal composition. Science. 2013;339:580–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Kalluri R, Weinberg RA. The basics of epithelial-mesenchymal transition. J Clin Investig. 2010;120:1786–1786.

    Article  CAS  PubMed Central  Google Scholar 

  172. Brooks SA, Lomax-Browne HJ, Carter TM, Kinch CE, Hall DM. Molecular interactions in cancer cell metastasis. Acta Histochem. 2010;112:3–25.

    Article  CAS  PubMed  Google Scholar 

  173. Verma RP, Hansch C. Matrix metalloproteinases (MMPs): chemical-biological functions and (Q)SARs. Bioorg Med Chem. 2007;15:2223–68.

    Article  CAS  PubMed  Google Scholar 

  174. Palumbo JS, Talmage KE, Massari JV, La Jeunesse CM, Flick MJ, Kombrinck KW, Hu Z, Barney KA, Degen JL. Tumor cell-associated tissue factor and circulating hemostatic factors cooperate to increase metastatic potential through natural killer cell-dependent and-independent mechanisms. Blood. 2007;110:133–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Labelle M, Begum S, Hynes RO. Direct signaling between platelets and cancer cells induces an epithelial-mesenchymal-like transition and promotes metastasis. Cancer Cell. 2011;20:576–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Hong Y, Fang F, Zhang Q. Circulating tumor cell clusters: What we know and what we expect (Review). Int J Oncol. 2016;49:2206–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Aceto N, Bardia A, Miyamoto DT, Donaldson MC, Wittner BS, Spencer JA, Yu M, Pely A, Engstrom A, Zhu H, et al. Circulating tumor cell clusters are oligoclonal precursors of breast cancer metastasis. Cell. 2014;158:1110–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Singh M, Manoranjan B, Mahendram S, McFarlane N, Venugopal C, Singh SK. Brain metastasis-initiating cells: survival of the fittest. Int J Mol Sci. 2014;15:9117–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Stoletov K, Kato H, Zardouzian E, Kelber J, Yang J, Shattil S, Klemke R. Visualizing extravasation dynamics of metastatic tumor cells. J Cell Sci. 2010;123:2332–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Beerling E, Seinstra D, de Wit E, Kester L, van der Velden D, Maynard C, Schafer R, van Diest P, Voest E, van Oudenaarden A, et al. Plasticity between epithelial and mesenchymal states unlinks EMT from metastasis-enhancing stem cell capacity. Cell Rep. 2016;14:2281–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Jolly MK, Ware KE, Gilja S, Somarelli JA, Levine H. EMT and MET: necessary or permissive for metastasis? Mol Oncol. 2017;11:755–69.

    Article  PubMed  PubMed Central  Google Scholar 

  182. Tran HD, Luitel K, Kim M, Zhang K, Longmore GD, Tran DD. Transient SNAIL1 expression is necessary for metastatic competence in breast cancer. Cancer Res. 2014;74:6330–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Ni T, Li XY, Na L, An T, Weiss SJ. Snail1-dependent p53 repression regulates expansion and activity of tumour-initiating cells in breast cancer. Nat Cell Biol. 2016;18:1221.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Xu Y, Lee DK, Feng Z, Xu Y, Bu W, Li Y, Liao L, Xu J. Breast tumor cell-specific knockout of Twist1 inhibits cancer cell plasticity, dissemination, and lung metastasis in mice. Proc Natl Acad Sci USA. 2017;114:11494–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Fischer KR, Durrans A, Lee S, Sheng J, Li F, Wong ST, Choi H, El Rayes T, Ryu S, Troeger J, et al. Epithelial-to-mesenchymal transition is not required for lung metastasis but contributes to chemoresistance. Nature. 2015;527:472–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Zheng X, Carstens JL, Kim J, Scheible M, Kaye J, Sugimoto H, Wu CC, LeBleu VS, Kalluri R. Epithelial-to-mesenchymal transition is dispensable for metastasis but induces chemoresistance in pancreatic cancer. Nature. 2015;527:525–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Li W, Kang Y. Probing the fifty shades of EMT in metastasis. Trends Cancer. 2016;2:65–7.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Aiello NM, Brabletz T, Kang Y, Nieto MA, Weinberg RA, Stanger BZ. Upholding a role for EMT in pancreatic cancer metastasis. Nature. 2017;547:E7-e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Ye X, Brabletz T, Kang Y, Longmore GD, Nieto MA, Stanger BZ, Yang J, Weinberg RA. Upholding a role for EMT in breast cancer metastasis. Nature. 2017;547:E1-e3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Ye X, Tam WL, Shibue T, Kaygusuz Y, Reinhardt F, Ng Eaton E, Weinberg RA. Distinct EMT programs control normal mammary stem cells and tumour-initiating cells. Nature. 2015;525:256–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Li Y, Lv Z, Zhang S, Wang Z, He L, Tang M, Pu W, Zhao H, Zhang Z, Shi Q, et al. Genetic fate mapping of transient cell Fate reveals N-cadherin activity and function in tumor metastasis. Dev Cell. 2020;54(593–607): e595.

    Google Scholar 

  192. Varga J, Greten FR. Cell plasticity in epithelial homeostasis and tumorigenesis. Nat Cell Biol. 2017;19:1133–41.

    Article  CAS  PubMed  Google Scholar 

  193. Tan TZ, Miow QH, Miki Y, Noda T, Mori S, Huang RY, Thiery JP. Epithelial-mesenchymal transition spectrum quantification and its efficacy in deciphering survival and drug responses of cancer patients. EMBO Mol Med. 2014;6:1279–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. George JT, Jolly MK, Xu S, Somarelli JA, Levine H. Survival outcomes in cancer patients predicted by a partial emt gene expression scoring metric. Cancer Res. 2017;77:6415–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Aiello NM, Maddipati R, Norgard RJ, Balli D, Li J, Yuan S, Yamazoe T, Black T, Sahmoud A, Furth EE, et al. EMT subtype influences epithelial plasticity and mode of cell migration. Dev Cell. 2018;45(681–695): e684.

    Google Scholar 

  196. Luond F, Sugiyama N, Bill R, Bornes L, Hager C, Tang F, Santacroce N, Beisel C, Ivanek R, Burglin T, et al. Distinct contributions of partial and full EMT to breast cancer malignancy. Dev Cell. 2021;56(3203–3221): e3211.

    Google Scholar 

  197. Dong J, Hu Y, Fan X, Wu X, Mao Y, Hu B, Guo H, Wen L, Tang F. Single-cell RNA-seq analysis unveils a prevalent epithelial/mesenchymal hybrid state during mouse organogenesis. Genome Biol. 2018;19:31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  198. Puram SV, Tirosh I, Parikh AS, Patel AP, Yizhak K, Gillespie S, Rodman C, Luo CL, Mroz EA, Emerick KS, et al. Single-cell transcriptomic analysis of primary and metastatic tumor ecosystems in head and neck cancer. Cell. 2017;171(1611–1624): e1624.

    Google Scholar 

  199. Jolly MK, Boareto M, Huang B, Jia D, Lu M, Ben-Jacob E, Onuchic JN, Levine H. Implications of the hybrid epithelial/mesenchymal phenotype in metastasis. Front Oncol. 2015;5:155.

    Article  PubMed  PubMed Central  Google Scholar 

  200. Lu M, Jolly MK, Levine H, Onuchic JN, Ben-Jacob E. MicroRNA-based regulation of epithelial-hybrid-mesenchymal fate determination. Proc Natl Acad Sci USA. 2013;110:18144–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Tian XJ, Zhang H, Xing J. Coupled reversible and irreversible bistable switches underlying TGFβ-induced epithelial to mesenchymal transition. Biophys J. 2013;105:1079–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Zhang J, Tian XJ, Zhang H, Teng Y, Li R, Bai F, Elankumaran S, Xing J. TGF-β-induced epithelial-to-mesenchymal transition proceeds through stepwise activation of multiple feedback loops. Sci Signal. 2014;7:ra91.

    PubMed  Google Scholar 

  203. Jolly MK, Jia D, Boareto M, Mani SA, Pienta KJ, Ben-Jacob E, Levine H. Coupling the modules of EMT and stemness: A tunable “stemness window” model. Oncotarget. 2015;6:25161–74.

    Article  PubMed  PubMed Central  Google Scholar 

  204. Celià-Terrassa T, Bastian C, Liu DD, Ell B, Aiello NM, Wei Y, Zamalloa J, Blanco AM, Hang X, Kunisky D, et al. Hysteresis control of epithelial-mesenchymal transition dynamics conveys a distinct program with enhanced metastatic ability. Nat Commun. 2018;9:5005.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  205. Bakir B, Chiarella AM, Pitarresi JR, Rustgi AK. EMT, MET, plasticity, and tumor metastasis. Trends Cell Biol. 2020;30:764–76.

    Article  PubMed  PubMed Central  Google Scholar 

  206. Loh CY, Chai JY, Tang TF, Wong WF, Sethi G, Shanmugam MK, Chong PP, Looi CY. The E-cadherin and N-cadherin switch in epithelial-to-mesenchymal transition: signaling, therapeutic implications, and challenges. Cells. 2019;8:1118.

    Article  CAS  PubMed Central  Google Scholar 

  207. Keller L, Werner S, Pantel K. Biology and clinical relevance of EpCAM. Cell Stress. 2019;3:165–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Goossens S, Vandamme N, Van Vlierberghe P, Berx G. EMT transcription factors in cancer development re-evaluated: beyond EMT and MET. Biochim Biophys Acta Rev Cancer. 2017;1868:584–91.

    Article  CAS  PubMed  Google Scholar 

  209. Kallergi G, Markomanolaki H, Giannoukaraki V, Papadaki MA, Strati A, Lianidou ES, Georgoulias V, Mavroudis D, Agelaki S. Hypoxia-inducible factor-1alpha and vascular endothelial growth factor expression in circulating tumor cells of breast cancer patients. Breast Cancer Res. 2009;11:R84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  210. Ota I, Li XY, Hu Y, Weiss SJ. Induction of a MT1-MMP and MT2-MMP-dependent basement membrane transmigration program in cancer cells by Snail1. Proc Natl Acad Sci USA. 2009;106:20318–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Dhar M, Lam JN, Walser T, Dubinett SM, Rettig MB, Di Carlo D. Functional profiling of circulating tumor cells with an integrated vortex capture and single-cell protease activity assay. Proc Natl Acad Sci USA. 2018;115:9986–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Lou XL, Sun J, Gong SQ, Yu XF, Gong R, Deng H. Interaction between circulating cancer cells and platelets: clinical implication. Chin J Cancer Res. 2015;27:450–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Cao Z, Livas T, Kyprianou N. Anoikis and EMT: Lethal “Liaisons” during Cancer Progression. Crit Rev Oncog. 2016;21:155–68.

    Article  PubMed  PubMed Central  Google Scholar 

  214. Zhao R, Cai Z, Li S, Cheng Y, Gao H, Liu F, Wu S, Liu S, Dong Y, Zheng L, et al.Expression and clinical relevance of epithelial and mesenchymal markers in circulating tumor cells from colorectal cancer.Oncotarget.2017; 8:9293-9302.

  215. Chebouti I, Kasimir-Bauer S, Buderath P, Wimberger P, Hauch S, Kimmig R, Kuhlmann JD. EMT-like circulating tumor cells in ovarian cancer patients are enriched by platinum-based chemotherapy. Oncotarget. 2017;8:48820–31.

    Article  PubMed  PubMed Central  Google Scholar 

  216. Ning Y, Zhang W, Hanna DL, Yang D, Okazaki S, Berger MD, Miyamoto Y, Suenaga M, Schirripa M, El-Khoueiry A, Lenz HJ. Clinical relevance of EMT and stem-like gene expression in circulating tumor cells of metastatic colorectal cancer patients. Pharmacogenomics J. 2018;18:29–34.

    Article  CAS  PubMed  Google Scholar 

  217. Kallergi G, Agelaki S, Kalykaki A, Stournaras C, Mavroudis D, Georgoulias V. Phosphorylated EGFR and PI3K/Akt signaling kinases are expressed in circulating tumor cells of breast cancer patients. Breast Cancer Res. 2008;10:R80.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  218. Aceto N, Toner M, Maheswaran S, Haber DA. En route to metastasis: circulating tumor cell clusters and epithelial-to-mesenchymal transition. Trends Cancer. 2015;1:44–52.

    Article  PubMed  Google Scholar 

  219. Milsom CC, Yu JL, Mackman N, Micallef J, Anderson GM, Guha A, Rak JW. Tissue factor regulation by epidermal growth factor receptor and epithelial-to-mesenchymal transitions: effect on tumor initiation and angiogenesis. Cancer Res. 2008;68:10068–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Francart ME, Vanwynsberghe AM, Lambert J, Bourcy M, Genna A, Ancel J, Perez-Boza J, Noel A, Birembaut P, Struman I, et al. Vimentin prevents a miR-dependent negative regulation of tissue factor mRNA during epithelial-mesenchymal transitions and facilitates early metastasis. Oncogene. 2020;39:3680–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Mani SA, Guo W, Liao MJ, Eaton EN, Ayyanan A, Zhou AY, Brooks M, Reinhard F, Zhang CC, Shipitsin M, et al. The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell. 2008;133:704–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Chen C, Zhao S, Karnad A, Freeman JW. The biology and role of CD44 in cancer progression: therapeutic implications. J Hematol Oncol. 2018;11:64.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  223. Kang JH, Choi MY, Cui YH, Kaushik N, Uddin N, Yoo KC, Kim MJ, Lee SJ. Regulation of FBXO4-mediated ICAM-1 protein stability in metastatic breast cancer. Oncotarget. 2017;8:83100–13.

    Article  PubMed  PubMed Central  Google Scholar 

  224. Weingarten C, Jenudi Y, Tshuva RY, Moskovich D, Alfandari A, Hercbergs A, Davis PJ, Ellis M, Ashur-Fabian O. The interplay between epithelial-mesenchymal transition (EMT) and the thyroid hormones-alphavbeta3 axis in ovarian cancer. Horm Cancer. 2018;9:22–32.

    Article  CAS  PubMed  Google Scholar 

  225. Wang PC, Weng CC, Hou YS, Jian SF, Fang KT, Hou MF, Cheng KH. Activation of VCAM-1 and its associated molecule CD44 leads to increased malignant potential of breast cancer cells. Int J Mol Sci. 2014;15:3560–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Szczerba BM, Castro-Giner F, Vetter M, Krol I, Gkountela S, Landin J, Scheidmann MC, Donato C, Scherrer R, Singer J, et al. Neutrophils escort circulating tumour cells to enable cell cycle progression. Nature. 2019;566:553–7.

    Article  CAS  PubMed  Google Scholar 

  227. Chockley PJ, Keshamouni VG. Immunological consequences of epithelial-mesenchymal transition in tumor progression. J Immunol. 2016;197:691–8.

    Article  CAS  PubMed  Google Scholar 

  228. Lu Y, Dong B, Xu F, Xu Y, Pan J, Song J, Zhang J, Huang Y, Xue W. CXCL1-LCN2 paracrine axis promotes progression of prostate cancer via the Src activation and epithelial-mesenchymal transition. Cell Commun Signal. 2019;17:118.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  229. Li S, Cong X, Gao H, Lan X, Li Z, Wang W, Song S, Wang Y, Li C, Zhang H, et al. Tumor-associated neutrophils induce EMT by IL-17a to promote migration and invasion in gastric cancer cells. J Exp Clin Cancer Res. 2019;38:6.

    Article  PubMed  PubMed Central  Google Scholar 

  230. Terry S, Savagner P, Ortiz-Cuaran S, Mahjoubi L, Saintigny P, Thiery JP, Chouaib S. New insights into the role of EMT in tumor immune escape. Mol Oncol. 2017;11:824–46.

    Article  PubMed  PubMed Central  Google Scholar 

  231. Romeo E, Caserta CA, Rumio C, Marcucci F. The vicious cross-talk between tumor cells with an EMT phenotype and cells of the immune system. Cells. 2019;8:460.

    Article  CAS  PubMed Central  Google Scholar 

  232. Abalde-Cela S, Piairo P, Dieguez L. The significance of circulating tumour cells in the clinic. Acta Cytol. 2019;63:466–78.

    Article  PubMed  Google Scholar 

  233. Cabel L, Proudhon C, Gortais H, Loirat D, Coussy F, Pierga JY, Bidard FC. Circulating tumor cells: clinical validity and utility. Int J Clin Oncol. 2017;22:421–30.

    Article  PubMed  Google Scholar 

  234. Schochter F, Friedl TWP, deGregorio A, Krause S, Huober J, Rack B, Janni W. Are circulating tumor cells (CTCs) ready for clinical use in breast cancer? An overview of completed and ongoing trials using CTCs for clinical treatment decisions. Cells. 2019;8:1412.

    Article  CAS  PubMed Central  Google Scholar 

  235. Scher HI, Heller G, Molina A, Attard G, Danila DC, Jia X, Peng W, Sandhu SK, Olmos D, Riisnaes R, et al. Circulating tumor cell biomarker panel as an individual-level surrogate for survival in metastatic castration-resistant prostate cancer. J Clin Oncol. 2015;33:1348–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Sastre J, Vidaurreta M, Gomez A, Rivera F, Massuti B, Lopez MR, Abad A, Gallen M, Benavides M, Aranda E, et al. Prognostic value of the combination of circulating tumor cells plus KRAS in patients with metastatic colorectal cancer treated with chemotherapy plus bevacizumab. Clin Colorectal Cancer. 2013;12:280–6.

    Article  CAS  PubMed  Google Scholar 

  237. Mego M, Karaba M, Minarik G, Benca J, Silvia J, Sedlackova T, Manasova D, Kalavska K, Pindak D, Cristofanilli M, et al. Circulating tumor cells with epithelial-to-mesenchymal transition phenotypes associated with inferior outcomes in primary breast cancer. Anticancer Res. 2019;39:1829–37.

    Article  CAS  PubMed  Google Scholar 

  238. Cierna Z, Mego M, Janega P, Karaba M, Minarik G, Benca J, Sedlackova T, Cingelova S, Gronesova P, Manasova D, et al. Matrix metalloproteinase 1 and circulating tumor cells in early breast cancer. BMC Cancer. 2014;14:472.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  239. Manjunath Y, Upparahalli SV, Avella DM, Deroche CB, Kimchi ET, Staveley-O’Carroll KF, Smith CJ, Li G, Kaifi JT. PD-L1 expression with epithelial mesenchymal transition of circulating tumor cells is associated with poor survival in curatively resected non-small cell lung cancer. Cancers (Basel). 2019;11:806.

    Article  CAS  Google Scholar 

  240. Alsuliman A, Colak D, Al-Harazi O, Fitwi H, Tulbah A, Al-Tweigeri T, Al-Alwan M, Ghebeh H. Bidirectional crosstalk between PD-L1 expression and epithelial to mesenchymal transition: significance in claudin-low breast cancer cells. Mol Cancer. 2015;14:149.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  241. Ikeda M, Koh Y, Teraoka S, Sato K, Kanai K, Hayata A, Tokudome N, Akamatsu H, Ozawa Y, Akamatsu K, et al. Detection of AXL expression in circulating tumor cells of lung cancer patients using an automated microcavity array system. Cancer Med. 2020;9:2122–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Papadaki MA, Stoupis G, Theodoropoulos PA, Mavroudis D, Georgoulias V, Agelaki S. Circulating tumor cells with stemness and epithelial-to-mesenchymal transition features are chemoresistant and predictive of poor outcome in metastatic breast cancer. Mol Cancer Ther. 2019;18:437–47.

    Article  CAS  PubMed  Google Scholar 

  243. Guan X, Ma F, Li C, Wu S, Hu S, Huang J, Sun X, Wang J, Luo Y, Cai R, et al. The prognostic and therapeutic implications of circulating tumor cell phenotype detection based on epithelial-mesenchymal transition markers in the first-line chemotherapy of HER2-negative metastatic breast cancer. Cancer Commun (Lond). 2019;39:1.

    Article  Google Scholar 

  244. Brabletz T, Jung A, Spaderna S, Hlubek F, Kirchner T. Opinion: migrating cancer stem cells—An integrated concept of malignant tumour progression. Nat Rev Cancer. 2005;5:744–9.

    Article  CAS  PubMed  Google Scholar 

  245. Blanpain C, Horsley V, Fuchs E. Epithelial stem cells: turning over new leaves. Cell. 2007;128:445–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Ginestier C, Hur MH, Charafe-Jauffret E, Monville F, Dutcher J, Brown M, Jacquemier J, Viens P, Kleer CG, Liu S, et al. ALDH1 is a marker of normal and malignant human mammary stem cells and a predictor of poor clinical outcome. Cell Stem Cell. 2007;1:555–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. Gupta PB, Chaffer CL, Weinberg RA. Cancer stem cells: mirage or reality? Nat Med. 2009;15:1010–2.

    Article  CAS  PubMed  Google Scholar 

  248. Liu S, Cong Y, Wang D, Sun Y, Deng L, Liu Y, Martin-Trevino R, Shang L, McDermott SP, Landis MD, et al. Breast cancer stem cells transition between epithelial and mesenchymal states reflective of their normal counterparts. Stem Cell Rep. 2014;2:78–91.

    Article  CAS  Google Scholar 

  249. Tellez CS, Juri DE, Do K, Bernauer AM, Thomas CL, Damiani LA, Tessema M, Leng S, Belinsky SA. EMT and stem cell-like properties associated with miR-205 and miR-200 epigenetic silencing are early manifestations during carcinogen-induced transformation of human lung epithelial cells. Cancer Res. 2011;71:3087–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Li W, Ma H, Zhang J, Zhu L, Wang C, Yang Y. Unraveling the roles of CD44/CD24 and ALDH1 as cancer stem cell markers in tumorigenesis and metastasis. Sci Rep. 2017;7:13856.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  251. Najafi M, Mortezaee K, Ahadi R. Cancer stem cell (a)symmetry & plasticity: tumorigenesis and therapy relevance. Life Sci. 2019;231: 116520.

    Article  CAS  PubMed  Google Scholar 

  252. Plaks V, Kong N, Werb Z. The cancer stem cell niche: how essential is the niche in regulating stemness of tumor cells? Cell Stem Cell. 2015;16:225–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Battula VL, Evans KW, Hollier BG, Shi Y, Marini FC, Ayyanan A, Wang RY, Brisken C, Guerra R, Andreeff M, Mani SA. Epithelial-mesenchymal transition-derived cells exhibit multilineage differentiation potential similar to mesenchymal stem cells. Stem Cells. 2010;28:1435–45.

    Article  CAS  PubMed  Google Scholar 

  254. Wellner U, Schubert J, Burk UC, Schmalhofer O, Zhu F, Sonntag A, Waldvogel B, Vannier C, Darling D, zur Hausen A, et al. The EMT-activator ZEB1 promotes tumorigenicity by repressing stemness-inhibiting microRNAs. Nat Cell Biol. 2009;11:1487–95.

    Article  CAS  PubMed  Google Scholar 

  255. Siddique HR, Saleem M. Role of BMI1, a stem cell factor, in cancer recurrence and chemoresistance: preclinical and clinical evidences. Stem Cells. 2012;30:372–8.

    Article  CAS  PubMed  Google Scholar 

  256. Song LB, Li J, Liao WT, Feng Y, Yu CP, Hu LJ, Kong QL, Xu LH, Zhang X, Liu WL, et al. The polycomb group protein Bmi-1 represses the tumor suppressor PTEN and induces epithelial-mesenchymal transition in human nasopharyngeal epithelial cells. J Clin Invest. 2009;119:3626–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Xu LF, Zhang J, Tian ZQ. Wu YZ [Epigenetics in cancer stem cells]. Yi Chuan. 2013;35:1049–57.

    Article  CAS  PubMed  Google Scholar 

  258. Valk-Lingbeek ME, Bruggeman SW, van Lohuizen M. Stem cells and cancer; the polycomb connection. Cell. 2004;118:409–18.

    Article  CAS  PubMed  Google Scholar 

  259. Shimono Y, Zabala M, Cho RW, Lobo N, Dalerba P, Qian D, Diehn M, Liu H, Panula SP, Chiao E, et al. Downregulation of miRNA-200c links breast cancer stem cells with normal stem cells. Cell. 2009;138:592–603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Brabletz S, Bajdak K, Meidhof S, Burk U, Niedermann G, Firat E, Wellner U, Dimmler A, Faller G, Schubert J, Brabletz T. The ZEB1/miR-200 feedback loop controls Notch signalling in cancer cells. EMBO J. 2011;30:770–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Meurette O, Mehlen P. Notch signaling in the tumor microenvironment. Cancer Cell. 2018;34:536–48.

    Article  CAS  PubMed  Google Scholar 

  262. Hill L, Browne G, Tulchinsky E. ZEB/miR-200 feedback loop: at the crossroads of signal transduction in cancer. Int J Cancer. 2013;132:745–54.

    Article  CAS  PubMed  Google Scholar 

  263. Nuti SV, Mor G, Li P, Yin G. TWIST and ovarian cancer stem cells: implications for chemoresistance and metastasis. Oncotarget. 2014;5:7260–71.

    Article  PubMed  PubMed Central  Google Scholar 

  264. Lambert AW, Weinberg RA. Linking EMT programmes to normal and neoplastic epithelial stem cells. Nat Rev Cancer. 2021;21:325–38.

    Article  CAS  PubMed  Google Scholar 

  265. Hwang WL, Jiang JK, Yang SH, Huang TS, Lan HY, Teng HW, Yang CY, Tsai YP, Lin CH, Wang HW, Yang MH. MicroRNA-146a directs the symmetric division of Snail-dominant colorectal cancer stem cells. Nat Cell Biol. 2014;16:268–80.

    Article  CAS  PubMed  Google Scholar 

  266. Guo W, Keckesova Z, Donaher J, Shibue T, Tischler V, Reinhardt F, Itzkovitz S, Noske A, Zürrer-Härdi U, Bell G. Slug and Sox9 cooperatively determine the mammary stem cell state. Cell. 2012;148:1015–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  267. Leng Z, Li Y, Zhou G, Lv X, Ai W, Li J, Hou L. Kruppel-like factor 4 regulates stemness and mesenchymal properties of colorectal cancer stem cells through the TGF-beta1/Smad/snail pathway. J Cell Mol Med. 2020;24:1866–77.

    Article  CAS  PubMed  Google Scholar 

  268. Ocana OH, Corcoles R, Fabra A, Moreno-Bueno G, Acloque H, Vega S, Barrallo-Gimeno A, Cano A, Nieto MA. Metastatic colonization requires the repression of the epithelial-mesenchymal transition inducer Prrx1. Cancer Cell. 2012;22:709–24.

    Article  CAS  PubMed  Google Scholar 

  269. Schmidt JM, Panzilius E, Bartsch HS, Irmler M, Beckers J, Kari V, Linnemann JR, Dragoi D, Hirschi B, Kloos UJ, et al. Stem-cell-like properties and epithelial plasticity arise as stable traits after transient Twist1 activation. Cell Rep. 2015;10:131–9.

    Article  CAS  PubMed  Google Scholar 

  270. Celia-Terrassa T, Meca-Cortes O, Mateo F, Martinez de Paz A, Rubio N, Arnal-Estape A, Ell BJ, Bermudo R, Diaz A, Guerra-Rebollo M, et al. Epithelial-mesenchymal transition can suppress major attributes of human epithelial tumor-initiating cells. J Clin Invest. 2012;122:1849–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  271. Tsai JH, Donaher JL, Murphy DA, Chau S, Yang J. Spatiotemporal regulation of epithelial-mesenchymal transition is essential for squamous cell carcinoma metastasis. Cancer Cell. 2012;22:725–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  272. Tsuji T, Ibaragi S, Shima K, Hu MG, Katsurano M, Sasaki A, Hu GF. Epithelial-mesenchymal transition induced by growth suppressor p12CDK2-AP1 promotes tumor cell local invasion but suppresses distant colony growth. Cancer Res. 2008;68:10377–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Grosse-Wilde A, Fouquier d’Herouel A, McIntosh E, Ertaylan G, Skupin A, Kuestner RE, del Sol A, Walters KA, Huang S. Stemness of the hybrid Epithelial/Mesenchymal State in Breast Cancer and Its Association with Poor Survival. PLoS ONE. 2015;10: e0126522.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  274. Ruscetti M, Quach B, Dadashian EL, Mulholland DJ, Wu H. Tracking and functional characterization of epithelial-mesenchymal transition and mesenchymal tumor cells during prostate cancer metastasis. Cancer Res. 2015;75:2749–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Strauss R, Li ZY, Liu Y, Beyer I, Persson J, Sova P, Moller T, Pesonen S, Hemminki A, Hamerlik P, et al. Analysis of epithelial and mesenchymal markers in ovarian cancer reveals phenotypic heterogeneity and plasticity. PLoS ONE. 2011;6: e16186.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Jolly MK, Huang B, Lu M, Mani SA, Levine H, Ben-Jacob E. Towards elucidating the connection between epithelial-mesenchymal transitions and stemness. J R Soc Interface. 2014;11:20140962.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  277. Ombrato L, Malanchi I. The EMT universe: space between cancer cell dissemination and metastasis initiation. Crit Rev Oncog. 2014;19:349–61.

    Article  PubMed  Google Scholar 

  278. Singh A, Settleman J. EMT, cancer stem cells and drug resistance: an emerging axis of evil in the war on cancer. Oncogene. 2010;29:4741–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  279. Dongre A, Rashidian M, Reinhardt F, Bagnato A, Keckesova Z, Ploegh HL, Weinberg RA. Epithelial-to-mesenchymal transition contributes to immunosuppression in breast carcinomas. Cancer Res. 2017;77:3982–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  280. Kudo-Saito C, Shirako H, Takeuchi T, Kawakami Y. Cancer metastasis is accelerated through immunosuppression during Snail-induced EMT of cancer cells. Cancer Cell. 2009;15:195–206.

    Article  CAS  PubMed  Google Scholar 

  281. Akalay I, Janji B, Hasmim M, Noman MZ, Andre F, De Cremoux P, Bertheau P, Badoual C, Vielh P, Larsen AK, et al. Epithelial-to-mesenchymal transition and autophagy induction in breast carcinoma promote escape from T-cell-mediated lysis. Cancer Res. 2013;73:2418–27.

    Article  CAS  PubMed  Google Scholar 

  282. Sanjabi S, Oh SA, Li MO. Regulation of the Immune Response by TGF-beta: from conception to autoimmunity and infection. Cold Spring Harb Perspect Biol. 2017;9: a022236.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  283. Chen L, Gibbons DL, Goswami S, Cortez MA, Ahn YH, Byers LA, Zhang X, Yi X, Dwyer D, Lin W, et al. Metastasis is regulated via microRNA-200/ZEB1 axis control of tumour cell PD-L1 expression and intratumoral immunosuppression. Nat Commun. 2014;5:5241.

    Article  CAS  PubMed  Google Scholar 

  284. Noman MZ, Janji B, Abdou A, Hasmim M, Terry S, Tan TZ, Mami-Chouaib F, Thiery JP, Chouaib S. The immune checkpoint ligand PD-L1 is upregulated in EMT-activated human breast cancer cells by a mechanism involving ZEB-1 and miR-200. Oncoimmunology. 2017;6: e1263412.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  285. Lou Y, Diao L, Cuentas ER, Denning WL, Chen L, Fan YH, Byers LA, Wang J, Papadimitrakopoulou VA, Behrens C, et al. Epithelial-mesenchymal transition is associated with a distinct tumor microenvironment including elevation of inflammatory signals and multiple immune checkpoints in lung adenocarcinoma. Clin Cancer Res. 2016;22:3630–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  286. Redfern AD, Spalding LJ, Thompson EW. The Kraken Wakes: induced EMT as a driver of tumour aggression and poor outcome. Clin Exp Metastasis. 2018;35:285–308.

    Article  PubMed  Google Scholar 

  287. Dudas J, Ladanyi A, Ingruber J, Steinbichler TB, Riechelmann H. Epithelial to mesenchymal transition: a mechanism that fuels cancer radio/chemoresistance. Cells. 2020;9:428.

    Article  CAS  PubMed Central  Google Scholar 

  288. Jonckheere S, Adams J, De Groote D, Campbell K, Berx G, Goossens S. Epithelial-mesenchymal transition (EMT) as a therapeutic target. Cells Tissues Organs. 2022;211:157–82.

    Article  CAS  PubMed  Google Scholar 

  289. Xia L, Tan S, Zhou Y, Lin J, Wang H, Oyang L, Tian Y, Liu L, Su M, Wang H, et al. Role of the NFkappaB-signaling pathway in cancer. Onco Targets Ther. 2018;11:2063–73.

    Article  PubMed  PubMed Central  Google Scholar 

  290. Skrypek N, Goossens S, De Smedt E, Vandamme N, Berx G. Epithelial-to-mesenchymal transition: epigenetic reprogramming driving cellular plasticity. Trends Genet. 2017;33:943–59.

    Article  CAS  PubMed  Google Scholar 

  291. Zhang L, Liao Y, Tang L. MicroRNA-34 family: a potential tumor suppressor and therapeutic candidate in cancer. J Exp Clin Cancer Res. 2019;38:53.

    Article  PubMed  PubMed Central  Google Scholar 

  292. Pattabiraman DR, Bierie B, Kober KI, Thiru P, Krall JA, Zill C, Reinhardt F, Tam WL, Weinberg RA. Activation of PKA leads to mesenchymal-to-epithelial transition and loss of tumor-initiating ability. Science. 2016;351:aad3680.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  293. Dongre A, Rashidian M, Eaton EN, Reinhardt F, Thiru P, Zagorulya M, Nepal S, Banaz T, Martner A, Spranger S, Weinberg RA. Direct and indirect regulators of epithelial-mesenchymal transition-mediated immunosuppression in breast carcinomas. Cancer Discov. 2021;11:1286–305.

    Article  CAS  PubMed  Google Scholar 

  294. Li M, Wang Y, Li M, Wu X, Setrerrahmane S, Xu H. Integrins as attractive targets for cancer therapeutics. Acta Pharm Sin B. 2021;11:2726–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  295. Almokadem S, Belani CP. Volociximab in cancer. Expert Opin Biol Ther. 2012;12:251–7.

    Article  CAS  PubMed  Google Scholar 

  296. Stoeltzing O, Liu W, Reinmuth N, Fan F, Parry GC, Parikh AA, McCarty MF, Bucana CD, Mazar AP, Ellis LM. Inhibition of integrin alpha5beta1 function with a small peptide (ATN-161) plus continuous 5-FU infusion reduces colorectal liver metastases and improves survival in mice. Int J Cancer. 2003;104:496–503.

    Article  CAS  PubMed  Google Scholar 

  297. Mullamitha SA, Ton NC, Parker GJ, Jackson A, Julyan PJ, Roberts C, Buonaccorsi GA, Watson Y, Davies K, Cheung S, et al. Phase I evaluation of a fully human anti-alphav integrin monoclonal antibody (CNTO 95) in patients with advanced solid tumors. Clin Cancer Res. 2007;13:2128–35.

    Article  CAS  PubMed  Google Scholar 

  298. Smith JW. Cilengitide Merck. Curr Opin Investig Drugs. 2003;4:741–5.

    CAS  PubMed  Google Scholar 

  299. Alba J, Barcia R, Gutiérrez-Berzal J, Ramos-Martínez JI. Could inhibition of metalloproteinases be used to block the process of metastasis? Cell Biochem Funct. 2022;40:600–7.

    Article  CAS  PubMed  Google Scholar 

  300. Miyoshi A, Kitajima Y, Sumi K, Sato K, Hagiwara A, Koga Y, Miyazaki K. Snail and SIP1 increase cancer invasion by upregulating MMP family in hepatocellular carcinoma cells. Br J Cancer. 2004;90:1265–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  301. Sun L, Diamond ME, Ottaviano AJ, Joseph MJ, Ananthanarayan V, Munshi HG. Transforming growth factor-beta 1 promotes matrix metalloproteinase-9-mediated oral cancer invasion through snail expression. Mol Cancer Res. 2008;6:10–20.

    Article  CAS  PubMed  Google Scholar 

  302. Shen M, Xu Z, Xu W, Jiang K, Zhang F, Ding Q, Xu Z, Chen Y. Inhibition of ATM reverses EMT and decreases metastatic potential of cisplatin-resistant lung cancer cells through JAK/STAT3/PD-L1 pathway. J Exp Clin Cancer Res. 2019;38:149.

    Article  PubMed  PubMed Central  Google Scholar 

  303. Lin G, Yu B, Liang Z, Li L, Qu S, Chen K, Zhou L, Lu Q, Sun Y, Zhu X. Silencing of c-jun decreases cell migration, invasion, and EMT in radioresistant human nasopharyngeal carcinoma cell line CNE-2R. Onco Targets Ther. 2018;11:3805–15.

    Article  PubMed  PubMed Central  Google Scholar 

  304. Kim MY, Oskarsson T, Acharyya S, Nguyen DX, Zhang XH, Norton L, Massague J. Tumor self-seeding by circulating cancer cells. Cell. 2009;139:1315–26.

    Article  PubMed  PubMed Central  Google Scholar 

  305. Vijay GV, Zhao N, Den Hollander P, Toneff MJ, Joseph R, Pietila M, Taube JH, Sarkar TR, Ramirez-Pena E, Werden SJ, et al. GSK3beta regulates epithelial-mesenchymal transition and cancer stem cell properties in triple-negative breast cancer. Breast Cancer Res. 2019;21:37.

    Article  PubMed  PubMed Central  Google Scholar 

  306. Williams ED, Gao D, Redfern A, Thompson EW. Controversies around epithelial-mesenchymal plasticity in cancer metastasis. Nat Rev Cancer. 2019;19:716–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  307. De Las RJ, Brozovic A, Izraely S, Casas-Pais A, Witz IP, Figueroa A. Cancer drug resistance induced by EMT: novel therapeutic strategies. Arch Toxicol. 2021;95:2279–97.

    Article  CAS  Google Scholar 

  308. Taki M, Abiko K, Baba T, Hamanishi J, Yamaguchi K, Murakami R, Yamanoi K, Horikawa N, Hosoe Y, Nakamura E, et al. Snail promotes ovarian cancer progression by recruiting myeloid-derived suppressor cells via CXCR2 ligand upregulation. Nat Commun. 2018;9:1685.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  309. Katsura A, Tamura Y, Hokari S, Harada M, Morikawa M, Sakurai T, Takahashi K, Mizutani A, Nishida J, Yokoyama Y, et al. ZEB1-regulated inflammatory phenotype in breast cancer cells. Mol Oncol. 2017;11:1241–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  310. Taki M, Abiko K, Ukita M, Murakami R, Yamanoi K, Yamaguchi K, Hamanishi J, Baba T, Matsumura N, Mandai M. Tumor Immune Microenvironment during Epithelial-Mesenchymal Transition. Clin Cancer Res. 2021;27:4669–79.

    Article  CAS  PubMed  Google Scholar 

  311. Rhim AD, Mirek ET, Aiello NM, Maitra A, Bailey JM, McAllister F, Reichert M, Beatty GL, Rustgi AK, Vonderheide RH, et al. EMT and dissemination precede pancreatic tumor formation. Cell. 2012;148:349–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  312. Ishay-Ronen D, Diepenbruck M, Kalathur RKR, Sugiyama N, Tiede S, Ivanek R, Bantug G, Morini MF, Wang J, Hess C, Christofori G. Gain fat-lose metastasis: converting invasive breast cancer cells into adipocytes inhibits cancer metastasis. Cancer Cell. 2019;35(17–32): e16.

    Google Scholar 

  313. Jia D, George JT, Tripathi SC, Kundnani DL, Lu M, Hanash SM, Onuchic JN, Jolly MK, Levine H. Testing the gene expression classification of the EMT spectrum. Phys Biol. 2019;16: 025002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  314. Shen M, Jiang YZ, Wei Y, Ell B, Sheng X, Esposito M, Kang J, Hang X, Zheng H, Rowicki M, et al. Tinagl1 suppresses triple-negative breast cancer progression and metastasis by simultaneously inhibiting integrin/FAK and EGFR signaling. Cancer Cell. 2019;35(64–80): e67.

    Google Scholar 

Download references

Acknowledgements

The figures in this review were created by BioRender.

Funding

This work was supported by the National Science Foundation for Excellent Young Scholars (32122052) and National Natural Science Foundation Regional Innovation and Development (No. U19A2003).

Author information

Authors and Affiliations

Authors

Contributions

X.W. conceived and supervised the manuscript. Y. H. and W. H. wrote and revised the manuscript, and Y.H. completed all figures and table. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Xiawei Wei.

Ethics declarations

Ethics approval and consent to participate

No ethics approval was required for this review that did not involve patients or patient data.

Consent for publication

All authors consent to publication.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, Y., Hong, W. & Wei, X. The molecular mechanisms and therapeutic strategies of EMT in tumor progression and metastasis. J Hematol Oncol 15, 129 (2022). https://doi.org/10.1186/s13045-022-01347-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13045-022-01347-8

Keywords