Next Article in Journal
Structure–Activity Relationships and Antiplasmodial Potencies of Novel 3,4-Disubstituted 1,2,5-Oxadiazoles
Previous Article in Journal
MiR-148a-3p within HucMSC-Derived Extracellular Vesicles Suppresses Hsp90b1 to Prevent Fibroblast Collagen Synthesis and Secretion in Silica-Induced Pulmonary Fibrosis
Previous Article in Special Issue
Duality of Valproic Acid Effects on Inflammation, Oxidative Stress and Autophagy in Human Eosinophilic Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Involvement of Epithelial-Mesenchymal Transition (EMT) in Autoimmune Diseases

Department of Rheumatology, Hôpital Erasme, Université Libre de Bruxelles, 1070 Brussels, Belgium
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(19), 14481; https://doi.org/10.3390/ijms241914481
Submission received: 1 August 2023 / Revised: 16 September 2023 / Accepted: 18 September 2023 / Published: 23 September 2023

Abstract

:
Epithelial-mesenchymal transition (EMT) is a complex reversible biological process characterized by the loss of epithelial features and the acquisition of mesenchymal features. EMT was initially described in developmental processes and was further associated with pathological conditions including metastatic cascade arising in neoplastic progression and organ fibrosis. Fibrosis is delineated by an excessive number of myofibroblasts, resulting in exuberant production of extracellular matrix (ECM) proteins, thereby compromising organ function and ultimately leading to its failure. It is now well acknowledged that a significant number of myofibroblasts result from the conversion of epithelial cells via EMT. Over the past two decades, evidence has accrued linking fibrosis to many chronic autoimmune and inflammatory diseases, including systemic sclerosis (SSc), rheumatoid arthritis (RA), systemic lupus erythematosus (SLE), Sjögren’s syndrome (SS), and inflammatory bowel diseases (IBD). In addition, chronic inflammatory states observed in most autoimmune and inflammatory diseases can act as a potent trigger of EMT, leading to the development of a pathological fibrotic state. In the present review, we aim to describe the current state of knowledge regarding the contribution of EMT to the pathophysiological processes of various rheumatic conditions.

1. Introduction

Epithelial-mesenchymal transition (EMT) is a complex reversible biological process that takes place in epithelial cells and is characterized by the loss of epithelial and the acquisition of mesenchymal features [1,2]. Even though it was initially considered a binary process, a growing body of evidence indicates that EMT occurs in a stepwise manner with distinct states expressing different proportions of epithelial and mesenchymal markers. These intermediate states are defined as hybrid or partial EMT as they display attributes of both epithelial and mesenchymal cells [3,4,5]. In addition, the transitioning cells along the EMT spectrum keep the ability to revert at any moment to a more epithelial state by undergoing the reverse process known as mesenchymal–epithelial transition (MET) [6,7]. A broad variety of signals from the extracellular environment are able to trigger EMT in epithelial cells, causing dramatic alterations in their transcriptional and epigenetic landscapes, further eliciting the repression of epithelial junctions and the reorganization of the cytoskeleton [1,8,9]. As a result, cells lose their polarity, detach themselves from the basal membrane and adopt a spindle-shaped morphology similar to fibroblasts, thus increasing their motility and enabling the apparition of invasive properties [9,10,11].
EMT prevails in both physiological and pathological settings. It was initially described in the early 1980s by embryologists as a fine-tuned process that governs embryonic development [12]. It is now well established that EMT plays a pivotal role in several developmental processes including gastrulation, neural crest development, somite dissociation, and palate and lip fusion [13,14,15,16]. In adults, EMT can also be transiently activated as a physiological response to injury. It has been demonstrated that epithelial cells adjacent to the wound undertake EMT, allowing them to move while maintaining loose contact. Once the cells have arrived at the wounded site, epithelial barrier integrity is restored as the cells revert back to an epithelial phenotype via MET [17]. Unfortunately, EMT can also be erroneously triggered by more dramatic pathological conditions. Both organ fibrosis and neoplastic progression were found to be driven by EMT [1,10,18]. Over the past two decades, evidence has piled up to link EMT to various tumor functions, including tumor stemness, tumor cell migration, intravasation to the blood and lymphatic vessels, metastasis, and resistance to therapy [19,20,21,22]. Moreover, a great deal of attention has been paid to the contribution of EMT to chronic fibrotic diseases such as end-stage liver disease, chronic kidney disease, idiopathic pulmonary fibrosis (IPF), or heart failure [23,24]. Fibrosis is delineated by an excessive number of myofibroblasts, resulting in exuberant production of extracellular matrix (ECM) proteins (i.e., collagen and fibronectin), thereby compromising organ function and ultimately leading to organ failure [25,26,27]. It is now well acknowledged that a significant percentage of myofibroblasts result from the conversion of epithelial cells via EMT [28,29]. As fibrosis has been linked to many rheumatic diseases, including systemic sclerosis (SSc) [30], rheumatoid arthritis (RA) [31], systemic lupus erythematosus (SLE) [32], and Sjögren’s syndrome (SS) [33], it is therefore suspected that chronic inflammatory states could act as a potent trigger of EMT, leading to the development of a pathological fibrotic state in these rheumatic diseases [1,4,10,34,35].
Even though several lines of evidence seems to confirm the occurrence of EMT in rheumatic diseases [36,37], its precise contribution to their pathophysiology remains unclear. In the present review, we therefore aim to describe the current state of knowledge regarding the contribution of EMT to the pathophysiological processes of various rheumatic autoimmune diseases.

2. Biological Mechanisms Underlying EMT

EMT is a tremendously complex biological process. Despite the quantum leaps made to unearth the different regulatory mechanisms underlying EMT, the exact regulation of this process is not yet fully understood. A broad range of signals stemming from the stromal microenvironment bind to their cognate receptors on epithelial cells and are able to trigger several distinct intracellular signaling pathways which can display cross-talk to some extent. These intracellular signaling events ultimately converge on activating EMT via the expression of core EMT-transcription factors (TFs) in various combination that lead to the repression of epithelial genes and the activation of mesenchymal ones, accounting for the loss of epithelial and the gain of mesenchymal traits. In addition, a process similar to EMT has been observed and described in cell types other than epithelial cells, suggesting that the frontiers of EMT might be looser than previously thought and that EMT might not be restricted only to epithelia cells [38]. This EMT-like process has been widely described in non-epithelial cancers such as melanoma, sarcoma, and leukemia [39,40,41] and in non-cancerous conditions where it aroused in non-epithelial cells such as endothelium [42], mesothelium, [43] or fibroblast-like synoviocytes (FLSs) [44]. Several authors have proposed other terminologies to avoid confusion with EMT arising in epithelial cells and called these processes either endothelial–mesenchymal transition (endoMT) [42] for endothelial cells, mesothelial–mesenchymal transition (MMT) [43] for mesothelial cells, or EMT-like for other cell types than epithelial cells [45].

2.1. Morphological Modification during EMT

Epithelial cells are closely held together by different types of lateral cell-cell junctions (i.e., adherens junctions, desmosomes, gap junctions and tight junctions). Epithelial cells form polarized sheets and harbor an apico-basal polarity by anchoring to the basement membrane via hemidesmosomes, which contain α6β4 integrin, and are linked to cytokeratin intermediate filaments, a compound of the cytoskeleton enabling epithelial cells to resist numerous physical stresses, inside the cell [11,19]. The loss of E-cadherin (a key component of adherens junctions) is often considered a prototypical event of EMT and is rapidly followed by the repression of other cell junction proteins such as zona occludens 1 (ZO-1), occludin, and claudin [46,47,48]. Consequently, the dissolution of epithelial cell junctions elicits a loss of apico-basal polarity and triggers cytoskeletal modifications with a switch from cytokeratin-rich to vimentin-rich networks, leading to the adoption of a spindle-shaped morphology [48]. These events permit cells to gain mobility and to migrate into surrounding tissues along a secreted matrix of fibronectin [10]. As mentioned before, the cells are pushed towards the EMT spectrum in a progressive and reversible manner, from a complete epithelial to complete mesenchymal state ultimately characterized by mesenchymal markers, such as vimentin, fibronectin, α-smooth muscle actin (α-SMA), N-cadherin, fibroblast specific protein 1 (FSP-1), and various matrix metalloproteases (MMPs) and the loss of epithelial markers such as E-cadherin, ZO-1, occludin, and several cytokeratins (Figure 1) [10,11,19].

2.2. Transcriptional Regulation of EMT

EMT is an elaborate process that is tightly regulated at the transcriptional level. Over the years, serious efforts have been made to disentangle this complex system. It has become clear that a great deal of transcription factors (TFs) are involved in the gene regulatory networks controlling EMT. Some of these have identified as master regulators or core EMT-TFs, as they are conserved across various biological conditions and across species, whereas other TFs may have more restricted functions depending on the tissue or the biological background [49,50], suggesting that the transcriptional regulation of EMT involves multiple layers of control [20,50]. Consequently, EMT core-TFs tend to control one another’s expression along with additional TFs that further define and drive EMT progression. Altogether, EMT-TFs act synergistically with one another, using some common pathways to coordinate the repression of epithelial and activation of mesenchymal phenotypes [51]. Core EMT-TFs include the zinc-finger E-box-binding homeobox factors zinc finger E-box binding homeobox 1 (ZEB1), ZEB2, snail family transcriptional repressor 1 SNAI1 (also known as SNAIL), SNAI2 (also known as SLUG), and the basic helix–loop–helix factors (bHLH) twist family BHLH transcription factor 1 (TWIST1) and TWIST2 [52,53].

2.2.1. SNAI Family

SNAI1 and SNAI2 belong to the SNAIL family of zinc-finger TFs. SNAI1 and SNAI2 are potent repressors of E-cadherin (CDH1) gene transcription as they bind to its promoter, where they recruit several corepressor complexes [54,55,56]. SNAI1 is also involved in the repression of multiple epithelial genes in charge of the maintenance of epithelial phenotype, including genes encoding claudin and occludin, major components of tight junctions [57]. In addition, SNAI1 cooperates with TWIST1 in the induction of ZEB1 expression [58] and with ETS1, itself ultimately activated by mitogen-activated protein kinase (MAPK) signaling, to induce MMP expression, a large family of endopeptidase involved in ECM remodeling [59]. SNAI1 can also interact with the small mothers against the decapentaplegic homolog (SMAD)3–SMAD4 complex, further entailing the transformation of growth factor (TGF)-β-mediated repression of E-cadherin and occludin [60].

2.2.2. ZEB Family

ZEB1 and ZEB2 are the two members of the ZEB family of zinc finger TFs. Similar to SNAI1, ZEB1 binds to CDH1 promoter and represses transcription via the recruitment of corepressor complexes [61,62]. ZEB1 mediate the transcriptional activation of the genes encoding vimentin and N-cadherin via the recruitment of coactivator complexes to their respective promoters [49,63]. Additionally, ZEB2 together with SNAI1 increase the expression of MMPs [64,65].

2.2.3. BHLH Family

TWIST1 and TWIST2 both belong to the bHLH class of transcriptional regulators. As with SNAI1 or ZEB1, TWIST1 suppresses the expression of the epithelial gene CDH1 and activates the expression of N-cadherin, a mesenchymal gene [66] due to the recruitment of co-repressor complexes at promoter sites [67]. Additionally, TWIST1 expression can be induced under hypoxic conditions by hypoxia-inducible factor 1α (HIF1α), further promoting EMT [68].

2.2.4. Epigenetic Regulation

Epigenetic modifications have been newly discovered to play a crucial role in EMT related to neoplastic and fibrotic diseases, adding even more complexity to this fine-tuned process. Epigenetic modifications encompass multiples processes including histone deacetylation, lysine methylation, long noncoding ribonucleic acids (lncRNAs) and micro-RNA (miRNAs). The miR-200 family is among the most studied miRNAs in cancer and IPF [69,70]. miRNAs are a class of small (17–22 nucleotides) non-coding RNA molecules that negatively regulate gene expression by targeting messenger RNA (mRNA) resulting from gene transcription [71]. The miR-200 family is among the most studied miRNAs in cancer and IPF and inhibits the expression of ZEB factors [72,73]. In turn, ZEB1 and ZEB2 bind to the promoters of miR-200, thereby allowing for a reciprocal feedback loop controlling EMT [72,74]. Evidence showed a dramatic reduction of miR-200 in cells undergoing EMT [69,70].

2.3. Signaling Pathways Regulating EMT

TGF-β signaling is considered as the prototypical pathway involved in EMT and has been extensively studied. However, EMT can also be induced via several other intracellular signaling pathways resulting from the encounter between epithelial cells and specific ligands. These pathways are not well compartmentalized, and some overlap exists to a certain extent between them (Figure 2).

2.3.1. TGF-β: SMAD-Dependent and SMAD-Independent Signaling Pathways

TGF-β1 is a member of the TGF-β superfamily of cytokines and signals via a tetrameric complex consisting of two TGFβ receptor type I (TβRI) and two TGFβ receptor type II (TβRII), a transmembrane serine-threonine kinase receptor. The TGF-β1 intracellular signaling pathway is subdivided into a canonical SMAD-dependent and a non-canonical SMAD-independent pathway [75].
The binding of TGF-β1 to its receptor leads to the phosphorylation and activation of cytoplasmic SMAD2 and SMAD3, which further assemble with SMAD4 to migrate to the nucleus, where they activate the transcription of several EMT-TFs such as SNAI1, SNAI2, TWIST1, and ZEB1 [66,76]. SMAD6 and SMAD7 are negative regulators of TGF-β1 signaling and prevent SMAD2 and SMAD3 from binding to TβRI [77,78]. Additionally, it has been demonstrated that SMAD proteins interact with other signaling pathways such as WNT and NOTCH pathways via their interaction with β-catenin and NOTCH receptor intracellular domain (NOTCH-ICD) [79].
TGF-β1 can also activate non-SMAD signaling pathways. It activates several signaling pathways such as phosphoinositide 3-kinase (PI3K)–protein kinase B (AKT), extracellular signal-regulated kinase (ERK) MAPK, p38 MAPK, and c-Jun N-terminal kinase (JNK) pathways [11,80,81]. Several studies have established that TGFβ1 enables the activation of p38 MAPK and JNK pathways following the association of TNF receptor-associated factor 6 (TRAF6) with the TGF-β receptor complex [82,83]. In addition, TGF-β is also competent to activate the mammalian target of rapamycin (mTOR) complex [81] and nuclear factor-κB (NF-κB) via its collaboration with the PI3K–AKT pathway [84]. All those pathways ultimately converge in the activation of core EMT-TFs, further entailing the activation of mesenchymal genes and the repression of epithelial ones.

2.3.2. WNT Signaling Pathway

The WNT/glycogen synthase kinase-3β (GSK-3β)/β-catenin axis is an evolutionarily conserved signaling pathway that plays a pivotal role in cellular homeostasis and regulates miscellaneous processes including embryological development, cell proliferation, differentiation, apoptosis, and EMT [85]. WNT ligands bind to and activate Frizzled receptors and low-density lipoprotein receptor-related protein (LRP5/6), which promote the inhibition of GSK-3β via the cytosolic proteins disheveled (DVL) [86]. This enables the release of β-catenin from the GSK3β-axis inhibition protein (AXIN)–adenomatous polyposis coli protein (APC) complex and its translocation to the nucleus, thereby activating T cell factor (TCF) and lymphoid enhancer binding factor 1 (LEF-1) TFs, which promotes the expression of various EMT-associated genes such as SNAI1 [87,88].

2.3.3. NOTCH Signaling Pathway

NOTCH signaling is a highly conserved intercellular communication pathway involved in developmental processes, cancer, and fibrosis [89,90,91]. Delta-like (DLL)1, DLL3, DLL4, Jagged (JAG)1, and JAG2 are the cognate ligands of the NOTCH receptors (NOTCH 1–4). Upon activation of the NOTCH receptor, γ-secretase mediates the cleavage of the active NICD, ensuing its migration to the nucleus to function as a transcriptional co-activator that complex with CSL (CBF-1/Suppression of hairless/Lag1), a DNA-binding protein, to stimulate the expression of EMT TFs [92,93]. In addition, NOTCH signaling can regulate expression of SNAI2 both directly [94,95] and indirectly via the induction of HIF-1α [96]. NOTCH-ICD can also stabilize cytoplasmic β-catenin and activate other pathways, such as ERK and NF-κB, that induce the SNAI1/2 and LEF-1 TFs [97,98,99].

2.3.4. Hedgehog Signaling Pathway

Hedgehog signaling is conserved across species and plays a critical role in embryonic development [100,101]. In adults, it was linked to EMT in various cancers and fibrotic diseases [102,103]. Members of the Hedgehog (HH) ligand family include Sonic (SHH), Desert Hedgehog (DHH) and Indian Hedgehog (IHH) [104], which bind to their transmembrane protein receptors, patched (PTCH)1 and PTCH2 [105]. The activation of PTCH1/2 releases Smoothened (SMO), which ultimately leads to the activation of GLI TFs, fostering the transcription of target genes such as PTCH, WNT, and SNAI1 [106,107,108].

2.3.5. Receptor Tyrosine Kinases (RTKs) Signaling Pathway

Receptor tyrosine kinases (RTKs) belong to a large family of cell surface receptors that play a significant role in cell–cell communication, controlling complex biological functions, including cell growth, motility, differentiation, and metabolism [109,110]. Since the discovery of the first RTK in the early 1990′s, a wide variety of growth factors, cytokines, and hormones have been identified as ligands of their cognate RTKs, triggering receptor dimerization followed by receptor activation via its auto-phosphorylation. This in turn enables the activation of the PI3K–AKT, ERK–MAPK, p38 MAPK, and JNK pathways [111]. In particular, growth factors such as epidermal growth factor (EGF), fibroblast growth factor (FGF), hepatocyte growth factor (HGF), and vascular endothelial growth factor (VEGF) have been identified as potent inducers of EMT [112,113,114]. These signaling cascades activate TFs that bind to the promoters of core EMT-TFs, such as SNAI1/2, ZEB1/2, and TWIST1 [115,116]. Additionally, RTK signaling displays some cooperation with other distinct pathways. For example, AKT activation can either induce SNAI1 expression via NF-κB activation [117] or β-catenin activation via inhibition of GSK3β of WNT signaling [118]. TGF-β signaling can also enhance EMT responses initiated by growth factors such as FGF or EGF via the SMAD-independent pathway [119].

2.3.6. JAK/STAT Signaling Pathway

Janus kinase (JAK)/signal transducers and activators of transcription (STAT) signaling is critical in regulating embryonic development and inflammatory responses and has been linked to malignancies and autoimmune diseases [120,121,122]. The JAK family consists of intracellular non-receptor tyrosine kinase proteins that are activated by several cytokines or growth factors, ensuing the dimerization and activation of the transcription protein STATs. Several lines of evidence have shown that STAT dimers activate the transcription of genes encoding EMT-TFs (i.e., SNAI1, ZEB1, JUNB, and TWIST1) [123,124,125].

2.4. Extracellular Elements Regulating EMT

2.4.1. Inflammation

It is now well recognized that inflammation contributes to EMT initiation. Proinflammatory cytokines such as tumor necrosis factor (TNF)-α, and interleukin (IL)-1 are potent inducers of the NF-κB pathway entailing the secretion of various cytokines, chemokines, and adhesion molecules, all of which contribute to the recruitment of immune cells, further perpetuating a pro-inflammatory environment [126]. Additionally, several studies have linked the activation of the NF-κB pathway to EMT via its direct activation of core EMT-TFs including SNAI1, ZEB1, and ZEB2 [117,127,128,129]. SNAI1 can in return activate the transcription of pro-inflammatory cytokines genes as a positive feedback loop response [3,130,131]. In cancer and fibrotic diseases, IL-6 and IL-17 were able to promote EMT via the JAK-STAT signaling pathway and induce SNAI1 expression [132,133,134]. In addition, IL-17 has been shown to promote EMT in various conditions such as lung cancer, IBD, and IPF [135,136,137]. In particular, evidence suggests that IL-17 induces EMT in A549 alveolar epithelial cells via the TGF-β mediated SMAD signaling pathway [137].

2.4.2. Hypoxia-Induced EMT

Hypoxia is considered an EMT permissive state and has been linked to numerous conditions such as cancer, fibrosis, and autoimmune diseases [138,139,140]. Under hypoxic conditions, prolyl hydroxylase (PHD) inactivation lead to HIF-1α accumulation, further driving expression of core EMT-TFs such as SNAI1, TWIST1, ZEB1, and ZEB2 [68,141,142,143]. Furthermore, evidence has shown that hypoxia is a potent inducer of the TGF-β SMAD-dependent and PI3K/AKT pathways via HIF-1α accumulation, causing the apparition of EMT in renal fibrosis and IPF, respectively [144,145]. It was also linked in several studies with the TGF-β SMAD-dependent [146], NF-κB [147], and NOTCH signaling pathways [96].

2.4.3. Extracellular Matrix Induced EMT

ECM itself plays an important part in EMT. EMT generates multiple modifications to the composition and the mechanical properties of ECM, and in return, ECM influences EMT in a feedforward manner. These interactions have mostly been studied in many areas of cancer research field, as ECM remodeling and increased matrix stiffness observed during EMT has been linked to tumor invasion and metastasis [148]. Type I collagen enables EMT via the binding and activation of receptors such as discoidin domain receptors (DDR)1, DDR2 and integrins alpha 2 beta 1. Those receptors activate NF-κB that ultimately promote expression of SNAI1/2 and LEF1 [149,150,151]. In addition, MMPs, a large family of endoproteases that promote the degeneration of various components of the ECM, have been linked to EMT through their capacity of remodeling of ECM further facilitating migration and invasion of cells. Conversely, the remodeling of ECM by MMPs can also modulate EMT by activating several signaling pathways from the ECM and modulating the bioavailability of several ligands [152]. More specifically, MMP-3 has been shown to directly induce EMT via the activation of Rac1 GTPase–reactive oxygen species (ROS) signaling, which promotes expression of SNAI1 and further supports EMT progression [153].

3. EMT in Various Rheumatic Diseases

3.1. Rheumatoid Arthritis

RA is a chronic systemic inflammatory autoimmune disease that primarily affects joint inflammation, pain, and stiffness and can, if left untreated, lead to irreversible joint destruction and functional disability [154]. The pathophysiology of RA arises from complex interactions with multiple genetic, environmental, and immunological factors, further ensuing a breakdown of immune tolerance and the development of chronic synovial inflammation [155]. Despite the progress made in the understanding of the disease, the exact mechanisms underlying RA physiopathology still remain elusive to this day. Current treatment strategies targeting inflammation involve traditional disease-modifying anti-rheumatic drugs (DMARS) as well as biologic agents that target pro-inflammatory cytokines, including TNF-α or IL-6, B cells, or the activation of T cells. Despite this multiplicity of drugs, several observational studies still reported 16 to 21% of patients with refractory disease, suggesting the need for treatments targeting pathways other than inflammation [156].
One of the most preeminent features of RA is the striking morphological and functional transformations occurring in the synovium, including synovial lining hyperplasia and inflammatory cell infiltration. These modifications can ultimately lead to the formation of an invasive tumor-like tissue called pannus that enables invasion of the periarticular bone at the cartilage–bone junction, further hastening the development of bone erosion and cartilage destruction [157,158]. The pannus is constituted of synovial tissue hyperplasia when the lining size increases from a physiological one to three cells layers up to 10–15 cell layers [159] in which neovascularization processes [160] are observed together with the accumulation of inflammatory infiltrates including T cells, macrophages, B cells, plasma cells, dendritic cells, and mast cells [161,162]. RA fibroblast-like synoviocytes (RA-FLSs) mainly account for this significant tissue hyperplasia [163,164] and share many biological properties with tumor cells, including tumor-like proliferation, migration, and invasion as well as an increased resistance to apoptosis, even in the absence of inflammatory stimulus [165,166,167]. In addition, aberrantly activated RA-FLSs exhibit promigratory adhesion molecules, upregulated expression of proto-oncogenes and increased production of pro-inflammatory cytokines and MMPs (Figure 3) [168,169].
Over the past 15 years, several lines of evidence seem to indicate that RA-FLSs may, at least in part, undergo EMT-like process (Table 1) [31,170,171]. This conclusion was indeed supported by immunohistological analysis of both healthy and RA synovium that showed the presence of α-SMA, a myofibroblast marker responsible for collagen accumulation in fibrosis exclusively in the synovial lining layer of RA patients [31,171,172]. Furthermore, exposure to synovial fluid from RA patients induced the expression of α-SMA in healthy FLSs, suggesting the presence of an EMT-like process. Additionally, concomitant upregulation of α-SMA and podoplanin, a glycoprotein physiologically expressed in the lymphatic endothelium and associated with EMT in invasive cancers [172,173], were found in RA-FLS, whereas both markers were absent in osteoarthritis (OA) FLS (Figure 3). In vitro treatment with pro-inflammatory cytokines known to promote EMT-like behavior (i.e., IL-1β, TNF-α, and TGF-β1) induced expression of both α-SMA and podoplanin in OA-FLS, thus evoking the role of podoplanin in the activation of RA-FLS leading to their increased migratory potential [172]. Mechanistic in vitro studies showed that TGF-β1 enhanced migration and invasion by inducing EMT-like processes via the activation of SMAD2/3 in RA-FLSs, which led to the upregulation of core EMT-TF SNAI1. The inhibition of SMAD2/3 blocked EMT and limited the migration and invasion induced by TGF-β1, therefore suggesting that the TGF-β1 SMAD-dependent signaling pathway is involved in EMT-like process and contributes to migration and invasion of RA-FLSs [31]. Moreover, several studies evaluated the effect of hypoxia on RA-FLSs, as hypoxia is a well-defined inducer of EMT and is commonly found in synovial tissues of RA patients [174]. These studies suggested that the activation of the PI3K/AKT/HIF-1α and the NF-κB/HIF-1α pathways plays a significant role in hypoxia-induced EMT-like process occurring in RA-FLSs [44,175]. HIF-1α was also shown to directly regulate the expression of NOTCH-1 and NOTCH-3 under hypoxic conditions, further mediating the activation of RA-FLSs [176]. In addition, a synergetic effect of IL-17A and hypoxia was demonstrated via upregulation of the expression of MMP2 and MMP9 via the activation of the NF-κB/HIF-1α pathway [175]. The use of Baicalein, an anticancer drug inducing cancer cell apoptosis, successfully inhibited FLS proliferation and migration via inhibition of the PI3K/Akt/mTOR pathway in a human synovial sarcoma cell line (SW982), a model of in vitro RA-FLSs. The upregulation of E-cadherin together with the downregulation of vimentin, MMP2, MMP9, and SNAI1 suggests that Baicalein promotes MET, the reverse process of EMT [177]. A pre-clinical study using collagen-induced arthritis (CIA) in rats also identified Snail as a potent inhibitor of Phosphatase and TENsin homolog (PTEN) and activator of Pi3k-Akt pathways and was linked to the development of the invasive phenotype of synovial cells [178]. CircCDKN2B−AS_006, a circular RNA, was found to be upregulated in the synovial biopsies from RA patients. A pre-clinical study on CIA rats confirmed the role of circCDKN2B−AS_006 in the proliferative and invasive phenotype of RA-FLSs, whereas its knockdown alleviated the severity of arthritis. Mechanistic analysis concluded that circCDKN2B−AS_006 promoted EMT-like processes in RA-FLSs by inducing the expression of runt-related transcription factor 1 (Runx1), ultimately activating the Wnt/β−catenin signaling pathway (Table 1) [170].

3.2. Sjögren’s Syndrome

SS is a chronic systemic autoimmune disease defined by lymphoplasmacytic infiltration of the exocrine glands subsequently leading to xerostomia and keratoconjunctivitis sicca, also referred as sicca syndrome. The disease is classified either as primary SS (pSS) when it occurs in the absence of other autoimmune diseases or secondary SS (sSS) when associated with other underlying autoimmune disorders such as SLE, RA and SSc [179]. Beyond the typical sicca syndrome, systemic manifestations resulting from the lymphocytic infiltration of organs can be present in up to 40% of cases including ILD, neurological manifestations and even an increased risk of lymphoma [179,180]. Over the past decades, several lines of evidence have pointed toward the pivotal role of salivary gland epithelial cells (SGECs) in the pathogenesis of pSS, especially in the initiation and progression towards autoimmunity, justifying the increasingly widespread use of the term “autoimmune epithelitis” [181]. There are consistent data indicating that the conjuncture of genetic predisposal, environmental insults, and hormonal disequilibrium may lead to the activation of the resting epithelium, the upregulation of toll-like receptors (TLRs) such as TLR-2, 3, 4, 7, 8, and 9 [182,183,184,185], leading to the release of alarmins and pro-inflammatory cytokines such as interferon (IFN), TNF-α, IL-6 and IL-17 further promoting downstream autoimmune inflammation [184,186], ultimately resulting in the atrophy and fibrosis of the salivary glands (SG) (Figure 4) [187,188].
It was demonstrated that SG fibrosis is positively correlated with biopsy focus score and ocular surface damage [189], which led to the prevailing hypothesis suggesting that chronic exposure to pro-inflammatory mediators may drive SG fibrosis via an EMT-dependent fibrotic process (Table 2) [190,191]. Indeed, compared to SG samples from healthy subjects, immunohistochemical staining analysis of SG from SS patients consistently showed the presence of SGECs with a myofibroblast-like phenotype characterized by a spindle-shaped morphology; a strong positivity for mesenchymal markers such as vimentin, collagen type I, SNAI1, and phosphorylated SMAD2/3 and SMAD4; and a decreased expression of the epithelial marker E-cadherin [33,192]. More interestingly, a significant increase in podoplanin was also found in immunostaining of SGs from pSS patients and was associated with higher numbers of lymphatic capillaries, as well as with the presence of myoepithelial cells neighboring the ductal epithelium [193]. In addition, upon exogenous TGF-β1 treatment, SGECs from healthy donors cultured in vitro were able to recapitulate the abnormalities observed in pSS SG samples, including increased gene expression of mesenchymal markers such as phosphorylated SMAD2/3, SMAD4, SNAI1, vimentin, and collagen type I and the reduction of E-cadherin expression. In contrast, the use of the specific TβR1 inhibitor SB-431542 in healthy SGECs treated with TGF-β1 showed a significant reduction in vimentin and collagen type I expression and E-cadherin levels similar to healthy SGECs [33]. Other studies highlighted the contribution of other cytokines such as IL-6, IL-17, and IL-22 to the EMT-dependent fibrotic process taking place in SGs of SS patients. Treatments of primary cultures of SGECs obtained from healthy donors with IL-6, IL-17, and IL-22 enabled all prototypical changes related to EMT, such as the apparition of a spindle-shaped morphology, together with the upregulation of vimentin and collagen type I and downregulated E-cadherin [191,192,194]. A mechanistic study confirmed the role of IL-17 in the activation of the canonical TGF-β1/SMAD2/3 and the noncanonical TGF-β/ERK1/2 pathways in human SGECs derived from healthy subjects. Additionally, co-treatment with IL-17 and specific TβR1 inhibitor SB431542, or ERK1/2 inhibitor U0126, abrogated the corresponding morphological changes and EMT markers expression in healthy SGECs [194]. More recently, ETS1 and MMP9 were found to be overexpressed in SGECs from patients with pSS. Further functional analysis demonstrated that ETS1 directly induced expression of MMP9 by exercising its TF activity and affected the expression of EMT markers in immortalized SGECs (iSGECs) from pSS [195]. In addition, the presence of EMT in SGEC from pSS patients was further strengthened by RNA sequencing data showing an over-representation of a pathway involving MMPs (collagen type III Alpha 1 chain (COL3A1), COL1A2 and A Disintegrin And Metalloproteinase with Thrombospondin Motifs (ADAMTS)) and core-EMT TF such as ZEB2 (Table 2) [196].

3.3. Systemic Lupus Erythematosus

SLE is a chronic autoimmune connective tissue disease that affects mainly young woman of childbearing age [197]. It is defined by a chronic and excessive immune response against self-antigens resulting in the production of autoantibodies and immune complexes (IC), further causing inflammation and tissue damage in multiple organs such as the joints, skin, kidneys, lungs, and brain [198]. Lupus nephritis (LN) is a common, yet devastating complication which affects 20–60% of SLE patients over time and is still associated with significant morbidity and mortality [199]. Even though LN is usually characterized by glomerulonephritis, it is frequently associated with tubular, interstitial, and capillary lesions [200] with reported incidence of tubulointerstitial fibrosis as high as 50% [201], which led to the suspicion of the presence of an EMT-dependent fibrosis process in tubular epithelial cells derived from human renal biopsies collected from LN [202]. Interestingly, two preclinical studies investigated the origin of interstitial α-SMA fibroblasts by using kidney fibrosis mouse model and concluded that EMT programming contributed to the generation of fibroblasts in 5% and 36% of patients, respectively [29,203].
It is now well established that TGF-β1 is a potent inducer of pathological EMT-dependent fibrosis occurring during chronic kidney disease [202,204] via the downregulation of pro-epithelial markers, including E-cadherin and ZO-1; the production of MMP2 and MMP9, which alter the tubular basement membranes’ integrity and induce marked morphological changes accompanied by the de novo synthesis of the mesenchymal marker α-SMA; and the acquisition of migratory capacities facilitating the invasion of the interstitium (Table 3) [205,206]. Studies have demonstrated a significant increase in the immune expression of TGF-β1 and TGF-β3 in renal biopsies of patients with LN [207,208]. It is suspected that anti-double-stranded DNA antibodies bind to mesangial cells, leading to the secretion of pro-inflammatory cytokines and to an increase in the synthesis of profibrotic TGF-β1 and fibronectin [208,209]. In addition, limited studies have demonstrated the potential involvement of the Th17/IL-17 axis in EMT arising in tubular epithelial cells [210,211]. In cultured proximal tubular epithelial cells (HK-2 cell line), IL-17A treatment elicited the expression of collagen I and III and α-SMA and the loss of E-cadherin via a TGF-β1-dependent pathway [210]. miR-130b-3p, a member of miRNA family, was significantly upregulated in serum of patients with early stage LN compared to healthy controls and positively correlated with 24-h proteinuria and renal chronicity index. Upon stimulation of TGF-β1, miR-130b-3p transfected renal tubular cell line (HK-2) enable the decrease of E-cadherin expression and increase of α-SMA, whereas the opposite effects were obtained with miR-130b-3p inhibitors therefore preventing TGF-β1 induced EMT [212]. Recently, neutrophil extracellular traps (NETs) were more abundantly found in lung biopsies of SLE patients compared to controls. RNA sequencing arising from these samples indicated that NETosis promotes EMT in lung epithelial cells from SLE patients. These data were further confirmed in vitro by demonstrating that stimulation with NETs significantly upregulated the expression of α-SMA, Twist1, and Snai1 while downregulating the expression of E-cadherin [213]. In addition, NET was also identified in renal biopsies from patients with active LN and correlated with the severity of proteinuria and with glomerular endoMT. These results were further investigated in cultured endothelial cells (ECs), where NET could induce endoMT through the degradation of vascular–endothelial (VE)–cadherin, leading to the activation of β-catenin signaling, which ultimately resulted in an increased expression of SNAI1, MMP9 and α-SMA. In MRL/lpr mice, NET also correlated with proteinuria and was associated with the glomerular upregulation of endoMT markers, including Snail1 and MMP9 (Table 3) [214].
Several pre-clinical studies were performed using the lupus-prone model of MRL/Lpr mice and underlined the role of several molecules in EMT-dependent fibrosis and their potential interest as new therapeutic targets. As an example, targeting oncostatin M (OSM), a member of the IL-6 cytokine family highly expressed in renal tissue of LN mice, improved tubulointerstitial fibrosis with increase expression of E-cadherin and decrease of mesenchymal markers such as α-SMA, fibronectin (FN) and Stat1/Stat3, indicating that OSM may activate EMT via the Jak-Stat signaling pathway [215]. Similarly, treatment targeting fractalkine (FKN), a chemokine (C-X3-C motif) ligand 1 known to be involved in the disease progression of LN, also improved renal function and alleviated tubule-interstitial fibrosis. Mesenchymal markers such as α-SMA and vimentin were reduced, together with members of the Wnt signaling pathways, including Wnt-4 and β-catenin. Those results were confirmed in human FKN-depleted HK-2 cell line. In addition, inhibition of the Wnt/β-catenin pathway by XAV939 alienated the effects of FKN overexpression [216]. In addition the expression of Stim1 in the kidney tissues of LN mice was significantly increased compared to control mice and was positively correlated with fibronectin, urine protein, and serum creatinine levels [217]. Yap1 expression was also significantly higher in the renal tissues of MRL/lpr mice compared to normal mice, and was positively correlated with FN and with the degree of renal function injury [218]. Recently, treatment with iguratimod, an anti-inflammatory small molecule drug approved in Japan for treatment of RA that lowers the production of various cytokines including TNF-α, IL-1β, IL-6, IL-8 and IL-17, successfully alleviated tubulo-interstitial lesions and fibrosis using MRL/lpr mice. In addition, in vitro co-treatment with TGF-β1 and iguratimod impeded morphological changes associated with fibrosis, downregulated E-cadherin and upregulated fibronectin [219]. Jieduquyuziyin, a traditional Chinese medicine used for LN, displayed therapeutic actions in MRL/lpr mice, where it improved renal function and alleviated renal fibrosis. Kidney analysis of Jieduquyuziyin-treated mice showed E-cadherin upregulation together with the downregulation of vimentin, α-SMA, TGF-β1 and phosphorylated Smad2/3. These data suggest that the therapeutic effect of Jieduquyuziyin could be related to the inhibition of EMT due to the inhibition of the TGF-β1/Smad2/3 signaling pathway (Table 3) [220].

3.4. Systemic Sclerosis

SSc is a rare and heterogeneous autoimmune connective tissue disease characterized by fibroproliferative vasculopathy, auto-immunity and aberrant fibrosis of the skin and internal organs, causing progressive impairment and failure of the affected organs. Even though the complexity of the disease translates into a wide spectrum of clinical manifestations, SSc is classically divided in two distinct subsets based on the pattern of skin involvement [221,222]. Limited cutaneous SSc (lcSSc) is usually characterized by limited skin fibrosis and vascular manifestations, whereas diffuse cutaneous SSc (dcSSc) is dominated by a rapidly progressing and widespread fibrosis of the skin and internal organs, often associated with a disastrous prognosis despite the considerable progress made in the management of patients and the understanding of the underlying pathophysiology [223]. The prevailing hypothesis suggests the presence of a complex interplay between vascular cells, immune cells and fibroblasts. In individuals with a permissive genetic background, triggering events such as viral infection or exposure to environmental toxins lead to endothelial cell damage and apoptosis, thus resulting in the release of pro-inflammatory cytokines, damage-associated molecular patterns (DAMPs) and ROS and the subsequent activation of the innate and adaptive immune system [224,225,226]. Activated immune cells lead to these release of a wide variety of pro-inflammatory and pro-fibrogenic cytokines such as IL-4, IL-6, IL-13, TNF-α and TGF-β together with the secretion of autoantibodies by activated B cells, entailing further tissue damage and the apparition of activated myofibroblasts [222,227]. These activated myofibroblasts, which are normally absent in healthy tissues, display a marked resistance to apoptosis, enhanced ECM synthesis, constitutive secretion of pro-fibrotic cytokines and are suspected to arise from a wide variety of cells ranging from resident fibroblasts to pericytes, adipocytes, endothelial and epithelial cells through processes of transdifferentiation including EMT and endoMT [228,229,230].
The presence of EMT in skin and lungs of SSc patients is supported by several lines of evidence (Table 4). Firstly, treatment with TGF-β and/or TNF-α, two well-known inducers of EMT which are overexpressed in SSc patients, were able to induce EMT in normal human epidermal keratinocytes (NHEK cells) and was characterized by morphological changes including apparition of spindle-shaped cells, downregulation of epithelial markers such as E-cadherin and ZO-1, and upregulation of mesenchymal markers such as vimentin, fibronectin, and MMPs. Treatment with SB431542, a SMAD inhibitor, significantly prevented EMT and could even reverse established EMT phenotype [231]. In addition, immunostainings performed on skin biopsies of patients with morphea, a form of localized scleroderma, suggested the involvement of EMT based on the upregulation of mesenchymal markers such as TGF-β1, α-SMA, fibronectin and SNAI1 and the downregulation of E-cadherin [232]. The expression of core EMT-TFs such as SNAI1 and TWIST1 was also reported in biopsies of patients with dcSSc [233]. Furthermore, skin biopsies from dcSSc patients showed increased expression of phosphorylated SMAD2/3 (active form) and SNAI1, suggesting that the TGF-β SMAD-dependent signaling pathway is involved in the fibrotic skin of dcSSc patients [230]. Similarly, human alveolar epithelial cells type II exposed to TGF-β also displayed the expression of EMT markers and acquired the capacity to produce collagen [234]. Several other EMT-related pathways, such as the WNT and NOTCH pathways have been shown to be involved in SSc [235,236,237]. Micro-RNA 138 (MiR-138), a well-established repressor of EMT in several types of cancers [238], was found to be significantly decreased in the sera of patients with dcSSc and lcSSc compared to healthy controls, further pointing to the involvement of EMT in SSc pathophysiology [239]. In addition, the presence of an EMT-like process arising in endothelial cells, which is also referred to as endoMT, has also consistently been described in skin tissues taken from SSc patients [240,241,242,243]. Immunohistochemical analysis of skin biopsies from SSc showed significant upregulation of α-SMA, HIF-1α and VEGF-α compared with healthy controls. These observations were further reproduced by culturing human endothelial cells under hypoxic conditions [240]. Lymphatic endothelial cells were also linked with endoMT during histological analyses of skin sections from patients with SSc hybrid cells coexpressing lymphatic vessel endothelial hyaluronan receptor-1 LYVE-1, a specific marker of lymphatic endothelial cells, and α-SMA was found exclusively in the fibrotic skin of SSc patients. The culturing of HdLy-MVECs with SSc serum or profibrotic TGFβ1 led to the acquisition of a myofibroblast-like morphofunctional phenotype, as well as the downregulation of lymphatic endothelial cell-specific markers and the parallel upregulation of myofibroblast markers [242]. The distribution of OSM receptor β (OSMRβ) was significantly increased in dermal EC and in fibroblasts of SSc patients compared to controls. In vitro treatment with OSM performed in human dermal endothelial cells, inducing cell migration and cell proliferation, proinflammatory cytokines, such as IL-6 and IL-33. The effects of OSM were mediated by OSMRβ and STAT3 SNAIL1, TGFβ3, ET-1, VE-cadherin and CD31, which also increased the levels of αSMA and TGFβ1, -2, -3 at (Table 4) [244].
Several studies underlined the role of promising molecules in EMT-dependent fibrosis and their potential interest as new therapeutic targets. Brachyury, a TF recently identified as an a potent inducer of EMT in human carcinoma cell lines [245], was overexpressed in SSc dermal fibroblasts both in vivo and in vitro compared to healthy controls. The silencing of brachyury reduced the expression of type I collagen in normal and SSc dermal fibroblasts, but did not decrease the levels of major disease-related cytokines [246]. Similarly, plasma levels of SPRT4-IT1, a lncRNA reported to be involved in EMT process in various cancers, was higher in diffuse than limited SSc and were positively correlated with modified Rodnan skin score (mRSS) [247]. In addition, E3 ubiquitin ligase was linked to EMT through the TGF-β SMAD-dependent signaling pathways [248]. Recently, LG283, a curcumin derivatives drug, exhibited a promising antagonistic activity on fibrosis and vascular injury in cultured human dermal fibroblasts and in the bleomycin-induced skin fibrosis mouse model through the inhibition of TGF-β/Smad/Snai1 pathway. Treatment with LG283 was able to downregulate the expression of Col1A2, α-SMA, FN, and phosphorylated Smad3 together with core EMT-TFs Snai1 and Snai2 [249]. Furthermore, several drugs also revealed encouraging results in that they attenuated fibrosis by hindering endoMT. Linagliptin, a specific dipeptidyl peptidase 4 (DPP4) inhibitor, alleviated pulmonary fibrosis in a bleomycin-induced mouse model of SSc. Treatment with linagliptin could mitigate the migration capacity of EC, decrease inflammatory cytokines, such as TNF-α and IL-6, and was a potent inhibitor of the Akt/mTOR pathway, resulting in higher levels of VE-cadherin and decreased levels of α-SMA, collagen I and Snai1 and Twist [250]. Macitentan, the antagonist of the endothelin-1 (ET-1) receptor, prevented endoMT and fibroblast accumulation in EC from SSc patients cultivated in vitro [241].
Table 4. Recapitulation table of findings regarding EMT/endoMT occurrence in SSc.
Table 4. Recapitulation table of findings regarding EMT/endoMT occurrence in SSc.
Cellular TypeModelSignaling
Pathway
CytokineTFsPhenotypical FeatureRef.
NHEK Human/in vitro TGF-β, TNF-α spindle-shaped cells, ⇧ vimentin, ⇧ FN, ⇧ MMPs, ⇩ E-cadherin, ⇩ ZO-1[231]
Skin biopsy (keratinocytes)Human morphea TGF-β1SNAI1⇧ α-SMA, ⇧ FN, ⇩ E-cadherin[232]
Skin biopsy (keratinocytes)Human dcSSc SNAI1, TWIST1 [233]
Skin biopsy (keratinocytes)Human dcSScTGF-β/SMAD SNAI1⇧ phosphorylated SMAD2/3[230]
alveolar epithelial cells type II Human/in vitroTGF-β
WNT/β-catenin
⇧ collagen type I, ⇧ vimentin[234,237]
KeratinocytesHuman in vitroWNT/β-catenin [235]
Skin biopsy (keratinocytes)HumanNOTCH [236]
Skin biopsy
(EC)
Human/in vitroHIF-1α HIF-1α⇧ α-SMA, ⇧ VEGF-α[240]
Skin biopsy
(lymphatic EC)
Human/in vitro TGFβ1 ⇧ α-SMA, ⇩ LYVE-1[242]
Skin biopsy
(EC)
Human/in vitroOSM/ OSMRβ/SNAI1 TGFβ3 SNAI1 ⇧ ET-1, ⇧ α-SMA, ⇩ VE-cadherin [244]
Dermal fibroblastsHuman/in vitro Brachyury⇧ collagen type I[246]
PlasmaHuman lcSSc SPRT4-IT1⇧ mRSS[247]
SerumHuman dcSSc/ccSSc ⇩ MiR-138[239]
Dermal fibroblastsHuma in vitro + bleomycine miceTGF-β/SMAD/SNAI1 SNAI1, SNAI2⇧ phosphorylated SMAD3, ⇧ col1A2, ⇧ α-SMA, ⇧ FN[249]
ECbleomycine mice Akt/mTOR pathway/Snai1 TNF-α, IL-6 Snai1, Twist⇩VE-cadherin and ⇧ α-SMA, ⇧ collagen[250]
ECHuman dcSScET-1TGF-β1 ⇧ α-SMA, ⇧ col1A1[241]
Abbreviations: α-SMA: α-smooth muscle actin; dcSSc: diffuse cutaneous SSc; EC: endothelial cells; EMT: epithelial-mesenchymal transition; endoMT: endothelial-mesenchymal transition; FN: fibronectin; lcSSc: Limited cutaneous SSc; LYVE-1: lymphatic vessel endothelial hyaluronan receptor-1 LYVE-1; MiR-138: Micro-RNA 138; MMPs: matrix metalloproteases; mRSS: modified Rodnan skin score; NHEK: normal human epidermal keratinocytes; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SSc: systemic sclerosis; TFs: transcription factors; TGF-β: transforming growth factor-β; TNF: tumor necrosis factor; TWIST1: Twist Family BHLH Transcription Factor 1; VE: vascular-endothelial; ZO-1: zona occludens 1; ⇧: upregulated; ⇩: downregulated.

4. Conclusions

In this review, the current knowledge about the entanglement of EMT in the pathogenesis of several autoimmune rheumatic disease was portrayed. Even though several lines of evidence seem to confirm the involvement of EMT in rheumatic diseases, its precise contribution to their pathophysiology remains unclear, and there are, unfortunately, lots of shortcomings and scant data in the literature. However, it seems that EMT contributes to some extent to the fibrogenic process stemming from rheumatic diseases. In addition, several proinflammatory cytokines have been linked to EMT detected in rheumatic diseases, suggesting that inflammation is an important prerequisite for triggering EMT. However, whether EMT can further contribute to disease progression or actually contribute to the pathophysiology of autoimmune diseases by interfering with the immune system still remains a big missing piece of the jigsaw. Fundamentally, further research is required to better comprehend its exact contribution to the physiopathology of rheumatic diseases and could pave the way for exciting new therapies.

Author Contributions

Writing—original draft preparation, J.S.; writing—review and editing, J.S. and M.S.S.; supervision, M.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAMTSA Disintegrin And Metalloproteinase with Thrombospondin Motifs
AKTprotein kinase B
APCadenomatous polyposis coli protein
α-SMAα-smooth muscle actin
AXINaxis inhibition protein
bHLHbasic helix–loop–helix factors
CDH1E-cadherin
CIAcollagen-induced arthritis
COL3A1 collagen type III Alpha 1 chain
C-X3-C motifchemokine ligand 1
DAMPsdamage-associated molecular patterns
dcSScdiffuse cutaneous SSc
DDRdiscoidin domain receptors
DHHdesert hh
DLLDelta-like
DMARSdisease-modifying anti-rheumatic drugs
DPP4dipeptidyl peptidase 4
ECendothelial cells
ECMextracellular matrix
EGFepidermal growth factor
EMTepithelial-mesenchymal transition
EndoMT endothelial-mesenchymal transition
ERKextracellular signal– regulated kinase
ET-1endothelin-1
FDCfollicular dendritic cells
FGFfibroblast growth factor
FLS fibroblast-like synoviocytes
FNfibronectin
FSP-1fibroblast specific protein 1
GSK-3βglycogen synthase kinase-3β
HGFhepatocyte growth factor
HHHedgehog
HIF1α hypoxia-inducible factor 1α
IBDinflammatory bowel diseases
ICimmune complexes
IFNinterferon
IHHindian Hedgehog
ILinterleukin
IPFidiopathic pulmonary fibrosis
iSGECimmortalized SGEC
JAGjagged
JAKjanus kinase
JNKc-Jun N-terminal kinase
lcSScLimited cutaneous SSc lcSSc
LEF-1lymphoid enhancer binding factor 1
LNlupus nephritis
lncRNAslong noncoding ribonucleic acids
LRP5/6low-density lipoprotein receptor-related protein
LYVE-1lymphatic vessel endothelial hyaluronan receptor-1
MAPKmitogen-activated protein kinase
MEKMAP kinase kinase
METmesenchymal–epithelial transition
miRNAsmicro-RNA
MiR-138 Micro-RNA 138
MMPs matrix metalloproteases
MMT mesothelial-mesenchymal transition
mRNAmessenger RNA
mRSS modified Rodnan skin score
mTORmammalian target of rapamycin
NET neutrophil extracellular traps
NF-κBnuclear factor-κB
NHEK normal human epidermal keratinocytes
NOTCH-ICDNOTCH receptor intracellular domain
OAosteoarthritis
OSMoncostatin M
PHDprolyl hydroxylases
PI3Kphosphoinositide 3-kinase
pSSprimary SS
PTCHpatched
PTENPhosphatase and TENsin homolog
RArheumatoid arthritis
RA-FLSsRA fibroblast-like synoviocytes
ROSreactive oxygen species
RTKsreceptor tyrosine kinases
RUNX1runt-related transcription factor 1
SGsalivary glands
SGECsalivary gland epithelial cells
SHHsonic hh
SLEsystemic lupus erythematosus
SMADhomolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma
SMOsmoothened
SNAI1Snail Family Transcriptional Repressor 1
SSSjögren’s syndrome
SScsystemic sclerosis
sSSsecondary SS
STATsignal transducers and activators of transcription
TβRITGFβ receptor type I
TβRIITGFβ receptor type II
TCFT cell factor
TECtubular-epithelial cell
TFstranscription factors
TGF-βtransforming growth factor-β
TLRtoll-like receptors
TNFtumor necrosis factor
TRAF6TNF receptor-associated factor 6
TWIST1Twist Family BHLH Transcription Factor 1
VEvascular-endothelial
VEGFvascular endothelial growth factor
ZEB1Zinc Finger E-Box Binding Homeobox 1
ZEB2Zinc Finger E-Box Binding Homeobox 2
ZO-1zona occludens 1

References

  1. Thiery, J.P.; Acloque, H.; Huang, R.Y.; Nieto, M.A. Epithelial-mesenchymal transitions in development and disease. Cell 2009, 139, 871–890. [Google Scholar] [CrossRef]
  2. Nieto, M.A. Epithelial-Mesenchymal Transitions in development and disease: Old views and new perspectives. Int. J. Dev. Biol. 2009, 53, 1541–1547. [Google Scholar] [CrossRef]
  3. Grande, M.T.; Sanchez-Laorden, B.; Lopez-Blau, C.; De Frutos, C.A.; Boutet, A.; Arevalo, M.; Rowe, R.G.; Weiss, S.J.; Lopez-Novoa, J.M.; Nieto, M.A. Snail1-induced partial epithelial-to-mesenchymal transition drives renal fibrosis in mice and can be targeted to reverse established disease. Nat. Med. 2015, 21, 989–997. [Google Scholar] [CrossRef]
  4. Nieto, M.A.; Huang, R.Y.; Jackson, R.A.; Thiery, J.P. Emt: 2016. Cell 2016, 166, 21–45. [Google Scholar] [CrossRef]
  5. Sheng, L.; Zhuang, S. New Insights Into the Role and Mechanism of Partial Epithelial-Mesenchymal Transition in Kidney Fibrosis. Front. Physiol. 2020, 11, 569322. [Google Scholar] [CrossRef] [PubMed]
  6. Samavarchi-Tehrani, P.; Golipour, A.; David, L.; Sung, H.K.; Beyer, T.A.; Datti, A.; Woltjen, K.; Nagy, A.; Wrana, J.L. Functional genomics reveals a BMP-driven mesenchymal-to-epithelial transition in the initiation of somatic cell reprogramming. Cell Stem Cell 2010, 7, 64–77. [Google Scholar] [CrossRef]
  7. Li, R.; Liang, J.; Ni, S.; Zhou, T.; Qing, X.; Li, H.; He, W.; Chen, J.; Li, F.; Zhuang, Q.; et al. A mesenchymal-to-epithelial transition initiates and is required for the nuclear reprogramming of mouse fibroblasts. Cell Stem Cell 2010, 7, 51–63. [Google Scholar] [CrossRef]
  8. Tam, W.L.; Weinberg, R.A. The epigenetics of epithelial-mesenchymal plasticity in cancer. Nat. Med. 2013, 19, 1438–1449. [Google Scholar] [CrossRef]
  9. Thiery, J.P.; Sleeman, J.P. Complex networks orchestrate epithelial-mesenchymal transitions. Nat. Rev. Mol. Cell Biol. 2006, 7, 131–142. [Google Scholar] [CrossRef] [PubMed]
  10. Kalluri, R.; Weinberg, R.A. The basics of epithelial-mesenchymal transition. J. Clin. Investig. 2009, 119, 1420–1428. [Google Scholar] [CrossRef]
  11. Lamouille, S.; Xu, J.; Derynck, R. Molecular mechanisms of epithelial-mesenchymal transition. Nat. Rev. Mol. Cell Biol. 2014, 15, 178–196. [Google Scholar] [CrossRef]
  12. Greenburg, G.; Hay, E.D. Epithelia suspended in collagen gels can lose polarity and express characteristics of migrating mesenchymal cells. J. Cell Biol. 1982, 95, 333–339. [Google Scholar] [CrossRef]
  13. Nakaya, Y.; Sheng, G. Epithelial to mesenchymal transition during gastrulation: An embryological view. Dev. Growth Differ. 2008, 50, 755–766. [Google Scholar] [CrossRef]
  14. Duband, J.L.; Thiery, J.P. Appearance and distribution of fibronectin during chick embryo gastrulation and neurulation. Dev. Biol. 1982, 94, 337–350. [Google Scholar] [CrossRef]
  15. Sela-Donenfeld, D.; Kalcheim, C. Regulation of the onset of neural crest migration by coordinated activity of BMP4 and Noggin in the dorsal neural tube. Development 1999, 126, 4749–4762. [Google Scholar] [CrossRef]
  16. Lim, J.; Thiery, J.P. Epithelial-mesenchymal transitions: Insights from development. Development 2012, 139, 3471–3486. [Google Scholar] [CrossRef]
  17. Arnoux, V.; Nassour, M.; L’Helgoualc’h, A.; Hipskind, R.A.; Savagner, P. Erk5 controls Slug expression and keratinocyte activation during wound healing. Mol. Biol. Cell 2008, 19, 4738–4749. [Google Scholar] [CrossRef]
  18. Chapman, H.A. Epithelial-mesenchymal interactions in pulmonary fibrosis. Annu. Rev. Physiol. 2011, 73, 413–435. [Google Scholar] [CrossRef]
  19. Shibue, T.; Weinberg, R.A. EMT, CSCs, and drug resistance: The mechanistic link and clinical implications. Nat. Rev. Clin. Oncol. 2017, 14, 611–629. [Google Scholar] [CrossRef]
  20. De Craene, B.; Berx, G. Regulatory networks defining EMT during cancer initiation and progression. Nat. Rev. Cancer 2013, 13, 97–110. [Google Scholar] [CrossRef]
  21. Brabletz, T. To differentiate or not—Routes towards metastasis. Nat. Rev. Cancer 2012, 12, 425–436. [Google Scholar] [CrossRef] [PubMed]
  22. Singh, A.; Settleman, J. EMT, cancer stem cells and drug resistance: An emerging axis of evil in the war on cancer. Oncogene 2010, 29, 4741–4751. [Google Scholar] [CrossRef]
  23. Bataller, R.; Brenner, D.A. Liver fibrosis. J. Clin. Investig. 2005, 115, 209–218. [Google Scholar] [CrossRef]
  24. Wynn, T.A. Integrating mechanisms of pulmonary fibrosis. J. Exp. Med. 2011, 208, 1339–1350. [Google Scholar] [CrossRef]
  25. Gabbiani, G. The myofibroblast in wound healing and fibrocontractive diseases. J. Pathol. 2003, 200, 500–503. [Google Scholar] [CrossRef]
  26. Kalluri, R.; Zeisberg, M. Fibroblasts in cancer. Nat. Rev. Cancer 2006, 6, 392–401. [Google Scholar] [CrossRef]
  27. Wynn, T.A. Cellular and molecular mechanisms of fibrosis. J. Pathol. 2008, 214, 199–210. [Google Scholar] [CrossRef]
  28. Radisky, D.C.; Kenny, P.A.; Bissell, M.J. Fibrosis and cancer: Do myofibroblasts come also from epithelial cells via EMT? J. Cell. Biochem. 2007, 101, 830–839. [Google Scholar] [CrossRef]
  29. Iwano, M.; Plieth, D.; Danoff, T.M.; Xue, C.; Okada, H.; Neilson, E.G. Evidence that fibroblasts derive from epithelium during tissue fibrosis. J. Clin. Investig. 2002, 110, 341–350. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Xu, L.; Chen, Q.; Guan, T.; Lin, N.; Xu, D.; Lu, L.; Dai, Q.; Song, X. Scorpion Venom Polypeptide Inhibits Pulmonary Epithelial-Mesenchymal Transition in Systemic Sclerosis-Interstitial Lung Disease Model Mice by Intervening TGF-beta1/Smad Signaling Pathway. Evid. Based Complement. Alter. Nat. Med. 2022, 2022, 6557486. [Google Scholar] [CrossRef]
  31. Zhu, D.; Zhao, J.; Lou, A.; Huang, Q.; OuYang, Q.; Zhu, J.; Fan, M.; He, Y.; Ren, H.; Yang, M. Transforming growth factor beta1 promotes fibroblast-like synoviocytes migration and invasion via TGF-beta1/Smad signaling in rheumatoid arthritis. Mol. Cell. Biochem. 2019, 459, 141–150. [Google Scholar] [CrossRef] [PubMed]
  32. Hsieh, C.; Chang, A.; Brandt, D.; Guttikonda, R.; Utset, T.O.; Clark, M.R. Predicting outcomes of lupus nephritis with tubulointerstitial inflammation and scarring. Arthritis Care Res. 2011, 63, 865–874. [Google Scholar] [CrossRef]
  33. Sisto, M.; Lorusso, L.; Ingravallo, G.; Tamma, R.; Ribatti, D.; Lisi, S. The TGF-beta1 Signaling Pathway as an Attractive Target in the Fibrosis Pathogenesis of Sjogren’s Syndrome. Mediat. Inflamm. 2018, 2018, 1965935. [Google Scholar] [CrossRef] [PubMed]
  34. Wynn, T.A.; Ramalingam, T.R. Mechanisms of fibrosis: Therapeutic translation for fibrotic disease. Nat. Med. 2012, 18, 1028–1040. [Google Scholar] [CrossRef]
  35. Lopez-Novoa, J.M.; Nieto, M.A. Inflammation and EMT: An alliance towards organ fibrosis and cancer progression. EMBO Mol. Med. 2009, 1, 303–314. [Google Scholar] [CrossRef] [PubMed]
  36. Jose Alcaraz, M. New potential therapeutic approaches targeting synovial fibroblasts in rheumatoid arthritis. Biochem. Pharmacol. 2021, 194, 114815. [Google Scholar] [CrossRef]
  37. Sisto, M.; Ribatti, D.; Lisi, S. Molecular Mechanisms Linking Inflammation to Autoimmunity in Sjogren’s Syndrome: Identification of New Targets. Int. J. Mol. Sci. 2022, 23, 13229. [Google Scholar] [CrossRef]
  38. 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, 341–352. [Google Scholar] [CrossRef]
  39. Park, C.S.; Yoshihara, H.; Gao, Q.; Qu, C.; Iacobucci, I.; Ghate, P.S.; Connelly, J.P.; Pruett-Miller, S.M.; Wagner, B.; Robinson, C.G.; et al. Stromal-induced epithelial-mesenchymal transition induces targetable drug resistance in acute lymphoblastic leukemia. Cell Rep. 2023, 42, 112804. [Google Scholar] [CrossRef]
  40. Qi, Y.; Wang, N.; He, Y.; Zhang, J.; Zou, H.; Zhang, W.; Gu, W.; Huang, Y.; Lian, X.; Hu, J.; et al. Transforming growth factor-beta1 signaling promotes epithelial-mesenchymal transition-like phenomena, cell motility, and cell invasion in synovial sarcoma cells. PLoS ONE 2017, 12, e0182680. [Google Scholar] [CrossRef] [PubMed]
  41. Pedri, D.; Karras, P.; Landeloos, E.; Marine, J.C.; Rambow, F. Epithelial-to-mesenchymal-like transition events in melanoma. FEBS J. 2022, 289, 1352–1368. [Google Scholar] [CrossRef]
  42. Derada Troletti, C.; Fontijn, R.D.; Gowing, E.; Charabati, M.; van Het Hof, B.; Didouh, I.; van der Pol, S.M.A.; Geerts, D.; Prat, A.; van Horssen, J.; et al. Inflammation-induced endothelial to mesenchymal transition promotes brain endothelial cell dysfunction and occurs during multiple sclerosis pathophysiology. Cell Death Dis. 2019, 10, 45. [Google Scholar] [CrossRef]
  43. Keshava, S.; Owens, S.; Qin, W.; Jeffers, A.; Kyei, P.; Komatsu, S.; Kleam, J.; Ikebe, M.; Idell, S.; Tucker, T.A. The mTORC2/SGK1/NDRG1 Signaling Axis Is Critical for Mesomesenchymal Transition of Pleural Mesothelial Cells and the Progression of Pleural Fibrosis. 2023. Available online: https://doi.org/10.1165/rcmb.2023-0131OC (accessed on 20 August 2023).
  44. Li, G.Q.; Zhang, Y.; Liu, D.; Qian, Y.Y.; Zhang, H.; Guo, S.Y.; Sunagawa, M.; Hisamitsu, T.; Liu, Y.Q. PI3 kinase/Akt/HIF-1alpha pathway is associated with hypoxia-induced epithelial-mesenchymal transition in fibroblast-like synoviocytes of rheumatoid arthritis. Mol. Cell. Biochem. 2013, 372, 221–231. [Google Scholar] [CrossRef] [PubMed]
  45. Aharonov, A.; Shakked, A.; Umansky, K.B.; Savidor, A.; Genzelinakh, A.; Kain, D.; Lendengolts, D.; Revach, O.Y.; Morikawa, Y.; Dong, J.; et al. ERBB2 drives YAP activation and EMT-like processes during cardiac regeneration. Nat. Cell Biol. 2020, 22, 1346–1356. [Google Scholar] [CrossRef] [PubMed]
  46. van Roy, F.; Berx, G. The cell-cell adhesion molecule E-cadherin. Cell. Mol. Life Sci. 2008, 65, 3756–3788. [Google Scholar] [CrossRef]
  47. Loh, C.Y.; Chai, J.Y.; Tang, T.F.; Wong, W.F.; Sethi, G.; Shanmugam, M.K.; Chong, P.P.; Looi, C.Y. The E-Cadherin and N-Cadherin Switch in Epithelial-to-Mesenchymal Transition: Signaling, Therapeutic Implications, and Challenges. Cells 2019, 8, 1118. [Google Scholar] [CrossRef]
  48. Kyuno, D.; Takasawa, A.; Kikuchi, S.; Takemasa, I.; Osanai, M.; Kojima, T. Role of tight junctions in the epithelial-to-mesenchymal transition of cancer cells. Biochim. Biophys. Acta Biomembr. 2021, 1863, 183503. [Google Scholar] [CrossRef] [PubMed]
  49. 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, 415–428. [Google Scholar] [CrossRef]
  50. Zheng, H.; Kang, Y. Multilayer control of the EMT master regulators. Oncogene 2014, 33, 1755–1763. [Google Scholar] [CrossRef] [PubMed]
  51. Xie, L.; Law, B.K.; Chytil, A.M.; Brown, K.A.; Aakre, M.E.; Moses, H.L. Activation of the Erk pathway is required for TGF-beta1-induced EMT in vitro. Neoplasia 2004, 6, 603–610. [Google Scholar] [CrossRef]
  52. Yang, J.; Mani, S.A.; Donaher, J.L.; Ramaswamy, S.; Itzykson, R.A.; Come, C.; Savagner, P.; Gitelman, I.; Richardson, A.; Weinberg, R.A. Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 2004, 117, 927–939. [Google Scholar] [CrossRef]
  53. Peinado, H.; Portillo, F.; Cano, A. Transcriptional regulation of cadherins during development and carcinogenesis. Int. J. Dev. Biol. 2004, 48, 365–375. [Google Scholar] [CrossRef]
  54. Cano, A.; Perez-Moreno, M.A.; Rodrigo, I.; Locascio, A.; Blanco, M.J.; del Barrio, M.G.; Portillo, F.; Nieto, M.A. The transcription factor snail controls epithelial-mesenchymal transitions by repressing E-cadherin expression. Nat. Cell Biol. 2000, 2, 76–83. [Google Scholar] [CrossRef] [PubMed]
  55. Herranz, N.; Pasini, D.; Diaz, V.M.; 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, 4772–4781. [Google Scholar] [CrossRef]
  56. Bolos, V.; Peinado, H.; Perez-Moreno, M.A.; Fraga, M.F.; 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. [Google Scholar] [CrossRef]
  57. Ikenouchi, J.; Matsuda, M.; Furuse, M.; Tsukita, S. Regulation of tight junctions during the epithelium-mesenchyme transition: Direct repression of the gene expression of claudins/occludin by Snail. J. Cell Sci. 2003, 116, 1959–1967. [Google Scholar] [CrossRef] [PubMed]
  58. Dave, N.; Guaita-Esteruelas, S.; Gutarra, S.; Frias, A.; Beltran, M.; Peiro, S.; de Herreros, A.G. Functional cooperation between Snail1 and twist in the regulation of ZEB1 expression during epithelial to mesenchymal transition. J. Biol. Chem. 2011, 286, 12024–12032. [Google Scholar] [CrossRef]
  59. Jorda, 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, 3371–3385. [Google Scholar] [CrossRef] [PubMed]
  60. Vincent, T.; Neve, E.P.; Johnson, J.R.; Kukalev, A.; Rojo, F.; Albanell, J.; Pietras, K.; Virtanen, I.; Philipson, L.; Leopold, P.L.; et al. A SNAIL1-SMAD3/4 transcriptional repressor complex promotes TGF-beta mediated epithelial-mesenchymal transition. Nat. Cell Biol. 2009, 11, 943–950. [Google Scholar] [CrossRef]
  61. Remacle, J.E.; Kraft, H.; Lerchner, W.; Wuytens, G.; Collart, C.; Verschueren, K.; Smith, J.C.; Huylebroeck, D. New mode of DNA binding of multi-zinc finger transcription factors: δEF1 family members bind with two hands to two target sites. EMBO J. 1999, 18, 5073–5084. [Google Scholar] [CrossRef]
  62. Eger, A.; Aigner, K.; Sonderegger, S.; Dampier, B.; Oehler, S.; Schreiber, M.; Berx, G.; Cano, A.; Beug, H.; Foisner, R. DeltaEF1 is a transcriptional repressor of E-cadherin and regulates epithelial plasticity in breast cancer cells. Oncogene 2005, 24, 2375–2385. [Google Scholar] [CrossRef]
  63. Sanchez-Tillo, E.; Lazaro, A.; Torrent, R.; Cuatrecasas, M.; Vaquero, E.C.; Castells, A.; Engel, P.; Postigo, A. ZEB1 represses E-cadherin and induces an EMT by recruiting the SWI/SNF chromatin-remodeling protein BRG1. Oncogene 2010, 29, 3490–3500. [Google Scholar] [CrossRef] [PubMed]
  64. 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, 1265–1273. [Google Scholar] [CrossRef]
  65. Miyoshi, A.; Kitajima, Y.; Kido, S.; Shimonishi, T.; Matsuyama, S.; Kitahara, K.; Miyazaki, K. Snail accelerates cancer invasion by upregulating MMP expression and is associated with poor prognosis of hepatocellular carcinoma. Br. J. Cancer 2005, 92, 252–258. [Google Scholar] [CrossRef]
  66. Xu, J.; Lamouille, S.; Derynck, R. TGF-beta-induced epithelial to mesenchymal transition. Cell Res. 2009, 19, 156–172. [Google Scholar] [CrossRef]
  67. Yang, M.H.; Hsu, D.S.; Wang, H.W.; Wang, H.J.; Lan, H.Y.; Yang, W.H.; Huang, C.H.; Kao, S.Y.; Tzeng, C.H.; Tai, S.K.; et al. Bmi1 is essential in Twist1-induced epithelial-mesenchymal transition. Nat. Cell Biol. 2010, 12, 982–992. [Google Scholar] [CrossRef] [PubMed]
  68. Yang, M.H.; Wu, M.Z.; Chiou, S.H.; Chen, P.M.; Chang, S.Y.; Liu, C.J.; Teng, S.C.; Wu, K.J. Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat. Cell Biol. 2008, 10, 295–305. [Google Scholar] [CrossRef] [PubMed]
  69. Humphries, B.; Yang, C. The microRNA-200 family: Small molecules with novel roles in cancer development, progression and therapy. Oncotarget 2015, 6, 6472–6498. [Google Scholar] [CrossRef]
  70. Moimas, S.; Salton, F.; Kosmider, B.; Ring, N.; Volpe, M.C.; Bahmed, K.; Braga, L.; Rehman, M.; Vodret, S.; Graziani, M.L.; et al. miR-200 family members reduce senescence and restore idiopathic pulmonary fibrosis type II alveolar epithelial cell transdifferentiation. ERJ Open Res. 2019, 5, 00138–2019. [Google Scholar] [CrossRef] [PubMed]
  71. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef] [PubMed]
  72. 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, 582–589. [Google Scholar] [CrossRef] [PubMed]
  73. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Khew-Goodall, Y.; Goodall, G.J. The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  74. Bracken, C.P.; Gregory, P.A.; Kolesnikoff, N.; Bert, A.G.; Wang, J.; Shannon, M.F.; Goodall, G.J. A double-negative feedback loop between ZEB1-SIP1 and the microRNA-200 family regulates epithelial-mesenchymal transition. Cancer Res. 2008, 68, 7846–7854. [Google Scholar] [CrossRef]
  75. Batlle, E.; Massague, J. Transforming Growth Factor-beta Signaling in Immunity and Cancer. Immunity 2019, 50, 924–940. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, L.; Yang, T.; Lu, D.W.; Zhao, H.; Feng, Y.L.; Chen, H.; Chen, D.Q.; Vaziri, N.D.; Zhao, Y.Y. Central role of dysregulation of TGF-beta/Smad in CKD progression and potential targets of its treatment. Biomed. Pharmacother. 2018, 101, 670–681. [Google Scholar] [CrossRef]
  77. Yan, X.; Chen, Y.G. Smad7: Not only a regulator, but also a cross-talk mediator of TGF-beta signalling. Biochem. J. 2011, 434, 1–10. [Google Scholar] [CrossRef]
  78. Yan, X.; Liao, H.; Cheng, M.; Shi, X.; Lin, X.; Feng, X.H.; Chen, Y.G. Smad7 Protein Interacts with Receptor-regulated Smads (R-Smads) to Inhibit Transforming Growth Factor-beta (TGF-beta)/Smad Signaling. J. Biol. Chem. 2016, 291, 382–392. [Google Scholar] [CrossRef]
  79. Luo, K. Signaling Cross Talk between TGF-beta/Smad and Other Signaling Pathways. Cold Spring Harb. Perspect. Biol. 2017, 9, 1–28. [Google Scholar] [CrossRef]
  80. Geng, X.Q.; Ma, A.; He, J.Z.; Wang, L.; Jia, Y.L.; Shao, G.Y.; Li, M.; Zhou, H.; Lin, S.Q.; Ran, J.H.; et al. Ganoderic acid hinders renal fibrosis via suppressing the TGF-beta/Smad and MAPK signaling pathways. Acta Pharmacol. Sin. 2020, 41, 670–677. [Google Scholar] [CrossRef]
  81. Bakin, A.V.; Tomlinson, A.K.; Bhowmick, N.A.; Moses, H.L.; Arteaga, C.L. Phosphatidylinositol 3-kinase function is required for transforming growth factor beta-mediated epithelial to mesenchymal transition and cell migration. J. Biol. Chem. 2000, 275, 36803–36810. [Google Scholar] [CrossRef]
  82. Yamashita, M.; Fatyol, K.; Jin, C.; Wang, X.; Liu, Z.; Zhang, Y.E. TRAF6 mediates Smad-independent activation of JNK and p38 by TGF-beta. Mol. Cell 2008, 31, 918–924. [Google Scholar] [CrossRef] [PubMed]
  83. Sorrentino, A.; Thakur, N.; Grimsby, S.; Marcusson, A.; von Bulow, V.; Schuster, N.; Zhang, S.; Heldin, C.H.; Landstrom, M. The type I TGF-beta receptor engages TRAF6 to activate TAK1 in a receptor kinase-independent manner. Nat. Cell Biol. 2008, 10, 1199–1207. [Google Scholar] [CrossRef] [PubMed]
  84. Gonzalez, D.M.; Medici, D. Signaling mechanisms of the epithelial-mesenchymal transition. Sci. Signal. 2014, 7, re8. [Google Scholar] [CrossRef]
  85. MacDonald, B.T.; Tamai, K.; He, X. Wnt/beta-catenin signaling: Components, mechanisms, and diseases. Dev. Cell 2009, 17, 9–26. [Google Scholar] [CrossRef] [PubMed]
  86. Niehrs, C. The complex world of WNT receptor signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 767–779. [Google Scholar] [CrossRef]
  87. Tan, R.J.; Zhou, D.; Zhou, L.; Liu, Y. Wnt/beta-catenin signaling and kidney fibrosis. Kidney Int. Suppl. 2014, 4, 84–90. [Google Scholar] [CrossRef] [PubMed]
  88. Zhou, B.P.; Deng, J.; Xia, W.; Xu, J.; Li, Y.M.; Gunduz, M.; Hung, M.C. Dual regulation of Snail by GSK-3beta-mediated phosphorylation in control of epithelial-mesenchymal transition. Nat. Cell Biol. 2004, 6, 931–940. [Google Scholar] [CrossRef]
  89. Schweisguth, F. Regulation of notch signaling activity. Curr. Biol. 2004, 14, R129–R138. [Google Scholar] [CrossRef]
  90. Radtke, F.; Raj, K. The role of Notch in tumorigenesis: Oncogene or tumour suppressor? Nat. Rev. Cancer 2003, 3, 756–767. [Google Scholar] [CrossRef]
  91. Edeling, M.; Ragi, G.; Huang, S.; Pavenstadt, H.; Susztak, K. Developmental signalling pathways in renal fibrosis: The roles of Notch, Wnt and Hedgehog. Nat. Rev. Nephrol. 2016, 12, 426–439. [Google Scholar] [CrossRef]
  92. Kopan, R. Notch: A membrane-bound transcription factor. J. Cell Sci. 2002, 115 Pt 6, 1095–1097. [Google Scholar] [CrossRef]
  93. Jeong, M.H.; Kim, H.R.; Park, Y.J.; Chung, K.H. Akt and Notch pathways mediate polyhexamethylene guanidine phosphate-induced epithelial-mesenchymal transition via ZEB2. Toxicol. Appl. Pharmacol. 2019, 380, 114691. [Google Scholar] [CrossRef]
  94. Niessen, K.; Fu, Y.; Chang, L.; Hoodless, P.A.; McFadden, D.; Karsan, A. Slug is a direct Notch target required for initiation of cardiac cushion cellularization. J. Cell Biol. 2008, 182, 315–325. [Google Scholar] [CrossRef] [PubMed]
  95. Leong, K.G.; Niessen, K.; Kulic, I.; Raouf, A.; Eaves, C.; Pollet, I.; Karsan, A. Jagged1-mediated Notch activation induces epithelial-to-mesenchymal transition through Slug-induced repression of E-cadherin. J. Exp. Med. 2007, 204, 2935–2948. [Google Scholar] [CrossRef]
  96. Sahlgren, C.; Gustafsson, M.V.; Jin, S.; Poellinger, L.; Lendahl, U. Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc. Natl. Acad. Sci. USA 2008, 105, 6392–6397. [Google Scholar] [CrossRef]
  97. Miele, L. Notch signaling. Clin. Cancer Res. 2006, 12, 1074–1079. [Google Scholar] [CrossRef] [PubMed]
  98. Li, L.; Zhang, J.; Xiong, N.; Li, S.; Chen, Y.; Yang, H.; Wu, C.; Zeng, H.; Liu, Y. Notch-1 signaling activates NF-kappaB in human breast carcinoma MDA-MB-231 cells via PP2A-dependent AKT pathway. Med. Oncol. 2016, 33, 33. [Google Scholar] [CrossRef] [PubMed]
  99. Christopoulos, P.F.; Gjolberg, T.T.; Kruger, S.; Haraldsen, G.; Andersen, J.T.; Sundlisaeter, E. Targeting the Notch Signaling Pathway in Chronic Inflammatory Diseases. Front. Immunol. 2021, 12, 668207. [Google Scholar] [CrossRef]
  100. Ingham, P.W.; McMahon, A.P. Hedgehog signaling in animal development: Paradigms and principles. Genes Dev. 2001, 15, 3059–3087. [Google Scholar] [CrossRef]
  101. Varjosalo, M.; Taipale, J. Hedgehog: Functions and mechanisms. Genes Dev. 2008, 22, 2454–2472. [Google Scholar] [CrossRef]
  102. Fabian, S.L.; Penchev, R.R.; St-Jacques, B.; Rao, A.N.; Sipila, P.; West, K.A.; McMahon, A.P.; Humphreys, B.D. Hedgehog-Gli pathway activation during kidney fibrosis. Am. J. Pathol. 2012, 180, 1441–1453. [Google Scholar] [CrossRef]
  103. Hanna, A.; Shevde, L.A. Hedgehog signaling: Modulation of cancer properies and tumor mircroenvironment. Mol. Cancer 2016, 15, 24. [Google Scholar] [CrossRef]
  104. Katoh, Y.; Katoh, M. Comparative genomics on Sonic hedgehog orthologs. Oncol. Rep. 2005, 14, 1087–1090. [Google Scholar] [CrossRef] [PubMed]
  105. van den Heuvel, M.; Ingham, P.W. smoothened encodes a receptor-like serpentine protein required for hedgehog signalling. Nature 1996, 382, 547–551. [Google Scholar] [CrossRef] [PubMed]
  106. Li, X.; Deng, W.; Nail, C.D.; Bailey, S.K.; Kraus, M.H.; Ruppert, J.M.; Lobo-Ruppert, S.M. Snail induction is an early response to Gli1 that determines the efficiency of epithelial transformation. Oncogene 2006, 25, 609–621. [Google Scholar] [CrossRef] [PubMed]
  107. Taipale, J.; Beachy, P.A. The Hedgehog and Wnt signalling pathways in cancer. Nature 2001, 411, 349–354. [Google Scholar] [CrossRef]
  108. Briscoe, J.; Therond, P.P. The mechanisms of Hedgehog signalling and its roles in development and disease. Nat. Rev. Mol. Cell Biol. 2013, 14, 416–429. [Google Scholar] [CrossRef]
  109. Robinson, D.R.; Wu, Y.M.; Lin, S.F. The protein tyrosine kinase family of the human genome. Oncogene 2000, 19, 5548–5557. [Google Scholar] [CrossRef]
  110. Ullrich, A.; Schlessinger, J. Signal transduction by receptors with tyrosine kinase activity. Cell 1990, 61, 203–212. [Google Scholar] [CrossRef]
  111. Lemmon, M.A.; Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell 2010, 141, 1117–1134. [Google Scholar] [CrossRef]
  112. Iderzorig, T.; Kellen, J.; Osude, C.; Singh, S.; Woodman, J.A.; Garcia, C.; Puri, N. Comparison of EMT mediated tyrosine kinase inhibitor resistance in NSCLC. Biochem. Biophys. Res. Commun. 2018, 496, 770–777. [Google Scholar] [CrossRef]
  113. Grimminger, F.; Gunther, A.; Vancheri, C. The role of tyrosine kinases in the pathogenesis of idiopathic pulmonary fibrosis. Eur. Respir. J. 2015, 45, 1426–1433. [Google Scholar] [CrossRef]
  114. Tsai, J.H.; Yang, J. Epithelial-mesenchymal plasticity in carcinoma metastasis. Genes Dev. 2013, 27, 2192–2206. [Google Scholar] [CrossRef] [PubMed]
  115. Gulhati, P.; Bowen, K.A.; Liu, J.; Stevens, P.D.; Rychahou, P.G.; Chen, M.; Lee, E.Y.; Weiss, H.L.; O’Connor, K.L.; Gao, T.; et al. mTORC1 and mTORC2 regulate EMT, motility, and metastasis of colorectal cancer via RhoA and Rac1 signaling pathways. Cancer Res. 2011, 71, 3246–3256. [Google Scholar] [CrossRef] [PubMed]
  116. Doehn, U.; Hauge, C.; Frank, S.R.; Jensen, C.J.; Duda, K.; Nielsen, J.V.; Cohen, M.S.; Johansen, J.V.; Winther, B.R.; Lund, L.R.; et al. RSK is a principal effector of the RAS-ERK pathway for eliciting a coordinate promotile/invasive gene program and phenotype in epithelial cells. Mol. Cell 2009, 35, 511–522. [Google Scholar] [CrossRef] [PubMed]
  117. Julien, S.; Puig, I.; Caretti, E.; Bonaventure, J.; Nelles, L.; van Roy, F.; Dargemont, C.; de Herreros, A.G.; Bellacosa, A.; Larue, L. Activation of NF-kappaB by Akt upregulates Snail expression and induces epithelium mesenchyme transition. Oncogene 2007, 26, 7445–7456. [Google Scholar] [CrossRef]
  118. Wang, H.; Fang, R.; Wang, X.F.; Zhang, F.; Chen, D.Y.; Zhou, B.; Wang, H.S.; Cai, S.H.; Du, J. Stabilization of Snail through AKT/GSK-3beta signaling pathway is required for TNF-alpha-induced epithelial-mesenchymal transition in prostate cancer PC3 cells. Eur. J. Pharmacol. 2013, 714, 48–55. [Google Scholar] [CrossRef]
  119. Yamaguchi, K.; Shirakabe, K.; Shibuya, H.; Irie, K.; Oishi, I.; Ueno, N.; Taniguchi, T.; Nishida, E.; Matsumoto, K. Identification of a member of the MAPKKK family as a potential mediator of TGF-beta signal transduction. Science 1995, 270, 2008–2011. [Google Scholar] [CrossRef]
  120. Darnell, J.E., Jr. STATs and gene regulation. Science 1997, 277, 1630–1635. [Google Scholar] [CrossRef] [PubMed]
  121. O’Shea, J.J.; Schwartz, D.M.; Villarino, A.V.; Gadina, M.; McInnes, I.B.; Laurence, A. The JAK-STAT pathway: Impact on human disease and therapeutic intervention. Annu. Rev. Med. 2015, 66, 311–328. [Google Scholar] [CrossRef] [PubMed]
  122. Banerjee, S.; Biehl, A.; Gadina, M.; Hasni, S.; Schwartz, D.M. JAK-STAT Signaling as a Target for Inflammatory and Autoimmune Diseases: Current and Future Prospects. Drugs 2017, 77, 521–546. [Google Scholar] [CrossRef]
  123. Yadav, A.; Kumar, B.; Datta, J.; Teknos, T.N.; Kumar, P. IL-6 promotes head and neck tumor metastasis by inducing epithelial-mesenchymal transition via the JAK-STAT3-SNAIL signaling pathway. Mol. Cancer Res. 2011, 9, 1658–1667. [Google Scholar] [CrossRef]
  124. Xiong, H.; Hong, J.; Du, W.; Lin, Y.W.; Ren, L.L.; Wang, Y.C.; Su, W.Y.; Wang, J.L.; Cui, Y.; Wang, Z.H.; et al. Roles of STAT3 and ZEB1 proteins in E-cadherin down-regulation and human colorectal cancer epithelial-mesenchymal transition. J. Biol. Chem. 2012, 287, 5819–5832. [Google Scholar] [CrossRef]
  125. Milara, J.; Hernandez, G.; Ballester, B.; Morell, A.; Roger, I.; Montero, P.; Escriva, J.; Lloris, J.M.; Molina-Molina, M.; Morcillo, E.; et al. The JAK2 pathway is activated in idiopathic pulmonary fibrosis. Respir. Res. 2018, 19, 24. [Google Scholar] [CrossRef] [PubMed]
  126. Blackwell, T.S.; Christman, J.W. The role of nuclear factor-kappa B in cytokine gene regulation. Am. J. Respir. Cell Mol. Biol. 1997, 17, 3–9. [Google Scholar] [CrossRef]
  127. Bachelder, R.E.; Yoon, S.O.; Franci, C.; de Herreros, A.G.; Mercurio, A.M. Glycogen synthase kinase-3 is an endogenous inhibitor of Snail transcription: Implications for the epithelial-mesenchymal transition. J. Cell Biol. 2005, 168, 29–33. [Google Scholar] [CrossRef] [PubMed]
  128. Barbera, M.J.; Puig, I.; Dominguez, D.; Julien-Grille, S.; Guaita-Esteruelas, S.; Peiro, S.; Baulida, J.; Franci, C.; Dedhar, S.; Larue, L.; et al. Regulation of Snail transcription during epithelial to mesenchymal transition of tumor cells. Oncogene 2004, 23, 7345–7354. [Google Scholar] [CrossRef]
  129. Chua, H.L.; Bhat-Nakshatri, P.; Clare, S.E.; Morimiya, A.; Badve, S.; Nakshatri, H. NF-kappaB represses E-cadherin expression and enhances epithelial to mesenchymal transition of mammary epithelial cells: Potential involvement of ZEB-1 and ZEB-2. Oncogene 2007, 26, 711–724. [Google Scholar] [CrossRef] [PubMed]
  130. Hsu, D.S.; Wang, H.J.; Tai, S.K.; Chou, C.H.; Hsieh, C.H.; Chiu, P.H.; Chen, N.J.; Yang, M.H. Acetylation of snail modulates the cytokinome of cancer cells to enhance the recruitment of macrophages. Cancer Cell 2014, 26, 534–548. [Google Scholar] [CrossRef]
  131. Lovisa, S.; LeBleu, V.S.; Tampe, B.; Sugimoto, H.; Vadnagara, K.; Carstens, J.L.; Wu, C.C.; Hagos, Y.; Burckhardt, B.C.; Pentcheva-Hoang, T.; et al. Epithelial-to-mesenchymal transition induces cell cycle arrest and parenchymal damage in renal fibrosis. Nat. Med. 2015, 21, 998–1009. [Google Scholar] [CrossRef] [PubMed]
  132. Heinrich, P.C.; Behrmann, I.; Haan, S.; Hermanns, H.M.; Muller-Newen, G.; Schaper, F. Principles of interleukin (IL)-6-type cytokine signalling and its regulation. Biochem. J. 2003, 374 Pt 1, 1–20. [Google Scholar] [CrossRef]
  133. Wu, X.; Tao, P.; Zhou, Q.; Li, J.; Yu, Z.; Wang, X.; Li, J.; Li, C.; Yan, M.; Zhu, Z.; et al. IL-6 secreted by cancer-associated fibroblasts promotes epithelial-mesenchymal transition and metastasis of gastric cancer via JAK2/STAT3 signaling pathway. Oncotarget 2017, 8, 20741–20750. [Google Scholar] [CrossRef] [PubMed]
  134. She, Y.X.; Yu, Q.Y.; Tang, X.X. Role of interleukins in the pathogenesis of pulmonary fibrosis. Cell Death Discov. 2021, 7, 52. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, H.J.; Zhang, Y.N.; Zhou, H.; Guan, L.; Li, Y.; Sun, M.J. IL-17A Promotes Initiation and Development of Intestinal Fibrosis Through EMT. Dig. Dis. Sci. 2018, 63, 2898–2909. [Google Scholar] [CrossRef] [PubMed]
  136. Gu, K.; Li, M.M.; Shen, J.; Liu, F.; Cao, J.Y.; Jin, S.; Yu, Y. Interleukin-17-induced EMT promotes lung cancer cell migration and invasion via NF-kappaB/ZEB1 signal pathway. Am. J. Cancer Res. 2015, 5, 1169–1179. [Google Scholar] [PubMed]
  137. Wang, T.; Liu, Y.; Zou, J.F.; Cheng, Z.S. Interleukin-17 induces human alveolar epithelial to mesenchymal cell transition via the TGF-beta1 mediated Smad2/3 and ERK1/2 activation. PLoS ONE 2017, 12, e0183972. [Google Scholar] [CrossRef]
  138. Emami Nejad, A.; Najafgholian, S.; Rostami, A.; Sistani, A.; Shojaeifar, S.; Esparvarinha, M.; Nedaeinia, R.; Haghjooy Javanmard, S.; Taherian, M.; Ahmadlou, M.; et al. The role of hypoxia in the tumor microenvironment and development of cancer stem cell: A novel approach to developing treatment. Cancer Cell Int. 2021, 21, 62. [Google Scholar] [CrossRef]
  139. Lokmic, Z.; Musyoka, J.; Hewitson, T.D.; Darby, I.A. Hypoxia and hypoxia signaling in tissue repair and fibrosis. Int. Rev. Cell Mol. Biol. 2012, 296, 139–185. [Google Scholar] [CrossRef] [PubMed]
  140. Watts, E.R.; Walmsley, S.R. Inflammation and Hypoxia: HIF and PHD Isoform Selectivity. Trends Mol. Med. 2019, 25, 33–46. [Google Scholar] [CrossRef] [PubMed]
  141. Imai, T.; Horiuchi, A.; Wang, C.; Oka, K.; Ohira, S.; Nikaido, T.; Konishi, I. Hypoxia attenuates the expression of E-cadherin via up-regulation of SNAIL in ovarian carcinoma cells. Am. J. Pathol. 2003, 163, 1437–1447. [Google Scholar] [CrossRef]
  142. Krishnamachary, B.; Zagzag, D.; Nagasawa, H.; Rainey, K.; Okuyama, H.; Baek, J.H.; Semenza, G.L. Hypoxia-inducible factor-1-dependent repression of E-cadherin in von Hippel-Lindau tumor suppressor-null renal cell carcinoma mediated by TCF3, ZFHX1A, and ZFHX1B. Cancer Res. 2006, 66, 2725–2731. [Google Scholar] [CrossRef]
  143. Evans, A.J.; Russell, R.C.; Roche, O.; Burry, T.N.; Fish, J.E.; Chow, V.W.; Kim, W.Y.; Saravanan, A.; Maynard, M.A.; Gervais, M.L.; et al. VHL promotes E2 box-dependent E-cadherin transcription by HIF-mediated regulation of SIP1 and snail. Mol. Cell Biol. 2007, 27, 157–169. [Google Scholar] [CrossRef] [PubMed]
  144. Wei, X.; Zhu, X.; Jiang, L.; Huang, X.; Zhang, Y.; Zhao, D.; Du, Y. Recent advances in understanding the role of hypoxia-inducible factor 1alpha in renal fibrosis. Int. Urol. Nephrol. 2020, 52, 1287–1295. [Google Scholar] [CrossRef] [PubMed]
  145. Aquino-Galvez, A.; Gonzalez-Avila, G.; Jimenez-Sanchez, L.L.; Maldonado-Martinez, H.A.; Cisneros, J.; Toscano-Marquez, F.; Castillejos-Lopez, M.; Torres-Espindola, L.M.; Velazquez-Cruz, R.; Rodriguez, V.H.O.; et al. Dysregulated expression of hypoxia-inducible factors augments myofibroblasts differentiation in idiopathic pulmonary fibrosis. Respir. Res. 2019, 20, 130. [Google Scholar] [CrossRef] [PubMed]
  146. Furuta, C.; Miyamoto, T.; Takagi, T.; Noguchi, Y.; Kaneko, J.; Itoh, S.; Watanabe, T.; Itoh, F. Transforming growth factor-beta signaling enhancement by long-term exposure to hypoxia in a tumor microenvironment composed of Lewis lung carcinoma cells. Cancer Sci. 2015, 106, 1524–1533. [Google Scholar] [CrossRef]
  147. Koong, A.C.; Chen, E.Y.; Giaccia, A.J. Hypoxia causes the activation of nuclear factor kappa B through the phosphorylation of I kappa B alpha on tyrosine residues. Cancer Res. 1994, 54, 1425–1430. [Google Scholar]
  148. Wei, S.C.; Fattet, L.; Tsai, J.H.; Guo, Y.; Pai, V.H.; Majeski, H.E.; Chen, A.C.; Sah, R.L.; Taylor, S.S.; Engler, A.J.; et al. Matrix stiffness drives epithelial-mesenchymal transition and tumour metastasis through a TWIST1-G3BP2 mechanotransduction pathway. Nat. Cell Biol. 2015, 17, 678–688. [Google Scholar] [CrossRef]
  149. Medici, D.; Nawshad, A. Type I collagen promotes epithelial-mesenchymal transition through ILK-dependent activation of NF-kappaB and LEF-1. Matrix Biol. 2010, 29, 161–165. [Google Scholar] [CrossRef]
  150. Valles, A.M.; Boyer, B.; Tarone, G.; Thiery, J.P. Alpha 2 beta 1 integrin is required for the collagen and FGF-1 induced cell dispersion in a rat bladder carcinoma cell line. Cell Adhes. Commun. 1996, 4, 187–199. [Google Scholar] [CrossRef]
  151. Walsh, L.A.; Nawshad, A.; Medici, D. Discoidin domain receptor 2 is a critical regulator of epithelial-mesenchymal transition. Matrix Biol. 2011, 30, 243–247. [Google Scholar] [CrossRef]
  152. Nistico, P.; Bissell, M.J.; Radisky, D.C. Epithelial-mesenchymal transition: General principles and pathological relevance with special emphasis on the role of matrix metalloproteinases. Cold Spring Harb. Perspect. Biol. 2012, 4, a011908. [Google Scholar] [CrossRef]
  153. Stallings-Mann, M.L.; Waldmann, J.; Zhang, Y.; Miller, E.; Gauthier, M.L.; Visscher, D.W.; Downey, G.P.; Radisky, E.S.; Fields, A.P.; Radisky, D.C. Matrix metalloproteinase induction of Rac1b, a key effector of lung cancer progression. Sci. Transl. Med. 2012, 4, 142ra195. [Google Scholar] [CrossRef] [PubMed]
  154. Feldmann, M.; Brennan, F.M.; Maini, R.N. Rheumatoid arthritis. Cell 1996, 85, 307–310. [Google Scholar] [CrossRef] [PubMed]
  155. Firestein, G.S. Evolving concepts of rheumatoid arthritis. Nature 2003, 423, 356–361. [Google Scholar] [CrossRef]
  156. Melville, A.R.; Kearsley-Fleet, L.; Buch, M.H.; Hyrich, K.L. Understanding Refractory Rheumatoid Arthritis: Implications for a Therapeutic Approach. Drugs 2020, 80, 849–857. [Google Scholar] [CrossRef]
  157. Mousavi, M.J.; Karami, J.; Aslani, S.; Tahmasebi, M.N.; Vaziri, A.S.; Jamshidi, A.; Farhadi, E.; Mahmoudi, M. Transformation of fibroblast-like synoviocytes in rheumatoid arthritis; from a friend to foe. Auto Immun. Highlights 2021, 12, 3. [Google Scholar] [CrossRef]
  158. Huber, L.C.; Distler, O.; Tarner, I.; Gay, R.E.; Gay, S.; Pap, T. Synovial fibroblasts: Key players in rheumatoid arthritis. Rheumatology 2006, 45, 669–675. [Google Scholar] [CrossRef] [PubMed]
  159. Mor, A.; Abramson, S.B.; Pillinger, M.H. The fibroblast-like synovial cell in rheumatoid arthritis: A key player in inflammation and joint destruction. Clin. Immunol. 2005, 115, 118–128. [Google Scholar] [CrossRef] [PubMed]
  160. Bodolay, E.; Koch, A.E.; Kim, J.; Szegedi, G.; Szekanecz, Z. Angiogenesis and chemokines in rheumatoid arthritis and other systemic inflammatory rheumatic diseases. J. Cell. Mol. Med. 2002, 6, 357–376. [Google Scholar] [CrossRef] [PubMed]
  161. Godfrey, H.P.; Ilardi, C.; Engber, W.; Graziano, F.M. Quantitation of human synovial mast cells in rheumatoid arthritis and other rheumatic diseases. Arthritis Rheum. 1984, 27, 852–856. [Google Scholar] [CrossRef]
  162. Ghadially, F.N.; Roy, S. Ultrastructure of synovial membrane in rheumatoid arthritis. Ann. Rheum. Dis. 1967, 26, 426–443. [Google Scholar] [CrossRef] [PubMed]
  163. Meyer, L.H.; Franssen, L.; Pap, T. The role of mesenchymal cells in the pathophysiology of inflammatory arthritis. Best Pract. Res. Clin. Rheumatol. 2006, 20, 969–981. [Google Scholar] [CrossRef] [PubMed]
  164. Noss, E.H.; Brenner, M.B. The role and therapeutic implications of fibroblast-like synoviocytes in inflammation and cartilage erosion in rheumatoid arthritis. Immunol. Rev. 2008, 223, 252–270. [Google Scholar] [CrossRef] [PubMed]
  165. Muller-Ladner, U.; Kriegsmann, J.; Franklin, B.N.; Matsumoto, S.; Geiler, T.; Gay, R.E.; Gay, S. Synovial fibroblasts of patients with rheumatoid arthritis attach to and invade normal human cartilage when engrafted into SCID mice. Am. J. Pathol. 1996, 149, 1607–1615. [Google Scholar] [PubMed]
  166. Baier, A.; Meineckel, I.; Gay, S.; Pap, T. Apoptosis in rheumatoid arthritis. Curr. Opin. Rheumatol. 2003, 15, 274–279. [Google Scholar] [CrossRef]
  167. Ma, K.; Zhang, C.; Li, W. Fascin1 mediated release of pro-inflammatory cytokines and invasion/migration in rheumatoid arthritis via the STAT3 pathway. Cell Cycle 2021, 20, 2210–2220. [Google Scholar] [CrossRef] [PubMed]
  168. Karouzakis, E.; Gay, R.E.; Gay, S.; Neidhart, M. Epigenetic control in rheumatoid arthritis synovial fibroblasts. Nat. Rev. Rheumatol. 2009, 5, 266–272. [Google Scholar] [CrossRef]
  169. Tolboom, T.C.; van der Helm-Van Mil, A.H.; Nelissen, R.G.; Breedveld, F.C.; Toes, R.E.; Huizinga, T.W. Invasiveness of fibroblast-like synoviocytes is an individual patient characteristic associated with the rate of joint destruction in patients with rheumatoid arthritis. Arthritis Rheum. 2005, 52, 1999–2002. [Google Scholar] [CrossRef] [PubMed]
  170. Xu, Y.; Zai, Z.; Lu, Z.; Zhang, T.; Wang, L.; Qian, X.; Tao, J.; Peng, X.; Zhang, Y.; Chen, F. Circular RNA CircCDKN2B-AS_006 Promotes the Tumor-like Growth and Metastasis of Rheumatoid Arthritis Synovial Fibroblasts by Targeting the miR-1258/RUNX1 Axis. Int. J. Mol. Sci. 2023, 24, 5880. [Google Scholar] [CrossRef]
  171. Steenvoorden, M.M.; Tolboom, T.C.; van der Pluijm, G.; Lowik, C.; Visser, C.P.; DeGroot, J.; Gittenberger-DeGroot, A.C.; DeRuiter, M.C.; Wisse, B.J.; Huizinga, T.W.; et al. Transition of healthy to diseased synovial tissue in rheumatoid arthritis is associated with gain of mesenchymal/fibrotic characteristics. Arthritis Res. Ther. 2006, 8, R165. [Google Scholar] [CrossRef]
  172. Ekwall, A.K.; Eisler, T.; Anderberg, C.; Jin, C.; Karlsson, N.; Brisslert, M.; Bokarewa, M.I. The tumour-associated glycoprotein podoplanin is expressed in fibroblast-like synoviocytes of the hyperplastic synovial lining layer in rheumatoid arthritis. Arthritis Res. Ther. 2011, 13, R40. [Google Scholar] [CrossRef]
  173. Quintanilla, M.; Montero-Montero, L.; Renart, J.; Martin-Villar, E. Podoplanin in Inflammation and Cancer. Int. J. Mol. Sci. 2019, 20, 707. [Google Scholar] [CrossRef] [PubMed]
  174. Hitchon, C.A.; El-Gabalawy, H.S.; Bezabeh, T. Characterization of synovial tissue from arthritis patients: A proton magnetic resonance spectroscopic investigation. Rheumatol. Int. 2009, 29, 1205–1211. [Google Scholar] [CrossRef]
  175. Li, G.; Zhang, Y.; Qian, Y.; Zhang, H.; Guo, S.; Sunagawa, M.; Hisamitsu, T.; Liu, Y. Interleukin-17A promotes rheumatoid arthritis synoviocytes migration and invasion under hypoxia by increasing MMP2 and MMP9 expression through NF-kappaB/HIF-1alpha pathway. Mol. Immunol. 2013, 53, 227–236. [Google Scholar] [CrossRef]
  176. Chen, J.; Cheng, W.; Li, J.; Wang, Y.; Chen, J.; Shen, X.; Su, A.; Gan, D.; Ke, L.; Liu, G.; et al. Notch-1 and Notch-3 Mediate Hypoxia-Induced Activation of Synovial Fibroblasts in Rheumatoid Arthritis. Arthritis Rheumatol. 2021, 73, 1810–1819. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, X.; Guan, X.; Piao, Y.; Che, X.; Si, M.; Jin, J. Baicalein Induces Apoptosis of Rheumatoid Arthritis Synovial Fibroblasts through Inactivation of the PI3K/Akt/mTOR Pathway. Evid. Based Complement. Alter. Nat. Med. 2022, 2022, 3643265. [Google Scholar] [CrossRef] [PubMed]
  178. Lauzier, A.; Lavoie, R.R.; Charbonneau, M.; Gouin-Boisvert, B.; Harper, K.; Dubois, C.M. Snail Is a Critical Mediator of Invadosome Formation and Joint Degradation in Arthritis. Am. J. Pathol. 2016, 186, 359–374. [Google Scholar] [CrossRef]
  179. Parisis, D.; Chivasso, C.; Perret, J.; Soyfoo, M.S.; Delporte, C. Current State of Knowledge on Primary Sjogren’s Syndrome, an Autoimmune Exocrinopathy. J. Clin. Med. 2020, 9, 2299. [Google Scholar] [CrossRef]
  180. Mariette, X.; Criswell, L.A. Primary Sjogren’s Syndrome. N. Engl. J. Med. 2018, 378, 931–939. [Google Scholar] [CrossRef]
  181. Kiripolsky, J.; McCabe, L.G.; Kramer, J.M. Innate immunity in Sjogren’s syndrome. Clin. Immunol. 2017, 182, 4–13. [Google Scholar] [CrossRef] [PubMed]
  182. Deshmukh, U.S.; Nandula, S.R.; Thimmalapura, P.R.; Scindia, Y.M.; Bagavant, H. Activation of innate immune responses through Toll-like receptor 3 causes a rapid loss of salivary gland function. J. Oral Pathol. Med. 2009, 38, 42–47. [Google Scholar] [CrossRef] [PubMed]
  183. Kawakami, A.; Nakashima, K.; Tamai, M.; Nakamura, H.; Iwanaga, N.; Fujikawa, K.; Aramaki, T.; Arima, K.; Iwamoto, N.; Ichinose, K.; et al. Toll-like receptor in salivary glands from patients with Sjogren’s syndrome: Functional analysis by human salivary gland cell line. J. Rheumatol. 2007, 34, 1019–1026. [Google Scholar] [PubMed]
  184. Spachidou, M.P.; Bourazopoulou, E.; Maratheftis, C.I.; Kapsogeorgou, E.K.; Moutsopoulos, H.M.; Tzioufas, A.G.; Manoussakis, M.N. Expression of functional Toll-like receptors by salivary gland epithelial cells: Increased mRNA expression in cells derived from patients with primary Sjogren’s syndrome. Clin. Exp. Immunol. 2007, 147, 497–503. [Google Scholar] [CrossRef]
  185. Zheng, L.; Zhang, Z.; Yu, C.; Yang, C. Expression of Toll-like receptors 7, 8, and 9 in primary Sjogren’s syndrome. Oral Surg. Oral Med. Oral Pathol. Oral Radiol. Endod. 2010, 109, 844–850. [Google Scholar] [CrossRef] [PubMed]
  186. Brito-Zeron, P.; Baldini, C.; Bootsma, H.; Bowman, S.J.; Jonsson, R.; Mariette, X.; Sivils, K.; Theander, E.; Tzioufas, A.; Ramos-Casals, M. Sjogren syndrome. Nat. Rev. Dis. Primers 2016, 2, 16047. [Google Scholar] [CrossRef] [PubMed]
  187. Koski, H.; Konttinen, Y.T.; Gu, X.H.; Hietanen, J.; Malmstrom, M. Transforming growth factor beta 2 in labial salivary glands in Sjogren’s syndrome. Ann. Rheum. Dis. 1995, 54, 744–747. [Google Scholar] [CrossRef] [PubMed]
  188. Skopouli, F.N.; Li, L.; Boumba, D.; Stefanaki, S.; Hanel, K.; Moutsopoulos, H.M.; Krilis, S.A. Association of mast cells with fibrosis and fatty infiltration in the minor salivary glands of patients with Sjogren’s syndrome. Clin. Exp. Rheumatol. 1998, 16, 63–65. [Google Scholar]
  189. Leehan, K.M.; Pezant, N.P.; Rasmussen, A.; Grundahl, K.; Moore, J.S.; Radfar, L.; Lewis, D.M.; Stone, D.U.; Lessard, C.J.; Rhodus, N.L.; et al. Minor salivary gland fibrosis in Sjogren’s syndrome is elevated, associated with focus score and not solely a consequence of aging. Clin. Exp. Rheumatol. 2018, 36 (Suppl. 112), 80–88. [Google Scholar]
  190. Zavadil, J.; Bottinger, E.P. TGF-beta and epithelial-to-mesenchymal transitions. Oncogene 2005, 24, 5764–5774. [Google Scholar] [CrossRef]
  191. Sisto, M.; Tamma, R.; Ribatti, D.; Lisi, S. IL-6 Contributes to the TGF-beta1-Mediated Epithelial to Mesenchymal Transition in Human Salivary Gland Epithelial Cells. Arch. Immunol. Ther. Exp. 2020, 68, 27. [Google Scholar] [CrossRef] [PubMed]
  192. Sisto, M.; Lorusso, L.; Tamma, R.; Ingravallo, G.; Ribatti, D.; Lisi, S. Interleukin-17 and -22 synergy linking inflammation and EMT-dependent fibrosis in Sjogren’s syndrome. Clin. Exp. Immunol. 2019, 198, 261–272. [Google Scholar] [CrossRef]
  193. Yazisiz, V.; Behlul, A.; Basak, S.T.; Borlu, F. Expression of podoplanin in minor salivary glands increases in primary Sjogren syndrome. J. Rheumatol. 2013, 40, 2100–2102. [Google Scholar] [CrossRef] [PubMed]
  194. Sisto, M.; Lorusso, L.; Ingravallo, G.; Ribatti, D.; Lisi, S. TGFbeta1-Smad canonical and -Erk noncanonical pathways participate in interleukin-17-induced epithelial-mesenchymal transition in Sjogren’s syndrome. Lab. Investig. 2020, 100, 824–836. [Google Scholar] [CrossRef] [PubMed]
  195. Noll, B.; Mougeot, F.B.; Brennan, M.T.; Mougeot, J.C. Regulation of MMP9 transcription by ETS1 in immortalized salivary gland epithelial cells of patients with salivary hypofunction and primary Sjogren’s syndrome. Sci. Rep. 2022, 12, 14552. [Google Scholar] [CrossRef] [PubMed]
  196. Riviere, E.; Chivasso, C.; Pascaud, J.; Bechara, R.; Ly, B.; Delporte, C.; Mariette, X.; Nocturne, G. Hyperosmolar environment and salivary gland epithelial cells increase extra-cellular matrix remodeling and lymphocytic infiltration in Sjogren’s syndrome. Clin. Exp. Immunol. 2023, 212, 39–51. [Google Scholar] [CrossRef]
  197. Dorner, T.; Furie, R. Novel paradigms in systemic lupus erythematosus. Lancet 2019, 393, 2344–2358. [Google Scholar] [CrossRef] [PubMed]
  198. Rahman, A.; Isenberg, D.A. Systemic lupus erythematosus. N. Engl. J. Med. 2008, 358, 929–939. [Google Scholar] [CrossRef]
  199. Kidney Disease: Improving Global Outcomes Glomerular Diseases Work, G. KDIGO 2021 Clinical Practice Guideline for the Management of Glomerular Diseases. Kidney Int. 2021, 100, S1–S276. [CrossRef]
  200. Pamfil, C.; Makowska, Z.; De Groof, A.; Tilman, G.; Babaei, S.; Galant, C.; Montigny, P.; Demoulin, N.; Jadoul, M.; Aydin, S.; et al. Intrarenal activation of adaptive immune effectors is associated with tubular damage and impaired renal function in lupus nephritis. Ann. Rheum. Dis. 2018, 77, 1782–1789. [Google Scholar] [CrossRef]
  201. Zhang, C.; Chan, C.C.Y.; Cheung, K.F.; Chau, M.K.M.; Yap, D.Y.H.; Ma, M.K.M.; Chan, K.W.; Yung, S.; Chan, T.M. Effect of mycophenolate and rapamycin on renal fibrosis in lupus nephritis. Clin. Sci. 2019, 133, 1721–1744. [Google Scholar] [CrossRef]
  202. Hills, C.E.; Squires, P.E. The role of TGF-beta and epithelial-to mesenchymal transition in diabetic nephropathy. Cytokine Growth Factor Rev. 2011, 22, 131–139. [Google Scholar] [CrossRef] [PubMed]
  203. LeBleu, V.S.; Taduri, G.; O’Connell, J.; Teng, Y.; Cooke, V.G.; Woda, C.; Sugimoto, H.; Kalluri, R. Origin and function of myofibroblasts in kidney fibrosis. Nat. Med. 2013, 19, 1047–1053. [Google Scholar] [CrossRef]
  204. Liu, Y. Epithelial to mesenchymal transition in renal fibrogenesis: Pathologic significance, molecular mechanism, and therapeutic intervention. J. Am. Soc. Nephrol. 2004, 15, 1–12. [Google Scholar] [CrossRef]
  205. Kizu, A.; Medici, D.; Kalluri, R. Endothelial-mesenchymal transition as a novel mechanism for generating myofibroblasts during diabetic nephropathy. Am. J. Pathol. 2009, 175, 1371–1373. [Google Scholar] [CrossRef]
  206. Loeffler, I.; Liebisch, M.; Wolf, G. Collagen VIII influences epithelial phenotypic changes in experimental diabetic nephropathy. Am. J. Physiol. Renal. Physiol. 2012, 303, F733–F745. [Google Scholar] [CrossRef] [PubMed]
  207. Yamamoto, T.; Noble, N.A.; Cohen, A.H.; Nast, C.C.; Hishida, A.; Gold, L.I.; Border, W.A. Expression of transforming growth factor-beta isoforms in human glomerular diseases. Kidney Int. 1996, 49, 461–469. [Google Scholar] [CrossRef]
  208. Rocha, L.P.; Xavier, S.C.; Helmo, F.R.; Machado, J.R.; Ramalho, F.S.; Dos Reis, M.A.; Correa, R.R. Epithelial-mesenchymal transition in pediatric nephropathies. Pathol. Res. Pract. 2016, 212, 1157–1166. [Google Scholar] [CrossRef]
  209. Yung, S.; Zhang, Q.; Zhang, C.Z.; Chan, K.W.; Lui, S.L.; Chan, T.M. Anti-DNA antibody induction of protein kinase C phosphorylation and fibronectin synthesis in human and murine lupus and the effect of mycophenolic acid. Arthritis Rheum. 2009, 60, 2071–2082. [Google Scholar] [CrossRef] [PubMed]
  210. Liu, L.; Li, F.G.; Yang, M.; Wang, L.; Chen, Y.; Wang, L.; Ji, W.; Fan, J.M. Effect of pro-inflammatory interleukin-17A on epithelial cell phenotype inversion in HK-2 cells in vitro. Eur. Cytokine Netw. 2016, 27, 27–33. [Google Scholar] [CrossRef]
  211. Dudas, P.L.; Sague, S.L.; Elloso, M.M.; Farrell, F.X. Proinflammatory/profibrotic effects of interleukin-17A on human proximal tubule epithelium. Nephron. Exp. Nephrol. 2011, 117, e114–e123. [Google Scholar] [CrossRef]
  212. Wang, W.; Mou, S.; Wang, L.; Zhang, M.; Shao, X.; Fang, W.; Lu, R.; Qi, C.; Fan, Z.; Cao, Q.; et al. Up-regulation of Serum MiR-130b-3p Level is Associated with Renal Damage in Early Lupus Nephritis. Sci. Rep. 2015, 5, 12644. [Google Scholar] [CrossRef] [PubMed]
  213. Lin, H.; Liu, J.; Li, N.; Zhang, B.; Nguyen, V.D.; Yao, P.; Feng, J.; Liu, Q.; Chen, Y.; Li, G.; et al. NETosis promotes chronic inflammation and fibrosis in systemic lupus erythematosus and COVID-19. Clin. Immunol. 2023, 254, 109687. [Google Scholar] [CrossRef]
  214. Pieterse, E.; Rother, N.; Garsen, M.; Hofstra, J.M.; Satchell, S.C.; Hoffmann, M.; Loeven, M.A.; Knaapen, H.K.; van der Heijden, O.W.H.; Berden, J.H.M.; et al. Neutrophil Extracellular Traps Drive Endothelial-to-Mesenchymal Transition. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 1371–1379. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, Q.; Du, Y.; Li, K.; Zhang, W.; Feng, X.; Hao, J.; Li, H.; Liu, S. Anti-OSM Antibody Inhibits Tubulointerstitial Lesion in a Murine Model of Lupus Nephritis. Mediat. Inflamm. 2017, 2017, 3038514. [Google Scholar] [CrossRef]
  216. Fu, D.; Senouthai, S.; Wang, J.; You, Y. FKN Facilitates HK-2 Cell EMT and Tubulointerstitial Lesions via the Wnt/beta-Catenin Pathway in a Murine Model of Lupus Nephritis. Front. Immunol. 2019, 10, 784. [Google Scholar] [CrossRef] [PubMed]
  217. Zhang, F.; Zhou, X.; Cui, D.; Zhang, W.; Lai, J.; Li, X.; Ruan, Y.; Xie, Y.; Shi, M.; Xiao, Y.; et al. The role of Stim1 in the progression of lupus nephritis in mice. Int. J. Clin. Exp. Pathol. 2020, 13, 3021–3032. [Google Scholar]
  218. Xie, Y.; Ruan, Y.; Zou, H.; Wang, Y.; Wu, X.; Li, X.; Lai, J.; Shi, M.; Xiao, Y.; Wang, Y.; et al. YAP1 Overexpression Is Associated with Kidney Dysfunction in Lupus Nephritis. Pathobiology 2021, 88, 412–423. [Google Scholar] [CrossRef]
  219. Xue, L.; Xu, J.; Lu, W.; Fu, J.; Liu, Z. Iguratimod alleviates tubulo-interstitial injury in mice with lupus. Ren. Fail. 2022, 44, 636–647. [Google Scholar] [CrossRef]
  220. Wu, S.; Ji, L.; Fan, X.; Fang, S.; Bao, J.; Yuan, X.; Fan, Y.; Xie, G. Jieduquyuzishen Prescription Attenuates Renal Fibrosis in MRL/lpr Mice via Inhibiting EMT and TGF-beta1/Smad2/3 Pathway. Evid. Based Complement. Alter. Nat. Med. 2022, 2022, 4987323. [Google Scholar] [CrossRef]
  221. Varga, J.; Abraham, D. Systemic sclerosis: A prototypic multisystem fibrotic disorder. J. Clin. Investig. 2007, 117, 557–567. [Google Scholar] [CrossRef]
  222. Desbois, A.C.; Cacoub, P. Systemic sclerosis: An update in 2016. Autoimmun. Rev. 2016, 15, 417–426. [Google Scholar] [CrossRef]
  223. Nihtyanova, S.I.; Schreiber, B.E.; Ong, V.H.; Rosenberg, D.; Moinzadeh, P.; Coghlan, J.G.; Wells, A.U.; Denton, C.P. Prediction of pulmonary complications and long-term survival in systemic sclerosis. Arthritis Rheumatol. 2014, 66, 1625–1635. [Google Scholar] [CrossRef]
  224. Kahaleh, M.B. Raynaud phenomenon and the vascular disease in scleroderma. Curr. Opin. Rheumatol. 2004, 16, 718–722. [Google Scholar] [CrossRef]
  225. Toledo, D.M.; Pioli, P.A. Macrophages in Systemic Sclerosis: Novel Insights and Therapeutic Implications. Curr. Rheumatol. Rep. 2019, 21, 31. [Google Scholar] [CrossRef]
  226. Wei, J.; Bhattacharyya, S.; Tourtellotte, W.G.; Varga, J. Fibrosis in systemic sclerosis: Emerging concepts and implications for targeted therapy. Autoimmun. Rev. 2011, 10, 267–275. [Google Scholar] [CrossRef]
  227. Cutolo, M.; Soldano, S.; Smith, V. Pathophysiology of systemic sclerosis: Current understanding and new insights. Expert Rev. Clin. Immunol. 2019, 15, 753–764. [Google Scholar] [CrossRef] [PubMed]
  228. Khalil, N.; Bereznay, O.; Sporn, M.; Greenberg, A.H. Macrophage production of transforming growth factor beta and fibroblast collagen synthesis in chronic pulmonary inflammation. J. Exp. Med. 1989, 170, 727–737. [Google Scholar] [CrossRef] [PubMed]
  229. van Caam, A.; Vonk, M.; van den Hoogen, F.; van Lent, P.; van der Kraan, P. Unraveling SSc Pathophysiology; The Myofibroblast. Front. Immunol. 2018, 9, 2452. [Google Scholar] [CrossRef]
  230. Nikitorowicz-Buniak, J.; Denton, C.P.; Abraham, D.; Stratton, R. Partially Evoked Epithelial-Mesenchymal Transition (EMT) Is Associated with Increased TGFbeta Signaling within Lesional Scleroderma Skin. PLoS ONE 2015, 10, e0134092. [Google Scholar] [CrossRef] [PubMed]
  231. O’Kane, D.; Jackson, M.V.; Kissenpfennig, A.; Spence, S.; Damkat-Thomas, L.; Tolland, J.P.; Smyth, A.E.; Denton, C.P.; Stuart Elborn, J.; McAuley, D.F.; et al. SMAD inhibition attenuates epithelial to mesenchymal transition by primary keratinocytes in vitro. Exp. Dermatol. 2014, 23, 497–503. [Google Scholar] [CrossRef]
  232. Takahashi, M.; Akamatsu, H.; Yagami, A.; Hasegawa, S.; Ohgo, S.; Abe, M.; Iwata, Y.; Arima, M.; Mizutani, H.; Nakata, S.; et al. Epithelial-mesenchymal transition of the eccrine glands is involved in skin fibrosis in morphea. J. Dermatol. 2013, 40, 720–725. [Google Scholar] [CrossRef]
  233. Nakamura, M.; Tokura, Y. Expression of SNAI1 and TWIST1 in the eccrine glands of patients with systemic sclerosis: Possible involvement of epithelial-mesenchymal transition in the pathogenesis. Br. J. Dermatol. 2011, 164, 204–205. [Google Scholar] [CrossRef]
  234. Goldmann, T.; Zissel, G.; Watz, H.; Dromann, D.; Reck, M.; Kugler, C.; Rabe, K.F.; Marwitz, S. Human alveolar epithelial cells type II are capable of TGFbeta-dependent epithelial-mesenchymal-transition and collagen-synthesis. Respir. Res. 2018, 19, 138. [Google Scholar] [CrossRef] [PubMed]
  235. Lee, D.W.; Lee, W.J.; Cho, J.; Yun, C.O.; Roh, H.; Chang, H.P.; Roh, T.S.; Lee, J.H.; Lew, D.H. Inhibition of Wnt signaling pathway suppresses radiation-induced dermal fibrosis. Sci. Rep. 2020, 10, 13594. [Google Scholar] [CrossRef]
  236. Zmorzynski, S.; Styk, W.; Filip, A.A.; Krasowska, D. The Significance of NOTCH Pathway in the Development of Fibrosis in Systemic Sclerosis. Ann. Dermatol. 2019, 31, 365–371. [Google Scholar] [CrossRef]
  237. Lam, A.P.; Flozak, A.S.; Russell, S.; Wei, J.; Jain, M.; Mutlu, G.M.; Budinger, G.R.; Feghali-Bostwick, C.A.; Varga, J.; Gottardi, C.J. Nuclear beta-catenin is increased in systemic sclerosis pulmonary fibrosis and promotes lung fibroblast migration and proliferation. Am. J. Respir. Cell Mol. Biol. 2011, 45, 915–922. [Google Scholar] [CrossRef] [PubMed]
  238. Yeh, M.; Oh, C.S.; Yoo, J.Y.; Kaur, B.; Lee, T.J. Pivotal role of microRNA-138 in human cancers. Am. J. Cancer Res. 2019, 9, 1118–1126. [Google Scholar] [PubMed]
  239. Bayati, P.; Poormoghim, H.; Mojtabavi, N. Aberrant expression of miR-138 as a novel diagnostic biomarker in systemic sclerosis. Biomark. Insights 2022, 17, 11772719221135442. [Google Scholar] [CrossRef]
  240. Mao, J.; Liu, J.; Zhou, M.; Wang, G.; Xiong, X.; Deng, Y. Hypoxia-induced interstitial transformation of microvascular endothelial cells by mediating HIF-1alpha/VEGF signaling in systemic sclerosis. PLoS ONE 2022, 17, e0263369. [Google Scholar] [CrossRef]
  241. Cipriani, P.; Di Benedetto, P.; Ruscitti, P.; Capece, D.; Zazzeroni, F.; Liakouli, V.; Pantano, I.; Berardicurti, O.; Carubbi, F.; Pecetti, G.; et al. The Endothelial-mesenchymal Transition in Systemic Sclerosis Is Induced by Endothelin-1 and Transforming Growth Factor-beta and May Be Blocked by Macitentan, a Dual Endothelin-1 Receptor Antagonist. J. Rheumatol. 2015, 42, 1808–1816. [Google Scholar] [CrossRef]
  242. Rosa, I.; Romano, E.; Fioretto, B.S.; El Aoufy, K.; Bellando-Randone, S.; Matucci-Cerinic, M.; Manetti, M. Lymphatic Endothelial-to-Myofibroblast Transition: A Potential New Mechanism Underlying Skin Fibrosis in Systemic Sclerosis. Cells 2023, 12, 2195. [Google Scholar] [CrossRef]
  243. Chiu, Y.H.; Spierings, J.; van Laar, J.M.; de Vries-Bouwstra, J.K.; van Dijk, M.; Goldschmeding, R. Association of endothelial to mesenchymal transition and cellular senescence with fibrosis in skin biopsies of systemic sclerosis patients: A cross-sectional study. Clin. Exp. Rheumatol. 2023, 41, 1612–1617. [Google Scholar] [CrossRef] [PubMed]
  244. Marden, G.; Wan, Q.; Wilks, J.; Nevin, K.; Feeney, M.; Wisniacki, N.; Trojanowski, M.; Bujor, A.; Stawski, L.; Trojanowska, M. The role of the oncostatin M/OSM receptor beta axis in activating dermal microvascular endothelial cells in systemic sclerosis. Arthritis Res. Ther. 2020, 22, 179. [Google Scholar] [CrossRef]
  245. Fernando, R.I.; Litzinger, M.; Trono, P.; Hamilton, D.H.; Schlom, J.; Palena, C. The T-box transcription factor Brachyury promotes epithelial-mesenchymal transition in human tumor cells. J. Clin. Investig. 2010, 120, 533–544. [Google Scholar] [CrossRef] [PubMed]
  246. Teng, L.; Yao, G.; Gao, R.; Wang, Y.; Lin, W.; Jin, Y.; Wang, Y.; Chen, H.; Wang, Z.; Chen, J. TGF-beta-mediated upregulation of brachyury contributes to constitutively up-regulated type I collagen expression in skin fibroblasts: Possible role in systemic sclerosis. Eur. J. Dermatol. 2019, 29, 359–365. [Google Scholar] [CrossRef] [PubMed]
  247. Abd-Elmawla, M.A.; Hassan, M.; Elsabagh, Y.A.; Alnaggar, A.; Senousy, M.A. Deregulation of long noncoding RNAs ANCR, TINCR, HOTTIP and SPRY4-IT1 in plasma of systemic sclerosis patients: SPRY4-IT1 as a novel biomarker of scleroderma and its subtypes. Cytokine 2020, 133, 155124. [Google Scholar] [CrossRef]
  248. Huang, X.L.; Zhang, L.; Duan, Y.; Wang, Y.J.; Zhao, J.H.; Wang, J. E3 ubiquitin ligase: A potential regulator in fibrosis and systemic sclerosis. Cell. Immunol. 2016, 306–307, 1–8. [Google Scholar] [CrossRef] [PubMed]
  249. Utsunomiya, A.; Chino, T.; Kasamatsu, H.; Hasegawa, T.; Utsunomiya, N.; Luong, V.H.; Matsushita, T.; Sasaki, Y.; Ogura, D.; Niwa, S.I.; et al. The compound LG283 inhibits bleomycin-induced skin fibrosis via antagonizing TGF-beta signaling. Arthritis Res. Ther. 2022, 24, 94. [Google Scholar] [CrossRef] [PubMed]
  250. Pei, B.; Zhang, N.; Pang, T.; Sun, G. Linagliptin ameliorates pulmonary fibrosis in systemic sclerosis mouse model via inhibition of endothelial-to-mesenchymal transition. Mol. Cell. Biochem. 2022, 477, 995–1007. [Google Scholar] [CrossRef]
Figure 1. Morphological modifications during EMT. Epithelial cells are closely held together by different types of cell–cell junctions (i.e., adherens junctions and tight junctions) and form polarized sheets harboring an apico-basal polarity. The loss of E-cadherin (a key component of adherens junctions) is often considered the first event initiating EMT and is rapidly followed by the repression of other cell junction proteins such as zona occludens 1 (ZO-1), occludin, and claudin. Consequently, the dissolution of epithelial cell junctions elicits a loss of apico-basal polarity and triggers cytoskeletal modifications leading to the acquisition of a spindle-shaped morphology. The cells are pushed towards the EMT spectrum in a progressive and reversible manner, from a complete epithelial to complete mesenchymal state. Abbreviations: α-SMA: α-smooth muscle actin; CDH1: E-cadherin; EMT: epithelial–mesenchymal transition; FSP-1: fibroblast specific protein 1; MET: mesenchymal–epithelial transition; SNAI1: Snail family transcriptional repressor 1; TFs: transcription factors; TWIST1: twist family BHLH transcription factor 1; ZEB1: zinc finger E-box binding homeobox 1; ZO-1: zona occludens 1.
Figure 1. Morphological modifications during EMT. Epithelial cells are closely held together by different types of cell–cell junctions (i.e., adherens junctions and tight junctions) and form polarized sheets harboring an apico-basal polarity. The loss of E-cadherin (a key component of adherens junctions) is often considered the first event initiating EMT and is rapidly followed by the repression of other cell junction proteins such as zona occludens 1 (ZO-1), occludin, and claudin. Consequently, the dissolution of epithelial cell junctions elicits a loss of apico-basal polarity and triggers cytoskeletal modifications leading to the acquisition of a spindle-shaped morphology. The cells are pushed towards the EMT spectrum in a progressive and reversible manner, from a complete epithelial to complete mesenchymal state. Abbreviations: α-SMA: α-smooth muscle actin; CDH1: E-cadherin; EMT: epithelial–mesenchymal transition; FSP-1: fibroblast specific protein 1; MET: mesenchymal–epithelial transition; SNAI1: Snail family transcriptional repressor 1; TFs: transcription factors; TWIST1: twist family BHLH transcription factor 1; ZEB1: zinc finger E-box binding homeobox 1; ZO-1: zona occludens 1.
Ijms 24 14481 g001
Figure 2. Signaling pathways regulating EMT. The encounter between specific ligands and epithelial cells results in the activation of several intracellular signaling pathways that ultimately lead to the expression of core EMT-TFs such as ZEB1/2, SNAI1/2, and TWIST1 that act pleiotropically to induce EMT. These pathways are not well compartmentalized, and some overlap exists to a certain extent between them. The canonical WNT pathway is activated upon the binding of WNT ligands to the Frizzled family of membrane receptors, ensuing the release of β-catenin from the GSK3β–AXIN-APC complex. This enables its translocation to the nucleus, thereby activating TCF and LEF-1 TFs, which promotes the expression of EMT-associated genes. The NOTCH pathway is activated upon binding of DDL or JAG ligands to the NOTCH receptor, leading to the release of the active NOTCH-ICD, which translocates to the nucleus to function as a transcriptional co-activator. The Hedgehog signaling pathway is activated by members of the Hedgehog (HH) ligand family that bind to PTCH receptor and activate SMO, which recruits GLI TFs, further entailing the transcription of EMT target genes. TGFβ activates the TGFβ family of receptors that trigger the phosphorylation and activation of cytoplasmic SMAD2 and SMAD3, which further assemble with SMAD4 to migrate to the nucleus, where they activate the transcription EMT-TFs. SMAD also interacts with β-catenin and NOTCH-ICD, enabling cross-talk between the TGF-β, WNT, and NOTCH pathways. The TGFβ pathway also collaborates with the PI3K–AKT pathway, which in turn triggers the activation of the mTOR and NF-κB and the RAS–RAF–MEK–ERK signaling axis. These pathways are also triggered by the binding of several growth factors to their cognate receptors. The binding of several cytokines to their receptors triggers the phosphorylation and activation of JAKs and STATs. STAT dimers activate the transcription of genes encoding core EMT-TFs. Abbreviations: AKT: protein kinase B; AXIN: axis inhibition protein; DHH: desert hh; DLL: Delta-like; EGF: epidermal growth factor; EMT: epithelial–mesenchymal transition; ERK: extracellular signal-regulated kinase; FGF: fibroblast growth factor; GSK-3β: glycogen synthase kinase-3β; HGF: hepatocyte growth factor; IHH: Indian Hedgehog; JAG: jagged; JAK: Janus kinase; LEF-1: lymphoid enhancer binding factor 1; MAPK: mitogen-activated protein kinase; MEK: MAP kinase; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor-κB; NOTCH-ICD: NOTCH receptor intracellular domain; PI3K: phosphoinositide 3-kinase; PTCH: patched; RTKs: receptor tyrosine kinases; SHH: sonic hh; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SMO: smoothened; SNAI1: snail family transcriptional repressor 1; STAT: signal transducers and activators of transcription; TβRI: TGFβ receptor type I; TβRII: TGFβ receptor type II; TCF: T cell factor; TFs: transcription factors; TGF-β: transforming growth factor-β; TWIST1: twist family BHLH transcription factor 1; VEGF: vascular endothelial growth factor; ZEB1: zinc finger E-box binding homeobox 1.Created with Biorender.com (accessed on 20 August 2023).
Figure 2. Signaling pathways regulating EMT. The encounter between specific ligands and epithelial cells results in the activation of several intracellular signaling pathways that ultimately lead to the expression of core EMT-TFs such as ZEB1/2, SNAI1/2, and TWIST1 that act pleiotropically to induce EMT. These pathways are not well compartmentalized, and some overlap exists to a certain extent between them. The canonical WNT pathway is activated upon the binding of WNT ligands to the Frizzled family of membrane receptors, ensuing the release of β-catenin from the GSK3β–AXIN-APC complex. This enables its translocation to the nucleus, thereby activating TCF and LEF-1 TFs, which promotes the expression of EMT-associated genes. The NOTCH pathway is activated upon binding of DDL or JAG ligands to the NOTCH receptor, leading to the release of the active NOTCH-ICD, which translocates to the nucleus to function as a transcriptional co-activator. The Hedgehog signaling pathway is activated by members of the Hedgehog (HH) ligand family that bind to PTCH receptor and activate SMO, which recruits GLI TFs, further entailing the transcription of EMT target genes. TGFβ activates the TGFβ family of receptors that trigger the phosphorylation and activation of cytoplasmic SMAD2 and SMAD3, which further assemble with SMAD4 to migrate to the nucleus, where they activate the transcription EMT-TFs. SMAD also interacts with β-catenin and NOTCH-ICD, enabling cross-talk between the TGF-β, WNT, and NOTCH pathways. The TGFβ pathway also collaborates with the PI3K–AKT pathway, which in turn triggers the activation of the mTOR and NF-κB and the RAS–RAF–MEK–ERK signaling axis. These pathways are also triggered by the binding of several growth factors to their cognate receptors. The binding of several cytokines to their receptors triggers the phosphorylation and activation of JAKs and STATs. STAT dimers activate the transcription of genes encoding core EMT-TFs. Abbreviations: AKT: protein kinase B; AXIN: axis inhibition protein; DHH: desert hh; DLL: Delta-like; EGF: epidermal growth factor; EMT: epithelial–mesenchymal transition; ERK: extracellular signal-regulated kinase; FGF: fibroblast growth factor; GSK-3β: glycogen synthase kinase-3β; HGF: hepatocyte growth factor; IHH: Indian Hedgehog; JAG: jagged; JAK: Janus kinase; LEF-1: lymphoid enhancer binding factor 1; MAPK: mitogen-activated protein kinase; MEK: MAP kinase; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor-κB; NOTCH-ICD: NOTCH receptor intracellular domain; PI3K: phosphoinositide 3-kinase; PTCH: patched; RTKs: receptor tyrosine kinases; SHH: sonic hh; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SMO: smoothened; SNAI1: snail family transcriptional repressor 1; STAT: signal transducers and activators of transcription; TβRI: TGFβ receptor type I; TβRII: TGFβ receptor type II; TCF: T cell factor; TFs: transcription factors; TGF-β: transforming growth factor-β; TWIST1: twist family BHLH transcription factor 1; VEGF: vascular endothelial growth factor; ZEB1: zinc finger E-box binding homeobox 1.Created with Biorender.com (accessed on 20 August 2023).
Ijms 24 14481 g002
Figure 3. Hypothetical contribution of EMT-like process to the pathophysiology of RA. Upon stimuli such as inflammation (IL-1β, TNF-α and IL-17) or hypoxia, several signaling pathways become activated in normal FLS present in the synovial lining. These EMT signaling pathways culminate with the production of several EMT-related TFs such as SNAI1, HIF-1α and Runx1 leading to the acquisition of more mesenchymal features. These activated FLS called RA-FLS gain migratory and invasive properties together with an increased resistance to apoptosis resulting in the generation of hyperplastic synovial lining. Abbreviations: AKT: protein kinase B; α-SMA: α-smooth muscle actin; CIA: collagen-induced arthritis; EMT: epithelial-mesenchymal transition; HIF1α: hypoxia-inducible factor 1α; IL: interleukin; MMPs: matrix metalloproteases; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor-κB; PI3K: phosphoinositide 3-kinase; RA: rheumatoid arthritis; RA-FLSs: RA fibroblast-like synoviocytes; RUNX1: runt-related transcription factor 1; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail family transcriptional repressor 1; TFs: transcription factors; TGF-β: transforming growth factor-β; TNF: tumor necrosis factor. Created with Biorender.com (accessed on 20 August 2023).
Figure 3. Hypothetical contribution of EMT-like process to the pathophysiology of RA. Upon stimuli such as inflammation (IL-1β, TNF-α and IL-17) or hypoxia, several signaling pathways become activated in normal FLS present in the synovial lining. These EMT signaling pathways culminate with the production of several EMT-related TFs such as SNAI1, HIF-1α and Runx1 leading to the acquisition of more mesenchymal features. These activated FLS called RA-FLS gain migratory and invasive properties together with an increased resistance to apoptosis resulting in the generation of hyperplastic synovial lining. Abbreviations: AKT: protein kinase B; α-SMA: α-smooth muscle actin; CIA: collagen-induced arthritis; EMT: epithelial-mesenchymal transition; HIF1α: hypoxia-inducible factor 1α; IL: interleukin; MMPs: matrix metalloproteases; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor-κB; PI3K: phosphoinositide 3-kinase; RA: rheumatoid arthritis; RA-FLSs: RA fibroblast-like synoviocytes; RUNX1: runt-related transcription factor 1; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail family transcriptional repressor 1; TFs: transcription factors; TGF-β: transforming growth factor-β; TNF: tumor necrosis factor. Created with Biorender.com (accessed on 20 August 2023).
Ijms 24 14481 g003
Figure 4. Hypothetical contribution of EMT to the pathophysiology of pSS. Upon stimuli such as inflammation (IL-6, IL-17 and IL-22), several signaling pathways become activated in normal SGEC present in salivary glands. These EMT signaling pathways culminate with the production of several EMT-related TFs such as SNAI1, ZEB1 and ETS1 leading to the acquisition of mesenchymal features. This results in salivary gland fibrosis with the appearance of spindle-shaped morphology of SGEC together with the expression of mesenchymal markers such as vimentin, podoplanin, MMP9 and several collagen types. Abbreviations: ADAMTS: A Disintegrin And Metalloproteinase with Thrombospondin Motifs; COL3A1: collagen type III Alpha 1 chain; EMT: epithelial-mesenchymal transition; ERK: extracellular signal–regulated kinase; IL: interleukin; iSGEC: immortalized SGEC; MMPs: matrix metalloproteases; SGEC: salivary gland epithelial cells; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail family transcriptional repressor; SS: Sjögren’s syndrome; TFs: transcription factors; TGF-β: transforming growth factor-β; ZEB2: zinc finger e-box binding homeobox 2. Created with Biorender.com (accessed on 20 August 2023).
Figure 4. Hypothetical contribution of EMT to the pathophysiology of pSS. Upon stimuli such as inflammation (IL-6, IL-17 and IL-22), several signaling pathways become activated in normal SGEC present in salivary glands. These EMT signaling pathways culminate with the production of several EMT-related TFs such as SNAI1, ZEB1 and ETS1 leading to the acquisition of mesenchymal features. This results in salivary gland fibrosis with the appearance of spindle-shaped morphology of SGEC together with the expression of mesenchymal markers such as vimentin, podoplanin, MMP9 and several collagen types. Abbreviations: ADAMTS: A Disintegrin And Metalloproteinase with Thrombospondin Motifs; COL3A1: collagen type III Alpha 1 chain; EMT: epithelial-mesenchymal transition; ERK: extracellular signal–regulated kinase; IL: interleukin; iSGEC: immortalized SGEC; MMPs: matrix metalloproteases; SGEC: salivary gland epithelial cells; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail family transcriptional repressor; SS: Sjögren’s syndrome; TFs: transcription factors; TGF-β: transforming growth factor-β; ZEB2: zinc finger e-box binding homeobox 2. Created with Biorender.com (accessed on 20 August 2023).
Ijms 24 14481 g004
Table 1. Recapitulation table of findings regarding EMT-like occurrence in RA.
Table 1. Recapitulation table of findings regarding EMT-like occurrence in RA.
Cellular TypeModelSignaling
Pathway
CytokineTFsPhenotypical FeatureRef.
RA-FLShuman/in vitro tumor-like [165,166,167,168,169]
RA-FLShuman/in vitro ⇧ α-SMA[31,171,172]
RA-FLShuman/in vitroTGF-β1IL-1β, TNF-α ⇧ podoplanin[172]
RA-FLS human/in vitroTGF-β1/SMAD2/3 SNAI1invasion[31]
RA-FLShuman/in vitroPI3K/AKT/HIF-1α HIF-1αinvasion[165]
RA-FLShuman/in vitroNF-κB/HIF-αIL-17 ⇧ MMP2, ⇧ MMP9, invasion[44,175]
RA-FLSin vitro SW982 human cell linePI3K/Akt/mTOR SNAI1⇧ vimentin, ⇧ MMP2, ⇧ MMP9, ⇩ E-cadherin[177]
RA-FLShuman/in vitro + CIA ratsNOTCH2/3 HIF-αinvasion, resistance to apoptosis[176]
RA-FLSCIA rats Pi3k-/Akt Snai1invasion[178]
RA-FLSCIA ratsWnt/β−catenin Runx1invasion[170]
Abbreviations: AKT: protein kinase B; α-SMA: α-smooth muscle actin; CIA: collagen-induced arthritis; EMT: epithelial-mesenchymal transition; HIF1α: hypoxia-inducible factor 1α; IL: interleukin; MMPs: matrix metalloproteases; mTOR: mammalian target of rapamycin; NF-κB: nuclear factor-κB; PI3K: phosphoinositide 3-kinase; RA: rheumatoid arthritis; RA-FLSs: RA fibroblast-like synoviocytes; Ref.: references; RUNX1: runt-related transcription factor 1; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail Family Transcriptional Repressor 1; TFs: transcription factors; TGF-β: transforming growth factor-β; TNF: tumor necrosis factor; ⇧: upregulated; ⇩: downregulated.
Table 2. Recapitulation table of findings regarding EMT occurrence in pSS.
Table 2. Recapitulation table of findings regarding EMT occurrence in pSS.
Cellular TypeModelSignaling
Pathway
CytokineTFsPhenotypical FeatureRef.
SGEChuman/in vitroTGF-β1/SMAD SNAI1spindle-shaped, ⇧ vimentin, ⇧ collagen type I, ⇩ E-cadherin[33]
SGEChuman/in vitroTGF-β1IL-6 spindle-shaped, ⇧ vimentin, ⇧ collagen type I, ⇩ E-cadherin[181]
SGEChuman/in vitro IL-17, IL-22 spindle-shaped, ⇧ vimentin, ⇧ collagen type I, ⇩ E-cadherin[182]
SGEChuman/in vitro podoplanin[193]
SGEChuman/in vitroTGF-β1/SMAD2/3
TGF-β1/ERK1/2
IL-17 spindle-shaped, ⇧ vimentin, ⇧ collagen type I, ⇩ E-cadherin[184]
SGEC/iSGEChuman/in vitro ETS1⇧ MMP9[195]
SGEChuman/in vitro ZEB2⇧ COL3A1, COL1A2 and ADAMTS[196]
Abbreviations: ADAMTS: A Disintegrin And Metalloproteinase with Thrombospondin Motifs; COL3A1: collagen type III Alpha 1 chain; EMT: epithelial-mesenchymal transition; ERK: extracellular signal–regulated kinase; IL: interleukin; iSGEC: immortalized SGEC; MMPs: matrix metalloproteases; Ref.: references; SGEC: salivary gland epithelial cells; SMAD: homolog of the Drosophila protein, mothers against decapentaplegic (Mad) and the Caenorhabditis elegans protein Sma; SNAI1: Snail Family Transcriptional Repressor 1; SS: Sjögren’s syndrome; TFs: transcription factors; TGF-β: transforming growth factor-β; ZEB2: Zinc Finger E-Box Binding Homeobox 2; ⇧: upregulated; ⇩: downregulated.
Table 3. Recapitulation table of findings regarding EMT or endoMT occurrence in SLE.
Table 3. Recapitulation table of findings regarding EMT or endoMT occurrence in SLE.
Cellular TypeModelSignaling PathwayCytokineTFsPhenotypical FeatureRef.
glomeruli/TEChuman LN biopsy TGF-β1/β3 [207,208]
mesangial cellshuman LN biopsy TGF-β1 ⇧ FN[197]
TEChuman/in vitro Th17/IL-17A ⇧ FN, ⇩ E-cadherin[210,211]
TEChuman/in vitro IL-17A, TGF-β1 ⇧ collagen I and III, ⇧ α-SMA ⇩ E-cadherin[210]
serumhuman LN TGF-β1 ⇧ miR-130b-3p, ⇧ α-SMA, ⇩ E-cadherin[212]
lung Human/lung biopsy SLENETs Twist1, Snai1⇧ α-SMA, ⇩ E-cadherin[213]
Glomeruli/EChuman LN biopsy/in vitro EC/ MRL/Lpr miceNETs/β-catenin/Snai1 Snai1⇧ α-SMA, ⇧ MMP9, ⇩ VE-cadherin[214]
TECMRL/Lpr miceJak/Stat/OSMOSMStat1/Stat3⇧ α-SMA, ⇧ FN, ⇩ E-cadherin[215]
TECMRL/lpr miceWnt/β-cateninFKNβ-catenin⇧ α-SMA, ⇧ vimentin, ⇧ Wnt-4[216]
TECMRL/lpr miceStim1 ⇧ FN, ⇧ urine protein, ⇧ serum creatinine[217]
TECMRL/lpr miceYap1 ⇧ FN, ⇧ renal injury[218]
TECMRL/lpr mice/in vitro TGF-β1 ⇧ FN, ⇩ E-cadherin[219]
TECMRL/lpr miceTGF-β1/Smad2/3 ⇧ α-SMA, ⇧ phosphorylated SMAD2/3, ⇩ E-cadherin[220]
Abbreviations: α-SMA: α-smooth muscle actin; EC: endothelial cells; EMT: epithelial–mesenchymal transition; endoMT: endothelial-mesenchymal transition; FN: fibronectin; IL: interleukin; JAK: janus kinase; LN: lupus nephritis; NET: neutrophil extracellular traps; OSM: oncostatin M; Ref.: references; SLE: systemic lupus erythematosus; SMAD: mothers against decapentaplegic homolog; SNAI1: Snail family transcriptional repressor 1; STAT: signal transducers and activators of transcription; TEC: tubular–epithelial cell; TFs: transcription factors; TGF-β: transforming growth factor-β; TWIST1: twist family BHLH transcription factor 1.; VE: vascular-endothelial; ⇧: upregulated; ⇩: downregulated.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sarrand, J.; Soyfoo, M.S. Involvement of Epithelial-Mesenchymal Transition (EMT) in Autoimmune Diseases. Int. J. Mol. Sci. 2023, 24, 14481. https://doi.org/10.3390/ijms241914481

AMA Style

Sarrand J, Soyfoo MS. Involvement of Epithelial-Mesenchymal Transition (EMT) in Autoimmune Diseases. International Journal of Molecular Sciences. 2023; 24(19):14481. https://doi.org/10.3390/ijms241914481

Chicago/Turabian Style

Sarrand, Julie, and Muhammad S. Soyfoo. 2023. "Involvement of Epithelial-Mesenchymal Transition (EMT) in Autoimmune Diseases" International Journal of Molecular Sciences 24, no. 19: 14481. https://doi.org/10.3390/ijms241914481

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

Article Metrics

Back to TopTop