EphA2 Signaling in tumors: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 3 by Mario Cioce.

The Eph receptors represent the largest group among Receptor Tyrosine kinase (RTK) families. The Eph/ephrin signaling axis plays center stage during development, and the deep perturbation of signaling consequent to its dysregulation in cancer reveals the multiplicity and complexity underlying its function. In the last decades, they have emerged as key players in solid tumors, including colorectal cancer (CRC). EphA2 is involved in tumor progression and resistance to therapy.

  • EphA2
  • EGFR
  • TKI
  • ephrins
  • CRC
  • CSCs
  • drug resistance
  • cetuximab
  • intra-tumor heterogeneity
  • inter-tumor heterogeneity

1. Introduction

1.1. General Structure of Eph Receptors and Ephrin Ligands

The EphA2 receptor belongs to the Eph (erythropoietin-producing human hepatocellular) superfamily, the largest among tyrosine kinase receptor families [1]. Eph receptors are classified into Eph-A and Eph-B subfamilies depending on their sequence homologies and binding affinity for their cognate ephrin ligands. Although Eph receptors preferentially bind ligands of the same class, cross-binding has been described [for a review, [2][3][4] (Figure 1). All Eph receptors contain an extracellular region, with a conserved N-terminal globular ligand-binding domain (LBD), a cysteine-rich domain which comprises a Sushi and an epidermal growth factor (EGF)-like domain and two fibronectin type-III repeats (FN1 and FN2). The intracellular region contains a juxtamembrane region (JM), a tyrosine kinase domain, a sterile alpha motif (SAM) domain, and a PDZ (Post-synaptic density protein-95, Drosophila disc large tumor suppressor (Dlg), Zona occludens-1) domain-binding motif that are responsible for the interaction with effector molecules. The receptor homo- and hetero-oligomerization involves the extracellular domains (LBD and cys-rich domain) [5][6]. The SAM domain is involved in receptor-receptor interactions, possibly aiding homo- or hetero-oligomerization. The ectodomain and the intracellular domain are linked by a transmembrane helix (TM) [for reviews, [2][3] (Figure 1). Ephrins ligands are divided into ephrin-A and ephrin-B subclasses [5][7]. Ephrin-A proteins (A1–A6) are anchored to the extracellular cell membrane via a glycosyl phosphatidylinositol (GPI) linkage that could be released to activate EphA receptors at distance [8]. Ephrin-B members (B1–B3) are transmembrane proteins containing a cytoplasmic domain with several conserved tyrosine residues and a terminal PDZ-binding motif allowing the interaction with proteins involved in cytoskeleton organization and cell adhesion (Figure 1). Thus, Eph-ephrin signaling is transduced either directly (in the case of ephrin-Bs) or by interaction with intracellular proteins (like Fyn) or other transmembrane proteins (like the neurotrophin receptor p75) (as for ephrin-As [9] (Figure 1).

Figure 1. The structure of Eph receptors and their ligands is shown. Eph receptors are consisting of an extracellular structure consisting of an ephrin binding domain connected to two fibronectin type-III repeats by a cysteine-rich EGF-like motif. The juxtamembrane region connects the extracellular portion of the receptor to the intracellular kinase domain that is linked to a sterile alpha motif (SAM) domain and PDZ-binding motif. Eph ligands (ephrin-A/B) are composed of a GPI-anchored receptor binding domain in the case of the ephrin-A type and a receptor-binding domain connected by a juxtamembrane domain to a cytoplasmic domain and a PDZ interaction motif, in the case of ephrin-B. Eph-Ephrin signaling is transduced either directly (in the case of ephrin-Bs) or by interaction with Fyn (as has been observed with ephrin-As). Ligand binding likely initiates clustering, aided by receptor-receptor interactions mediated by the SAM domain and by the PDZ (Post-synaptic density protein-95, Drosophila disc large tumor suppressor (Dlg), Zona occludens-1)-domain-binding motif. The formed complexes mediate bi-directional signaling called ephrin “reverse” and Eph “forward” signaling.

1.2. General Features of Eph-Ephrin Signaling

Since both Eph receptors and ephrins are anchored to the plasma membrane, the Eph-ephrin signaling is intrinsically bidirectional. The forward signaling is consequent to ligand binding and clustering of the receptors on the expressing cells. Trans-phosphorylation of clustered Eph receptors in the juxtamembrane domain enables efficient kinase activity [10][11]. Phosphorylation of the conserved tyrosine in the activation loop appears to be less critical for Eph receptor activation than it generally is for RTKs, mainly contributing to its maximal activity [10][12]. The reverse signaling takes place in the ligand expressing cells [13]. Both repulsive and attractive effects can be consequent to Eph-ephrin binding between cells: additionally, an initial cell-cell adhesion event mediated by the Eph-ephrin interaction may switch to a repulsive one in a time-dependent way, as the effect of cleavage of the membrane-bound ephrin or internalization of the receptor-ligand complex [14]. Ephrins can also attenuate forward signaling by Eph receptors co-expressed in the same cell [15] and also receptor delivery in extracellular vesicles to ligand expressing cancer cells has been shown to be functionally relevant, in cancer settings and in response to stress [16]. The Eph receptors also display non-catalytic functions. For instance, there are two pseudo-kinases (i.e., EphA10 and EphB6) within this large family and their function may be involved in tumorigenesis and resistance to therapy. Such a complexity supports a high adaptive potential, allowing for switching the Eph-ephrin signaling according to changes of both intracellular and extracellular stimuli.

2. EphA2 Signaling

2.1. EphA2 Signaling in Normal Cells

The EphA2 receptor is a 130-kDa transmembrane glycoprotein identified in the early 90′ in Hela cells during a screen for RTKs [17]. EphA2 potentially interacts with any ephrin-A ligand, with the most frequent partner being the membrane-bound, GPI-anchored ephrin-A1, this latter discovered in 1994 [18]. After ligand binding and trans-tyrosine phosphorylation, EphA2 forms a complex with c-Cbl, to be targeted to endosomes and degraded. About 35% of the receptor is recycled back to the plasma membrane [19]. The EphA2 forward signaling is executed through ligand-instigated binding of downstream adaptors and signaling partners [20]. In fact, as for other RTKs, phosphorylation of the tyrosine residues creates docking sites for SH2/SH3 containing-proteins such as Fyn, Src, Nck, Crk, RasGAP, LMW-PTP, PI3K, and the adapter proteins Grb2, Grb10, and SLAP. EphA2 modulates cytoskeletal organization through Rho/Rac GTPases [21]. Most of these proteins affect depolymerization of the actin cytoskeleton while others modulate cell adhesion [22] with important consequences on vascular assembly, angiogenesis, and cell migration [23]. The reverse signaling elicited by EphA2 (acting as a ligand) on the ephrin expressing cell is poorly characterized and known to be mediated, for ephrinAs, by the src family kinase Fyn. Reverse signaling can mediate cell adhesion or repulsion and modulates axon guidance and synaptogenesis in the developing brain [24]. Forward signaling by EphA2 has inhibitory effects on cell proliferation, through Ras/MAPK [25]. Erk inhibition takes place through activation of GAPs and/or inhibition of GEFs [2][26]. EphA2 kinase-dependent signaling thus suppresses the AKT–mTORC1 and RAS–ERK oncogenic pathways and inhibits cell adhesion and migration [27][28] (Figure 2A). For instance, ligand-bound EphA2 attenuated Erk activation in primary keratinocytes and hepatoma cells [29]; Ephrin-A/EphA signaling suppressed Erk activation induced by IGF-1 in myoblasts, facilitating myogenic differentiation [30]. In neurons, EphA-dependent Erk inhibition suppressed the effects of the TrkB RTK on growth cone motility [31][32].

Figure 2. EphA2 signaling in normal (A) and cancer (B) cells. (A) In untransformed cells, EphA2 is engaged by its ligands, mainly EphrinA1 and highly tyrosine-phosphorylated. This mediates cell adhesion/repulsion through activation of Rac/Rho GTPAses and RASGap. Ligand binding also mediates inhibition of MAPK and AKT. Upon ligand binding, the EphA2 is targeted to endosomes in a CBL-mediated process. (B) In cancer cells, unliganded and overexpressed EphA2 is mainly phosphorylated in ser897 by PI3k/AKT, ERK/RSK, PKA, and PKC, in response to oncogenic stimuli The Akt-mTORC1, Raf-MEK-ERK, and Pyk2-Src-ERK signaling pathways were identified as the downstream signaling of the EphA2 non-canonical pathway. S897-phosphorylated EphA2 recruits Ephexin4 that in turn acts on RhoG to promote cell migration and anoikis resistance (this latter effect through a RhoG-AKT pathway). Further, FAK-integrin mediates cell adhesion and migration and may promote CSC features, including drug resistance (please also see Figure 3). The phospho-tyrosine content of EphA2 is also reduced by the LMW-PTPase, frequently overexpressed in cancer. The pro-tumorigenic contribution of EphA2 may thus derive from ligand independency, overexpression, reduced phospho-tyrosine content, and increased serine/threonine phosphorylation. Additionally, ligand-stimulated EphA2 negatively modulates the recycling of EGFR, by inhibiting AKT/PIKfyve, thus reducing the amount of available EGFR on the plasma membrane and migration. On the other hand, such feedback is attenuated in transformed cells, where EGFR levels in the plasma membrane are increased and this correlates with ligand independency of EphA2 and activation of motile responses to EGF.

2.2. EphA2 in Tissue Patterning

We believe that a short journey into the role of EphA2 in tissue patterning and homeostasis may turn useful to better illustrate its involvement in tumor progression, apparently tumor-context specific but obeying signaling principles common to tissue patterning and repair. Eph receptors and ephrins have a key role in cell positioning, cell motility, cell differentiation, control of tissue morphogenesis and patterning, development of the vascular system (for a review, [2][33]). In fact, during tissue patterning, Eph receptors engagement by their ligands impedes cell mixing during tissue development and is essential to create functional topographic domains driving the formation of distinct cellular compartments [34][35][36]. For instance, forward signaling by EphA2 and its ligands aids in establishing synaptic connections in the developing nervous system by modulating growth cone guidance and axon branching [3]. Additionally, Ephrin-A2 reverse signaling inhibited the proliferation of neural progenitor cells, thus negatively modulating neurogenesis [37][38]. In developing mammary glands, EphA2 is important for promoting branching morphogenesis in vivo, as being expressed in mammary progenitor cells [39]. Ephs or ephrins may also cooperate with cell junctional modules (tight junctions and adherens junctions) to facilitate cell sorting processes and preserve the epithelial integrity and physiology in embryonal and adult tissues [40][41]. In normal colon epithelia, several studies have shown a decreasing gradient of EphB2 expression from the base to the top of the crypt, whereas EphA2 expression was observed in the differentiated compartment of the crypt apical columnar cells [42]. In fact, EphA2 is implicated in the repair of the gut epithelia [43] and of kidney epithelia during ischemia-reperfusion injury [44]. All these repair processes imply activation or reactivation of embryonal programs, like EMT or MET. Not coincidentally, as above mentioned, those programs are frequently reactivated in cancer cells [45][46].

3. Molecular Determinants of EphA2 Signaling in Tumors

In all the tumor settings studied, the role of EphA2, which ranges from a tumor-suppressive to a pro-tumorigenic one, depends on a number of intrinsic and extrinsic factors, some of which have been recently determined: its subcellular localization, the levels of expression, the presence of the ligand and the crosstalk with other receptors, such as the Epidermal Growth Factor Receptor (EGFR).

3.1. Intracellular Localization of EphA2

In non-neoplastic epithelia, EphA2 is localized to sites of cell-cell contact, in an E-cadherin-dependent way [47]. In absence of E-cadherin, associated with reduced cell-cell contacts and pro-metastatic behavior of the cancer cells, EphA2 was redistributed to membrane ruffles where it cannot engage with membrane-bound ligand ephrin-A1 on adjacent cells, thus reducing the tumor-suppressive juxtacrine signaling [40][47]. Also in cells lacking Claudin 4, another event associated with acquired pro-tumorigenic potential, EphA2 was increased and mislocalized and this correlated with increased oncogenic signaling [48].

3.2. Expression Levels of EphA2

EphA2 is highly expressed in many cancers with important prognostic implications. Elevated EphA2 expression positively correlated with poor prognosis, improved metastatic potential, and reduced overall survival of patients, in a tumor context-specific functioning (Table 1). Notably, the EphA2 is rarely mutated or amplified in cancer tissues [49][50]. However, expression of EphA2 may be modulated by p53, Ras and negatively modulated by estrogens [47][51][52]. In established cell line cultures, EphA2 expression was higher in cancer cells than in untransformed ones: increased staining intensity was observed, for example, in a large fraction of breast carcinoma cells (an average of 87%) when compared to benign mammary epithelial cells (an average of 3%) [53]. Related to this, overexpression of EphA2 was sufficient to transform immortalized mammary epithelial cells [54]. Additionally, EphA2 is present in GBM cells in a mainly non–tyrosine-phosphorylated state [55].

Table 1. Examples of EphA2 overexpression in human malignancies, with its significance and the number of cases analyzed.

Cancer Type

mRNA

Protein

Linked to

Cases (n)

Ref

Esophageal Squamous Cell Carcinoma

Protein

loco regional metastases; pathological grade; reduced OS

80

Miyazaki et al., 2002 [56]

Gastric Cancer

Protein

cancer recurrence (in association with YAP)

47

Huang et al., 2020 [57]

Prostate cancer

Protein

pathological grading

93

Zeng et al., 2003 [58]

Colorectal cancer

mRNA

protein

CSC markers (CD44 and Lgr5); reduced OS

338

Dunne et al., 2016 [59]

Colorectal cancer

mRNA

poor prognosis and response to cetuximab

226

Strimpakos et al., 2013 [60]

Colorectal cancer

mRNA

tumor progression and poor OS (EphA2 with miR-423-5p, CREB1, ADAMTS14)

1663 (TGCA)

De Robertis et al., 2018 [42]

Colorectal cancer

mRNA

worse PFS despite EGFRhigh (cetuximab-treated patients)

80 (TGCA)

De Robertis et al., 2017 [61]

Ovarian carcinoma

Protein

aggressive features and median survival

79

Thaker et al., 2004 [62]

Ovarian cancer

mRNA

protein

poor survival

118

Han et al., 2005 [63]

Epithelial Ovarian Cancer

 

poor survival (stronger when combined with p53null status)

79

Merritt et al., 2006 [64]

Endometrial cancer

Protein

higher pathological grade and clinical stage; shorter disease-specific survival (DSS)

139

Merritt et al., 2011 [65]

Cervical carcinoma

mRNA

decreased overall survival (OS)

206

Wu et al., 2004 [66]

Head and neck squamous cell carcinoma

mRNA

protein

higher clinical stage, recurrence, and lymph node metastasis; reduced disease-free survival (DFS) and OS

98

Liu et al., 2011 [67]

Glioblastoma

mRNA

protein

increased pathological grade; reduced OS

21

Liu et al., 2006 [68]

Malignant glioma

protein

decreased DFS and OS (oppositely to EphrinA1)

78

Li et al., 2010 [69]

Glioblastoma multiforme

protein

Reduced OS

40

Wang et al., 2008 [70]

Renal Cell Carcinoma

protein

increased pathological grade, reduced DFS and OS

34

Herrem et al., 2005 [71]

Renal Cell Carcinoma

protein

reduced OS

62

Xu et al., 2014 [72]

Non-Small-Cell-Lung-Cancer

protein

smoking history; reduced PFS and OS

279

Brannan et al., 2009 [73]

Non-Small-Cell-Lung-Cancer

protein

reduced overall survival (Stronger when associated with PKR)

218

Guo et al., 2013 [74]

Non-Small-Cell-Lung-Cancer

protein

brain metastases; reduced OS

270

Kinch et al., 2003 [75]

Hepatocellular carcinoma

mRNA

protein

higher pathological grade; and reduced OS

40

Cui et al., 2010 [76]

Hepatocellular carcinoma

protein

decreased OS

129

Yang et al., 2009 [77]

Gastric cancer

protein

higher in high-risk macroscopic grade 3 and 4 tumors

49

Nakamura et al., 2005 [78]

3.3. Ligand-Dependent EphA2 Signaling

A conspicuous amount of evidence suggests that ligand-mediated activation of EphA2 has tumor-suppressive functions. For instance, inverse expression of ephrin-A1 and EphA2 in human breast cancer cell lines was a frequent finding [79][80]. When tumors were grown in vivo, EphA2 appeared to be poorly activated by the endogenous ephrin-A [28]. Consistent with the previous observations, regulation of EphA2 expression in GBM by Fc-ephrin-A1 stimulation resulted in the loss of self-renewal ability and decreased proliferation in vitro and in vivo [81][82]. Ephrin-A1 ligand-induced EphA2 phosphorylation induces receptor endocytosis and the CBL ubiquitin-ligase mediated proteasome degradation [19][83]. Induction of ephrins may represent per se a mechanism for silencing Eph signaling. For example, during mouse ESC differentiation, FGF4 reduces EphA2 signaling, by transcriptionally inducing its ligands. This correlated with increased tyrosine phosphorylation and reduced Ser/Thr phosphorylation of EphA2 and reduced expression of pluripotency core factors, thereby leading to ESC differentiation [82]. Differently to other RTKs, activation of Eph receptors by ephrins does not increase cell proliferation or transform murine fibroblasts: conversely, it rather inhibited the Ras/MAPK and attenuated mitogen-activated protein kinase (MAPK) activation by platelet-derived growth factor (PDGF), epidermal growth factor (EGF) and vascular endothelial growth factor (VEGF), in a range of cell lines [84]. In cancer cells, including PTEN deficient prostate cancer cells and glioma cells, ephrin-dependent EphA2 activation led to rapid dephosphorylation of Akt at T308 and S473 residues leading in some cases to mTORC1 inactivation and decreased cell growth and migration [85][86][87] (Figure 2A). However, the effect of ephrin-A1 on EphA2 expressing cells may also be cell type-specific and transformation status-dependent: for instance, ephrin-A1 treatment inhibited proliferation of prostate cancer cells but failed to do so in fibroblasts [84]. Progranulin, a recently discovered EphA2 ligand, induced transient activation of MAPK in both untransformed HUVECs and transformed prostate cancer cells, but sustained activation of AKT was observed only in the latter cancer cells [88]. Altogether, this suggests that the dichotomic view (tumor suppression vs tumor promotion based on ligand availability) is too simple. Ephrin-driven forward signaling suppressed AKT activation in an ephrin-dependent way [85][86][87] ligand-bound EphA2 suppressed the recycling of EGFR to the plasma membrane, causing EGFR accumulation at the endosomes and thereby attenuating EGFR-induced cell migration. This happened in both Mouse Embryo Fibroblasts (MEFs) and in triple-negative breast cancer cells (MDA-MB-231) and was due to reduced PIKfyve activation in early endosomes following EphA2-mediated inhibition of AKT [89][90] (Figure 2A). In keeping with a tumor-suppressive role for ligand-bound EphA2, forward signaling elicited by ephrin-A ligands from normal cells on EphA2 expressing, RasV12 positive cells caused repulsion and segregation of the transformed cells [91].

3.4. Ligand-Independent Activation of EphA2

Low juxtacrine signaling and/or insufficient levels of ephrinA1 on cancer cells reduce EphA2 tyrosine phosphorylation [55] and this leads to attenuated internalization and degradation of EphA2 receptor, with a relative increase of EphA2 levels. Concomitantly, when the ephrinA1-mediated inhibition of AKT is removed, EGFR recycling to the plasma membrane is reduced and the EphA2 ligand-independent effect is switched on by phosphorylation on S897 (Figure 2B). Phosphorylation of the S897 residue (among the 25 ser/thr residues in EphA2) in the region linking the kinase domain with the SAM domain is the main target for “non-canonical” ephrin-independent and/or kinase-independent EphA2 signaling [86] and this activated multiple mechanisms, encompassing the downstream activation of Akt–mTORC1, Raf–MEK–ERK, and Pyk2–Src–ERK [92] (Figure 2B). Additionally, the association between EphA2 and FAK resulted in integrin-mediated adhesion, cell spreading, and migration [93] (Figure 2B). Further, unliganded, EphA2 destabilized adherent junctions via Rho-GTP activation, by inhibiting p190 RhoGAP (a Rho-GTP inhibitor) through activating the low molecular weight phospho-tyrosine phosphatase (LMW-PTP) [94]. LMW-PTP by itself may decrease the phospho-tyrosine content of EphA2 [95], possibly when activated by stress signals [96] (Figure 2B). LMW-PTP, overexpressed in many cancers, has overlapping functions with EphA2, including cell motility and resistance to therapy [97][98]. S897-phosphorylated EphA2 recruited Ephexin4 to promote cell migration and anoikis resistance via RhoG and Rac [99]. RhoG may also activate the PI3K/Akt signaling pathway to promote cell proliferation and survival independently of the activation of Rac [100][101]. RhoG-mediated activation of PI3K and Akt also suppressed anoikis [102]. Anoikis is an apoptotic modality induced by the detachment of adherent cells from the extracellular matrix and its suppression is a feature of metastatic cells [103] (Figure 2). Phosphorylation of the S897 residue in the region linking the kinase domain with the SAM domain is thus the main target for the “non-canonical” ephrin-independent and/or kinase-independent EphA2 signaling [86]. AKT, RSK, PKA, and PKC phosphorylated EphA2-S897, and this increased cell migration/invasion and metastasis and promoted cancer stem cell-like features [104][105][106] (Figure 2B). Structurally, unliganded EphA2 forms predominantly dimers rather than high-order oligomeric structures [107]. There is also evidence that the unliganded EphA2 receptor JM + kinase region may interact with phosphatidylinositol phosphates (PIPs), even if the physiological relevance of this remains to be addressed [108].

3.5. Tumor Context Modulates EphA2 Signaling

Besides these general mechanisms, EphA2 is endowed with tumor-context specific functions, described below. In gastric cancer cell lines, ligand-independent EphA2 activation upregulated N-cadherin and Snail, and the Wnt/β-catenin targets TCF4, Cyclin-D1, and c-Myc, thereby triggering epithelial-to-mesenchymal transition (EMT) [109]. EMT is a complex process during which tumor cells progressively acquire mesenchymal features (such as resistance to stress and acquisition of migratory ability and metabolic resilience). In detail, overexpressed EphA2 was shown to bind to Wnt-1 and to promote beta-catenin nuclear accumulation. This upregulated c-MYC that, in turn, promoted further EphA2 increase in a feed-forward manner, by binding to the EphA2 promoter [110]. As mentioned before, in breast cancer cells Ephexin4, a guanine nucleotide exchange factor (GEF) for RhoG, interacted with S897-phosphorylated EphA2 and mediated ephrin-independent cell migration, invasion, and resistance to anoikis (Figure 2B). In glioblastoma (GBM), stimulation of the cells with EGF induced MEK- and RSK-dependent EphA2 S897 phosphorylation [111]. Miao and coworkers found that EphA2 S897 phosphorylation was present mainly in grade IV human glioma specimens, in regions enriched for pS473-Akt signal and invasive cells [86]. S897 phosphorylation of EphA2 has also been involved in determining the aggressiveness of thyroid cancer cells and shown to be mediated by ERK1/2 activation downstream of oncogenes like RET (RET/PTC), KRAS (G12R), or BRAFV600E [112]. The same EphA2 residue is phosphorylated by ionizing radiation in a MEK/ERK/RSK-dependent manner, mediated by increased ROS, in multiple cancer cell lines [113].

Regarding the events downstream of EphA2 S897, the Akt–mammalian target of ra-pamycin complex 1(mTORC1), Raf–MEK–ERK, and Pyk2–Src–ERK pathways were shown to be downstream effectors of the S897 EphA2 pathway in cholangiocarcinoma cells [114]. In prostate cancer and GBM, EphA2 S897 expression induced amoeboid motil-ity, which correlated with the induction of stemness markers, increased clonogenic poten-tial and tumour growth [81][115][116]. The EphA2 S897 increased in glucose starvation conditions in GBM cells and this correlated with cell survival and ROS-mediated ERK-RSK activation, induced by the cystine/glutamate antiporter xCT [117]. Thus, the S897 phosphorylation of EphA2 may work as a stress rheostat, transducing adaptive responses and thereby influencing tumor progression.

Notably, the “simple” abrogation of tyrosine phosphorylation in EphA2 may repre-sent “per se” an oncogenic signal. For instance, reintroduction of pY772A EphA2 in EphA2 knock-down naso-pharyngeal-carcinoma (NPC) cells increased cell proliferation, anchorage-independent growth in vitro and tumor growth in vivo. Mechanistically, EphA2-Y772A triggered activation (rather than inhibition) of Shp2/Erk-1/2 signaling pathway in the NPC cells, the latter involving binding of GAB1 and GRB2 as well [118]. In support of this, expression of kinase-deficient variants of EphA2 in breast cancer cells led to decreased tumor volume and increased tumor cell apoptosis [119].

A number of EphA2 mutations interfering with ephrin binding or kinase activity in cancer tissues such as intrahepatic cholangiocarcinoma (ICC) is being growingly recognized [120]. For instance, an EphA2 A859D Y772 dead mutant, exhibiting lower levels of phosphorylated Y772 and suppressed degradation through CBL, was recently identified in squamous cell carcinoma (SSC) and malignant pleural mesothelioma (MPM) speci-mens [121]. Contrariwise, tyrosine kinase activity of overexpressed EphA2 was also shown as required for the S897 phosphorylation via ERK to stimulate GBM cell prolifera-tion [111], thus showing the limit of a dichotomist view of canonical vs non-canonical EphA2 signaling and suggesting a more complex scenarios where both signaling modalities are highly interconnected [see also [27]]. For example, in Ewing sarcoma (ES) EphA2 promoted angiogenesis via ligand- (and caveolin-1)-dependent signaling [122], while enhancing tumorigenicity, migration and invasion in vitro and in vivo, in an S897 dependent manner [123].

EphA2 promotes resistance to a broad range of anticancer agents, including conventional agents like cisplatin and paclitaxel, EGFR Tyrosine Kinase inhibitors, BRAF inhibitors and EGFR-blocking antibodies. It does so by activating tumor speciifc pathways and by impinging on key processes such as the Epithelial to Mesenchymal Transition (EMT) (see cover figure, please).

References

  1. Tuzi, N.L.; Gullick, W.J. Eph, the largest known family of putative growth factor receptors. Br. J. Cancer 1994, 69, 417–421.
  2. Lisabeth, E.M.; Falivelli, G.; Pasquale, E.B. Eph Receptor Signaling and Ephrins. Cold Spring Harb. Perspect. Biol. 2013, 5, a009159.
  3. Pasquale, E.B. Eph-Ephrin Bidirectional Signaling in Physiology and Disease. Cell 2008, 133, 38–52.
  4. Pasquale, E.B. Journal club. A biologist is gratified to find reconciliation for a conflicted receptor. Nature 2009, 461, 149.
  5. Himanen, J.P.; Goldgur, Y.; Miao, H.; Myshkin, E.; Guo, H.; Buck, M.; Nguyen, M.; Rajashankar, K.R.; Wang, B.; Nikolov, D.B. Ligand recognition by A-class Eph receptors: Crystal structures of the EphA2 ligand-binding domain and the EphA2/ephrin-A1 complex. EMBO Rep. 2009, 10, 722–728.
  6. Janes, P.W.; Griesshaber, B.; Atapattu, L.; Nievergall, E.; Hii, L.L.; Mensinga, A.; Chheang, C.; Day, B.W.; Boyd, A.W.; Bastiaens, P.I.; et al. Eph receptor function is modulated by heterooligomerization of A and B type Eph receptors. J. Cell Biol. 2011, 195, 1033–1045.
  7. Eph Nomenclature Committee. Unified Nomenclature for Eph Family Receptors and Their Ligands, the Ephrins. Cell 1997, 90, 403–404.
  8. Wykosky, J.; Palma, E.; Gibo, D.M.; Ringler, S.L.; Turner, C.P.; Debinski, W. Soluble monomeric EphrinA1 is released from tumor cells and is a functional ligand for the EphA2 receptor. Oncogene 2008, 27, 7260–7273.
  9. Egea, J.; Klein, R. Bidirectional Eph–ephrin signaling during axon guidance. Trends Cell Biol. 2007, 17, 230–238.
  10. Binns, K.L.; Taylor, P.P.; Sicheri, F.; Pawson, T.; Holland, S.J. Phosphorylation of Tyrosine Residues in the Kinase Domain and Juxtamembrane Region Regulates the Biological and Catalytic Activities of Eph Receptors. Mol. Cell. Biol. 2000, 20, 4791–4805.
  11. Wybenga-Groot, L.E.; Baskin, B.; Ong, S.H.; Tong, J.; Pawson, T.; Sicheri, F. Structural Basis for Autoinhibition of the EphB2 Receptor Tyrosine Kinase by the Unphosphorylated Juxtamembrane Region. Cell 2001, 106, 745–757.
  12. Singla, N.; Erdjument-Bromage, H.; Himanen, J.P.; Muir, T.W.; Nikolov, D.B. A Semisynthetic Eph Receptor Tyrosine Kinase Provides Insight into Ligand-Induced Kinase Activation. Chem. Biol. 2011, 18, 361–371.
  13. Murai, K.; Pasquale, E. ‘Eph’ective signaling: Forward, reverse and crosstalk. J. Cell Sci. 2003, 116, 2823–2832.
  14. Atapattu, L.; Lackmann, M.; Janes, P.W. The role of proteases in regulating Eph/ephrin signaling. Cell Adhes. Migr. 2014, 8, 294–307.
  15. Falivelli, G.; Lisabeth, E.M.; De La Torre, E.R.; Pérez-Tenorio, G.; Tosato, G.; Salvucci, O.; Pasquale, E.B. Attenuation of Eph Receptor Kinase Activation in Cancer Cells by Coexpressed Ephrin Ligands. PLoS ONE 2013, 8, e81445.
  16. Takasugi, M.; Okada, R.; Takahashi, A.; Chen, D.V.; Watanabe, S.; Hara, E. Small extracellular vesicles secreted from senescent cells promote cancer cell proliferation through EphA2. Nat. Commun. 2017, 8, 15729.
  17. Hirai, H.; Maru, Y.; Hagiwara, K.; Nishida, J.; Takaku, F. A novel putative tyrosine kinase receptor encoded by the eph gene. Science 1987, 238, 1717–1720.
  18. Bartley, T.D.; Hunt, R.W.; Welcher, A.A.; Boyle, W.J.; Parker, V.P.; Lindberg, R.A.; Lu, H.S.; Colombero, A.M.; Elliott, R.L.; Guthrie, B.A.; et al. B61 is a ligand for the ECK receptor protein-tyrosine kinase. Nature 1994, 368, 558–560.
  19. Walker-Daniels, J.; Riese, D.J., 2nd; Kinch, M.S. c-Cbl-dependent EphA2 protein degradation is induced by ligand binding. Mol. Cancer Res. 2002, 1, 79–87.
  20. Seiradake, E.; Schaupp, A.; Ruiz, D.D.T.; Kaufmann, R.; Mitakidis, N.; Harlos, K.; Aricescu, A.R.; Klein, R.; Jones, E.Y. Structurally encoded intraclass differences in EphA clusters drive distinct cell responses. Nat. Struct. Mol. Biol. 2013, 20, 958–964.
  21. Miao, H.; Wang, B. EphA receptor signaling—Complexity and emerging themes. Semin. Cell Dev. Biol. 2012, 23, 16–25.
  22. Miao, H.; Burnett, E.; Kinch, M.; Simon, E.; Wang, B. Activation of EphA2 kinase suppresses integrin function and causes focal-adhesion-kinase dephosphorylation. Nat. Cell Biol. 1999, 2, 62–69.
  23. Bin Fang, W.; Brantley-Sieders, D.M.; Hwang, Y.; Ham, A.-J.L.; Chen, J. Identification and Functional Analysis of Phosphorylated Tyrosine Residues within EphA2 Receptor Tyrosine Kinase. J. Biol. Chem. 2008, 283, 16017–16026.
  24. Xu, N.-J.; Henkemeyer, M. Ephrin reverse signaling in axon guidance and synaptogenesis. Semin. Cell Dev. Biol. 2012, 23, 58–64.
  25. Wang, Y.; Shang, Y.; Li, J.; Chen, W.; Li, G.; Wan, J.; Liu, W.; Zhang, M. Specific Eph receptor-cytoplasmic effector signaling mediated by SAM–SAM domain interactions. eLife 2018, 7, e35677.
  26. Noren, N.K.; Pasquale, E.B. Eph receptor–ephrin bidirectional signals that target Ras and Rho proteins. Cell. Signal. 2004, 16, 655–666.
  27. Barquilla, A.; Pasquale, E.B. Eph Receptors and Ephrins: Therapeutic Opportunities. Annu. Rev. Pharmacol. Toxicol. 2015, 55, 465–487.
  28. Pasquale, E.B. Eph receptors and ephrins in cancer: Bidirectional signalling and beyond. Nat. Rev. Cancer 2010, 10, 165–180.
  29. Guo, H.; Miao, H.; Gerber, L.; Singh, J.; Denning, M.F.; Gilliam, A.C.; Wang, B. Disruption of EphA2 Receptor Tyrosine Kinase Leads to Increased Susceptibility to Carcinogenesis in Mouse Skin. Cancer Res. 2006, 66, 7050–7058.
  30. Minami, M.; Koyama, T.; Wakayama, Y.; Fukuhara, S.; Mochizuki, N. EphrinA/EphA signal facilitates insulin-like growth factor-I–induced myogenic differentiation through suppression of the Ras/extracellular signal–regulated kinase 1/2 cascade in myoblast cell lines. Mol. Biol. Cell 2011, 22, 3508–3519.
  31. Meier, C.; Anastasiadou, S.; Knöll, B. Ephrin-A5 Suppresses Neurotrophin Evoked Neuronal Motility, ERK Activation and Gene Expression. PLoS ONE 2011, 6, e26089.
  32. Nie, D.-Y.; Di Nardo, A.; Han, J.M.; Baharanyi, H.; Kramvis, I.; Huynh, T.; Dabora, S.L.; Codeluppi, S.; Pandolfi, P.P.; Pasquale, E.B.; et al. Tsc2-Rheb signaling regulates EphA-mediated axon guidance. Nat. Neurosci. 2010, 13, 163–172.
  33. Niethamer, T.K.; Bush, J.O. Getting direction(s): The Eph/ephrin signaling system in cell positioning. Dev. Biol. 2019, 447, 42–57.
  34. Cayuso, J.; Xu, Q.; Wilkinson, D.G. Mechanisms of boundary formation by Eph receptor and ephrin signaling. Dev. Biol. 2015, 401, 122–131.
  35. Jørgensen, C.; Sherman, A.; Chen, G.I.; Pasculescu, A.; Poliakov, A.; Hsiung, M.; Larsen, B.; Wilkinson, D.G.; Linding, R.; Pawson, T. Cell-Specific Information Processing in Segregating Populations of Eph Receptor Ephrin-Expressing Cells. Science 2009, 326, 1502–1509.
  36. Wu, Z.; Ashlin, T.G.; Xu, Q.; Wilkinson, D.G. Role of forward and reverse signaling in Eph receptor and ephrin mediated cell segregation. Exp. Cell Res. 2019, 381, 57–65.
  37. Holmberg, J.; Armulik, A.; Senti, K.-A.; Edoff, K.; Spalding, K.L.; Momma, S.; Cassidy, R.M.; Flanagan, J.G.; Frisén, J. Ephrin-A2 reverse signaling negatively regulates neural progenitor proliferation and neurogenesis. Genes Dev. 2005, 19, 462–471.
  38. Jiao, J.-W.; Feldheim, D.A.; Chen, D.F. Ephrins as negative regulators of adult neurogenesis in diverse regions of the central nervous system. Proc. Natl. Acad. Sci. USA 2008, 105, 8778–8783.
  39. Vaught, D.B.; Chen, J.; Brantley-Sieders, D.M. Regulation of Mammary Gland Branching Morphogenesis by EphA2 Receptor Tyrosine Kinase. Mol. Biol. Cell 2009, 20, 2572–2581.
  40. Orsulic, S.; Kemler, R. Expression of Eph receptors and ephrins is differentially regulated by E-cadherin. J. Cell Sci. 2000, 113, 1793–1802.
  41. Tanaka, M.; Kamata, R.; Sakai, R. EphA2 Phosphorylates the Cytoplasmic Tail of Claudin-4 and Mediates Paracellular Permeability. J. Biol. Chem. 2005, 280, 42375–42382.
  42. De Robertis, M.; Mazza, T.; Fusilli, C.; LoIacono, L.; Poeta, M.L.; Sanchez, M.; Massi, E.; Lamorte, G.; Diodoro, M.G.; Pescarmona, E.; et al. EphB2 stem-related and EphA2 progression-related miRNA-based networks in progressive stages of CRC evolution: Clinical significance and potential miRNA drivers. Mol. Cancer 2018, 17, 169.
  43. Rosenberg, I.M.; Göke, M.; Kanai, M.; Reinecker, H.-C.; Podolsky, D.K. Epithelial cell kinase-B61: An autocrine loop modulating intestinal epithelial migration and barrier function. Am. J. Physiol. 1997, 273, G824–G832.
  44. Baldwin, C.; Chen, Z.W.; Bedirian, A.; Yokota, N.; Nasr, S.H.; Rabb, H.; Lemay, S. Upregulation of EphA2 during in vivo and in vitro renal ischemia-reperfusion injury: Role of Src kinases. Am. J. Physiol. Renal Physiol. 2006, 291, F960–F971.
  45. Dongre, A.; Weinberg, R.A. New insights into the mechanisms of epithelial–mesenchymal transition and implications for cancer. Nat. Rev. Mol. Cell Biol. 2019, 20, 69–84.
  46. Gooding, A.J.; Schiemann, W.P. Epithelial–Mesenchymal Transition Programs and Cancer Stem Cell Phenotypes: Mediators of Breast Cancer Therapy Resistance. Mol. Cancer Res. 2020, 18, 1257–1270.
  47. Zantek, N.D.; Azimi, M.; Fedor-Chaiken, M.; Wang, B.; Brackenbury, R.; Kinch, M.S. E-cadherin regulates the function of the EphA2 receptor tyrosine kinase. Cell Growth Differ. Mol. Biol. J. Am. Assoc. Cancer Res. 1999, 10, 629–638.
  48. Shang, X.; Lin, X.; Howell, S.B. Claudin-4 controls the receptor tyrosine kinase EphA2 pro-oncogenic switch through beta-catenin. Cell Commun. Signal. 2014, 12, 59.
  49. Faoro, L.; Singleton, P.A.; Cervantes, G.M.; Lennon, F.E.; Choong, N.W.; Kanteti, R.; Ferguson, B.D.; Husain, A.N.; Tretiakova, M.S.; Ramnath, N.; et al. EphA2 Mutation in Lung Squamous Cell Carcinoma Promotes Increased Cell Survival, Cell Invasion, Focal Adhesions, and Mammalian Target of Rapamycin Activation. J. Biol. Chem. 2010, 285, 18575–18585.
  50. Mudali, S.V.; Fu, B.; Lakkur, S.S.; Luo, M.; Embuscado, E.E.; Iacobuzio-Donahue, C.A. Patterns of EphA2 protein expression in primary and metastatic pancreatic carcinoma and correlation with genetic status. Clin. Exp. Metastasis 2006, 23, 357–365.
  51. Dohn, M.; Jiang, J.; Chen, X. Receptor tyrosine kinase EphA2 is regulated by p53-family proteins and induces apoptosis. Oncogene 2001, 20, 6503–6515.
  52. Zelinski, D.P.; Zantek, N.D.; Walker-Daniels, J.; Peters, M.A.; Taparowsky, E.J.; Kinch, M.S. Estrogen and Myc negatively regulate expression of the EphA2 tyrosine kinase. J. Cell. Biochem. 2002, 85, 714–720.
  53. Kinch, M.S.; Carles-Kinch, K. Overexpression and functional alterations of the EphA2 tyrosine kinase in cancer. Clin. Exp. Metastasis 2003, 20, 59–68.
  54. Zelinski, D.P.; Zantek, N.D.; Stewart, J.C.; Irizarry, A.R.; Kinch, M.S. EphA2 overexpression causes tumorigenesis of mam-mary epithelial cells. Cancer Res. 2001, 61, 2301–2306.
  55. Wykosky, J.; Gibo, D.M.; Stanton, C.; Debinski, W. EphA2 as a Novel Molecular Marker and Target in Glioblastoma Multiforme. Mol. Cancer Res. 2005, 3, 541–551.
  56. Miyazaki, T.; Kato, H.; Fukuchi, M.; Nakajima, M.; Kuwano, H. EphA2 overexpression correlates with poor prognosis in esophageal squamous cell carcinoma. Int. J. Cancer 2002, 103, 657–663.
  57. Huang, C.; Yuan, W.; Lai, C.; Zhong, S.; Yang, C.; Wang, R.; Mao, L.; Chen, Z.; Chen, Z. EphA2-to-YAP pathway drives gastric cancer growth and therapy resistance. Int. J. Cancer 2020, 146, 1937–1949.
  58. Zeng, G.; Hu, Z.; Kinch, M.S.; Pan, C.-X.; Flockhart, D.A.; Kao, C.; Gardner, T.A.; Zhang, S.; Li, L.; Baldridge, L.A.; et al. High-Level Expression of EphA2 Receptor Tyrosine Kinase in Prostatic Intraepithelial Neoplasia. Am. J. Pathol. 2003, 163, 2271–2276.
  59. Dunne, P.D.; Dasgupta, S.; Blayney, J.K.; McArt, D.G.; Redmond, K.L.; Weir, J.-A.; Bradley, C.A.; Sasazuki, T.; Shirasawa, S.; Wang, T.; et al. EphA2 Expression Is a Key Driver of Migration and Invasion and a Poor Prognostic Marker in Colorectal Cancer. Clin. Cancer Res. 2016, 22, 230–242.
  60. Strimpakos, A.; Pentheroudakis, G.; Kotoula, V.; De Roock, W.; Kouvatseas, G.; Papakostas, P.; Makatsoris, T.; Papamichael, D.; Andreadou, A.; Sgouros, J.; et al. The Prognostic Role of Ephrin A2 and Endothelial Growth Factor Receptor Pathway Mediators in Patients with Advanced Colorectal Cancer Treated with Cetuximab. Clin. Colorectal Cancer 2013, 12, 267–274.e2.
  61. De Robertis, M.; LoIacono, L.; Fusilli, C.; Poeta, M.L.; Mazza, T.; Sanchez, M.; Marchionni, L.; Signori, E.; Lamorte, G.; Vescovi, A.L.; et al. Dysregulation of EGFR Pathway in EphA2 Cell Subpopulation Significantly Associates with Poor Prognosis in Colorectal Cancer. Clin. Cancer Res. 2016, 23, 159–170.
  62. Thaker, P.H.; Deavers, M.; Celestino, J.; Thornton, A.; Fletcher, M.S.; Landen, C.N.; Kinch, M.S.; Kiener, P.A.; Sood, A.K. EphA2 Expression Is Associated with Aggressive Features in Ovarian Carcinoma. Clin. Cancer Res. 2004, 10, 5145–5150.
  63. Han, L.; Dong, Z.; Qiao, Y.; Kristensen, G.B.; Holm, R.; Nesland, J.M.; Suo, Z. The clinical significance of EphA2 and Ephrin A-1 in epithelial ovarian carcinomas. Gynecol. Oncol. 2005, 99, 278–286.
  64. Merritt, W.M.; Thaker, P.H.; Landen, C.N., Jr.; Deavers, M.; Fletcher, M.S.; Lin, Y.G.; Han, L.Y.; Kamat, A.A.; Gershenson, D.; Kinch, M.S.; et al. Analysis of EphA2 expression and mutant p53 in ovarian carcinoma. Cancer Biol. Ther. 2006, 5, 1357–1360.
  65. Merritt, W.M.; Kamat, A.A.; Hwang, J.-Y.; Bottsford-Miller, J.; Lu, C.; Lin, Y.G.; Coffey, D.; Spannuth, W.A.; Nugent, E.; Han, L.Y.; et al. Clinical and biological impact of EphA2 overexpression and angiogenesis in endometrial cancer. Cancer Biol. Ther. 2010, 10, 1306–1314.
  66. Wu, D.; Suo, Z.; Kristensen, G.B.; Li, S.; Troen, G.; Holm, R.; Nesland, J.M. Prognostic value of EphA2 and EphrinA-1 in squamous cell cervical carcinoma. Gynecol. Oncol. 2004, 94, 312–319.
  67. Liu, Y.; Zhang, X.; Qiu, Y.; Huang, D.; Zhang, S.; Xie, L.; Qi, L.; Yu, C.; Zhou, X.; Hu, G.; et al. Clinical significance of EphA2 expression in squamous-cell carcinoma of the head and neck. J. Cancer Res. Clin. Oncol. 2010, 137, 761–769.
  68. Liu, F.; Park, P.J.; Lai, W.; Maher, E.; Chakravarti, A.; Durso, L.; Jiang, X.; Yu, Y.; Brosius, A.; Thomas, M.; et al. A Genome-Wide Screen Reveals Functional Gene Clusters in the Cancer Genome and Identifies EphA2 as a Mitogen in Glioblastoma. Cancer Res. 2006, 66, 10815–10823.
  69. Li, X.; Wang, L.; Gu, J.-W.; Li, B.; Liu, W.; Wang, Y.-G.; Zhang, X.; Zhen, H.-N.; Fei, Z. Up-regulation of EphA2 and down-regulation of EphrinA1 are associated with the aggressive phenotype and poor prognosis of malignant glioma. Tumor Biol. 2010, 31, 477–488.
  70. Wang, L.-F.; Fokas, E.; Bieker, M.; Rose, F.; Rexin, P.; Zhu, Y.; Pagenstecher, A.; Engenhart-Cabillic, R.; An, H.-X. Increased expression of EphA2 correlates with adverse outcome in primary and recurrent glioblastoma multiforme patients. Oncol. Rep. 2008, 19, 151–156.
  71. Herrem, C.J.; Tatsumi, T.; Olson, K.S.; Shirai, K.; Finke, J.H.; Bukowski, R.M.; Zhou, M.; Richmond, A.L.; Derweesh, I.; Kinch, M.S.; et al. Expression of EphA2 is prognostic of disease-free interval and overall survival in surgically treated patients with renal cell carcinoma. Clin. Cancer Res. 2005, 11, 226–231.
  72. Xu, J.; Zhang, J.; Cui, L.; Zhang, H.; Zhang, S.; Bai, Y. High EphA2 protein expression in renal cell carcinoma is associated with a poor disease outcome. Oncol. Lett. 2014, 8, 687–692.
  73. Brannan, J.M.; Dong, W.; Prudkin, L.; Behrens, C.; Lotan, R.; Bekele, B.N.; Wistuba, I.; Johnson, F.M. Expression of the Receptor Tyrosine Kinase EphA2 Is Increased in Smokers and Predicts Poor Survival in Non–Small Cell Lung Cancer. Clin. Cancer Res. 2009, 15, 4423–4430.
  74. Guo, C.; Shao, R.; Correa, A.M.; Behrens, C.; Johnson, F.M.; Raso, M.G.; Prudkin, L.; Solis, L.M.; Nunez, M.I.; Fang, B.; et al. Prognostic Significance of Combinations of RNA-Dependent Protein Kinase and EphA2 Biomarkers for NSCLC. J. Thorac. Oncol. 2013, 8, 301–308.
  75. Kinch, M.S.; Moore, M.-B.; Harpole, D.H., Jr. Predictive value of the EphA2 receptor tyrosine kinase in lung cancer recurrence and survival. Clin. Cancer Res. 2003, 9, 613–618.
  76. Cui, X.-D.; Lee, M.-J.; Yu, G.-R.; Kim, I.-H.; Yu, H.-C.; Song, E.-Y.; Kim, D.-G. EFNA1 ligand and its receptor EphA2: Potential biomarkers for hepatocellular carcinoma. Int. J. Cancer 2009, 126, 940–949.
  77. Yang, P.; Yuan, W.; He, J.; Wang, J.; Yu, L.; Jin, X.; Hu, Y.; Liao, M.; Chen, Z.; Zhang, Y. Overexpression of EphA2, MMP-9, and MVD-CD34 in hepatocellular carcinoma: Implications for tumor progression and prognosis. Hepatol. Res. 2009, 39, 1169–1177.
  78. Nakamura, R.; Kataoka, H.; Sato, N.; Kanamori, M.; Ihara, M.; Igarashi, H.; Ravshanov, S.; Wang, Y.-J.; Li, Z.-Y.; Shimamura, T.; et al. EPHA2/EFNA1 expression in human gastric cancer. Cancer Sci. 2005, 96, 42–47.
  79. Macrae, M.; Neve, R.M.; Rodriguez-Viciana, P.; Haqq, C.; Yeh, J.; Chen, C.; Gray, J.W.; McCormick, F. A conditional feedback loop regulates Ras activity through EphA2. Cancer Cell 2005, 8, 111–118.
  80. Fox, B.P.; Kandpal, R.P. Invasiveness of breast carcinoma cells and transcript profile: Eph receptors and ephrin ligands as molecular markers of potential diagnostic and prognostic application. Biochem. Biophys. Res. Commun. 2004, 318, 882–892.
  81. Binda, E.; Visioli, A.; Giani, F.; Lamorte, G.; Copetti, M.; Pitter, K.L.; Huse, J.T.; Cajola, L.; Zanetti, N.; DiMeco, F.; et al. The EphA2 Receptor Drives Self-Renewal and Tumorigenicity in Stem-like Tumor-Propagating Cells from Human Glioblastomas. Cancer Cell 2012, 22, 765–780.
  82. Fernandez-Alonso, R.; Bustos, F.; Budzyk, M.; Kumar, P.; Helbig, A.O.; Hukelmann, J.; Lamond, A.I.; Lanner, F.; Zhou, H.; Petsalaki, E.; et al. Phosphoproteomics identifies a bimodal EPHA2 receptor switch that promotes embryonic stem cell differentiation. Nat. Commun. 2020, 11, 1357.
  83. Wang, Y.J.; Ota, S.; Kataoka, H.; Kanamori, M.; Li, Z.Y.; Band, H.; Tanaka, M.; Sugimura, H. Negative regulation of EphA2 receptor by Cbl. Biochem. Biophys. Res. Commun. 2002, 296, 214–220.
  84. Miao, H.; Wei, B.-R.; Peehl, D.M.; Li, Q.; Alexandrou, T.; Schelling, J.R.; Rhim, J.S.; Sedor, J.R.; Burnett, E.; Wang, B. Activation of EphA receptor tyrosine kinase inhibits the Ras/MAPK pathway. Nat. Cell Biol. 2001, 3, 527–530.
  85. Menges, C.W.; McCance, D.J. Constitutive activation of the Raf–MAPK pathway causes negative feedback inhibition of Ras–PI3K–AKT and cellular arrest through the EphA2 receptor. Oncogene 2007, 27, 2934–2940.
  86. Miao, H.; Li, D.-Q.; Mukherjee, A.; Guo, H.; Petty, A.; Cutter, J.; Basilion, J.P.; Sedor, J.; Wu, J.; Danielpour, D.; et al. EphA2 Mediates Ligand-Dependent Inhibition and Ligand-Independent Promotion of Cell Migration and Invasion via a Reciprocal Regulatory Loop with Akt. Cancer Cell 2009, 16, 9–20.
  87. Yang, N.-Y.; Fernandez, C.; Richter, M.; Xiao, Z.; Valencia, F.; Tice, D.A.; Pasquale, E.B. Crosstalk of the EphA2 receptor with a serine/threonine phosphatase suppresses the Akt-mTORC1 pathway in cancer cells. Cell. Signal. 2011, 23, 201–212.
  88. Neill, T.; Buraschi, S.; Goyal, A.; Sharpe, C.; Natkanski, E.; Schaefer, L.; Morrione, A.; Iozzo, R.V. EphA2 is a functional receptor for the growth factor progranulin. J. Cell Biol. 2016, 215, 687–703.
  89. Stallaert, W.; Bruggemann, Y.; Sabet, O.; Baak, L.; Gattiglio, M.; Bastiaens, P.I.H. Contact inhibitory Eph signaling suppresses EGF-promoted cell migration by decoupling EGFR activity from vesicular recycling. Sci. Signal. 2018, 11, eaat0114.
  90. Parri, M.; Buricchi, F.; Giannoni, E.; Grimaldi, G.; Mello, T.; Raugei, G.; Ramponi, G.; Chiarugi, P. EphrinA1 Activates a Src/Focal Adhesion Kinase-mediated Motility Response Leading to Rho-dependent Actino/Myosin Contractility. J. Biol. Chem. 2007, 282, 19619–19628.
  91. Hill, W.; Hogan, C. Normal epithelial cells trigger EphA2-dependent RasV12 cell repulsion at the single cell level. Small GTPases 2019, 10, 305–310.
  92. Zhou, Y.; Sakurai, H. Emerging and Diverse Functions of the EphA2 Noncanonical Pathway in Cancer Progression. Biol. Pharm. Bull. 2017, 40, 1616–1624.
  93. Beauchamp, A.; Debinski, W. Ephs and ephrins in cancer: Ephrin-A1 signalling. Semin. Cell Dev. Biol. 2012, 23, 109–115.
  94. Fang, W.B.; Ireton, R.C.; Zhuang, G.; Takahashi, T.; Reynolds, A.; Chen, J. Overexpression of EPHA2 receptor destabilizes adherens junctions via a RhoA-dependent mechanism. J. Cell Sci. 2008, 121, 358–368.
  95. Kikawa, K.D.; Vidale, D.R.; Van Etten, R.L.; Kinch, M.S. Regulation of the EphA2 Kinase by the Low Molecular Weight Tyrosine Phosphatase Induces Transformation. J. Biol. Chem. 2002, 277, 39274–39279.
  96. Lori, G.; Gamberi, T.; Paoli, P.; Caselli, A.; Pranzini, E.; Marzocchini, R.; Modesti, A.; Raugei, G. LMW-PTP modulates glucose metabolism in cancer cells. Biochim. Biophys. Acta BBA Gen. Subj. 2018, 1862, 2533–2544.
  97. Raugei, G.; Ramponi, G.; Chiarugi, P. Low molecular weight protein tyrosine phosphatases: Small, but smart. Cell. Mol. Life Sci. 2002, 59, 941–949.
  98. Malentacchi, F.; Marzocchini, R.; Gelmini, S.; Orlando, C.; Serio, M.; Ramponi, G.; Raugei, G. Up-regulated expression of low molecular weight protein tyrosine phosphatases in different human cancers. Biochem. Biophys. Res. Commun. 2005, 334, 875–883.
  99. Kawai, H.; Kobayashi, M.; Hiramoto-Yamaki, N.; Harada, K.; Negishi, M.; Katoh, H. Ephexin4-mediated promotion of cell migration and anoikis resistance is regulated by serine 897 phosphorylation of EphA2. FEBS Open Bio 2013, 3, 78–82.
  100. Murga, C.; Zohar, M.; Teramoto, H.; Gutkind, J.S. Rac1 and RhoG promote cell survival by the activation of PI3K and Akt, independently of their ability to stimulate JNK and NF-κB. Oncogene 2002, 21, 207–216.
  101. Fujimoto, S.; Negishi, M.; Katoh, H. RhoG Promotes Neural Progenitor Cell Proliferation in Mouse Cerebral Cortex. Mol. Biol. Cell 2009, 20, 4941–4950.
  102. Yamaki, N.; Negishi, M.; Katoh, H. RhoG regulates anoikis through a phosphatidylinositol 3-kinase-dependent mechanism. Exp. Cell Res. 2007, 313, 2821–2832.
  103. Simpson, C.D.; Anyiwe, K.; Schimmer, A.D. Anoikis resistance and tumor metastasis. Cancer Lett. 2008, 272, 177–185.
  104. Barquilla, A.; Lamberto, I.; Noberini, R.; Heynen-Genel, S.; Brill, L.M.; Pasquale, E.B. Protein kinase A can block EphA2 receptor–mediated cell repulsion by increasing EphA2 S897 phosphorylation. Mol. Biol. Cell 2016, 27, 2757–2770.
  105. Gehring, M.P.; Pasquale, E.B. Protein kinase C phosphorylates the EphA2 receptor on serine 892 in the regulatory linker connecting the kinase and SAM domains. Cell. Signal. 2020, 73, 109668.
  106. Harada, K.; Hiramoto-Yamaki, N.; Negishi, M.; Katoh, H. Ephexin4 and EphA2 mediate resistance to anoikis through RhoG and phosphatidylinositol 3-kinase. Exp. Cell Res. 2011, 317, 1701–1713.
  107. Singh, D.R.; Kanvinde, P.; King, C.; Pasquale, E.B.; Hristova, K. The EphA2 receptor is activated through induction of distinct, ligand-dependent oligomeric structures. Commun. Biol. 2018, 1, 15.
  108. Chavent, M.; Karia, D.; Kalli, A.C.; Domański, J.; Duncan, A.L.; Hedger, G.; Stansfeld, P.J.; Seiradake, E.; Jones, E.Y.; Sansom, M.S.P. Interactions of the EphA2 Kinase Domain with PIPs in Membranes: Implications for Receptor Function. Structure 2018, 26, 1025–1034.e2.
  109. Wen, Q.; Chen, Z.; Chen, Z.; Chen, J.; Wang, R.; Huang, C.; Yuan, W. EphA2 affects the sensitivity of oxaliplatin by inducing EMT in oxaliplatin-resistant gastric cancer cells. Oncotarget 2017, 8, 47998–48011.
  110. Peng, Q.; Chen, L.; Wu, W.; Wang, J.; Zheng, X.; Chen, Z.; Jiang, Q.; Han, J.; Wei, L.; Wang, L.; et al. EPH receptor A2 governs a feedback loop that activates Wnt/β-catenin signaling in gastric cancer. Cell Death Dis. 2018, 9, 1146.
  111. Hamaoka, Y.; Negishi, M.; Katoh, H. Tyrosine kinase activity of EphA2 promotes its S897 phosphorylation and glioblastoma cell proliferation. Biochem. Biophys. Res. Commun. 2018, 499, 920–926.
  112. Allocca, C.; Cirafici, A.M.; Laukkanen, M.O.; Castellone, M.D. Serine 897 Phosphorylation of EPHA2 Is Involved in Signaling of Oncogenic ERK1/2 Drivers in Thyroid Cancer Cells. Thyroid 2020.
  113. Graves, P.R.; Din, S.U.; Ashamalla, M.; Ashamalla, H.; Gilbert, T.S.K.; Graves, L.M. Ionizing radiation induces EphA2 S897 phosphorylation in a MEK/ERK/RSK-dependent manner. Int. J. Radiat. Biol. 2017, 93, 929–936.
  114. Cui, X.-D.; Lee, M.-J.; Kim, J.-H.; Hao, P.-P.; Liu, L.; Yu, G.-R.; Kim, D.-G. Activation of mammalian target of rapamycin complex 1 (mTORC1) and Raf/Pyk2 by growth factor-mediated Eph receptor 2 (EphA2) is required for cholangiocarcinoma growth and metastasis. Hepatology 2013, 57, 2248–2260.
  115. Taddei, M.L.; Parri, M.; Angelucci, A.; Bianchini, F.; Marconi, C.; Giannoni, E.; Raugei, G.; Bologna, M.; Calorini, L.; Chiarugi, P. EphA2 Induces Metastatic Growth Regulating Amoeboid Motility and Clonogenic Potential in Prostate Carcinoma Cells. Mol. Cancer Res. 2011, 9, 149–160.
  116. Taddei, M.L.; Parri, M.; Angelucci, A.; Onnis, B.; Bianchini, F.; Giannoni, E.; Raugei, G.; Calorini, L.; Rucci, N.; Teti, A.; et al. Kinase-Dependent and -Independent Roles of EphA2 in the Regulation of Prostate Cancer Invasion and Metastasis. Am. J. Pathol. 2009, 174, 1492–1503.
  117. Teramoto, K.; Katoh, H. The cystine/glutamate antiporter xCT is a key regulator of EphA2 S897 phosphorylation under glucose-limited conditions. Cell. Signal. 2019, 62, 109329.
  118. Xiang, Y.-P.; Xiao, T.; Li, Q.-G.; Lu, S.-S.; Zhu, W.; Liu, Y.-Y.; Qiu, J.-Y.; Song, Z.-H.; Huang, W.; Yi, H.; et al. Y772 phosphorylation of EphA2 is responsible for EphA2-dependent NPC nasopharyngeal carcinoma growth by Shp2/Erk-1/2 signaling pathway. Cell Death Dis. 2020, 11, 709.
  119. Bin Fang, W.; Brantley-Sieders, D.M.; Parker, M.A.; Reith, A.D.; Chen, J. A kinase-dependent role for EphA2 receptor in promoting tumor growth and metastasis. Oncogene 2005, 24, 7859–7868.
  120. Sheng, Y.; Wei, J.; Zhang, Y.; Gao, X.; Wang, Z.; Yang, J.; Yan, S.; Zhu, Y.; Zhang, Z.; Xu, D.; et al. Mutated EPHA2 is a target for combating lymphatic metastasis in intrahepatic cholangiocarcinoma. Int. J. Cancer 2019, 144, 2440–2452.
  121. Tan, Y.-H.C.; Srivastava, S.; Won, B.M.; Kanteti, R.; Arif, Q.; Husain, A.N.; Li, H.; Vigneswaran, W.T.; Pang, K.-M.; Kulkarni, P.; et al. EPHA2 mutations with oncogenic characteristics in squamous cell lung cancer and malignant pleural mesothelioma. Oncogenesis 2019, 8, 49.
  122. Sáinz-Jaspeado, M.; Huertas-Martinez, J.; Lagares-Tena, L.; Liberal, J.M.; Mateo-Lozano, S.; De Álava, E.; De Torres, C.; Mora, J.; Del Muro, X.G.; Tirado, O.M. EphA2-Induced Angiogenesis in Ewing Sarcoma Cells Works through bFGF Production and Is Dependent on Caveolin-1. PLoS ONE 2013, 8, e71449.
  123. Garcia-Monclús, S.; López-Alemany, R.; Almacellas-Rabaiget, O.; Herrero-Martín, D.; Huertas-Martinez, J.; Lagares-Tena, L.; Alba-Pavón, P.; Hontecillas-Prieto, L.; Mora, J.; De Álava, E.; et al. EphA2 receptor is a key player in the metastatic onset of Ewing sarcoma. Int. J. Cancer 2018, 143, 1188–1201.
More
Video Production Service