Next Article in Journal
Synchrotron-Based Pencil Beam Scanning Nozzle with an Integrated Mini-Ridge Filter: A Dosimetric Study to Optimize Treatment Delivery
Next Article in Special Issue
Status of Agents Targeting the HGF/c-Met Axis in Lung Cancer
Previous Article in Journal
Tracking Functional Tumor Cell Subpopulations of Malignant Glioma by Phasor Fluorescence Lifetime Imaging Microscopy of NADH
Previous Article in Special Issue
Role and Therapeutic Targeting of the HGF/MET Pathway in Glioblastoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Activated HGF-c-Met Axis in Head and Neck Cancer

1
Departments of Otolaryngology, University of Kansas Medical Center, Kansas City, KS 66160, USA
2
Anatomy & Cell Biology, University of Kansas Medical Center, Kansas City, KS 66160, USA
3
Cancer Biology, University of Kansas Medical Center, Kansas City, KS 66160, USA
*
Author to whom correspondence should be addressed.
Cancers 2017, 9(12), 169; https://doi.org/10.3390/cancers9120169
Submission received: 16 October 2017 / Revised: 6 December 2017 / Accepted: 7 December 2017 / Published: 12 December 2017
(This article belongs to the Special Issue Hepatocyte Growth Factor Pathway in Cancer)

Abstract

:
Head and neck squamous cell carcinoma (HNSCC) is a highly morbid disease. Recent developments including Food and Drug Administration (FDA) approved molecular targeted agent’s pembrolizumab and cetuximab show promise but did not improve the five-year survival which is currently less than 40%. The hepatocyte growth factor receptor; also known as mesenchymal–epithelial transition factor (c-Met) and its ligand hepatocyte growth factor (HGF) are overexpressed in head and neck squamous cell carcinoma (HNSCC); and regulates tumor progression and response to therapy. The c-Met pathway has been shown to regulate many cellular processes such as cell proliferation, invasion, and angiogenesis. The c-Met pathway is involved in cross-talk, activation, and perpetuation of other signaling pathways, curbing the cogency of a blockade molecule on a single pathway. The receptor and its ligand act on several downstream effectors including phospholipase C gamma (PLCγ), cellular Src kinase (c-Src), phosphotidylinsitol-3-OH kinase (PI3K) alpha serine/threonine-protein kinase (Akt), mitogen activate protein kinase (MAPK), and wingless-related integration site (Wnt) pathways. They are also known to cross-talk with other receptors; namely epidermal growth factor receptor (EGFR) and vascular endothelial growth factor receptor (VEGFR) and specifically contribute to treatment resistance. Clinical trials targeting the c-Met axis in HNSCC have been undertaken because of significant preclinical work demonstrating a relationship between HGF/c-Met signaling and cancer cell survival. Here we focus on HGF/c-Met impact on cellular signaling in HNSCC to potentiate tumor growth and disrupt therapeutic efficacy. Herein we summarize the current understanding of HGF/c-Met signaling and its effects on HNSCC. The intertwining of c-Met signaling with other signaling pathways provides opportunities for more robust and specific therapies, leading to better clinical outcomes.

Graphical Abstract

1. Introduction

Of all the neoplasms in the head and neck region, roughly 90% are classified as head and neck squamous cell carcinoma (HNSCC) [1]. The estimated new cases for HNSCC related deaths in the United States alone is nearly 50,000 [2]. Even with major advances in multimodality treatment, the five-year progression-free survival rate in HPV negative patients is less than 50% and overall survival rate for recurrent or metastatic HNSCC patients remains necessitous [3]. Major risk factors associated with HNSCC occurrence include human papilloma virus (HPV) infection, excessive alcohol consumption, tobacco use, chewing areca nut, poor oral hygiene, and genetic alterations leading to susceptibility to malignancies [4,5]. Treatment is stage and site dependent. Irradiation, chemotherapy, and surgery are the current therapeutic options.
Cetuximab, a human/mouse chimeric monoclonal antibody targeting (EGFR), is the only FDA-approved targeted therapy for EGFR in HNSCC [6]. However, combinatorial cetuximab treatment with radiation or chemotherapy yields only modest results (29.3 vs. 49 months and 7.4 vs. 10.1 months, respectively) [7]. Despite a growing body of evidence that suggest HNSCC is highly heterogeneous [8], EGFR overexpression has been identified as common in up to 90% of HNSCCs [9]. Resistance to cetuximab has been observed relative to hepatocyte growth factor/mesenchymal-epithelial transition factor (HGF/c-Met) signaling [10]. Immune-checkpoint inhibitors (that facilitate T-cell activation), nivolumab and pembrolizumab promise greater overall survival and extended time between recurrence [11]. Both agents were Food and Drug Administration (FDA) approved for advanced HNSCC in 2016. Although immune-checkpoint inhibitors point to the promise of targeted therapies in HNSCC, early data from clinical trials indicate limited efficacy with monotherapy in HNSCC compared to standard chemotherapy. Thus, there continues to be a need for more effective approaches for HNSCC. Receptor tyrosine kinases regulate pro-tumor signaling pathways in HNSCC and are an attractive therapeutic target.
The c-Met proto-oncogene is a membrane spanning receptor tyrosine kinase (RTK) whose ligand is the hepatocyte growth factor (HGF) [12]. Its aberrant activity is documented in HNSCC as well as many other cancers [13]. Its maladaptive signaling in HNSCC is associated with invasion, metastasis, and intertwining with other pathways. Met expression is associated with worse prognosis and lower overall survival [14]. Further, HGF/Met expression levels have been clinically shown to be inversely proportional to patient survival with late-stage nasopharyngeal carcinoma [15]. This review summarizes the current knowledge about HGF/c-Met pathway, as it is associated with HNSCC, and posits that c-Met is a valuable target in the treatment therein.

2. Brief HGF/c-Met History

Hepatocyte growth factor, also then named “Hepatotropin”, was isolated from serum in partially hepatectomized rats and described as a mitogen of liver regeneration in 1984 [16]. HGF was purified from rat platelets in 1984 [17] human plasma in 1988 [18,19] and rat liver in 1991 [20].
Scatter factor (SF) was first identified as a protein secreted by fibroblasts and smooth muscle and had properties that dispersed “scattered” uniform epithelial cells in culture by disrupting intracellular junctions [21,22,23]. Until the early 1990s, HGF and SF once thought to be two different ligands until they were discovered to be, in fact, identical structures serving the same receptor, c-Met [12,24,25,26,27,28,29]. Around the same time, HGF was being implicated as a mitogen, motogen, and morphogen (having the ability to promote mitosis, motility, morphological changes, respectively) [30], while also being recognized as an elicitor of angiogenesis [31], tumorigenesis [32], and invasion [23]. Independently, NIH 3T3 fibroblasts transfected with human Met cDNA displayed motogenic and invasive properties when treated with HGF [33].
HGF displayed cytotoxic effects in a few cell lines, deviating from known HGF activity [34]. With this in mind, Higashio et al. found that Sarcoma-180 was killed in the presence of soluble factors derived from human lung fibroblasts naming it tumor cytotoxic factor (TCF). The same group found the molecular makeup of TCF and HGF to be identical [35,36,37,38].
c-Met, the receptor for HGF, was originally discovered in osteosarcoma cell line that was chemically transformed in 1984 [39]. The chemical treatment yielded an oncogenic activation wherein chromosome 1 was rearranged and fused a region called tpr (translocated promoter region) [40,41,42]. Tpr-Met is highly constitutively active as a tyrosine kinase and can alter cells in vitro. Tpr-Met cDNA led to the identification of the full length c-Met receptor [43].
c-Met was shown to be overexpressed in the human gastric cancer line (GTL-16), and later in human thyroid carcinomas, human colorectal cancer, and with many other epithelial tumors in the early 90s [44,45,46]. Pertinent to this review, Matsumoto et al. showed that soluble factors from fibroblast conditioned media promoted invasion of HNSCC [47]. The group later identified the factors as HGF/SF in 1994 [48]. In 2000, the first report of direct MET involvement of tumor metastasis (found in HNSCC, but it was the first time the connection between MET and tumor growth was made) was noted by Di Renzo et al. in. They saw that clonal expansion of neoplastic cells harbored mutations that activated MET [49].

3. Role of HGF/c-Met Signaling in Homeostasis

In normal development and tissue regeneration HGF/c-Met is crucial. HGF/c-Met has physiological roles in cardiovascular remodeling [50], skeletal muscle regeneration [51], differentiation and migration during spermatogenesis [52], axonal growth [53], renal tubulogenesis [54], and blood cell growth and differentiation [55]. Recently, HGF has been found to be elevated during cerebral aneurism promoting healing of tissue damage, and promotes organ regeneration. Blood drawn from 16 patients suffering intracranial aneurysms was drawn from either aneurysm lumen or femoral artery. There was a five-fold increase in HGF plasma levels drawn from the site of injury. The same study evaluated mice with induced intracranial aneurism treated with a c-Met antagonist or control. The c-Met antagonist treated mice experienced decreased survival compared to control [56].
Loreto et al. (2010) studied c-Met expression during salivary gland morphogenesis. A sample of 12 human embryonic salivary glands were studied at different stages of development using immunohistochemical staining. While all stages of salivary gland development demonstrated staining for c-Met, there was stronger expression in the earlier stages of development [57]. HGF/c-Met has been studied in the embryogenesis of liver development [58], tooth development [59], mammary gland morphogenesis [60,61,62], skeletal myogenesis [14], neural induction [63], chemo attraction of motor neurons [64,65,66], and hepatic bile duct formation [67]. It is also involved in regulation of differentiating events during development [68,69,70].
The study of cancer compared to wound healing and tissue regeneration has long been thought as apposite, with many parallels being readily identified. Additionally, many cancer therapies interrupt wound healing [71]. Likewise, many signals in normal embryogenesis are dysregulated in cancers [72]. HGF/c-Met is important to both normative function and cancer biology, suggesting a value in appreciating both.

4. Ligand Expression, Secretion, and Activation

HGF expression is normally found in the vicinity of epithelial and endothelial cells expressing the c-Met receptor. While much remains to be discovered about HGF formation and secretion, it has been reported that granulocytes (particularly neutrophils) store inactive pro-HGF [73], and coagulation factor Xa was shown to induce secretion of mature HGF [74]. HGF is secreted by stromal cells as an inactive 92 kDa single-chain polypeptide known as pro-HGF. Pro-HGF consists of an amino-terminal heparin binding domain, four Kringle domains, and a serine protease-like domain [75]. While pro-HGF can bind c-Met, it does so with no activation of the receptor. Maturation of pro-HGF occurs in the ECM [29]. Pro-HGF is cleaved between the fourth Kringle domain Arg494 and serine proteinase domain Val495 to its active heterodimer by serine proteases [76]. Though pro-HGF can bind to c-Met with low affinity, it does not induce receptor activation. HGF is processed by serum or cellular proteases including HGFA, factor XIIa and XIa, matriptase, hepsin, transmembrane protease serine 13 (TMPRSS13), human airway trypsin-like protease (HAT), urokinase-type plasminogen activator (uPA), and tissue-type plasminogen activator (tPA) [77]. SPINT1 and SPINT2, are genes encoding serine protease inhibitors that inhibit pro-HGF cleavage adding another degree of regulation to the HGF/c-Met signaling axis [78,79]. Additionally, heparan-sulfate proteoglycans (HSPGs) bound to HGF, facilitate the binding of pro-HGF to the extracellular matrix proteins, protect against proteolytic cleavage, create stores for later use, enable clearance by endocytosis, and regulate tissue diffusion [80,81,82]. Normal breast ductal branching depends on HSPGs binding HGF in vivo [83]. In physionormative biology, HSPGs binding HGF seems to protect from aberrant signaling by sequestering and degrading the growth factor [84]. In fact, blocking heparanase, an enzyme that cleaves the HSPG side chain thereby interrupting their affinity for growth factors, actually reduces tumor growth and extends time to recurrence [85,86]. Contrarily, over-sulfated heparan (clinically given as a coagulating agent) can amplify many signaling pathways [87]. Activated HGF consists of an α-chain with an N-terminal, hairpin loop domain linked via disulfide bond to the fourth Kringle domain in the β-chain. The N terminal and first Kringle domain of HGF bind with high affinity to the Ig-like fold, plexins, transcription factor (IPT) domain, suggesting N and K1 as receptor agonist [88]. The SPH domain of HGF binds the sema domain of c-Met with lower affinity, but with the ability to pull two HGF bound c-Met complexes together for dimerization [89] (Figure 1).

5. Receptor Structure and Activation

Mesenchymal epithelial transition factor (c-Met) is a semaphorin, membrane spanning receptor tyrosine kinase that is encoded by the MET proto-oncogene [12,90]. From the met transcript, a 150 kDa polypeptide is produced that is partially glycosylated to become a 170 kDa precursor protein [91,92]. The Met precursor is additionally glycosylated and cleaved into a 50 kDa α-chain and a 140 kDa β-chain (Figure 1). The larger β-subunit contains an intracellular tyrosine kinase domain, a membrane-spanning domain, and a larger extracellular region. The α-subunit is exclusively extracellular. The two subunits are joined by di-sulfide bonds and together make up a mature α-β Met heterodimer [90,93,94].
The α-chain of mature, bivalent HGF binds c-Met β-chain with high affinity at the immunoglobulin-plexin-transcription (IPT) domain. Meanwhile the β-chain of HGF binds c-Met’s Sema domain with low affinity. Engineered proteins mimicking the IPT domains of Met show anti-tumor effects in mice; in vitro display inhibit HGF induced cell growth and invasion [95]. It is worth noting that pro-HGF binds c-Met with high affinity and can displace active HGF. HGF is diffused in low levels throughout the ECM, while pro-HGF is in higher quantities. HGF is supplied by stromal cells in a paracrine manner and is not controlled at the transcriptional level. High affinity pro-HGF allows its receptor to be “primed” to meet and environmental challenges. Activation can occur while pro-HGF is receptor-bound.
In physionormative settings, binding of HGF activates c-Met by phosphorylating two intracellular tyrosine residues, inducing catalytic activity [96]. The carboxy-terminal of the beta chain has two tyrosine residues that serve as docking sites for intracellular proteins. When phosphorylated, activation of c-Met in response to HGF is partly mediated by autophosphorylation of the two catalytic residues, Tyr1234 and Tyr1235, located within the activation loop of the tyrosine kinase activity receptor [28]. This is followed by phosphorylation of the residues Tyr1349 and Tyr1356 in the C-terminus, activating a docking site for Src homology-2 (SH2) domain, phosphotyrosine binding (PTB) domain, and Met binding domain (MDB) that all contain signal transducers [97,98,99].

5.1. Intracellular Signaling

Adapter proteins known to interact with the tyrosine residues include: growth factor receptor-bound protein 2 (Grb2), src homology/collagen (SHC), Grb2-associated adaptor protein (Gab1), and CT10 (chicken tumor virus No. 10) regulator of kinase/Crk-like (Crk/CRKL) with signal transducers phosphotidylinsitol-3-OH kinase (PI3K), signal transducer and activator of transcription-3 (Stat3), phospholipase C-γ (PLCγ), Ras guanine nucleotide exchange factor son-of-sevenless (SOS), Src kinase, and Src homology region-2 containing protein tyrosine phosphatase 2 (SHP2) [97,98,99,100,101,102,103,104,105]. Y1349 and Y1356 facilitate interplay with GRB2-associated-binding protein 1 (Gab1), Src, and SHC. Y1356 residue recruits Grb2, PI3K, PLC-γ, and SHP2 to the Met signaling complex [106]. Gab1 is a scaffolding protein adaptor that contains the Met-binding site. The interaction at that site with Met results in the phosphorylation of Gab1, which is responsible for the unique biological effects of HGF [99]. PI3K/AKT (alpha serine/threonine-protein kinase), mitogen-activated protein kinase (MAPK), and signal transducer and activators of transcription (STAT) pathways are downstream pathways affected by normal c-Met signaling (Figure 2).
Activation of MAPK will successively activate different protein kinases whose terminal effectors include extracellular signal-regulated kinases (Erk1 and Erk2), Jun amino-terminal kinases (JNK1, JNK2, and JNK3) and p38 (a class of Mapk’s). These proteins will activate cell cycle regulators producing cell proliferation and stimulating functional alterations in the cytoskeleton, important to cell migration and invasion. Inactivation of B-cell lymphoma 2 (Bcl-2) antagonist of cell death (BAD) and degradation of the pro-apoptotic protein p53 increases cell survival and apoptotic resistance by activation of the PI3K/AKT pathway [107]. STAT3 activation relocates the c-Met receptor to the nucleus from the plasma membrane, and begins activity as a transcription factor regulating genes for cell differentiation and proliferation [108].
The plasma membrane proves to be an important area of complex c-Met interaction with other surface proteins, contributing to dynamic c-Met biological responses—EGFR, α6β4 integrin, semaphorins in the plexin B family, and the variant of the hyaluronan receptor cluster of differentiation 44 (CD44) (links intracellular cytoskeleton and the extracellular matrix) [109,110].

5.2. Regulation of c-Met Signaling

After HGF binding, c-Met is endocytosed via clatherin coated vesicles to early peripheral endosomes. This is facilitated by protein kinase Cε (PKCε), that promotes both the transfer of active Erk to focal adhesions and HGF-induced cell migration. Next, PKCα mediates c-Met travel along the microtubule network towards the late perinuclear compartments. The accumulation of c-Met in endosomal compartments near the nucleus is a determining step of STAT3 activation [111].
Downregulation of c-Met involves trafficking and degradation in the lysosomes. Initiated by juxtamembrane domain site Y1003 association with casitas B-lineage lymphoma (CBL) and endocytic adaptors, c-Met then accumulates in multivesicular bodies that subsequently fuse with lysosomes, which then break down the proteins. Loss of Y1003 residue allows c-Met to evade lysomal degradation and become oncogenic [112]. Alternatively, c-Met can also be cleaved in either the extracellular domain or the intracellular domain. Extracellular cleavage occurs by disintegrin and metalloprotease (ADAM) which creates a fragment that sequesters the ligand and obstructs the receptor’s activity. Intracellular cleavage is mediated by γ-secretase that produces a fragment that is destroyed by the proteasome [113].
c-Met downregulation can also occur by its binding to protein tyrosine phosphatases (PTPs), which include receptor-type PTPs density enhanced phosphatase 1 (dEP1) and leukocyte common antigen-related molecule (LAR), and the nonreceptor PTPs PTP1B and T-cell protein tyrosine phosphatase (TCPTP). TCPTP modulates c-Met signaling by dephosphorylating the tyrosines in the kinase domain, and dEP1 downregulates c-Met signaling by dephosphorylating the docking tyrosines [114,115].

6. Aberrant Functions in HNSCC

6.1. c-MET Drives Tumorigenesis in HNSCC

Typical of most tyrosine kinase receptors, c-MET can be activated by one of three mechanisms. First, ligand binding triggers dimerization and transactivation of the receptor. Phosphorylated c-Met (p-Met) is overexpressed in HNSCC patients. Total c-Met is overexpressed in 78% of HNSCC cases, and 66% of HNSCC demonstrate phosphorylation at activating sites Y1230, Y1234, and Y1235. Activating sites in this case refers to tyrosines whose phosphorylation results in the recruitment of signaling molecules that go on to transduce a signal. Thus the phosphorylation of tyrosine results in an activation of a signaling cascade [116] (Figure 2). Both autocrine and paracrine mechanisms can trigger c-Met signaling. Glioblastoma, osteosarcoma, rhabdomyosarcoma, and breast carcinoma [107,117,118,119] cells demonstrate autocrine activation of the receptor. While HNSCC cell lines express c-Met, they do not secrete HGF. Rather HNSCC cancer-associated fibroblasts (CAF) have been shown to secrete HGF, suggesting an HGF-c-Met paracrine driven tumorigenesis [120].
Second, physical modifications in the receptor can sustain constitutive activation via somatic mutations. In particular, mutations in the kinase domain involving Y1235D and Y1230C are important activators of the c-Met pathway in HNSCC [49,121] (Figure 2). Other germ-line MET mutations occur in liver metastasis from colon cancer [122] and hepatocellular carcinoma [123] cause the gene to be over amplified. MET gene mutations occur in papillary renal cell carcinoma [124], childhood hepatocellular carcinoma [125], and lymph-node metastases of HNSCC [49]. Gain in MET copy number is present in 16% of HNSCC cases and is associated with c-Met overexpression and poorer outcomes for those individuals [14]. However, while not at high frequency, MET mutations have been identified on several domains including the sema, juxtamembrane, and kinase domains. HNSCC metastatic potential increases with Y1230C and Y1235D kinase domain mutations [49,121]. Y1235D mutation is also associated with impaired local HNSCC tumor control, interferes with response to radiation treatment, and worsens recurrence [126]. In HNSCC tumor tissue ligand binding sema domain sites T230M, E168D, N375S; the juxtamembrane domain sites T1010I and R988C; and the kinase domain sites T1275I and V1333I mutations have been described, but consequences of these mutations not yet determined. While not directly responsible for phorphsorylation, the mutation V1333I was present in 13.5% of HNSCC tumor tissue [116]. Juxtamembrane domain deletion, CBL E3 ubiquitin-ligase recruitment, and exon 14 skipping have been attributed to more rare mutations [127] (Figure 3).
Third, MET protein overexpression is common to thyroid [45,128], ovarian [129,130], pancreatic [131], prostate [132], renal-cell [133], hepatocellular [134], gastric [135], esophageal [136], breast [137], colorectal [138], and head and neck carcinomas [139]. Oligomerization occurs when a cognate, extracellular ligand binds to a monomeric c-Met receptor, stimulating autophosphorylation of the tyrosine residues of the kinase domain. The phosphorylated domain then initiates the c-Met signaling cascade [140]. Ligand overexpression was first reported nearly 20 years ago [141] and has since been shown to drive tumorigenesis in a paracrine manner [120]. HGF overexpression is reported in 58% of HNSCC tumors, lagging behind c-Met expression [10]. This lends to the idea that not only are ligand and receptor overexpression driving tumorigenesis, but also proteins required for concurrent activation are also importantly overexpressed.

6.2. c-MET Drives Metastasis in HNSCC

HGF/c-Met plays many roles in the early metastasis of epithelial cancers. These include delamination of epithelial cells and epithelial–mesenchymal transition necessary for non-local epithelial cell migration [142], basement membrane degradation and remodeling by urokinase and matrix metalloproteinase [143], activation of focal adhesion kinase and paxillin causing integrin-dependent migration [48], pre-metastatic niche formation via tumor-derived exosomes [14], tumor lymphangiogenesis [144,145], and hemangiogenesis [146,147].
Lymph node invasion is a common feature of clinical HNSCC and is highly predictive of patient mortality [148]. c-Met is highly expressed in lymph node metastasis in HNSCC, while MET gene amplification is low [149]. c-Met tends to be present in all stages of metastasis, it tends to be most expressed at N2 and N3 nodal metastasis [150,151]. HGF is elevated in HNSCC patients compared to healthy individuals [49]. Invasion occurs in normative physiology during embryogenesis. Many of the same regulatory and inductive properties are involved in invasion of cancer. Normal invasive growth signals can be dysregulated contributing to cell transformation and later tumor progression [23]. Mice orthotopically injected with Met knockdown (MetKD) HNSCC cells show no regional lymph node metastasis. Further, these mice show significant reduction in primary tumors and increased cellular apoptosis derived from MetKD [152].
In late stage cancer, primary tumor cells invade contiguous tissues, and some tumor cells travel to distant organs or lymph nodes where they may develop into new, secondary tumors [153]. Anoikis is a process of programmed cell death induced by the detachment of anchorage-dependent cells from their extracellular matrix (ECM) associations. Anchorage independent cell growth and attachment to an inappropriate matrix is typical of metastasis and invasion. Anoikis is the process by which programmed cell death is induced upon detachment from is appropriate matrix (or niche) HGF induces anoikis resistance by fibronectin signaling in HNSCC, disrupting integrin signaling [154]. MET activation prevents anoikis via ERK or Akt pathways. [155] HGF/c-Met signaling may help an abnormal cell survive temporary displacement of focal contacts by PI3K activation, a pro-survival signal [156]. Anoikis resistance is an important mode of action by which HNSCC develops nodal metastasis from the primary tumor site.

7. c-MET Pathway Crosstalk in HNSCC

7.1. Wnt/β-Catenin

Wnt/β-Catenin is a pathway important for pattern formation during embryogenesis and roles in cancer [157]. β-catenin-dependent transcription is mediated through c-Met in colon cancer cells, and will convert cells to a cell type with cancer stem cell properties [158]. Cancer stem-like (CSC) cells in HNSCC were inhibited with PF-2341066 (a c-Met inhibitor) and β-catenin was shown to be the downregulating factor contributing to CSC elimination. The Wnt pathway frizzled class receptor 8 (FZD8) expression rescued impaired HNSCC cells that were treated with a c-Met inhibitor. It was noted by the same group that FZD8 was upregulated by c-Met signaling through ERK/c-Fos cascade [159].

7.2. c-SRC

c-Src regulates many signaling cascades that control various biological outcomes. Inhibition of c-Src in cancer cells can result in anchorage-independent growth, survival, tumor vascularity, migration, metastasis, survival, and invasion. Activated c-Src mediates erlotinib resistance in HNSCC by stimulating c-Met independent of ligand [160]. Sen et al. earlier showed in both xenograft and in vitro models that combination inhibition of c-Met and c-Src resulted in synergistic cytotoxicity, enhanced apoptosis, and decreased tumor size [161].

7.3. TGF-β

Bhowmick et al. showed that transforming growth factor (TGF)-β type II receptor knockout in mice gave rise to prostate and gut epithelial tumors by activating c-Met through paracrine overexpression of HGF by stromal cells [162]. TGF-β is mediated by the transcription factor mothers against decapentaplegic homology (SMAD) that bind SMAD binding element (SBE) of target genes, regulating their expression. Smad deletion caused HGF upregulation, contributing to angiogenesis in mice. The HGF promoter of keratinocytes have one such SBE that allows binding of SMAD 1, 2, and 4 families [163]. In HNSCC, in vitro knockdown of Smad4 induced cetuximab resistance by activating TGF-β and c-Met pathways [164].

7.4. EGFR

EGFR and c-Met have in common similar downstream pathways: MAPK (RAS/Rapidly accelerating fibrosarcoma (RAF)/ MAP kinase kinases (MEK)/ERK) and PI3K/AKT/mechanistic target of rapamycin (mTOR) [165,166,167,168]. To eliminate these aberrant pathway modalities altogether, dual EGFR/c-Met blockade is of interest. Seiwert et al. showed that dual blockade of SU11274 (c-Met inhibitor) and erlotinib (EGFR inhibitor) in HNSCC lines produce greater-than-additive inhibition of cell growth via erbB3/AKT signaling [116]. Further, Lieu et al. demonstrated that cell lines treated with foretinib and erlotinib or lapatinib (EGFR/human epidermal growth factor (HER) 2 inhibitor) synergistically inhibited growth [169]. Even more, crizotinib (c-Met inhibitor) and gefitinib given in combination potentiated the effects of cell line invasion, wound healing, and proliferation; increased antitumor activity in vivo when compared to EGFR inhibition alone was shown by Xe et al. They also noted that in the absence of HGF, TGF-α (an EGFR ligand) activates c-Met when EGFR is blocked [170]. The intertwining of the RTK’s c-Met and EGFR may explain acquired or constitutive EGFR resistance. c-Met/EGFR co-expression is common. It is likely that c-Met/HGF expression is a common mechanism of EGFR treatment resistance in HNSCC [171]. Wheeler et al. further show that cell lines resistance to cetuximab highly express c-Met, EGFR, HER-2, and HER-3 [172].

8. c-MET Contribution to Therapeutic Resistance in HNSCC

The c-Met signaling axis has been implicated in acquired resistance to epidermal growth factor receptor (EGFR) targeting therapies in HNSCC. EGFR, like c-Met, is an RTK that is overexpressed in roughly 90% of HNSCC patients. EGFR overexpression correlates with regional lymph node metastasis and poor outcomes [173]. Acquired EGFR resistance and recurrence secondary to c-Met activation is a common feature in clinical trials [160,174]. Small molecule tyrosine kinase inhibitors (TKI’s), gefitinib or erlotinib, competitively bind the ATP-binding region EGFR, thereby selectively inhibiting the receptor function. Despite EGFR overexpression and the small molecule TKI’s in vitro effectiveness, it does not significantly improve clinical outcomes [175]. c-Met upregulation is also linked to cetuximab resistance in HNSCC patients [10]. EGFR only rarely exhibits activating mutations. Implicatively, EGFR resistance comes not from EGFR itself, but rather from co-activation of other RTKs [176]. c-Met knockdown in vitro sensitizes EGFR to erlotinib, and c-Met inhibition also shows erlotinib sensitization in a dominant-active c-Src subset of EGFR overexpressing HNSCC lines [160].

9. Co-Therapies in HNSCC Targeting c-Met

c-Met plays a major role in compensating for inhibition of RTK pathways that drive proliferation and metastasis in HNSCC. Therefore, targeting c-Met in conjunction with these pathways may lead to more effective therapeutic strategies. Here we discuss c-Met inhibitors used in combination with other targeted and nontargeted therapies in clinical and preclinical models of HNSCC.

9.1. Crizotinib (PF2341066)

Crizotinib is a class Ia small molecule dual inhibitor of c-Met and anaplastic lymphoma kinase (ALK) that was approved for therapeutic use in advanced non-small cell lung cancers positive for either c-ros oncogene 1 (ROS-1) or ALK. Crizotinib inhibits c-Met and ALK by directly binding the activation loop and stabilizing the autoinhibitory conformation of each kinase. Though crizotinib is a potent inhibitor of both kinases, it exhibits greater specificity against c-Met [177].
Preclinical studies using crizotinib alone show potent reduction of proliferation for HNSCC cell lines (cell IC50 4.1–4.7 μM) and a significant inhibition of both wound-closure and invasion in vitro. Crizotinib induced apoptosis and greatly decreased the tumor burden in a patient xenograft in vivo model [120]. In combination with the EGFR inhibitor gefitinib, crizotinib showed greater inhibition of cell survival in vitro as well as impaired wound closure and invasion. Similar results were obtained in vivo with enhanced reduction of patient xenograft tumor volume and the number of cells positive for proliferation markers [170]. Conversely, when crizotinib was combined with radiation therapy, inhibition of c-MET appeared to reduce radiosensitivity in vitro as measured by colony forming assay and in vivo using a mouse model [178]. To date, no clinical trials have been performed using crizotinib to treat HNSCC.

9.2. Capmatinib (INC-280)

To date, capmatinib is the only class Ib small molecule inhibitor of c-Met to be used in clinical trials for HNSCC. This class of inhibitors, like class Ia, interacts with the autoinhibitory loop of c-Met, yet they also take advantage of the smaller ATP-binding site of c-Met to generate greater specificity against this kinase [120]. Unfortunately, little in vitro data is available for the effects of capmatinib on head and neck cell lines; however, it is known to be a potent inhibitor of c-Met in a kinase assay (enzymatic IC50 0.13 nM) and proliferation of other tumor lines (cell IC50 1.2–12.4 nM). In tumor-bearing mice, capmatinib was able to inhibit tumor growth and even cause complete regression of some tumor lines without any noticeable toxicity [179]. An ongoing Phase I trial is aimed at determining safety and tolerability of capmatinib in advanced refractory c-Met-dependent solid tumors (NCT01324479). Another Phase Ib/II trial exploring the safety and efficacy of capmatinib treatment with cetuximab, an anti-EGFR antibody, in HNSCC and metastatic colorectal cancers positive for c-Met has recently been suspended for unknown reasons (NCT02205398).

9.3. Golvatinib (E7050)

Golvatinib is one of two class II c-Met inhibitors to be used in clinical trials. These inhibitors are large, hydrophobic molecules that occupy the back pocket of c-Met and are usually less specific . As such, golvatinib is a dual inhibitor against c-Met and vascular endothelial growth factor (VEGFR)-2 (enzymatic IC50 14 nM and 16 nM, respectively). In preclinical studies, it inhibited proliferation of a broad range of tumor cell lines (cell IC50 6.2 nM to 4.3 μM) and inhibited HUVEC growth in response to HGF and VEGF as an in vitro model of angiogenesis. Golvatinib also greatly reduced tumor burden in mouse xenograft models and blood vessel density in remaining tumors [180].
Golvatinib has been tested in two clinical trials for HNSCC to date. The first study is a Phase I/II trial targeting EGFR with cetuximab in the presence or absence of golvatinib and is currently recruiting (NCT01332266). The second clinical trial is meant to determine the efficacy of golvatinib treatment in conjunction with cepecitabine and cisplatin (NCT01355302). Unfortunately, this trial did not pass Phase I with five of the seven enrolled patients experiencing serious adverse effects, including supraventricular tachycardia, convulsion, and pulmonary embolism.

9.4. Foretinib

Foretinib is a class II inhibitor targeting c-Met, VEGFR2, Recepteur d’Origine Nantais (RON), kinase insert domain receptor (KDR), and Fms related tyrosine kinase 1 (Flt-1) with potent inhibition of c-Met kinase activity (enzymatic IC50 1.16 nM) [181]. In HNSCC cell lines, foretinib alone strongly inhibited cell growth (cell IC50 0.61–0.79 μM) and exhibited synergistic inhibition of cell proliferation when combined with the EGFR inhibitor erlotinib [169]. Foretinib has also recently been demonstrated to contribute to radiosensitization in models of esophageal squamous cell carcinoma. Following irradiation, foretinib-treated cells had diminished repair of DNA double-stranded breaks and underwent more apoptosis. This effect was confirmed in vivo as seen by reduced tumor burden in mouse xenografts [182]. It is unclear why inhibition of c-Met by foretinib increased radiosensitivity, whereas inhibition with crizotinib appeared to have the opposite effect, though this phenomenon is likely due to other kinases targeted by these two molecules.
A clinical trial examining the safety and efficacy of foretinib alone in treating recurrent, metastatic HNSCC was terminated after Phase I after none of the enrolled patients presented partial or complete response (NCT00725764). Though this trial did not progress, enrolled patients had a 50% disease stabilization rate, and 6 of the 14 patients showed shrinking tumor size [183]. No clinical trials to date have used foretinib in combination with other agents to treat HNSCC.

9.5. Ficlatuzimab

Ficlatuzimab is currently the only biological therapy undergoing trial that targets the c-Met/HGF axis in HNSCC. It is a humanized monoclonal antibody that sequesters HGF, thereby preventing the stimulation of c-Met by HGF. Given its mechanism of action, however, it is important to recognize that ficlatuzumab is unable to inhibit HGF-independent c-Met activation. Ficlatuzumab has been demonstrated to interfere with stromal contributions to proliferation, migration, and invasion in a preclinical in vitro model of HNSCC, as well as epithelial to mesenchymal transition (EMT) [184]. Ficlatuzumab is currently being tested in combination with cetuximab in a Phase Ib clinical trial of recurrent and metastatic HNSCC (NCT02277197) (Table 1).

10. Conclusions

Over three decades of research have been spent investigating the HGF/c-Met pathway and its clinical relevance. In HNSCC, c-Met receptor overexpression is commonly associated with poorer outcomes in patients. This is attributable to increased nodal metastases, cellular proliferation, and cell survival. Notably, c-Met contributes to treatment resistance by bypassing traditionally clinically inhibited signals such as EGF. However, Met inhibitors have yet to show marked improvement for patient outcomes. One of the hurdles to overcome is shifting promising multi-signal blockades from preclinical to clinical trials. Further, there is a need to develop c-Met-specific agents and use these in patients with appropriate biomarkers. Understanding the integration between HGF/c-Met axis and other oncogenic signaling pathways may lead to stronger treatment modalities in the near future (Table 2).

Acknowledgments

This study was supported in part by University of Kansas Cancer Center Pilot Award under CCSG 1-P30-CA168524-02, an NIH Clinical and Translational Science Award grant (UL1TR000001, formerly UL1RR033179) awarded to the University of Kansas Medical Center, and an internal Lied Basic Science Grant Program of the KUMC Research Institute.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

EGFREpidermal Growth Factor Receptor
c-METMet Tyrosine Kinase Receptor, mesenchymal epithelial transition factor
HGFHepatocyte Growth Factor
HNSCCHead and neck squamous cell carcinoma
HPVHuman Papilloma Virus
EGFEpidermal Growth Factor
mABmonoclonal antibody
TMETumor Microenvironment
SFScatter Factor
HSPGHeparan Sulfate Proteoglycan

References

  1. Marur, S.; Forastiere, A.A. Head and Neck Squamous Cell Carcinoma: Update on Epidemiology, Diagnosis, and Treatment. Mayo Clin. Proc. 2016, 91, 386–396. [Google Scholar] [CrossRef] [PubMed]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer Statistics, 2017. CA Cancer J. Clin. 2017, 67, 7–30. [Google Scholar] [CrossRef] [PubMed]
  3. Baxi, S.; Fury, M.; Ganly, I.; Rao, S.; Pfister, D.G. Ten years of progress in head and neck cancers. J. Natl. Compr. Cancer Netw. 2012, 10, 806–810. [Google Scholar] [CrossRef]
  4. Li, Y.C.; Chang, J.T.; Chiu, C.; Lu, Y.C.; Li, Y.L.; Chiang, C.H.; You, G.R.; Lee, L.Y.; Cheng, A.J. Areca nut contributes to oral malignancy through facilitating the conversion of cancer stem cells. Mol. Carcinog. 2016, 55, 1012–1023. [Google Scholar] [CrossRef] [PubMed]
  5. Galbiatti, A.L.; Padovani-Junior, J.A.; Maniglia, J.V.; Rodrigues, C.D.; Pavarino, E.C.; Goloni-Bertollo, E.M. Head and neck cancer: Causes, prevention and treatment. Braz. J. Otorhinolaryngol. 2013, 79, 239–247. [Google Scholar] [CrossRef] [PubMed]
  6. Vermorken, J.B.; Mesia, R.; Rivera, F.; Remenar, E.; Kawecki, A.; Rottey, S.; Erfan, J.; Zabolotnyy, D.; Kienzer, H.R.; Cupissol, D.; et al. Platinum-based chemotherapy plus cetuximab in head and neck cancer. N. Engl. J. Med. 2008, 359, 1116–1127. [Google Scholar] [CrossRef] [PubMed]
  7. Bonner, J.A.; Harari, P.M.; Giralt, J.; Cohen, R.B.; Jones, C.U.; Sur, R.K.; Raben, D.; Baselga, J.; Spencer, S.A.; Zhu, J.; et al. Radiotherapy plus cetuximab for locoregionally advanced head and neck cancer: 5-year survival data from a phase 3 randomised trial, and relation between cetuximab-induced rash and survival. Lancet Oncol. 2010, 11, 21–28. [Google Scholar] [CrossRef]
  8. Stransky, N.; Egloff, A.M.; Tward, A.D.; Kostic, A.D.; Cibulskis, K.; Sivachenko, A.; Kryukov, G.V.; Lawrence, M.S.; Sougnez, C.; McKenna, A.; et al. The mutational landscape of head and neck squamous cell carcinoma. Science 2011, 333, 1157–1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Grandis, J.R.; Tweardy, D.J. Elevated levels of transforming growth factor alpha and epidermal growth factor receptor messenger RNA are early markers of carcinogenesis in head and neck cancer. Cancer Res. 1993, 53, 3579–3584. [Google Scholar] [PubMed]
  10. Madoz-Gurpide, J.; Zazo, S.; Chamizo, C.; Casado, V.; Carames, C.; Gavin, E.; Cristobal, I.; Garcia-Foncillas, J.; Rojo, F. Activation of MET pathway predicts poor outcome to cetuximab in patients with recurrent or metastatic head and neck cancer. J. Transl. Med. 2015, 13, 282. [Google Scholar] [CrossRef] [PubMed]
  11. Ferris, R.L.; Blumenschein, G.J.; Fayette, J.; Guigay, J.; Colevas, A.D.; Licitra, L.; Harrington, K.; Kasper, S.; Vokes, E.E.; Even, C.; et al. Nivolumab for Recurrent Squamous-Cell Carcinoma of the Head and Neck. N. Engl. J. Med. 2016, 375, 1856–1867. [Google Scholar] [CrossRef] [PubMed]
  12. Bottaro, D.P.; Rubin, J.S.; Faletto, D.L.; Chan, A.M.; Kmiecik, T.E.; Vande Woude, G.F.; Aaronson, S.A. Identification of the hepatocyte growth factor receptor as the c-met proto-oncogene product. Science 1991, 251, 802–804. [Google Scholar] [CrossRef] [PubMed]
  13. Organ, S.L.; Tsao, M.-S. An overview of the c-MET signaling pathway. Ther. Adv. Med. Oncol. 2011, 3, S7–S19. [Google Scholar] [CrossRef] [PubMed]
  14. Cho, Y.A.; Kim, E.K.; Heo, S.J.; Cho, B.C.; Kim, H.R.; Chung, J.M.; Yoon, S.O. Alteration status and prognostic value of MET in head and neck squamous cell carcinoma. J. Cancer 2016, 7, 2197–2206. [Google Scholar] [CrossRef] [PubMed]
  15. Qian, C.-N.; Guo, X.; Cao, B.; Kort, E.J.; Lee, C.-C.; Chen, J.; Wang, L.-M.; Mai, W.-Y.; Min, H.-Q.; Hong, M.-H.; et al. Met Protein Expression Level Correlates with Survival in Patients with Late-stage Nasopharyngeal Carcinoma. Cancer Res. 2002, 62, 589–596. [Google Scholar] [PubMed]
  16. Nakamura, T.; Nawa, K.; Ichihara, A. Partial purification and characterization of hepatocyte growth factor from serum of hepatectomized rats. Biochem. Biophys. Res. Commun. 1984, 122, 1450–1459. [Google Scholar] [CrossRef]
  17. Nakamura, T.; Teramoto, H.; Ichihara, A. Purification and characterization of a growth factor from rat platelets for mature parenchymal hepatocytes in primary cultures. Proc. Natl. Acad. Sci. USA 1986, 83, 6489–6493. [Google Scholar] [CrossRef] [PubMed]
  18. Gohda, E.; Tsubouchi, H.; Nakayama, H.; Hirono, S.; Sakiyama, O.; Takahashi, K.; Miyazaki, H.; Hashimoto, S.; Daikuhara, Y. Purification and partial characterization of hepatocyte growth factor from plasma of a patient with fulminant hepatic failure. J. Clin. Investig. 1988, 81, 414–419. [Google Scholar] [CrossRef] [PubMed]
  19. Zarnegar, R.; Michalopoulos, G. Purification and biological characterization of human hepatopoietin A, a polypeptide growth factor for hepatocytes. Cancer Res. 1989, 49, 3314–3320. [Google Scholar] [PubMed]
  20. Asami, O.; Ihara, I.; Shimidzu, N.; Shimizu, S.; Tomita, Y.; Ichihara, A.; Nakamura, T. Purification and characterization of hepatocyte growth factor from injured liver of carbon tetrachloride-treated rats. J. Biochem. 1991, 109, 8–13. [Google Scholar] [PubMed]
  21. Stoker, M.; Perryman, M. An epithelial scatter factor released by embryo fibroblasts. J. Cell Sci. 1985, 77, 209–223. [Google Scholar] [PubMed]
  22. Stoker, M.; Gherardi, E.; Perryman, M.; Gray, J. Scatter factor is a fibroblast-derived modulator of epithelial cell mobility. Nature 1987, 327, 239–242. [Google Scholar] [CrossRef] [PubMed]
  23. Weidner, K.M.; Behrens, J.; Vandekerckhove, J.; Birchmeier, W. Scatter factor: Molecular characteristics and effect on the invasiveness of epithelial cells. J. Cell Biol. 1990, 111, 2097–2108. [Google Scholar] [CrossRef] [PubMed]
  24. Konishi, T.; Takehara, T.; Tsuji, T.; Ohsato, K.; Matsumoto, K.; Nakamura, T. Scatter factor from human embryonic lung fibroblasts is probably identical to hepatocyte growth factor. Biochem. Biophys. Res. Commun. 1991, 180, 765–773. [Google Scholar] [CrossRef]
  25. Furlong, R.A.; Takehara, T.; Taylor, W.G.; Nakamura, T.; Rubin, J.S. Comparison of biological and immunochemical properties indicates that scatter factor and hepatocyte growth factor are indistinguishable. J. Cell Sci. 1991, 100, 173–177. [Google Scholar] [PubMed]
  26. Hartmann, G.; Naldini, L.; Weidner, K.M.; Sachs, M.; Vigna, E.; Comoglio, P.M.; Birchmeier, W. A functional domain in the heavy chain of scatter factor/hepatocyte growth factor binds the c-Met receptor and induces cell dissociation but not mitogenesis. Proc. Natl. Acad. Sci. USA 1992, 89, 11574–11578. [Google Scholar] [CrossRef] [PubMed]
  27. Weidner, K.M.; Arakaki, N.; Hartmann, G.; Vandekerckhove, J.; Weingart, S.; Rieder, H.; Fonatsch, C.; Tsubouchi, H.; Hishida, T.; Daikuhara, Y. Evidence for the identity of human scatter factor and human hepatocyte growth factor. Proc. Natl. Acad. Sci. USA 1991, 88, 7001–7005. [Google Scholar] [CrossRef] [PubMed]
  28. Naldini, L.; Vigna, E.; Ferracini, R.; Longati, P.; Gandino, L.; Prat, M.; Comoglio, P.M. The tyrosine kinase encoded by the MET proto-oncogene is activated by autophosphorylation. Mol. Cell. Biol. 1991, 11, 1793–1803. [Google Scholar] [CrossRef] [PubMed]
  29. Naldini, L.; Vigna, E.; Narsimhan, R.P.; Gaudino, G.; Zarnegar, R.; Michalopoulos, G.K.; Comoglio, P.M. Hepatocyte growth factor (HGF) stimulates the tyrosine kinase activity of the receptor encoded by the proto-oncogene c-MET. Oncogene 1991, 6, 501–504. [Google Scholar] [PubMed]
  30. Gherardi, E.; Stoker, M. Hepatocyte growth factor—Scatter factor: Mitogen, motogen, and met. Cancer Cells 1991, 3, 227–232. [Google Scholar] [PubMed]
  31. Bussolino, F.; Di Renzo, M.F.; Ziche, M.; Bocchietto, E.; Olivero, M.; Naldini, L.; Gaudino, G.; Tamagnone, L.; Coffer, A.; Comoglio, P.M. Hepatocyte growth factor is a potent angiogenic factor which stimulates endothelial cell motility and growth. J. Cell Biol. 1992, 119, 629–641. [Google Scholar] [CrossRef] [PubMed]
  32. Rong, S.; Bodescot, M.; Blair, D.; Dunn, J.; Nakamura, T.; Mizuno, K.; Park, M.; Chan, A.; Aaronson, S.; Vande Woude, G.F. Tumorigenicity of the met proto-oncogene and the gene for hepatocyte growth factor. Mol. Cell. Biol. 1992, 12, 5152–5158. [Google Scholar] [CrossRef] [PubMed]
  33. Giordano, S.; Zhen, Z.; Medico, E.; Gaudino, G.; Galimi, F.; Comoglio, P.M. Transfer of motogenic and invasive response to scatter factor/hepatocyte growth factor by transfection of human MET protooncogene. Proc. Natl. Acad. Sci. USA 1993, 90, 649–653. [Google Scholar] [CrossRef] [PubMed]
  34. Tajima, H.; Matsumoto, K.; Nakamura, T. Hepatocyte growth factor has potent anti-proliferative activity in various tumor cell lines. FEBS Lett. 1991, 291, 229–232. [Google Scholar] [CrossRef]
  35. Higashio, K.; Shima, N.; Goto, M.; Itagaki, Y.; Nagao, M.; Yasuda, H.; Morinaga, T. Identity of a tumor cytotoxic factor from human fibroblasts and hepatocyte growth factor. Biochem. Biophys. Res. Commun. 1990, 170, 397–404. [Google Scholar] [CrossRef]
  36. Shima, N.; Itagaki, Y.; Nagao, M.; Yasuda, H.; Morinaga, T.; Higashio, K. A fibroblast-derived tumor cytotoxic factor/F-TCF (hepatocyte growth factor/HGF) has multiple functions in vitro. Cell Biol. Int. Rep. 1991, 15, 397–408. [Google Scholar] [CrossRef]
  37. Shima, N.; Nagao, M.; Ogaki, F.; Tsuda, E.; Murakami, A.; Higashio, K. Tumor cytotoxic factor/hepatocyte growth factor from human fibroblasts: Cloning of its cDNA, purification and characterization of recombinant protein. Biochem. Biophys. Res. Commun. 1991, 180, 1151–1158. [Google Scholar] [CrossRef]
  38. Shima, N.; Higashio, K. [Structure and biological property of fibroblast-derived tumor cytotoxic factor (F-TCF)]. Nihon Rinsho 1992, 50, 1962–1966. [Google Scholar] [PubMed]
  39. Cooper, C.S.; Park, M.; Blair, D.G.; Tainsky, M.A.; Huebner, K.; Croce, C.M.; Vande Woude, G.F. Molecular cloning of a new transforming gene from a chemically transformed human cell line. Nature 1984, 311, 29–33. [Google Scholar] [CrossRef] [PubMed]
  40. Takahashi, M.; Cooper, G.M. ret transforming gene encodes a fusion protein homologous to tyrosine kinases. Mol. Cell. Biol. 1987, 7, 1378–1385. [Google Scholar] [CrossRef] [PubMed]
  41. Dean, M.; Park, M.; Vande Woude, G.F. Characterization of the rearranged tpr-met oncogene breakpoint. Mol. Cell. Biol. 1987, 7, 921–924. [Google Scholar] [CrossRef] [PubMed]
  42. Rodrigues, G.A.; Park, M. Dimerization mediated through a leucine zipper activates the oncogenic potential of the met receptor tyrosine kinase. Mol. Cell. Biol. 1993, 13, 6711–6722. [Google Scholar] [CrossRef] [PubMed]
  43. Park, M.; Dean, M.; Cooper, C.S.; Schmidt, M.; O’Brien, S.J.; Blair, D.G.; Vande Woude, G.F. Mechanism of met oncogene activation. Cell 1986, 45, 895–904. [Google Scholar] [CrossRef]
  44. Ponzetto, C.; Giordano, S.; Peverali, F.; Della Valle, G.; Abate, M.L.; Vaula, G.; Comoglio, P.M. c-met is amplified but not mutated in a cell line with an activated met tyrosine kinase. Oncogene 1991, 6, 553–559. [Google Scholar] [PubMed]
  45. Di Renzo, M.F.; Olivero, M.; Ferro, S.; Prat, M.; Bongarzone, I.; Pilotti, S.; Belfiore, A.; Costantino, A.; Vigneri, R.; Pierotti, M.A.; et al. Overexpression of the c-MET/HGF receptor gene in human thyroid carcinomas. Oncogene 1992, 7, 2549–2553. [Google Scholar] [PubMed]
  46. Liu, C.; Park, M.; Tsao, M.S. Overexpression of c-met proto-oncogene but not epidermal growth factor receptor or c-erbB-2 in primary human colorectal carcinomas. Oncogene 1992, 7, 181–185. [Google Scholar] [PubMed]
  47. Matsumoto, K.; Horikoshi, M.; Rikimaru, K.; Enomoto, S. A study of an in vitro model for invasion of oral squamous cell carcinoma. J. Oral Pathol. Med. 1989, 18, 498–501. [Google Scholar] [CrossRef] [PubMed]
  48. Matsumoto, K.; Matsumoto, K.; Nakamura, T.; Kramer, R.H. Hepatocyte growth factor/scatter factor induces tyrosine phosphorylation of focal adhesion kinase (p125FAK) and promotes migration and invasion by oral squamous cell carcinoma cells. J. Biol. Chem. 1994, 269, 31807–31813. [Google Scholar]
  49. Di Renzo, M.F.; Olivero, M.; Martone, T.; Maffe, A.; Maggiora, P.; Stefani, A.D.; Valente, G.; Giordano, S.; Cortesina, G.; Comoglio, P.M. Somatic mutations of the MET oncogene are selected during metastatic spread of human HNSC carcinomas. Oncogene 2000, 19, 1547–1555. [Google Scholar] [CrossRef] [PubMed]
  50. Gallo, S.; Sala, V.; Gatti, S.; Crepaldi, T. Cellular and molecular mechanisms of HGF/Met in the cardiovascular system. Clin. Sci. 2015, 129, 1173–1193. [Google Scholar] [CrossRef] [PubMed]
  51. Webster, M.T.; Fan, C.M. c-MET regulates myoblast motility and myocyte fusion during adult skeletal muscle regeneration. PLoS ONE 2013, 8, e81757. [Google Scholar] [CrossRef] [PubMed]
  52. Depuydt, C.E.; Zalata, A.; de Potter, C.R.; van Emmelo, J.; Comhaire, F.H. The receptor encoded by the human C-MET oncogene is expressed in testicular tissue and on human spermatozoa. Mol. Hum. Reprod. 1996, 2, 2–8. [Google Scholar] [CrossRef] [PubMed]
  53. Li, P.P.; Madhavan, R.; Peng, H.B. Differential regulation of axonal growth and neuromuscular junction assembly by HGF/c-Met signaling. Dev. Dyn. 2012, 241, 1562–1574. [Google Scholar] [CrossRef] [PubMed]
  54. Montesano, R.; Matsumoto, K.; Nakamura, T.; Orci, L. Identification of a fibroblast-derived epithelial morphogen as hepatocyte growth factor. Cell 1991, 67, 901–908. [Google Scholar] [CrossRef]
  55. Fehlner-Gardiner, C.C.; Cao, H.; Jackson-Boeters, L.; Nakamura, T.; Elliott, B.E.; Uniyal, S.; Chan, B.M. Characterization of a functional relationship between hepatocyte growth factor and mouse bone marrow-derived mast cells. Differentiation 1999, 65, 27–42. [Google Scholar] [CrossRef] [PubMed]
  56. Pena-Silva, R.A.; Chalouhi, N.; Wegman-Points, L.; Ali, M.; Mitchell, I.; Pierce, G.L.; Chu, Y.; Ballas, Z.K.; Heistad, D.; Hasan, D. Novel role for endogenous hepatocyte growth factor in the pathogenesis of intracranial aneurysms. Hypertension 2015, 65, 587–593. [Google Scholar] [CrossRef] [PubMed]
  57. Loreto, C.; Caltabiano, R.; Musumeci, G.; Caltabiano, C.; Greco, M.G.; Leonardi, R. Hepatocyte growth factor receptor, c-Met, in human embryo salivary glands. An immunohistochemical study. Anat. Histol. Embryol. 2010, 39, 173–177. [Google Scholar] [CrossRef] [PubMed]
  58. Schmidt, C.; Bladt, F.; Goedecke, S.; Brinkmann, V.; Zschiesche, W.; Sharpe, M.; Gherardi, E.; Birchmeier, C. Scatter factor/hepatocyte growth factor is essential for liver development. Nature 1995, 373, 699–702. [Google Scholar] [CrossRef] [PubMed]
  59. Amano, O.; Matsumoto, K.; Nakamura, T.; Iseki, S. Expression and localization of hepatocyte growth factor in rat submandibular gland. Growth Factors 1994, 10, 145–151. [Google Scholar] [CrossRef] [PubMed]
  60. Soriano, J.V.; Pepper, M.S.; Nakamura, T.; Orci, L.; Montesano, R. Hepatocyte growth factor stimulates extensive development of branching duct-like structures by cloned mammary gland epithelial cells. J. Cell Sci. 1995, 108, 413–430. [Google Scholar] [PubMed]
  61. Yang, Y.; Spitzer, E.; Meyer, D.; Sachs, M.; Niemann, C.; Hartmann, G.; Weidner, K.M.; Birchmeier, C.; Birchmeier, W. Sequential requirement of hepatocyte growth factor and neuregulin in the morphogenesis and differentiation of the mammary gland. J. Cell Biol. 1995, 131, 215–226. [Google Scholar] [CrossRef] [PubMed]
  62. Niemann, C.; Brinkmann, V.; Spitzer, E.; Hartmann, G.; Sachs, M.; Naundorf, H.; Birchmeier, W. Reconstitution of mammary gland development in vitro: Requirement of c-met and c-erbB2 signaling for branching and alveolar morphogenesis. J. Cell Biol. 1998, 143, 533–545. [Google Scholar] [CrossRef] [PubMed]
  63. Streit, A.; Stern, C.D.; Thery, C.; Ireland, G.W.; Aparicio, S.; Sharpe, M.J.; Gherardi, E. A role for HGF/SF in neural induction and its expression in Hensen’s node during gastrulation. Development 1995, 121, 813–824. [Google Scholar] [PubMed]
  64. Ebens, A.; Brose, K.; Leonardo, E.D.; Hanson, M.G., Jr.; Bladt, F.; Birchmeier, C.; Barres, B.A.; Tessier-Lavigne, M. Hepatocyte growth factor/scatter factor is an axonal chemoattractant and a neurotrophic factor for spinal motor neurons. Neuron 1996, 17, 1157–1172. [Google Scholar] [CrossRef]
  65. Wong, V.; Glass, D.J.; Arriaga, R.; Yancopoulos, G.D.; Lindsay, R.M.; Conn, G. Hepatocyte growth factor promotes motor neuron survival and synergizes with ciliary neurotrophic factor. J. Biol. Chem. 1997, 272, 5187–5191. [Google Scholar] [CrossRef] [PubMed]
  66. Yamamoto, Y.; Livet, J.; Pollock, R.A.; Garces, A.; Arce, V.; deLapeyriere, O.; Henderson, C.E. Hepatocyte growth factor (HGF/SF) is a muscle-derived survival factor for a subpopulation of embryonic motoneurons. Development 1997, 124, 2903–2913. [Google Scholar] [PubMed]
  67. Johnson, M.; Koukoulis, G.; Matsumoto, K.; Nakamura, T.; Iyer, A. Hepatocyte growth factor induces proliferation and morphogenesis in nonparenchymal epithelial liver cells. Hepatology 1993, 17, 1052–1061. [Google Scholar] [CrossRef] [PubMed]
  68. Basler, K.; Hafen, E. Control of photoreceptor cell fate by the sevenless protein requires a functional tyrosine kinase domain. Cell 1988, 54, 299–311. [Google Scholar] [PubMed]
  69. Chabot, B.; Stephenson, D.A.; Chapman, V.M.; Besmer, P.; Bernstein, A. The proto-oncogene c-kit encoding a transmembrane tyrosine kinase receptor maps to the mouse W locus. Nature 1988, 335, 88–89. [Google Scholar] [CrossRef] [PubMed]
  70. Sonnenberg, E.; Godecke, A.; Walter, B.; Bladt, F.; Birchmeier, C. Transient and locally restricted expression of the ros1 protooncogene during mouse development. EMBO J. 1991, 10, 3693–3702. [Google Scholar] [PubMed]
  71. Schäfer, M.; Werner, S. Cancer as an overhealing wound: An old hypothesis revisited. Nat. Rev. Mol. Cell Biol. 2008, 9, 628. [Google Scholar] [CrossRef] [PubMed]
  72. Kelleher, F.C.; Fennelly, D.; Rafferty, M. Common critical pathways in embryogenesis and cancer. Acta Oncol. 2006, 45, 375–388. [Google Scholar] [CrossRef] [PubMed]
  73. Grenier, A.; Chollet-Martin, S.; Crestani, B.; Delarche, C.; El Benna, J.; Boutten, A.; Andrieu, V.; Durand, G.; Gougerot-Pocidalo, M.-A.; Aubier, M.; et al. Presence of a mobilizable intracellular pool of hepatocyte growth factor in human polymorphonuclear neutrophils. Blood 2002, 99, 2997–3004. [Google Scholar] [CrossRef] [PubMed]
  74. Ohnishi, T.; Kakimoto, K.; Bandow, K.; Lowenstein, C.J.; Daikuhara, Y.; Matsuguchi, T. Mature Hepatocyte Growth Factor/Scatter Factor on the Surface of Human Granulocytes Is Released by a Mechanism Involving Activated Factor Xa. J. Immunol. 2006, 176, 6945–6953. [Google Scholar] [CrossRef] [PubMed]
  75. Tashiro, K.; Hagiya, M.; Nishizawa, T.; Seki, T.; Shimonishi, M.; Shimizu, S.; Nakamura, T. Deduced primary structure of rat hepatocyte growth factor and expression of the mRNA in rat tissues. Proc. Natl. Acad. Sci. USA 1990, 87, 3200–3204. [Google Scholar] [CrossRef] [PubMed]
  76. Schwall, R.H.; Chang, L.Y.; Godowski, P.J.; Kahn, D.W.; Hillan, K.J.; Bauer, K.D.; Zioncheck, T.F. Heparin induces dimerization and confers proliferative activity onto the hepatocyte growth factor antagonists NK1 and NK2. J. Cell Biol. 1996, 133, 709–718. [Google Scholar] [CrossRef] [PubMed]
  77. Kawaguchi, M.; Kataoka, H. Mechanisms of Hepatocyte Growth Factor Activation in Cancer Tissues. Cancers 2014, 6, 1890–1904. [Google Scholar] [CrossRef] [PubMed]
  78. Kawaguchi, T.; Qin, L.; Shimomura, T.; Kondo, J.; Matsumoto, K.; Denda, K.; Kitamura, N. Purification and cloning of hepatocyte growth factor activator inhibitor type 2, a Kunitz-type serine protease inhibitor. J. Biol. Chem. 1997, 272, 27558–27564. [Google Scholar] [CrossRef] [PubMed]
  79. Shimomura, T.; Denda, K.; Kitamura, A.; Kawaguchi, T.; Kito, M.; Kondo, J.; Kagaya, S.; Qin, L.; Takata, H.; Miyazawa, K.; et al. Hepatocyte growth factor activator inhibitor, a novel Kunitz-type serine protease inhibitor. J. Biol. Chem. 1997, 272, 6370–6376. [Google Scholar] [CrossRef] [PubMed]
  80. Bernfield, M.; Gotte, M.; Park, P.W.; Reizes, O.; Fitzgerald, M.L.; Lincecum, J.; Zako, M. Functions of cell surface heparan sulfate proteoglycans. Ann. Rev. Biochem. 1999, 68, 729–777. [Google Scholar] [CrossRef] [PubMed]
  81. Delehedde, M.; Lyon, M.; Sergeant, N.; Rahmoune, H.; Fernig, D.G. Proteoglycans: Pericellular and cell surface multireceptors that integrate external stimuli in the mammary gland. J. Mammary Gland Biol. Neoplasia 2001, 6, 253–273. [Google Scholar] [CrossRef] [PubMed]
  82. Lander, A.D.; Nie, Q.; Wan, F.Y. Do morphogen gradients arise by diffusion? Dev. Cell 2002, 2, 785–796. [Google Scholar] [CrossRef]
  83. Garner, O.B.; Bush, K.T.; Nigam, K.B.; Yamaguchi, Y.; Xu, D.; Esko, J.D.; Nigam, S.K. Stage-Dependent Regulation of Mammary Ductal Branching by Heparan Sulfate and HGF-cMet Signaling. Dev. Biol. 2011, 355, 394–403. [Google Scholar] [CrossRef] [PubMed]
  84. Hartmann, G.; Prospero, T.; Brinkmann, V.; Ozcelik, Ö.; Winter, G.; Hepple, J.; Batley, S.; Bladt, F.; Sachs, M.; Birchmeier, C.; et al. Engineered mutants of HGF/SF with reduced binding to heparan sulphate proteoglycans, decreased clearance and enhanced activity in vivo. Curr. Biol. 1998, 8, 125–135. [Google Scholar] [CrossRef]
  85. Gross-Cohen, M.; Feld, S.; Doweck, I.; Neufeld, G.; Hasson, P.; Arvatz, G.; Barash, U.; Naroditsky, I.; Ilan, N.; Vlodavsky, I. Heparanase 2 attenuates head and neck tumor vascularity and growth. Cancer Res. 2016, 76, 2791–2801. [Google Scholar] [CrossRef] [PubMed]
  86. Levy-Adam, F.; Feld, S.; Cohen-Kaplan, V.; Shteingauz, A.; Gross, M.; Arvatz, G.; Naroditsky, I.; Ilan, N.; Doweck, I.; Vlodavsky, I. Heparanase 2 interacts with heparan sulfate with high affinity and inhibits heparanase activity. J. Biol. Chem. 2010, 285, 28010–28019. [Google Scholar] [CrossRef] [PubMed]
  87. Pan, J.; Qian, Y.; Zhou, X.; Lu, H.; Ramacciotti, E.; Zhang, L. Chemically oversulfated glycosaminoglycans are potent modulators of contact system activation and different cell signaling pathways. J. Biol. Chem. 2010, 285, 22966–22975. [Google Scholar] [CrossRef] [PubMed]
  88. Mazzone, M.; Basilico, C.; Cavassa, S.; Pennacchietti, S.; Risio, M.; Naldin, L.; Comoglio, P.M.; Michieli, P. An uncleavable form of pro-scatter factor suppresses tumor growth and dissemination in mice. J. Clin. Investig. 2004, 114, 1418–1432. [Google Scholar] [CrossRef] [PubMed]
  89. Naldini, L.; Vigna, E.; Bardelli, A.; Follenzi, A.; Galimi, F.; Comoglio, P.M. Biological activation of pro-HGF (hepatocyte growth factor) by urokinase is controlled by a stoichiometric reaction. J. Biol. Chem. 1995, 270, 603–611. [Google Scholar] [CrossRef] [PubMed]
  90. Park, M.; Dean, M.; Kaul, K.; Braun, M.J.; Gonda, M.A.; Vande Woude, G. Sequence of MET protooncogene cDNA has features characteristic of the tyrosine kinase family of growth-factor receptors. Proc. Natl. Acad. Sci. USA 1987, 84, 6379–6383. [Google Scholar] [CrossRef] [PubMed]
  91. Faletto, D.L.; Tsarfaty, I.; Kmiecik, T.E.; Gonzatti, M.; Suzuki, T.; Vande Woude, G.F. Evidence for non-covalent clusters of the c-met proto-oncogene product. Oncogene 1992, 7, 1149–1157. [Google Scholar] [PubMed]
  92. Gonzatti-Haces, M.; Park, M.; Dean, M.; Blair, D.G.; Vande Woude, G.F. The human met oncogene is a member of the tyrosine kinase family. Princess Takamatsu Symp. 1986, 17, 221–232. [Google Scholar] [PubMed]
  93. Chan, A.M.; King, H.W.; Deakin, E.A.; Tempest, P.R.; Hilkens, J.; Kroezen, V.; Edwards, D.R.; Wills, A.J.; Brookes, P.; Cooper, C.S. Characterization of the mouse met proto-oncogene. Oncogene 1988, 2, 593–599. [Google Scholar] [PubMed]
  94. Giordano, S.; Ponzetto, C.; Di Renzo, M.F.; Cooper, C.S.; Comoglio, P.M. Tyrosine kinase receptor indistinguishable from the c-met protein. Nature 1989, 339, 155–156. [Google Scholar] [CrossRef] [PubMed]
  95. Basilico, C.; Arnesano, A.; Galluzzo, M.; Comoglio, P.M.; Michieli, P. A high affinity hepatocyte growth factor-binding site in the immunoglobulin-like region of Met. J. Biol. Chem. 2008, 283, 21267–21277. [Google Scholar] [CrossRef] [PubMed]
  96. Longati, P.; Bardelli, A.; Ponzetto, C.; Naldini, L.; Comoglio, P.M. Tyrosines1234–1235 are critical for activation of the tyrosine kinase encoded by the MET proto-oncogene (HGF receptor). Oncogene 1994, 9, 49–57. [Google Scholar] [PubMed]
  97. Pelicci, G.; Giordano, S.; Zhen, Z.; Salcini, A.E.; Lanfrancone, L.; Bardelli, A.; Panayotou, G.; Waterfield, M.D.; Ponzetto, C.; Pelicci, P.G.; et al. The motogenic and mitogenic responses to HGF are amplified by the Shc adaptor protein. Oncogene 1995, 10, 1631–1638. [Google Scholar] [PubMed]
  98. Ponzetto, C.; Bardelli, A.; Zhen, Z.; Maina, F.; dalla Zonca, P.; Giordano, S.; Graziani, A.; Panayotou, G.; Comoglio, P.M. A multifunctional docking site mediates signaling and transformation by the hepatocyte growth factor/scatter factor receptor family. Cell 1994, 77, 261–271. [Google Scholar] [CrossRef]
  99. Weidner, K.M.; Di Cesare, S.; Sachs, M.; Brinkmann, V.; Behrens, J.; Birchmeier, W. Interaction between Gab1 and the c-Met receptor tyrosine kinase is responsible for epithelial morphogenesis. Nature 1996, 384, 173–176. [Google Scholar] [CrossRef] [PubMed]
  100. Garcia-Guzman, M.; Dolfi, F.; Zeh, K.; Vuori, K. Met-induced JNK activation is mediated by the adapter protein Crk and correlates with the Gab1-Crk signaling complex formation. Oncogene 1999, 18, 7775–7786. [Google Scholar] [CrossRef] [PubMed]
  101. Sakkab, D.; Lewitzky, M.; Posern, G.; Schaeper, U.; Sachs, M.; Birchmeier, W.; Feller, S.M. Signaling of hepatocyte growth factor/scatter factor (HGF) to the small GTPase Rap1 via the large docking protein Gab1 and the adapter protein CRKL. J. Biol. Chem. 2000, 275, 10772–10778. [Google Scholar] [CrossRef] [PubMed]
  102. Graziani, A.; Gramaglia, D.; Cantley, L.C.; Comoglio, P.M. The tyrosine-phosphorylated hepatocyte growth factor/scatter factor receptor associates with phosphatidylinositol 3-kinase. J. Biol. Chem. 1991, 266, 22087–22090. [Google Scholar] [PubMed]
  103. Boccaccio, C.; Ando, M.; Tamagnone, L.; Bardelli, A.; Michieli, P.; Battistini, C.; Comoglio, P.M. Induction of epithelial tubules by growth factor HGF depends on the STAT pathway. Nature 1998, 391, 285–288. [Google Scholar] [CrossRef] [PubMed]
  104. Graziani, A.; Gramaglia, D.; dalla Zonca, P.; Comoglio, P.M. Hepatocyte growth factor/scatter factor stimulates the Ras-guanine nucleotide exchanger. J. Biol. Chem. 1993, 268, 9165–9168. [Google Scholar] [PubMed]
  105. Fixman, E.D.; Fournier, T.M.; Kamikura, D.M.; Naujokas, M.A.; Park, M. Pathways downstream of Shc and Grb2 are required for cell transformation by the tpr-Met oncoprotein. J. Biol. Chem. 1996, 271, 13116–13122. [Google Scholar] [CrossRef] [PubMed]
  106. Furge, K.A.; Zhang, Y.W.; Vande Woude, G.F. Met receptor tyrosine kinase: Enhanced signaling through adapter proteins. Oncogene 2000, 19, 5582–5589. [Google Scholar] [CrossRef] [PubMed]
  107. Birchmeier, C.; Birchmeier, W.; Gherardi, E.; Vande Woude, G.F. Met, metastasis, motility and more. Nat. Rev. Mol. Cell. Biol. 2003, 4, 915–925. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, Y.W.; Wang, L.M.; Jove, R.; Vande Woude, G.F. Requirement of Stat3 signaling for HGF/SF-Met mediated tumorigenesis. Oncogene 2002, 21, 217–226. [Google Scholar] [CrossRef] [PubMed]
  109. Bertotti, A.; Comoglio, P.M.; Trusolino, L. Beta4 integrin activates a Shp2-Src signaling pathway that sustains HGF-induced anchorage-independent growth. J. Cell Biol. 2006, 175, 993–1003. [Google Scholar] [CrossRef] [PubMed]
  110. Orian-Rousseau, V.; Chen, L.; Sleeman, J.P.; Herrlich, P.; Ponta, H. CD44 is required for two consecutive steps in HGF/c-Met signaling. Genes Dev. 2002, 16, 3074–3086. [Google Scholar] [CrossRef] [PubMed]
  111. Kermorgant, S.; Parker, P.J. Receptor trafficking controls weak signal delivery: A strategy used by c-Met for STAT3 nuclear accumulation. J. Cell Biol. 2008, 182, 855–863. [Google Scholar] [CrossRef] [PubMed]
  112. Mak, H.H.; Peschard, P.; Lin, T.; Naujokas, M.A.; Zuo, D.; Park, M. Oncogenic activation of the Met receptor tyrosine kinase fusion protein, Tpr-Met, involves exclusion from the endocytic degradative pathway. Oncogene 2007, 26, 7213–7221. [Google Scholar] [CrossRef] [PubMed]
  113. Hammond, D.E.; Urbe, S.; Vande Woude, G.F.; Clague, M.J. Down-regulation of MET, the receptor for hepatocyte growth factor. Oncogene 2001, 20, 2761–2770. [Google Scholar] [CrossRef] [PubMed]
  114. Machide, M.; Hashigasako, A.; Matsumoto, K.; Nakamura, T. Contact inhibition of hepatocyte growth regulated by functional association of the c-Met/hepatocyte growth factor receptor and LAR protein-tyrosine phosphatase. J. Biol. Chem. 2006, 281, 8765–8772. [Google Scholar] [CrossRef] [PubMed]
  115. Sangwan, V.; Paliouras, G.N.; Abella, J.V.; Dube, N.; Monast, A.; Tremblay, M.L.; Park, M. Regulation of the Met receptor-tyrosine kinase by the protein-tyrosine phosphatase 1B and T-cell phosphatase. J. Biol. Chem. 2008, 283, 34374–34383. [Google Scholar] [CrossRef] [PubMed]
  116. Seiwert, T.Y.; Jagadeeswaran, R.; Faoro, L.; Janamanchi, V.; Nallasura, V.; El Dinali, M.; Yala, S.; Kanteti, R.; Cohen, E.E.W.; Lingen, M.W.; et al. The MET Receptor Tyrosine Kinase Is a Potential Novel Therapeutic Target for Head and Neck Squamous Cell Carcinoma. Cancer Res. 2009, 69, 3021–3031. [Google Scholar] [CrossRef] [PubMed]
  117. Ferracini, R.; Di Renzo, M.F.; Scotlandi, K.; Baldini, N.; Olivero, M.; Lollini, P.; Cremona, O.; Campanacci, M.; Comoglio, P.M. The Met/HGF receptor is over-expressed in human osteosarcomas and is activated by either a paracrine or an autocrine circuit. Oncogene 1995, 10, 739–749. [Google Scholar] [PubMed]
  118. Scotlandi, K.; Baldini, N.; Oliviero, M.; Di Renzo, M.F.; Martano, M.; Serra, M.; Manara, M.C.; Comoglio, P.M.; Ferracini, R. Expression of Met/hepatocyte growth factor receptor gene and malignant behavior of musculoskeletal tumors. Am. J. Pathol. 1996, 149, 1209–1219. [Google Scholar] [PubMed]
  119. Ferracini, R.; Olivero, M.; Di Renzo, M.F.; Martano, M.; De Giovanni, C.; Nanni, P.; Basso, G.; Scotlandi, K.; Lollini, P.L.; Comoglio, P.M. Retrogenic expression of the MET proto-oncogene correlates with the invasive phenotype of human rhabdomyosarcomas. Oncogene 1996, 12, 1697–1705. [Google Scholar] [PubMed]
  120. Knowles, L.M.; Stabile, L.P.; Egloff, A.M.; Rothstein, M.E.; Thomas, S.M.; Gubish, C.T.; Lerner, E.C.; Seethala, R.R.; Suzuki, S.; Quesnelle, K.M.; et al. HGF and c-Met Participate in Paracrine Tumorigenic Pathways in Head and Neck Squamous Cell Cancer. Clin. Cancer Res. 2009, 15, 3740–3750. [Google Scholar] [CrossRef] [PubMed]
  121. Ghadjar, P.; Blank-Liss, W.; Simcock, M.; Hegyi, I.; Beer, K.T.; Moch, H.; Aebersold, D.M.; Zimmer, Y. MET Y1253D-activating point mutation and development of distant metastasis in advanced head and neck cancers. Clin. Exp. Metastasis 2009, 26, 809–815. [Google Scholar] [CrossRef] [PubMed]
  122. Di Renzo, M.F.; Olivero, M.; Giacomini, A.; Porte, H.; Chastre, E.; Mirossay, L.; Nordlinger, B.; Bretti, S.; Bottardi, S.; Giordano, S. Overexpression and amplification of the met/HGF receptor gene during the progression of colorectal cancer. Clin. Cancer Res. 1995, 1, 147–154. [Google Scholar] [PubMed]
  123. Takeo, S.; Arai, H.; Kusano, N.; Harada, T.; Furuya, T.; Kawauchi, S.; Oga, A.; Hirano, T.; Yoshida, T.; Okita, K.; et al. Examination of oncogene amplification by genomic DNA microarray in hepatocellular carcinomas: Comparison with comparative genomic hybridization analysis. Cancer Genet. Cytogenet. 2001, 130, 127–132. [Google Scholar] [CrossRef]
  124. Schmidt, L.; Duh, F.M.; Chen, F.; Kishida, T.; Glenn, G.; Choyke, P.; Scherer, S.W.; Zhuang, Z.; Lubensky, I.; Dean, M.; et al. Germline and somatic mutations in the tyrosine kinase domain of the MET proto-oncogene in papillary renal carcinomas. Nat. Genet. 1997, 16, 68–73. [Google Scholar] [CrossRef] [PubMed]
  125. Park, W.S.; Dong, S.M.; Kim, S.Y.; Na, E.Y.; Shin, M.S.; Pi, J.H.; Kim, B.J.; Bae, J.H.; Hong, Y.K.; Lee, K.S.; et al. Somatic mutations in the kinase domain of the Met/hepatocyte growth factor receptor gene in childhood hepatocellular carcinomas. Cancer Res. 1999, 59, 307–310. [Google Scholar] [PubMed]
  126. Aebersold, D.M.; Landt, O.; Berthou, S.; Gruber, G.; Beer, K.T.; Greiner, R.H.; Zimmer, Y. Prevalence and clinical impact of Met Y1253D-activating point mutation in radiotherapy-treated squamous cell cancer of the oropharynx. Oncogene 2003, 22, 8519–8523. [Google Scholar] [CrossRef] [PubMed]
  127. Pilotto, S.; Gkountakos, A.; Carbognin, L.; Scarpa, A.; Tortora, G.; Bria, E. MET exon 14 juxtamembrane splicing mutations: Clinical and therapeutical perspectives for cancer therapy. Ann. Transl. Med. 2017, 5, 2. [Google Scholar] [CrossRef] [PubMed]
  128. Ruco, L.P.; Ranalli, T.; Marzullo, A.; Bianco, P.; Prat, M.; Comoglio, P.M.; Baroni, C.D. Expression of Met protein in thyroid tumours. J. Pathol. 1996, 180, 266–270. [Google Scholar] [CrossRef]
  129. Di Renzo, M.F.; Olivero, M.; Katsaros, D.; Crepaldi, T.; Gaglia, P.; Zola, P.; Sismondi, P.; Comoglio, P.M. Overexpression of the Met/HGF receptor in ovarian cancer. Int. J. Cancer 1994, 58, 658–662. [Google Scholar] [CrossRef] [PubMed]
  130. Tapper, J.; Kettunen, E.; El-Rifai, W.; Seppala, M.; Andersson, L.C.; Knuutila, S. Changes in gene expression during progression of ovarian carcinoma. Cancer Genet. Cytogenet. 2001, 128, 1–6. [Google Scholar] [CrossRef]
  131. Di Renzo, M.F.; Poulsom, R.; Olivero, M.; Comoglio, P.M.; Lemoine, N.R. Expression of the Met/hepatocyte growth factor receptor in human pancreatic cancer. Cancer Res. 1995, 55, 1129–1138. [Google Scholar] [PubMed]
  132. Humphrey, P.A.; Zhu, X.; Zarnegar, R.; Swanson, P.E.; Ratliff, T.L.; Vollmer, R.T.; Day, M.L. Hepatocyte growth factor and its receptor (c-MET) in prostatic carcinoma. Am. J. Pathol. 1995, 147, 386–396. [Google Scholar] [PubMed]
  133. Natali, P.G.; Prat, M.; Nicotra, M.R.; Bigotti, A.; Olivero, M.; Comoglio, P.M.; Di Renzo, M.F. Overexpression of the met/HGF receptor in renal cell carcinomas. Int. J. Cancer 1996, 69, 212–217. [Google Scholar] [CrossRef]
  134. Tavian, D.; De Petro, G.; Benetti, A.; Portolani, N.; Giulini, S.M.; Barlati, S. u-PA and c-MET mRNA expression is co-ordinately enhanced while hepatocyte growth factor mRNA is down-regulated in human hepatocellular carcinoma. Int. J. Cancer 2000, 87, 644–649. [Google Scholar] [CrossRef]
  135. Taniguchi, K.; Yonemura, Y.; Nojima, N.; Hirono, Y.; Fushida, S.; Fujimura, T.; Miwa, K.; Endo, Y.; Yamamoto, H.; Watanabe, H. The relation between the growth patterns of gastric carcinoma and the expression of hepatocyte growth factor receptor (c-met), autocrine motility factor receptor, and urokinase-type plasminogen activator receptor. Cancer 1998, 82, 2112–2122. [Google Scholar] [CrossRef]
  136. Porte, H.; Triboulet, J.P.; Kotelevets, L.; Carrat, F.; Prevot, S.; Nordlinger, B.; DiGioia, Y.; Wurtz, A.; Comoglio, P.; Gespach, C.; et al. Overexpression of stromelysin-3, BM-40/SPARC, and MET genes in human esophageal carcinoma: Implications for prognosis. Clin. Cancer Res. 1998, 4, 1375–1382. [Google Scholar] [PubMed]
  137. Camp, R.L.; Rimm, E.B.; Rimm, D.L. Met expression is associated with poor outcome in patients with axillary lymph node negative breast carcinoma. Cancer 1999, 86, 2259–2265. [Google Scholar] [CrossRef]
  138. Wielenga, V.J.; van der Voort, R.; Taher, T.E.; Smit, L.; Beuling, E.A.; van Krimpen, C.; Spaargaren, M.; Pals, S.T. Expression of c-Met and heparan-sulfate proteoglycan forms of CD44 in colorectal cancer. Am. J. Pathol. 2000, 157, 1563–1573. [Google Scholar] [CrossRef]
  139. Morello, S.; Olivero, M.; Aimetti, M.; Bernardi, M.; Berrone, S.; Di Renzo, M.F.; Giordano, S. MET receptor is overexpressed but not mutated in oral squamous cell carcinomas. J. Cell. Physiol. 2001, 189, 285–290. [Google Scholar] [CrossRef] [PubMed]
  140. Hays, J.L.; Watowich, S.J. Oligomerization-induced modulation of TPR-MET tyrosine kinase activity. J. Biol. Chem. 2003, 278, 27456–27463. [Google Scholar] [CrossRef] [PubMed]
  141. Marshall, D.D.; Kornberg, L.J. Overexpression of scatter factor and its receptor (c-met) in oral squamous cell carcinoma. Laryngoscope 1998, 108, 1413–1417. [Google Scholar] [PubMed]
  142. Bladt, F.; Riethmacher, D.; Isenmann, S.; Aguzzi, A.; Birchmeier, C. Essential role for the c-met receptor in the migration of myogenic precursor cells into the limb bud. Nature 1995, 376, 768–771. [Google Scholar] [CrossRef] [PubMed]
  143. Jeffers, M.; Rong, S.; Vande Woude, G.F. Enhanced tumorigenicity and invasion-metastasis by hepatocyte growth factor/scatter factor-met signalling in human cells concomitant with induction of the urokinase proteolysis network. Mol. Cell. Biol. 1996, 16, 1115–1125. [Google Scholar] [CrossRef] [PubMed]
  144. Jiang, W.G.; Davies, G.; Martin, T.A.; Parr, C.; Watkins, G.; Mansel, R.E.; Mason, M.D. The potential lymphangiogenic effects of hepatocyte growth factor/scatter factor in vitro and in vivo. Int. J. Mol. Med. 2005, 16, 723–728. [Google Scholar] [PubMed]
  145. Kajiya, K.; Hirakawa, S.; Ma, B.; Drinnenberg, I.; Detmar, M. Hepatocyte growth factor promotes lymphatic vessel formation and function. EMBO J. 2005, 24, 2885–2895. [Google Scholar] [CrossRef] [PubMed]
  146. Sengupta, S.; Gherardi, E.; Sellers, L.A.; Wood, J.M.; Sasisekharan, R.; Fan, T.P. Hepatocyte growth factor/scatter factor can induce angiogenesis independently of vascular endothelial growth factor. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 69–75. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, Y.W.; Su, Y.; Volpert, O.V.; Vande Woude, G.F. Hepatocyte growth factor/scatter factor mediates angiogenesis through positive VEGF and negative thrombospondin 1 regulation. Proc. Natl. Acad. Sci. USA 2003, 100, 12718–12723. [Google Scholar] [CrossRef] [PubMed]
  148. Goldson, T.M.; Han, Y.; Knight, K.B.; Weiss, H.L.; Resto, V.A. Clinicopathological predictors of lymphatic metastasis in HNSCC: Implications for molecular mechanisms of metastatic disease. J. Exp. Ther. Oncol. 2010, 8, 211–221. [Google Scholar] [PubMed]
  149. Choe, J.-Y.; Yun, J.Y.; Nam, S.-J.; Kim, J.E. Expression of c-Met Is Different along the Location and Associated with Lymph Node Metastasis of Head and Neck Carcinoma. Korean J. Pathol. 2012, 46, 515–522. [Google Scholar] [CrossRef] [PubMed]
  150. Cortesina, G.; Martone, T.; Galeazzi, E.; Olivero, M.; De Stefani, A.; Bussi, M.; Valente, G.; Comoglio, P.M.; Di Renzo, M.F. Staging of head and neck squamous cell carcinoma using the MET oncogene product as marker of tumor cells in lymph node metastases. Int. J. Cancer 2000, 89, 286–292. [Google Scholar] [CrossRef]
  151. Galeazzi, E.; Olivero, M.; Gervasio, F.C.; De Stefani, A.; Valente, G.; Comoglio, P.M.; Di Renzo, M.F.; Cortesina, G. Detection of MET oncogene/hepatocyte growth factor receptor in lymph node metastases from head and neck squamous cell carcinomas. Eur. Arch. Otorhinolaryngol. 1997, 254 (Suppl. 1), S138–S143. [Google Scholar] [CrossRef] [PubMed]
  152. Tao, X.; Hill, K.S.; Gaziova, I.; Sastry, S.K.; Qui, S.; Szaniszlo, P.; Fennewald, S.; Resto, V.A.; Elferink, L.A. Silencing Met receptor tyrosine kinase signaling decreased oral tumor growth and increased survival of nude mice. Oral Oncol. 2014, 50, 104–112. [Google Scholar] [CrossRef] [PubMed]
  153. Hanahan, D.; Weinberg, R.A. The Hallmarks of Cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef]
  154. Shen, C.-J.; Chan, S.-H.; Lee, C.-T.; Huang, W.-C.; Tsai, J.-P.; Chen, B.-K. Oleic acid-induced ANGPTL4 enhances head and neck squamous cell carcinoma anoikis resistance and metastasis via up-regulation of fibronectin. Cancer Lett. 2017, 386, 110–122. [Google Scholar] [CrossRef] [PubMed]
  155. Zeng, Q.; Chen, S.; You, Z.; Yang, F.; Carey, T.E.; Saims, D.; Wang, C.Y. Hepatocyte growth factor inhibits anoikis in head and neck squamous cell carcinoma cells by activation of ERK and Akt signaling independent of NFkappa B. J. Biol. Chem. 2002, 277, 25203–25208. [Google Scholar] [CrossRef] [PubMed]
  156. Baum, B.; Settleman, J.; Quinlan, M.P. Transitions between epithelial and mesenchymal states in development and disease. Semin. Cell Dev. Biol. 2008, 19, 294–308. [Google Scholar] [CrossRef] [PubMed]
  157. Komiya, Y.; Habas, R. Wnt signal transduction pathways. Organogenesis 2008, 4, 68–75. [Google Scholar] [CrossRef] [PubMed]
  158. Vermeulen, L.; De Sousa, E.M.F.; van der Heijden, M.; Cameron, K.; de Jong, J.H.; Borovski, T.; Tuynman, J.B.; Todaro, M.; Merz, C.; Rodermond, H.; et al. Wnt activity defines colon cancer stem cells and is regulated by the microenvironment. Nat. Cell Biol. 2010, 12, 468–476. [Google Scholar] [CrossRef] [PubMed]
  159. Sun, S.; Liu, S.; Duan, S.Z.; Zhang, L.; Zhou, H.; Hu, Y.; Zhou, X.; Shi, C.; Zhou, R.; Zhang, Z. Targeting the c-Met/FZD8 signaling axis eliminates patient-derived cancer stem-like cells in head and neck squamous carcinomas. Cancer Res. 2014, 74, 7546–7559. [Google Scholar] [CrossRef] [PubMed]
  160. Stabile, L.P.; He, G.; Lui, V.W.; Thomas, S.; Henry, C.; Gubish, C.T.; Joyce, S.; Quesnelle, K.M.; Siegfried, J.M.; Grandis, J.R. c-Src activation mediates erlotinib resistance in head and neck cancer by stimulating c-Met. Clin. Cancer Res. 2013, 19, 380–392. [Google Scholar] [CrossRef] [PubMed]
  161. Sen, B.; Peng, S.; Saigal, B.; Williams, M.D.; Johnson, F.M. Distinct interactions between c-Src and c-Met in mediating resistance to c-Src inhibition in head and neck cancer. Clin. Cancer Res. 2011, 17, 514–524. [Google Scholar] [CrossRef] [PubMed]
  162. Bhowmick, N.A.; Chytil, A.; Plieth, D.; Gorska, A.E.; Dumont, N.; Shappell, S.; Washington, M.K.; Neilson, E.G.; Moses, H.L. TGF-ß Signaling in Fibroblasts Modulates the Oncogenic Potential of Adjacent Epithelia. Science 2004, 303, 848–851. [Google Scholar] [CrossRef] [PubMed]
  163. Hoot, K.E.; Oka, M.; Han, G.; Bottinger, E.; Zhang, Q.; Wang, X.J. HGF upregulation contributes to angiogenesis in mice with keratinocyte-specific Smad2 deletion. J. Clin. Investig. 2010, 120, 3606–3616. [Google Scholar] [CrossRef] [PubMed]
  164. Cheng, H.; Fertig, E.; Ozawa, H.; Hatakeyama, H.; Considine, M.; Perez, J.; Ochs, M.; Chung, C.H. Abstract 1889: Smad4 inactivation correlates with EMT and cetuximab resistance in head and neck squamous cell carcinoma. Cancer Res. 2012, 72, 1889. [Google Scholar] [CrossRef]
  165. Spector, N.L.; Xia, W.; Burris, H., 3rd; Hurwitz, H.; Dees, E.C.; Dowlati, A.; O’Neil, B.; Overmoyer, B.; Marcom, P.K.; Blackwell, K.L.; et al. Study of the biologic effects of lapatinib, a reversible inhibitor of ErbB1 and ErbB2 tyrosine kinases, on tumor growth and survival pathways in patients with advanced malignancies. J. Clin. Oncol. 2005, 23, 2502–2512. [Google Scholar] [CrossRef] [PubMed]
  166. Sergina, N.V.; Rausch, M.; Wang, D.; Blair, J.; Hann, B.; Shokat, K.M.; Moasser, M.M. Escape from HER-family tyrosine kinase inhibitor therapy by the kinase-inactive HER3. Nature 2007, 445, 437–441. [Google Scholar] [CrossRef] [PubMed]
  167. Hsieh, A.C.; Moasser, M.M. Targeting HER proteins in cancer therapy and the role of the non-target HER3. Br. J. Cancer 2007, 97, 453–457. [Google Scholar] [CrossRef] [PubMed]
  168. Bertotti, A.; Burbridge, M.F.; Gastaldi, S.; Galimi, F.; Torti, D.; Medico, E.; Giordano, S.; Corso, S.; Rolland-Valognes, G.; Lockhart, B.P.; et al. Only a Subset of Met-Activated Pathways Are Required to Sustain Oncogene Addiction. Sci. Signal. 2009, 2, ra80. [Google Scholar] [CrossRef] [PubMed]
  169. Liu, L.; Shi, H.; Liu, Y.; Anderson, A.; Peterson, J.; Greger, J.; Martin, A.M.; Gilmer, T.M. Synergistic effects of foretinib with HER-targeted agents in MET and HER1- or HER2-coactivated tumor cells. Mol. Cancer Ther. 2011, 10, 518–530. [Google Scholar] [CrossRef] [PubMed]
  170. Xu, H.; Stabile, L.P.; Gubish, C.T.; Gooding, W.E.; Grandis, J.R.; Siegfried, J.M. Dual blockade of EGFR and c-Met abrogates redundant signaling and proliferation in head and neck carcinoma cells. Clin. Cancer Res. 2011, 17, 4425–4438. [Google Scholar] [CrossRef] [PubMed]
  171. Tepper, S.R.; Zuo, Z.; Khattri, A.; Heß, J.; Seiwert, T.Y. Growth factor expression mediates resistance to EGFR inhibitors in head and neck squamous cell carcinomas. Oral Oncol. 2016, 56, 62–70. [Google Scholar] [CrossRef] [PubMed]
  172. Wheeler, D.L.; Huang, S.; Kruser, T.J.; Nechrebecki, M.M.; Armstrong, E.A.; Benavente, S.; Gondi, V.; Hsu, K.T.; Harari, P.M. Mechanisms of acquired resistance to cetuximab: Role of HER (ErbB) family members. Oncogene 2008, 27, 3944–3956. [Google Scholar] [CrossRef] [PubMed]
  173. Rubin Grandis, J.; Melhem, M.F.; Gooding, W.E.; Day, R.; Holst, V.A.; Wagener, M.M.; Drenning, S.D.; Tweardy, D.J. Levels of TGF-alpha and EGFR protein in head and neck squamous cell carcinoma and patient survival. J. Natl. Cancer Inst. 1998, 90, 824–832. [Google Scholar] [CrossRef] [PubMed]
  174. Brand, T.M.; Iida, M.; Wheeler, D.L. Molecular mechanisms of resistance to the EGFR monoclonal antibody cetuximab. Cancer Biol. Ther. 2011, 11, 777–792. [Google Scholar] [CrossRef] [PubMed]
  175. Soulieres, D.; Senzer, N.N.; Vokes, E.E.; Hidalgo, M.; Agarwala, S.S.; Siu, L.L. Multicenter phase II study of erlotinib, an oral epidermal growth factor receptor tyrosine kinase inhibitor, in patients with recurrent or metastatic squamous cell cancer of the head and neck. J. Clin. Oncol. 2004, 22, 77–85. [Google Scholar] [CrossRef] [PubMed]
  176. Stommel, J.M.; Kimmelman, A.C.; Ying, H.; Nabioullin, R.; Ponugoti, A.H.; Wiedemeyer, R.; Stegh, A.H.; Bradner, J.E.; Ligon, K.L.; Brennan, C.; et al. Coactivation of receptor tyrosine kinases affects the response of tumor cells to targeted therapies. Science 2007, 318, 287–290. [Google Scholar] [CrossRef] [PubMed]
  177. Cui, J.J.; Tran-Dubé, M.; Shen, H.; Nambu, M.; Kung, P.-P.; Pairish, M.; Jia, L.; Meng, J.; Funk, L.; Botrous, I.; et al. Structure Based Drug Design of Crizotinib (PF-02341066), a Potent and Selective Dual Inhibitor of Mesenchymal—Epithelial Transition Factor (c-MET) Kinase and Anaplastic Lymphoma Kinase (ALK). J. Med. Chem. 2011, 54, 6342–6363. [Google Scholar] [CrossRef] [PubMed]
  178. Baschnagel, A.M.; Galoforo, S.; Thibodeau, B.J.; Ahmed, S.; Nirmal, S.; Akervall, J.; Wilson, G.D. Crizotinib Fails to Enhance the Effect of Radiation in Head and Neck Squamous Cell Carcinoma Xenografts. Anticancer Res. 2015, 35, 5973–5982. [Google Scholar] [PubMed]
  179. Liu, X.; Wang, Q.; Yang, G.; Marando, C.; Koblish, H.K.; Hall, L.M.; Fridman, J.S.; Behshad, E.; Wynn, R.; Li, Y.; et al. A Novel Kinase Inhibitor, INCB28060, Blocks c-MET-Dependent Signaling, Neoplastic Activities, and Cross-Talk with EGFR and HER-3. Clin. Cancer Res. 2011, 17, 7127–7138. [Google Scholar] [CrossRef] [PubMed]
  180. Nakagawa, T.; Tohyama, O.; Yamaguchi, A.; Matsushima, T.; Takahashi, K.; Funasaka, S.; Shirotori, S.; Asada, M.; Obaishi, H. E7050: A dual c-Met and VEGFR-2 tyrosine kinase inhibitor promotes tumor regression and prolongs survival in mouse xenograft models. Cancer Sci. 2010, 101, 210–215. [Google Scholar] [CrossRef] [PubMed]
  181. Li, S.; Zhao, Y.; Wang, K.; Gao, Y.; Han, J.; Cui, B.; Gong, P. Discovery of novel 4-(2-fluorophenoxy)quinoline derivatives bearing 4-oxo-1,4-dihydrocinnoline-3-carboxamide moiety as c-Met kinase inhibitors. Bioorg. Med. Chem. 2013, 21, 2843–2855. [Google Scholar] [CrossRef] [PubMed]
  182. Chen, G.-Z.; Dai, W.-S.; Zhu, H.-C.; Song, H.-M.; Yang, X.; Wang, Y.-D.; Min, H.; Lu, Q.; Liu, S.; Sun, X.-C.; et al. Foretinib Enhances the Radiosensitivity in Esophageal Squamous Cell Carcinoma by Inhibiting Phosphorylation of c-Met. J. Cancer 2017, 8, 983–992. [Google Scholar] [CrossRef] [PubMed]
  183. Seiwert, T.; Sarantopoulos, J.; Kallender, H.; McCallum, S.; Keer, H.N.; Blumenschein, G. Phase II trial of single-agent foretinib (GSK1363089) in patients with recurrent or metastatic squamous cell carcinoma of the head and neck. Investig. New Drugs 2013, 31, 417–424. [Google Scholar] [CrossRef] [PubMed]
  184. Kumar, D.; Kandl, C.; Hamilton, C.D.; Shnayder, Y.; Tsue, T.T.; Kakarala, K.; Ledgerwood, L.; Sun, X.S.; Huang, H.J.; Girod, D.; et al. Mitigation of tumor-associated fibroblast-facilitated head and neck cancer progression with anti-hepatocyte growth factor antibody ficlatuzumab. JAMA Otolaryngol. Head Neck Surg. 2015, 141, 1133–1139. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Pro-HGF (hepatocyte growth factor) is secreted as an inactive, single-chain precursor. Activation is stimulated by wounds, chemical agents, radiation, hypoxia, and oxidative stress among other cellular disruptions. It is activated by cleavage of the bond at Arg494-Val495 by a cellular protease to yield active, heterodimeric HGF with α (consisting of four Kringle domains and a hairpin loop) and β-chains linked by a disulfide bond. Extracellularly, the c-Met receptor consist of the sema domain (with disulfide linked α-β domains), the PSI (Plexin, Semaphorin, and Integrin) domain and four IPT (Ig-like, plexins, transcription factors) domains. Intracellular mesenchymal–epithelial transition factor (c-Met) domains include a kinase domain flanked by the juxtamembrane domain and the multifunctional docking site. High affinity binding occurs between the hairpin loop and first Kringle domain of the α-chain of HGF and the third and fourth IPT domain of c-Met. Low affinity binding occurs between the β-chain of HGF and the sema domain of c-Met.
Figure 1. Pro-HGF (hepatocyte growth factor) is secreted as an inactive, single-chain precursor. Activation is stimulated by wounds, chemical agents, radiation, hypoxia, and oxidative stress among other cellular disruptions. It is activated by cleavage of the bond at Arg494-Val495 by a cellular protease to yield active, heterodimeric HGF with α (consisting of four Kringle domains and a hairpin loop) and β-chains linked by a disulfide bond. Extracellularly, the c-Met receptor consist of the sema domain (with disulfide linked α-β domains), the PSI (Plexin, Semaphorin, and Integrin) domain and four IPT (Ig-like, plexins, transcription factors) domains. Intracellular mesenchymal–epithelial transition factor (c-Met) domains include a kinase domain flanked by the juxtamembrane domain and the multifunctional docking site. High affinity binding occurs between the hairpin loop and first Kringle domain of the α-chain of HGF and the third and fourth IPT domain of c-Met. Low affinity binding occurs between the β-chain of HGF and the sema domain of c-Met.
Cancers 09 00169 g001
Figure 2. Hepatocyte growth factor (HGF) binding c-Met activates signaling cascades including Src, CT10 (chicken tumor virus No. 10) regulator of kinase (Crk), mitogen-activated protein kinase/ extracellular signal-regulated kinases (Mapk/Erk), and phosphotidylinsitol-3-OH kinase (PI3K), leading to loss of tumor cell apoptosis, survival, transformation, invasion, and motility. Downregulation of c-Met are highlighted in blue. c-Met signaling that promotes cell survival and apoptotic resistance are highlighted in lavender. c-Met signaling that promotes transformation is highlighted in orange. c-Met signaling that promotes invasion and motility is highlighted in green. CBL: casitas B-lineage lymphoma; FAK: focal adhesion kinase; STAT3: signal transducer and activator of transcription-3; PKC: protein kinase c; PIP3:phosphatidylinositol (3,4,5)-trisphosphate; PLC: phospholipase C; JNK: jun n-terminal kinase; Gab1: Grb2-associated adaptor protein; Grb2: growth factor receptor-bound protein 2; Akt: alpha serine/threonine-protein kinase; SOS: Ras guanine nucleotide exchange factor son-of-sevenless; Shp2: Src kinase, and Src homology region-2 containing protein tyrosine phosphatase 2; Ras: rat sarcoma; Raf: rapidly accelerated fibrosarcoma; BAD: Bcl-2 associated death promoter; p53: cellular tumor antigen p53.
Figure 2. Hepatocyte growth factor (HGF) binding c-Met activates signaling cascades including Src, CT10 (chicken tumor virus No. 10) regulator of kinase (Crk), mitogen-activated protein kinase/ extracellular signal-regulated kinases (Mapk/Erk), and phosphotidylinsitol-3-OH kinase (PI3K), leading to loss of tumor cell apoptosis, survival, transformation, invasion, and motility. Downregulation of c-Met are highlighted in blue. c-Met signaling that promotes cell survival and apoptotic resistance are highlighted in lavender. c-Met signaling that promotes transformation is highlighted in orange. c-Met signaling that promotes invasion and motility is highlighted in green. CBL: casitas B-lineage lymphoma; FAK: focal adhesion kinase; STAT3: signal transducer and activator of transcription-3; PKC: protein kinase c; PIP3:phosphatidylinositol (3,4,5)-trisphosphate; PLC: phospholipase C; JNK: jun n-terminal kinase; Gab1: Grb2-associated adaptor protein; Grb2: growth factor receptor-bound protein 2; Akt: alpha serine/threonine-protein kinase; SOS: Ras guanine nucleotide exchange factor son-of-sevenless; Shp2: Src kinase, and Src homology region-2 containing protein tyrosine phosphatase 2; Ras: rat sarcoma; Raf: rapidly accelerated fibrosarcoma; BAD: Bcl-2 associated death promoter; p53: cellular tumor antigen p53.
Cancers 09 00169 g002
Figure 3. Common c-Met mutations in head and neck squamous cell carcinoma: T230M, E168D, and N375S affect the extracellular sema domain. Intracellularly, T1010I and R988C affect the juxtamedullary (JM) domain, while Y1235D, Y1230C, T1275I, and V1333I affect the kinase domain. PSI: plexin-semaphorin-integrin domain; IPT: immunoglobulin-plexin-transcription domain; KD: kinase domain; MDF: multifunctional docking site.
Figure 3. Common c-Met mutations in head and neck squamous cell carcinoma: T230M, E168D, and N375S affect the extracellular sema domain. Intracellularly, T1010I and R988C affect the juxtamedullary (JM) domain, while Y1235D, Y1230C, T1275I, and V1333I affect the kinase domain. PSI: plexin-semaphorin-integrin domain; IPT: immunoglobulin-plexin-transcription domain; KD: kinase domain; MDF: multifunctional docking site.
Cancers 09 00169 g003
Table 1. Inhibitors of the HGF/c-Met axis under investigation in HNSCC.
Table 1. Inhibitors of the HGF/c-Met axis under investigation in HNSCC.
CompoundClassIC50/KiCo-TreatmentsStage of Development in HNSCC
c-MetCell
CrizotinibIa2.0 nM (Ki)4.1–4.7 μMGefitinibPreclinical
RadiationPreclinical
CapmatinibIb0.13 nM (IC50)1.2–12.4 nMCetuximabPhase Ib/II (NCT02205398)
GolvatinibII14 nM (IC50)6.2 nM–4.3 μMCetuximabPhase I/II (NCT01332266)
Cisplatin and CapecitabinePhase I/II * (NCT01355302)
ForetinibII1.16 nM (IC50)0.61–0.79 μMErlotinibPreclinical
RadiationPreclinical
N/APhase I/II * (NCT00725764)
FiclatuzumabMonoclonal AntibodyN/ACetuximabPhase I (NCT02277197)
* Denotes the trial did not progress into phase II. c-Met: mesenchymal-epithelial transition factor; HNSCC: head and neck squamous cell carcinoma.
Table 2. Structures of small molecule inhibitors of c-Met.
Table 2. Structures of small molecule inhibitors of c-Met.
CrizotinibGolvatinib
Cancers 09 00169 i001 Cancers 09 00169 i002
CapmatinibForetinib
Cancers 09 00169 i003 Cancers 09 00169 i004

Share and Cite

MDPI and ACS Style

Arnold, L.; Enders, J.; Thomas, S.M. Activated HGF-c-Met Axis in Head and Neck Cancer. Cancers 2017, 9, 169. https://doi.org/10.3390/cancers9120169

AMA Style

Arnold L, Enders J, Thomas SM. Activated HGF-c-Met Axis in Head and Neck Cancer. Cancers. 2017; 9(12):169. https://doi.org/10.3390/cancers9120169

Chicago/Turabian Style

Arnold, Levi, Jonathan Enders, and Sufi Mary Thomas. 2017. "Activated HGF-c-Met Axis in Head and Neck Cancer" Cancers 9, no. 12: 169. https://doi.org/10.3390/cancers9120169

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