Next Article in Journal
Hepatocyte Growth Factor from a Clinical Perspective: A Pancreatic Cancer Challenge
Next Article in Special Issue
Hypermethylation of MAPK13 Promoter in Oesophageal Squamous Cell Carcinoma Is Associated with Loss of p38δ MAPK Expression
Previous Article in Journal
Supportive Management of Mucositis and Metabolic Derangements in Head and Neck Cancer Patients
Previous Article in Special Issue
Dichotomy in the Epigenetic Mark Lysine Acetylation is Critical for the Proliferation of Prostate Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting RTK Signaling Pathways in Cancer

The John van Geest Cancer Research Centre, School of Science and Technology, Nottingham Trent University, Clifton Lane, NG11 8NS Nottingham, UK
Cancers 2015, 7(3), 1758-1784; https://doi.org/10.3390/cancers7030860
Submission received: 4 May 2015 / Revised: 24 August 2015 / Accepted: 26 August 2015 / Published: 3 September 2015
(This article belongs to the Special Issue Cancer Cell Proliferation)

Abstract

:
The RAS/MAP kinase and the RAS/PI3K/AKT pathways play a key role in the regulation of proliferation, differentiation and survival. The induction of these pathways depends on Receptor Tyrosine Kinases (RTKs) that are activated upon ligand binding. In cancer, constitutive and aberrant activations of components of those pathways result in increased proliferation, survival and metastasis. For instance, mutations affecting RTKs, Ras, B-Raf, PI3K and AKT are common in perpetuating the malignancy of several types of cancers and from different tissue origins. Therefore, these signaling pathways became prime targets for cancer therapy. This review aims to provide an overview about the most frequently encountered mutations, the pathogenesis that results from such mutations and the known therapeutic strategies developed to counteract their aberrant functions.

1. Introduction

Receptor tyrosine kinases (RTKs) are a family of cell surface receptors, which act as receptors for growth factors, hormones, cytokines, neurotrophic factors and other extracellular signaling molecules. RTKs mediate key signaling pathways that are involved in cell proliferation, differentiation, survival and cell migration [1]. The RTK family comprises several subfamilies which include, among others, epidermal growth factor receptors (EGFRs), fibroblast growth factor receptors (FGFRs), insulin and insulin-like growth factor receptors (IR and IGFR), platelet-derived growth factor receptors (PDGFRs), vascular endothelial growth factor receptors (VEGFRs), hepatocyte growth factor receptors (HGFRs), and proto-oncogene c-KIT [2,3]. RTKs monomers are organized into an extracellular (N-terminal), a transmembrane and a cytoplasmic kinase domain. They are activated via ligand-induced dimerization that results in receptor auto-phosphorylation and tyrosine activation of RTKs’ substrates including phospholipase C-γ, mitogen-activated protein kinases and phosphatidylinositol 3-kinase [4,5,6] (Figure 1).
Figure 1. Schematic representation of Receptor Tyrosine Kinase and downstream signaling pathways. Receptor Tyrosine Kinases are auto-phosphorylated upon ligand binding, which results in the activation of Ras and induction of serine/threonine kinase Raf. Raf phosphorylates Mek1/2 which in turn phosphorylate and activate Erk1/2. Raf also activates MAP3 kinases that activate MKK4/7, MKKK3/6 and MEK5, which activates JNK1/2, p38 and ERK5, consecutively. MAP3Ks are also activated by inflammatory cytokines, oxidative stress and UV radiation. PI3K is activated by RTK autophosphorylation and results in the activation of Akt which also induces mTOR within the mTORC1 complex. Akt is also regulated by mTORC2 complex. PLCγ activation leads to Ca+2 mobilization and to the activation of PKC. These events play an essential role in proliferation, differentiation, survival and cell migration.
Figure 1. Schematic representation of Receptor Tyrosine Kinase and downstream signaling pathways. Receptor Tyrosine Kinases are auto-phosphorylated upon ligand binding, which results in the activation of Ras and induction of serine/threonine kinase Raf. Raf phosphorylates Mek1/2 which in turn phosphorylate and activate Erk1/2. Raf also activates MAP3 kinases that activate MKK4/7, MKKK3/6 and MEK5, which activates JNK1/2, p38 and ERK5, consecutively. MAP3Ks are also activated by inflammatory cytokines, oxidative stress and UV radiation. PI3K is activated by RTK autophosphorylation and results in the activation of Akt which also induces mTOR within the mTORC1 complex. Akt is also regulated by mTORC2 complex. PLCγ activation leads to Ca+2 mobilization and to the activation of PKC. These events play an essential role in proliferation, differentiation, survival and cell migration.
Cancers 07 00860 g001
Mutations that affect RTK signaling often lead to cell transformation, which is observed in a wide variety of malignancies. These mutations affect RTKs or components of downstream pathways such as MAP kinase and the PI3K/AKT. This results in increased cell proliferation, survival, invasion and metastasis. Therefore, targeting RTK signaling pathways remains a challenge for scientists and clinicians working in the cancer field. Several small molecule inhibitors and antibodies are being clinically developed to target RTKs, the MAP kinase and PI3K/AKT pathways. This review attempts to highlight the important role played by RTK signaling in carcinogenesis and the therapeutic strategies available, so far, to target these important cellular pathways.

2. Receptor Tyrosine Kinase Signaling and Cancer

2.1. Targeting Receptor Tyrosine Kinases (RTKs) in Cancer

Most RTKs are found mutated in a variety of cancers and from different tissue origins. This chapter discusses the role of RTKs in cancer and the therapeutic strategies developed to target them (Figure 2 and Table 1).
Figure 2. Schematic representation of the mode of action of RTKs inhibitors. In cancer therapy, RTKs are targeted using monoclonal antibodies that prevent ligand binding and therefore the activation of downstream signaling pathways. Tyrosine kinase inhibitors act on the tyrosine kinase domain of RTK, preventing receptors’ auto-phosphorylation and inhibiting signal transduction.
Figure 2. Schematic representation of the mode of action of RTKs inhibitors. In cancer therapy, RTKs are targeted using monoclonal antibodies that prevent ligand binding and therefore the activation of downstream signaling pathways. Tyrosine kinase inhibitors act on the tyrosine kinase domain of RTK, preventing receptors’ auto-phosphorylation and inhibiting signal transduction.
Cancers 07 00860 g002
Table 1. Examples of RTK targeted molecular cancer therapies being used clinically or subjected to clinical trials.
Table 1. Examples of RTK targeted molecular cancer therapies being used clinically or subjected to clinical trials.
TargetCompoundCancerReferences
EGFR family
HER2Trastuzumab (Herceptin)HER2-positive breast cancer[7,8,9]
EGFRCetuximab (Erbitux)Metastatic colorectal cancer
(RAS wild type)
Metastatic non-small-cell lung cancer
[10]
Panitumumab (Vectibix)[10,11,12]
Gefitinib (Iressa)[13,14,15]
Erlotinib (Tarceva)[13,16]
EGFR and HER2Lapatinib (TykerbHER2-positive breast cancer
(Trastuzumab-resistant)
[14,17]
AfatinibNSCLC
HER2-positive breast cancer
[18,19,20,21]
VEGFRSorafenib (Nexavar)Renal, liver and thyroid cancer[22,23,24]
Sunitinib (Sutent)Renal cell cancer
Gastrointestinal stromal tumor (GIST)
[25,26]
Bevacizumab (Avastin)Metastatic colorectal carcinoma[27]
PDGFRImatinib (Gleevec)GIST (KIT+)[28]
PDGFR and VEGFRSunitinibAngiogenesis[29,30,31,32]
Soratinib
Pazopanib
Nilotinib
FGFR and VEGFRBrivanib (BMS-540215)Human hepatocellular carcinoma model[33]
VEGFR, PDGFR, FLT-3, c-KIT and FGFRCHIR-258 (TKI-258)Multiple myelomas[34,35]
METSGX523MDCK and A549 cells and GTL16 xenograft models[36]
C-KITImatinib (Gleevec)GIST[37,38,39]

2.1.1. EGFR-Targeted Therapy

Gene mutations affecting EGFR members have been associated with several cancers [40]. In breast cancer, overexpression of HER2 (ERBB2) is found in approximately 10%–30% of patients and is associated with reduced survival [41]. Mutations affecting EGFR gene result in its overexpression in 30%–50% of glioblastomas [42,43], 25%–82% in colorectal cancer [44,45,46,47] and 5%–20% in non-small-cell lung cancer [13,48]. Therefore, molecular targeted therapeutics were developed against those receptors. Trastuzumab (Herceptin), a monoclonal antibody, is used to target the extracellular domain of the HER2 protein in HER2-positive breast cancer patients and has been shown to increase survival at early and late stages of breast cancer [7]. Cetuximab (Erbitux) and Panitumumab (Vectibix) are two other examples of monoclonal antibodies that are used to target the EGFR-ligand binding in the treatment of patients with metastatic colorectal cancer [11,12]. The benefit of cetuximab and Panitumumab was limited to patients with RAS wild-type tumors [10].
Lapatinib (Tykerb), a tyrosine kinase inhibitor, targets the ATP binding pocket of the kinase domain of EGFR and HER2 and has been used as an alternative treatment of HER2-positive breast cancer patients that developed resistance to Trastuzumab [8,9]. It has also been used in combination with chemotherapeutic compounds such as Capecitabine, and has been shown to reduce the risk of disease progression in women with advanced HER2-positive breast cancer who had received multiple previous treatments [17]. More recently, Lapatinib has been used in combination with letrozole (Femara) to treat postmenopausal women with Hormone receptor (HR) positive, HER2-positive metastatic breast cancer. This combination resulted in increased progression free survival in the HER2-positive population [14]. Gefitinib (Iressa) and Erlotinib (Tarceva), which are also tyrosine kinase inhibitors, have been used in treatment of patients with metastatic non-small-cell lung cancer. These drugs have been used in combination with chemotherapy and resulted in an improved and progression-free survivals [15,16]. Finally, Afatinib (Giotrif) is a novel ErbB family blocker that selectively blocks ErbB family members (EFGR, HER2, ErbB4 and ErbB3). Unlike Gefitinib and Erlotinib, Afatinib irreversibly (covalently) binds to proteins of ErbB family members and blocks their signaling pathways, thus promoting a sustained anti-proliferative activity [18,19]. This drug has been tested in several clinical trials and has been shown to extend progression free survival of patients with non-small cell lung carcinoma (NSCLC). However, this effect appears to be more beneficial to patients carrying EGFR del19 mutations [20]. Furthermore, and as Afatinib targets HER2, it is also being investigated for use in other HER2-positive cancers such as HER2-positive breast cancer [21].

2.1.2. VEGFR-Targeted Therapy

This family of receptors, which binds VEGF, plays a key role in vasculogenesis and angiogenesis and is critical to tumor-induced new vascular formation [49]. Several studies have reported elevated levels of VEGFR in several cancers and these correlated with metastasis and poor prognosis [50,51,52]. A number of VEGFR inhibitors have been developed with the aim of reducing angiogenesis and lymphangiogenesis associated with cancer progression [49]. Sorafenib (Nexavar), a small molecule inhibitor of tyrosine protein kinase, has been used for the treatment of renal cell, liver and thyroid cancers. An improved progression-free survival following Sorafenib treatment was reported in patients with advanced renal cell cancer and nonresponsive thyroid cancer [22,23]. In patients with liver cancer, an improvement of median overall survival was reported [24]. Sunitinib (Sutent, SU11248) is another VEGFR protein tyrosine kinase inhibitor, which has been shown to improve overall survival of patients with renal cell cancer and gastrointestinal stromal tumor [25,26]. Besides the use of small molecule inhibitors to target VEGFR, a monoclonal antibody (Bevacizumab, Avastin) has been used in combination with chemotherapy to treat patients with metastatic colorectal carcinoma. This resulted in improvement of patients’ survival [27].

2.1.3. PDGFR-Targeted Therapy

PDGF and PDGFRs have important functions in the regulation of cell growth and survival. Mutations within PDGFRα gene have been found in 5% of gastrointestinal stromal cancer (GIST). These mutations affect tyrosine kinase domains and juxtamembrane domain [53]. PDGFR genes were also involved in gene rearrangements found in certain leukemias [54]. In addition, amplifications of PDGFRα were reported in 5%–10% of glioblastoma multiforme, in oligodendrocytoma, esophageal squamous cell carcinoma and artery intimal sarcomas [55,56,57,58,59,60]. As for other dysfunctional RTKs, tyrosine kinase inhibitors have been developed to target directly PDGFR or as a secondary target. These small molecule inhibitors include imatinib, sunitinib, sorafenib, pazopanib and nilotinib. Imatinib (Gleevec), a well-known inhibitor of the oncogenic Bcr-abl fusion protein responsible for chronic myelogenous leukemia (CML), has been used to target PDGFR in gastrointestinal stromal tumors KIT positive. Although this treatment led to significant improvement of overall survival, many patients developed resistance to imatinib [28]. Other drugs such as sunitinib, soratinib, pazopanib and nilotinib were used to target multiple RTK receptors (e.g., PDGFR and VGFR) with the aim of inhibiting cell proliferation and angiogenesis to ensure maximum shrinkage of the tumor [29,30,31,32].

2.1.4. FGFR-Targeted Therapy

Several mutations affecting FGFR genes have been reported in the literature [61]. Amplifications of FGFR1 and 2 have been found in breast cancer [62,63,64,65,66,67,68,69,70] and in gastric cancer where these mutations were associated with poor prognosis [71,72]. FGFR1 amplifications were found in bladder cancer, oral squamous carcinoma and ovarian cancer [73,74,75]. Point mutations that affect FGFR1, 2 and 3 lead to the increase of receptors or constitutive activations and were observed in cancer of the prostate, bladder, breast, brain, lung, uterus, stomach, head and neck, colon and malignant melanoma [76,77,78,79,80,81,82,83,84,85,86]. Chromosomal translocations involving FGFR genes generate oncogenic protein fusions that are present in several hematopoietic malignancies such as multiple myelomas and myeloproliferative disorder syndrome [87,88,89,90,91,92]. Although several small molecule inhibitors of the FGFR tyrosine kinase are currently in clinical development, these molecules also target other RTKs such as VEGFR, PDGFR and c-Kit [33,34,35,93,94,95,96,97,98,99]. Examples of these inhibitors include Brivanib (BMS-540215), a dual effect inhibitor of FGFR and VEGFR that has been shown to affect tumor growth in mouse models of human hepatocellular carcinoma (HCC) [33]. CHIR-258 (TKI-258), a multiple target inhibitor (VEGFR, PDGFR, FLT-3, c-Kit and FGFR), is an effective inhibitor of multiple myelomas harboring the translocation t (4, 14) (p16; q32) that expresses wild type or activated FGFR3 [34,35].

2.1.5. MET-Targeted Therapy

MET is the receptor for the hepatocyte growth factor and is involved in cell growth, migration, invasion, metastasis and angiogenesis [100,101]. MET mutations and amplifications have been reported in many cancers such as neuroblastoma, glioblastomas, osteosarcomas, oesophageal and gastric colorectal cancers, multiple myelomas and T-cell leukemia. These alterations have been shown to be a driver of proliferation, invasion and metastasis and are associated with aggressive phenotype and poor prognosis [102,103,104,105,106,107,108,109,110,111,112]. Since the generation of the first c-MET inhibitor K252a [113], several inhibitors of MET have been tested clinically [114,115]. Among these inhibitors, SGX523 is a highly specific inhibitor of MET and has been shown to inhibit the growth of MDCK and A549 cells and GTL16 xenografts [36]. ARQ197 (ArQule) is also a MET-specific inhibitor that has been shown to inhibit the growth of the breast cancer cell line MDA-MB-231, the prostate cancer cell line PC3, the colon cancer cell line HT29 and the pancreatic cancer cell line PaCa2 [116,117]. Other inhibitors have a Broad spectrum kinase inhibitor effect such as MP470 which acts on MET, RET, KIT, PDGFR and FLT3 or XL880 and PF2341066 that act on VEGFR2 and ALK, respectively [114].

2.1.6. c-KIT-Targeted Therapy

c-Kit also known as CD117 or Mast/Stem Cell Growth Factor Receptor is a cell surface receptor of SCF (Stem Cell Factor). C-Kit activation by SCF initiates the activation of downstream pathways that are involved in the regulation of multiple cellular processes such as proliferation, survival, cell migration, hematopoiesis, stem cell maintenance, melanogenesis and gametogenesis. Mutations in c-KIT result in SCF-independent activation of downstream signaling pathways associated with increased proliferation and cell survival, mostly found in leukemia, gastrointestinal stromal tumors (GIST), testicular germ cell tumor (TGCT) and melanoma. The majority of oncogenic c-Kit mutations are located in the juxtamembrane region (e.g., c-KitV560G) or within the kinase domain (e.g., c-KitD816V) [118]. The tyrosine kinase inhibitor Imatinib (Gleevec), a well-known inhibitor of the oncogenic Bcr-abl fusion protein, has been used to target the juxtamembrane domain of c-KIT in GIST patients. However, secondary mutations occur in other parts of the receptor (such as exon 17) that renders the tyrosine kinase resistant to the inhibition of imatinib [37,38,119]. Although newer c-Kit drugs have been developed (dasatinib and PKC412) to overcome the resistance to imatinib, these drugs have made little impact. This is mainly due to the lack of well-validated inhibitors of the forms of KIT that carry certain types of mutations [39].

2.2. The RAS/MAP Kinase Pathway

This pathway is a central player for a multitude of physiological and pathological cellular processes such as growth, proliferation, differentiation, migration and apoptosis [120]. Its activation by RTKs triggers a cascade of phosphorylation involving downstream kinases which leads to the phosphorylation of target proteins in the nucleus and cytoplasm. The first part of the cascade relies on the activation of at least one of the four major MAP kinases: ERK1 and 2 (ERK1/2), ERK5, p38, and JNK. ERK1 and 2 are probably the most studied MAP kinases. Receptor tyrosine kinases frequently engage Erk1/2 by recruiting the RAS guanine exchange factor Sos to the plasma membrane. This factor is constitutively associated with the protein adapter Grb2, which brings Sos in close proximity to the small GTPase RAS resulting in a nucleotide exchange from GDP to GTP, a change of Ras protein conformation and activation of the serine/threonine kinase Raf. The next step involves Raf phosphorylation of Mek1/2 which in turn phosphorylate and activate Erk1/2 [121] (Figure 1).
The ERK5 pathway is the least studied pathway among the MAP kinase pathways. However, its emerging role as an important player in the regulation of tumour migration and invasion refocused the “spotlight” on this pathway [122]. ERK5 (also known as BMK1) is activated by various stimuli such as oxidative stress, growth factors and oncogenes and plays an important role in cell proliferation, survival, differentiation and embryonic development of the vascular system [122,123] (Figure 1). The role of ERK5 in cell proliferation was demonstrated by in vitro expression of a dominant-negative form that resulted in preventing HeLa cells from entering the S phase of the cell cycle [124]. Similar results were observed in other cancer cell lines, re-enforcing its role as a regulator of cell proliferation [122]. ERK5 involvement in cell survival has been shown using in vivo and in vitro experimental approaches. The inactivation of the MEK5/ERK5 pathways using a mek5−/− mouse model sensitized the mek5−/− mouse embryonic fibroblast (MEFs) to osmotic-stress-induced apoptosis [125]. In another model (mouse tumour xenograft model), the induced deletion of ERK5, significantly reduced tumour volume and vascular density, that were mediated by the pro-proliferative and pro-survival factors RSK (p90 ribosomal S6 kinase) and rpS6 (ribosomal protein) [126]. In vitro, knockdown of ERK5 using siRNA, triggered apoptosis and reduced chemo-resistance of HL-60 acute myeloid leukemia cells [126]. ERK5 also appears to play a role in prostate cancer invasion and metastasis. High levels of expression of ERK5 correlated with the presence of bony metastases and less favourable disease-specific survival in prostate cancer patients. This expression was associated with increased expression of the extracellular matrix proteinase MMP9 [127]. The role of ERK5 in controlling cell differentiation has been shown by its negative control of macrophage differentiation through negative regulation of the expression of macrophage colony stimulating factor receptor (M-CSFR) [128]. Finally, the role of the ERK5 pathway in neoangiogenesis has been evidenced by target depletion of ERK5 in xenograft tumour models of B16F10 melanoma and LL/2 Lewis lung cancer and which resulted in reduced mass and vascular density of the tumours [126].
The Jun N-terminal kinase (JNK) and the p38 MAPK pathways’ family members, also called stress activated protein kinase pathways, function in cell context and cell type specific manner [129] (Figure 1). JNK1, JNK2 and JNK3 are encoded by MAPK8, MAPK9 and MAPK10. Although JNK1 and JNK2 are ubiquitously expressed, JNK3 is mainly expressed in the brain and testis. These factors mediate their response through targeting AP1, a heterodimeric transcription factor composed of Jun and Fos family members and which plays an important role in several cellular processes including proliferation, differentiation and apoptosis. In human cancer, mutations affecting MAPK9 resulted in JNK1 high expression in liver and prostate cancers, while mutations of MAPK10 that led to a loss of function was associated with brain tumours [130,131]. Interestingly, in mouse models, the function of JNK1 and JNK2 in regulating cell proliferation appears to be complex. For instance, mice knockout experiments of JNK1 and JNK2 resulted in confusing results. JNK1 appears to have a tumor suppressor function, whereas JNK2 functions as a tumour promoter [132]. Moreover, JNK1 knockout, unlike JNK2, significantly decreased HCC (hepatocellular carcinoma) in the DEN (diethylnitrosamine)-induced HCC mouse model [133]. In a similar manner, JNK2 but not JNK1 knockout, prevented skin cancer formation that was induced by DMBA (7,12-dimethylbenz[a]anthracene) and PMA treatments [134,135]. A possible explanation to this might be linked to their ability to interact with JUN, a regulator of cell cycle progression [136]. Finally, higher expression levels of MKK4 and MKK7, two JNK MAP kinase activators, have been associated with high-grade prostate cancer [132].
The p38 family comprises four isoforms (p38α, p38β, p38γ and p38δ) which could have overlapping functions [137]. Although p38α is expressed in most tissues, p38β, p38γ and p38δ appear to be expressed in specific tissues such as the brain, skeletal muscle and endocrine glands [138,139]. The p38 MAPK pathway regulates the phosphorylation of several transcription factors, such as p53, activating transcription factor 2 (ATF2), ElK1; and protein kinases, including MAPK activated kinase 2 (MK2; also known as MAPK2), mitogen- and stress-activated protein kinase1 (MsK1), MAP kinase-interacting serine/threonine kinase 1 (MNK1) and MNK2). This wide spectrum of activities allows the p38 pathway to negatively regulate cell cycle progression at the G1/S and G2/M phases of cell cycle progression. This tumor suppressive role has been studied using mice disrupted in MEK3 and MEK6 genes, where these mice exhibit increased tumorigenic potential and Ras-induced transformation [140]. Although the p38 MAPK pathway is involved in tumor suppression, increased expression of the phosphorylated form of p38 correlated with malignancy of different types of cancers [138]. This is consistent with previous reports that support a role of p38 MAPKs in epithelial-mesenchymal transition (EMT), a key event associated with cell migration, invasion and tumor cells’ extravasation [140,141]. In contrast, p38 MAPK pathway appears to play a role in the resistance to anoikis, another important event in tumor migration and spreading, as it allows cancer cells to survive following a loss of contact with the extracellular matrix or their neighboring cells [142]. Overall, the p38 MAPK pathway appears to be expressed in several types of cancers and in this regard constitutes a potential target for cancer therapies.

Targeting the RAS/MAP Kinase Pathway in Cancer

Most of cancer associated mutations affecting the MAP kinase signaling pathway involve mutations in RAS and RAF genes [143,144]. Mutations in RAS family of genes (K-RAS, N-RAS, H-RAS) have been found in several types of cancers [145]. High frequency of KRAS mutations are found in pancreatic, large intestine, biliary tract, small intestine, lung, endometrial and ovarian cancers. High frequencies of NRAS mutations are associated with melanoma, nervous system, hematopoietic and lymphatic, and thyroid cancers. HRAS mutations have a high frequency in salivary gland, urinary tract, cervix, upper aerodigestive tract, penis, prostate and skin cancers.
Mutations associated with RAF family (A-Raf, B-Raf, and C-Raf) are mainly associated with the BRAF gene and are found in a variety of cancers with high frequency in malignant melanoma, thyroid and colon carcinoma [146]. This mutation is single-base missense substitution of valine with glutamic acid at codon 600 (V600E) of the kinase domain and is prevalent in melanomas and papillary thyroid carcinomas [147,148]. Although low frequency, ARAF mutations are found in ovary and large intestine tumors and CRAF mutations in ovary and lung tumors [145].
Small inhibitor molecules are being developed to target primarily Mek and Raf in patients with different types of cancers [145] (Figure 3 and Table 2). Sorafenib, PLX4720, PLX4032 and GSK2118436 are drugs which are being used to target B-RafV600E in malignant melanoma and other advanced malignancies. Other chemical inhibitors such as LErafAON (NeoPharm) and ISIS 5132 are being used to target C-Raf in ovarian and breast cancers but also in other malignancies. MEK inhibitors such as CI-1040, PD-0325901, AZD6244, RDEA119/BAY 86-9766, GDC-0973/XL581 and AZD8330/ARRY-424704 which are also being tested clinically target MEK and for a wide variety of cancers, while others such as GSK1120212 target, in addition to MEK, C-Raf, B-Raf V600E and BRAF wild type [145]. Finally, inhibitors of the JNK proteins are being investigated for potential clinical use. These include the ATP-competitive JNK inhibitor SP600125 and JNK peptide inhibitor (D-JNKI-1), which showed promising results in vitro and in vivo tumor models [132] (Figure 3). P38 pathway inhibitors are mainly developed for treatment of diseases such as rheumatoid arthritis and Crohn’s disease, however, the inhibitor LY2228820 dimesylate produced significant inhibition of the tumors’ growth in in vivo models of melanoma, NSCLC, glioma, myeloma, and ovarian and breast cancers [149].
Figure 3. Schematic representation of components of the MAP kinase pathway targeted by small molecule inhibitors. The inhibitors (small molecule inhibitors and/or peptides) have been developed to target Raf, MEK, JNK1/2, p38 and ERK5.
Figure 3. Schematic representation of components of the MAP kinase pathway targeted by small molecule inhibitors. The inhibitors (small molecule inhibitors and/or peptides) have been developed to target Raf, MEK, JNK1/2, p38 and ERK5.
Cancers 07 00860 g003
Table 2. Examples of MAP kinase and PI3K/AKT pathways’ factors targeted by molecular cancer therapies being used clinically or that are subject to clinical trials.
Table 2. Examples of MAP kinase and PI3K/AKT pathways’ factors targeted by molecular cancer therapies being used clinically or that are subject to clinical trials.
TargetCompoundCancerReferences
MAP Kinase pathway
BRAFV600ESorafenibMalignant melanoma[143,144,145,146]
PLX4720
PLX4032
GSK2118436
C-RAFLErafAON (NeoPharm)Ovarian and Breast cancer[143,144,145,146]
ISIS 5132
MEKCI-1040Various cancers[143,144,145,146]
PD-0325901
AZD6244
RDEA119/BAY 86-9766
GDC-0973/XL581
AZD8330/ARRY-424704
MAP Kinase pathway
C-RAF, MEKGSK1120212Various cancers[143,144,145,146]
B-RAFV600E
BRAF wild type
PI3K/AKT pathway
PI3K/mTORNVP-BEZ235Various cancers[150,151,152,153,154,155,156,157,158,159,160,161,162,163,164]
BGT226
XL765/SAR245409
SF1126
GDC-0980
PI-103
PF-04691502
PKI-587
GSK2126458

2.3. The PI3K/AKT Pathway

RTKs activation by growth factors, hormones, cytokines, neurotrophic factors and other extracellular signalling molecules trigger the activation of the lipid kinase PI3K, which phosphorylates phosphatidylinositol-4,5-bisphosphate (PIP2) on the plasma membrane and generates phosphatidylinositol-3,4,5-trisphosphate (PIP3) (Figure 1). The serine/threonine kinase Akt/PKB binds to PIP3, through its pleckstrin homology (PH) domain resulting in Akt translocation to the membrane and its partial phosphorylation by the phosphoinositide-dependent protein kinase 1 (PDK1) at Thr308. Akt is fully activated upon its phosphorylation at Ser473 by mTOR complex 2 (mTORC2) [165,166]. Following these series of activations, Akt phosphorylates several target proteins such as the glycogen synthase kinase 3α(GSK3α), mTOR, forkhead box O transcription factors (FoxO), MDM2, BCL2-interacting mediator of cell death (BIM) and BCL2-associated agonist of cell death (BAD), to facilitate cell survival and cell cycle entry [150,167]. Akt activation is negatively regulated by the Phosphatase and Tensin Homolog (PTEN), which dephosphorylates PIP3 preventing Akt translocation to the plasma membrane and thereby preventing its activation [150].

Targeting the PI3K/AKT Pathway in Cancer

Genetic mutations and amplifications affecting the different molecules associated with this pathway have been found in several cancers. The somatic mutation of AKT1 isoform associated with the substitution of glutamic acid by a lysine at amino acid 17 (E17K) of Akt1 have been reported in human breast, colorectal, ovarian cancers and squamous cell lung carcinoma [168,169]. This mutation activates Akt1 by promoting its oncogenic localization to the plasma membrane, which stimulates downstream signaling resulting in cell transformation. Furthermore, this mutation is also found in AKT3 gene and results in an oncogenic protein product (Akt3E17K) expressed in melanoma tumors [170]. Mutations in the catalytically active protein (PIK3CA/p100α) and the regulatory protein (p85α), which form the PI3K complex, have been reported in glioblastoma, ovarian, breast, colon, and endometrial cancers [171,172]. Mutations in the tumor suppressor PTEN, which negatively regulates Akt activation, have been reported in glioblastoma, melanoma, endometrial, prostate, breast and ovarian cancers [150]. These mutations are diverse and include insertion, substitutions and deletions. Genetic amplifications in PIK3CA, AKT1 and the AKT2 genes have also been reported. Other amplifications in the PIK3CA gene were found in squamous cell lung carcinoma, head and neck, cervical, gastric, and oesophageal cancers [173,174,175]. Amplifications in the AKT1 gene were associated with gastric cancer and in AKT2 gene with head and neck, pancreatic, ovary and breast cancers [173,176,177,178,179].
Several small molecule inhibitors of the PI3K/AKT pathway have been developed (Figure 4 and Table 2). Some of these drugs have a dual inhibitor activity or target a specific component of this pathway. NVP-BEZ235, BGT226, XL765/SAR245409, SF1126, GDC-0980, PI-103, PF-04691502, PKI-587, and GSK2126458 have a dual activity toward PI3K and mTOR [153]. These inhibitors, some of which are being tested clinically, target the isoforms of PI3K and the ATP-binding sites of mTORC1 and mTORC2 [152,153]. Small inhibitor molecules have been developed to target specifically PI3K; among them XL147, PX866, GDC0941, BKM120, CAL101 (targets p110δ) are at early or late clinical development for patients with advanced solid tumors and lymphomas [154,155,156,157,158,159,160]. Small molecule inhibitors targeting Akt which include Perifosine, GSK690693, VQD002 and MK2206 are also being tested clinically [150,160,161,162]. Finally, OSI027 and AZD8055, which target and inhibit the catalytic site of mTOR, have shown clinical utility in certain cancers [163,164].
Figure 4. Schematic representation of components of the PI3K/AKT pathway (PI3K, Akt and mTOR) targeted by small molecule inhibitors.
Figure 4. Schematic representation of components of the PI3K/AKT pathway (PI3K, Akt and mTOR) targeted by small molecule inhibitors.
Cancers 07 00860 g004

3. Concluding Remarks

Although significant progress has been made in developing small molecule inhibitors and monoclonal antibodies that target components of the RTK signaling pathways in cancer, substantial challenges prevent rapid and efficient therapies. In this regard, an important obstacle remains in the capacity of cancer cells to adapt to these inhibitors by developing resistance through the emergence of additional mutations. Therefore, complementary inhibitors have to be developed to overcome this resistance. A combination of inhibitors which target RTK, components of the MAP kinase (MEK or Raf inhibitors) or PI3K/AKT may have a better effect in cancer patients’ treatment. Finally, cancers have heterogeneous populations of cells which may react differently to those inhibitors and might play a role in chemo-resistance and cancer initiation and progression following chemotherapeutic treatments. Cancer stem cells are a well-known example of chemo-resistance that leads to relapse in cancer patients. For instance, malignant melanoma stem cells express the ATP-binding cassette transporter ABCB5 which plays an important role in drugs efflux, and, thereby, may attenuate the therapeutic efficiency of the inhibitors used in cancer therapy [180]. Therefore, a better understanding of tumor cellular heterogeneity would result in better therapeutic design and more efficient drugs.

Acknowledgments

This work was funded by the John and Lucille van Geest Foundation. I would like to thank David Boocock for his comments on the manuscript.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Lemmon, M.A.; Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell 2010, 141, 1117–1134. [Google Scholar] [CrossRef] [PubMed]
  2. Li, E.; Hristova, K. Role of receptor tyrosine kinase transmembrane domains in cell signaling and human pathologies. Biochemistry 2006, 45, 6241–6251. [Google Scholar] [CrossRef] [PubMed]
  3. Hubbard, S.R.; Miller, W.T. Receptor tyrosine kinases: Mechanisms of activation and signaling. Curr. Opin. Cell Biol. 2007, 19, 117–123. [Google Scholar] [CrossRef] [PubMed]
  4. Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell 2000, 103, 211–225. [Google Scholar] [CrossRef]
  5. Hubbard, S.R. Juxtamembrane autoinhibition in receptor tyrosine kinases. Nat. Rev. Mol. Cell Biol. 2004, 5, 464–471. [Google Scholar] [CrossRef] [PubMed]
  6. Ullrich, A.; Schlessinger, J. Signal transduction by receptors with tyrosine kinase activity. Cell 1990, 61, 203–212. [Google Scholar] [CrossRef]
  7. Hudis, C.A. Trastuzumab—Mechanism of action and use in clinical practice. N. Engl. J. Med. 2007, 357, 39–51. [Google Scholar] [CrossRef] [PubMed]
  8. Tripathy, D.; Slamon, D.J.; Cobleigh, M.; Arnold, A.; Saleh, M.; Mortimer, J.E.; Murphy, M.; Stewart, S.J. Safety of treatment of metastatic breast cancer with trastuzumab beyond disease progression. J. Clin. Oncol. 2004, 22, 1063–1070. [Google Scholar] [CrossRef] [PubMed]
  9. Montemurro, F.; Donadio, M.; Clavarezza, M.; Redana, S.; Jacomuzzi, M.E.; Valabrega, G.; Danese, S.; Vietti-Ramus, G.; Durando, A.; Venturini, M.; et al. Outcome of patients with HER2-positive advanced breast cancer progressing during trastuzumab-based therapy. Oncologist 2006, 11, 318–324. [Google Scholar] [CrossRef] [PubMed]
  10. Price, T.J.; Peeters, M.; Kim, T.W.; Li, J.; Cascinu, S.; Ruff, P.; Suresh, A.S.; Thomas, A.; Tjulandin, S.; Zhang, K.; et al. Panitumumab versus cetuximab in patients with chemotherapy-refractory wild-type KRAS exon 2 metastatic colorectal cancer (ASPECCT): A randomised, multicentre, open-label, non-inferiority phase 3 study. Lancet Oncol. 2014, 15, 569–579. [Google Scholar] [CrossRef]
  11. Messersmith, W.A.; Ahnen, D.J. Targeting EGFR in colorectal cancer. N. Engl. J. Med. 2008, 359, 1834–1836. [Google Scholar] [CrossRef] [PubMed]
  12. Douillard, J.Y.; Oliner, K.S.; Siena, S.; Tabernero, J.; Burkes, R.; Barugel, M.; Humblet, Y.; Bodoky, G.; Cunningham, D.; Jassem, J.; et al. Panitumumab-FOLFOX4 treatment and RAS mutations in colorectal cancer. N. Engl. J. Med. 2013, 369, 1023–1034. [Google Scholar] [CrossRef] [PubMed]
  13. Riely, G.J.; Pao, W.; Pham, D.; Li, A.R.; Rizvi, N.; Venkatraman, E.S.; Zakowski, M.F.; Kris, M.G.; Ladanyi, M.; Miller, V.A. Clinical course of patients with non-small cell lung cancer and epidermal growth factor receptor exon 19 and exon 21 mutations treated with gefitinib or erlotinib. Clin. Cancer Res. 2006, 12, 839–844. [Google Scholar] [CrossRef] [PubMed]
  14. Johnston, S.; Pippen, J., Jr.; Pivot, X.; Lichinitser, M.; Sadeghi, S.; Dieras, V.; Gomez, H.L.; Romieu, G.; Manikhas, A.; Kennedy, M.J.; et al. Lapatinib combined with letrozole versus letrozole and placebo as first-line therapy for postmenopausal hormone receptor-positive metastatic breast cancer. J. Clin. Oncol. 2009, 20, 5538–5546. [Google Scholar] [CrossRef] [PubMed]
  15. Mok, T.S.; Wu, Y.L.; Thongprasert, S.; Yang, C.H.; Chu, D.T.; Saijo, N.; Sunpaweravong, P.; Han, B.; Margono, B.; Ichinose, Y.; et al. Gefitinib or carboplatin–paclitaxel in pulmonary adenocarcinoma. N. Engl. J. Med. 2009, 361, 947–957. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, C.; Wu, Y.L.; Chen, G.; Feng, J.; Liu, X.Q.; Wang, C.; Zhang, S.; Wang, J.; Zhou, S.; Ren, S.; et al. Erlotinib versus chemotherapy as first-line treatment for patients with advanced EGFR mutation-positive non-small-cell lung cancer (OPTIMAL, CTONG-0802): A multicentre, open-label, randomised, phase 3 study. Lancet Oncol. 2011, 12, 735–742. [Google Scholar] [CrossRef]
  17. Geyer, C.E.; Forster, J.; Lindquist, D.; Chan, S.; Romieu, C.G.; Pienkowski, T.; Jagiello-Gruszfeld, A.; Crown, J.; Chan, A.; Kaufman, B.; et al. Lapatinib plus capecitabine for HER2-positive advanced breast cancer. N. Engl. J. Med. 2006, 355, 2733–2743. [Google Scholar] [CrossRef] [PubMed]
  18. Li, D.; Ambrogio, L.; Shimamura, T.; Kubo, S.; Takahashi, M.; Chirieac, L.R.; Padera, F.; Shapiro, G.I.; Baum, I.; Himmelsbach, F.; et al. BIBW2992, an irreversible EGFR/HER2 inhibitor highly effective in preclinical lung cancer models. Oncogene 2008, 27, 4702–4711. [Google Scholar] [CrossRef] [PubMed]
  19. Solca, F.; Dahl, G.; Zoephel, A.; Bader, G.; Sanderson, M.; Klein, C.; Kraemer, O.; Himmelsbach, F.; Haaksma, E.; Adolf, G.R. Target binding properties and cellular activity of afatinib (BIBW 2992), an irreversible ErbB family blocker. J. Pharmacol. Exp. Ther. 2012, 343, 342–350. [Google Scholar] [CrossRef] [PubMed]
  20. Yang, J.C.H.; Wu, Y.L.; Schuler, M.; Sebastian, M.; Popat, S.; Yamamoto, N.; Zhou, C.; Hu, C.P.; O’Byrne, K.; Sequist, L.V.; et al. Afatinib versus cisplatin-based chemotherapy for EGFR mutation-positive lung adenocarcinoma (LUX-Lung 3 and LUX-Lung 6): Analysis of overall survival data from two randomised, phase 3 trials. Lancet Oncol. 2015, 16, 141–151. [Google Scholar] [CrossRef]
  21. Lin, N.U.; Winer, E.P.; Wheatley, D.; Carey, L.A.; Houston, S.; Mendelson, D.; Munster, P.; Frakes, L.; Kelly, S.; Garcia, A.A.; et al. A phase II study of afatinib (BIBW 2992), an irreversible ErbB family blocker, in patients with HER2-positive metastatic breast cancer progressing after trastuzumab. Breast Cancer Res. Treat. 2012, 133, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  22. Escudier, B.; Eisen, T.; Stadler, W.M.; Szczylik, C.; Oudard, S.; Siebels, M.; Negrier, S.; Chevreau, C.; Solska, E.; Desai, A.A.; et al. Sorafenib in advanced clear-cell renal-cell carcinoma. N. Engl. J. Med. 2007, 356, 125–134. [Google Scholar] [CrossRef] [PubMed]
  23. Brose, M.S.; Nutting, C.M.; Jarzab, B.; Elisei, R.; Siena, S.; Bastholt, L.; de la Fouchardiere, C.; Pacini, F.; Paschke, R.; Shong, Y.K.; et al. Sorafenib in radioactive iodine-refractory, locally advanced or metastatic differentiated thyroid cancer: A randomised, double-blind, phase 3 trial. Lancet 2014, 384, 319–328. [Google Scholar] [CrossRef]
  24. Llovet, J.M.; Ricci, S.; Mazzaferro, V.; Hilgard, P.; Gane, E.; Blanc, J.F.; Cosme de Oliveira, A.; Santoro, A.; Raoul, J.-L.; Forner, A.; et al. Sorafenib in advanced hepatocellular carcinoma. N. Engl. J. Med. 2008, 359, 378–390. [Google Scholar] [CrossRef] [PubMed]
  25. Motzer, R.J.; Hutson, T.E.; Tomczak, P.; Michaelson, M.D.; Bukowski, R.M.; Oudard, S.; Negrier, N.; Szczylik, C.; Pili, R.; Bjarnason, G.A.; et al. Overall survival and updated results for sunitinib compared with interferon alfa in patients with metastatic renal cell carcinoma. J. Clin. Oncol. 2009, 27, 3584–3590. [Google Scholar] [CrossRef] [PubMed]
  26. Demetri, G.D.; van Oosterom, A.T.; Garrett, C.R.; Blackstein, M.E.; Shah, M.H.; Verweij, J.; McArthur, G.; Judson, I.R.; Heinrich, M.C.; Morgan, J.A.; et al. Efficacy and safety of sunitinib in patients with advanced gastrointestinal stromal tumour after failure of imatinib: A randomised controlled trial. Lancet 2006, 368, 1329–1338. [Google Scholar] [CrossRef]
  27. Hurwitz, H.; Fehrenbacher, L.; Novotny, W.; Cartwright, T.; Hainsworth, J.; Heim, W.; Berlin, J.; Baron, A.; Griffing, S.; Holmgren, E.; et al. Bevacizumab plus irinotecan, fluorouracil, and leucovorin for metastatic colorectal cancer. N. Engl. J. Med. 2004, 350, 2335–2342. [Google Scholar] [CrossRef] [PubMed]
  28. Sleijfer, S.; Wiemer, E.; Seynaeve, C.; Verweij, J. Improved insight into resistance mechanisms to imatinib in gastrointestinal stromal tumors: A basis for novel approaches and individualization of treatment. Oncologist 2007, 12, 719–726. [Google Scholar] [CrossRef] [PubMed]
  29. Motzer, R.J.; Michaelson, M.D.; Redman, B.G.; Hudes, G.R.; Wilding, G.; Figlin, R.A.; Ginsberg, M.S.; Kim, S.T.; Baum, S.M.; DePrimo, S.E.; et al. Activity of SU11248, a multitargeted inhibitor of vascular endothelial growth factor receptor and platelet-derived growth factor receptor, in patients with metastatic renal cell carcinoma. J. Clin. Oncol. 2006, 24, 16–24. [Google Scholar] [CrossRef] [PubMed]
  30. Van der Graaf, W.T.; Blay, J.Y.; Chawla, S.P.; Kim, D.W.; Bui-Nguyen, B.; Casali, P.G.; Schöffski, P.; Aglietta, M.; Staddon, A.P.; Beppu, Y.; et al. Pazopanib for metastatic soft-tissue sarcoma (PALETTE): A randomised, double-blind, placebo-controlled phase 3 trial. Lancet 2012, 379, 1879–1886. [Google Scholar] [CrossRef]
  31. Adams, V.R.; Leggas, M. Sunitinib malate for the treatment of metastatic renal cell carcinoma and gastrointestinal stromal tumors. Clin. Ther. 2007, 29, 1338–1353. [Google Scholar] [CrossRef] [PubMed]
  32. Cauchi, C.; Somaiah, N.; Engstrom, P.F.; Litwin, S.; Lopez, M.; Lee, J.; Davey, B.B.; von Mehren, M. Evaluation of nilotinib in advanced GIST previously treated with imatinib and sunitinib. Cancer Chemother. Pharmacol. 2012, 69, 977–982. [Google Scholar] [CrossRef] [PubMed]
  33. Huynh, H.; Ngo, V.C.; Fargnoli, J.; Ayers, M.; Soo, K.C.; Koong, H.N.; Thng, C.H.; Ong, H.S.; Chung, A.; Chow, P.; et al. Brivanib alaninate, a dual inhibitor of vascular endothelial growth factor receptor and fibroblast growth factor receptor tyrosine kinases, induces growth inhibition in mouse models of human hepatocellular carcinoma. Clin. Cancer Res. 2008, 14, 6146–6153. [Google Scholar] [CrossRef] [PubMed]
  34. Sarker, D.; Molife, R.; Evans, T.J.; Hardie, M.; Marriott, C.; Butzberger-Zimmerli, P.; Morrison, R.; Fox, J.A.; Heise, C.; Louie, S.; et al. A phase I pharmacokinetic and pharmacodynamic study of TKI258, an oral, multitargeted receptor tyrosine kinase inhibitor in patients with advanced solid tumors. Clin. Cancer Res. 2008, 14, 2075–2081. [Google Scholar] [CrossRef] [PubMed]
  35. Trudel, S.; Li, Z.H.; Wei, E.; Wiesmann, M.; Chang, H.; Chen, C.; Reece, D.; Heise, C.; Stewart, A.K. CHIR-258, a novel, multitargeted tyrosine kinase inhibitor for the potential treatment of t (4; 14) multiple myeloma. Blood 2008, 105, 2941–2948. [Google Scholar] [CrossRef] [PubMed]
  36. Buchanan, S.G.; Hendle, J.; Lee, P.S.; Smith, C.R.; Bounaud, P.Y.; Jessen, K.A.; Tang, C.M.; Huser, N.H.; Felce, J.D.; Froning, K.J.; et al. SGX523 is an exquisitely selective, ATP-competitive inhibitor of the MET receptor tyrosine kinase with antitumor activity in vivo. Mol. Cancer Ther. 2009, 8, 3181–3190. [Google Scholar] [CrossRef] [PubMed]
  37. Frost, M.J.; Ferrao, P.T.; Hughes, T.P.; Ashman, L.K. Juxtamembrane Mutant V560GKit Is More Sensitive to Imatinib (STI571) Compared with Wild-Type c-Kit Whereas the Kinase Domain Mutant D816VKit Is Resistant 1 Supported by a grant from the National Health and Medical Research Council of Australia (NHMRC). MF is recipient of an Australian Postgraduate Award. LKA is a NHMRC Principal Research Fellow. Imatinib was provided by Novartis. 1. Mol. Cancer Ther. 2002, 1, 1115–1124. [Google Scholar] [PubMed]
  38. Wardelmann, E.; Thomas, N.; Merkelbach-Bruse, S.; Pauls, K.; Speidel, N.; Büttner, R.; Bihl, H.; Leutner, C.C.; Heinicke, T.; Hohenberger, P. Acquired resistance to imatinib in gastrointestinal stromal tumours caused by multiple KIT mutations. Lancet Oncol. 2005, 6, 249–251. [Google Scholar] [CrossRef]
  39. Ashman, L.K.; Griffith, R. Therapeutic targeting of c-KIT in cancer. Expert Opin. Investig. Drugs 2013, 22, 103–115. [Google Scholar] [CrossRef] [PubMed]
  40. Takeuchi, K.; Ito, F. Receptor tyrosine kinases and targeted cancer therapeutics. Biol. Pharm. Bull. 2011, 34, 1774–1780. [Google Scholar] [CrossRef] [PubMed]
  41. Iqbal, N.; Iqbal, N. Human Epidermal Growth Factor Receptor 2 (HER2) in Cancers: Overexpression and Therapeutic Implications. Mol. Biol. Int. 2014. [Google Scholar] [CrossRef] [PubMed]
  42. Voldborg, B.R.; Damstrup, L.; Spang-Thomsen, M.; Poulsen, H.S. Epidermal growth factor receptor (EGFR) and EGFR mutations, function and possible role in clinical trials. Ann. Oncol. 1997, 8, 1197–1206. [Google Scholar] [CrossRef] [PubMed]
  43. Hatanpaa, K.J.; Burma, S.; Zhao, D.; Habib, A.A. Epidermal growth factor receptor in glioma: Signal transduction, neuropathology, imaging, and radioresistance. Neoplasia 2010, 12, 675–684. [Google Scholar] [CrossRef] [PubMed]
  44. Yarden, Y.; Sliwkowski, M.X. Untangling the ErbB signalling network. Nat. Rev. Mol. Cell Biol. 2001, 2, 127–137. [Google Scholar] [CrossRef] [PubMed]
  45. Radinsky, R.; Risin, S.; Fan, D.; Dong, Z.; Bielenberg, D.; Bucana, C.D.; Fidler, I.J. Level and function of epidermal growth factor receptor predict the metastatic potential of human colon carcinoma cells. Clin. Cancer Res. 1995, 1, 19–31. [Google Scholar] [PubMed]
  46. Goldstein, N.S.; Armin, M. Epidermal growth factor receptor immunohistochemical reactivity in patients with American Joint Committee on cancer stage IV colon adenocarcinoma. Cancer 2001, 92, 1331–1346. [Google Scholar] [CrossRef]
  47. McKay, J.A.; Murray, L.J.; Curran, S.; Ross, V.G.; Clark, C.; Murray, G.I.; Cassidy, J.; McLeod, H.L. Evaluation of the epidermal growth factor receptor (EGFR) in colorectal tumours and lymph node metastases. Eur. J. Cancer 2002, 38, 2258–2264. [Google Scholar] [CrossRef]
  48. Gazdar, A. Activating and resistance mutations of EGFR in non-small-cell lung cancer: Role in clinical response to EGFR tyrosine kinase inhibitors. Oncogene 2009, 28, S24–S31. [Google Scholar] [CrossRef] [PubMed]
  49. Koch, S.; Claesson-Welsh, L. Signal transduction by vascular endothelial growth factor receptors. Cold Spring Harbor Perspect. Med. 2012. [Google Scholar] [CrossRef] [PubMed]
  50. Dvorak, H.F. Vascular permeability factor/vascular endothelial growth factor: A critical cytokine in tumor angiogenesis and a potential target for diagnosis and therapy. J. Clin. Oncol. 2002, 20, 4368–4380. [Google Scholar] [CrossRef] [PubMed]
  51. Karkkainen, M.J.; Petrova, T.V. Vascular endothelial growth factor receptors in the regulation of angiogenesis and lymphangiogenesis. Oncogene 2002, 19, 5598–5605. [Google Scholar] [CrossRef] [PubMed]
  52. Nagy, J.A.; Vasile, E.; Feng, D.; Sundberg, C.; Brown, L.F.; Detmar, M.J.; Lawitts, J.A.; Benjamin, L.; Tan, X.; Manseau, E.J.; et al. Vascular permeability factor/vascular endothelial growth factor induces lymphangiogenesis as well as angiogenesis. J. Exp. Med. 2002, 196, 1497–1506. [Google Scholar] [CrossRef] [PubMed]
  53. Corless, C.L.; Schroeder, A.; Griffith, D.; Town, A.; McGreevey, L.; Harrell, P.; Shiraga, S.; Bainbridge, T.; Morich, J.; Heinrich, M.C. PDGFRA mutations in gastrointestinal stromal tumors: Frequency, spectrum and in vitro sensitivity to imatinib. J. Clin. Oncol. 2005, 23, 5357–5364. [Google Scholar] [CrossRef] [PubMed]
  54. Toffalini, F.; Demoulin, J.B. New insights into the mechanisms of hematopoietic cell transformation by activated receptor tyrosine kinases. Blood 2010, 116, 2429–2437. [Google Scholar] [CrossRef] [PubMed]
  55. Fleming, T.P.; Saxena, A.; Clark, W.C.; Robertson, J.T.; Oldfield, E.H.; Aaronson, S.A.; Ali, I.U. Amplification and/or overexpression of platelet-derived growth factor receptors and epidermal growth factor receptor in human glial tumors. Cancer Res. 1992, 52, 4550–4553. [Google Scholar] [PubMed]
  56. Kumabe, T.; Sohma, Y.; Kayama, T.; Yoshimoto, T.; Yamamoto, T. Amplification of alpha-platelet-derived growth factor receptor gene lacking an exon coding for a portion of the extracellular region in a primary brain tumor of glial origin. Oncogene 1992, 7, 627–633. [Google Scholar] [PubMed]
  57. Puputti, M.; Tynninen, O.; Sihto, H.; Blom, T.; Mäenpää, H.; Isola, J.; Paetau, A.; Joensuu, H.; Nupponen, N.N. Amplification of KIT, PDGFRA, VEGFR2, and EGFR in gliomas. Mol. Cancer Res. 2006, 4, 927–934. [Google Scholar] [CrossRef] [PubMed]
  58. Smith, J.S.; Wang, X.Y.; Qian, J.; Hosek, S.M.; Scheithauer, B.W.; Jenkins, R.B.; James, C.D. Amplification of the platelet-derived growth factor receptor-A (PDGFRA) gene occurs in oligodendrogliomas with grade IV anaplastic features. J. Neuropathol. Exp. Neurol. 2000, 59, 495–503. [Google Scholar] [PubMed]
  59. Arai, H.; Ueno, T.; Tangoku, A.; Yoshino, S.; Abe, T.; Kawauchi, S.; Oga, A.; Furuya, T.; Oka, M.; Sasaki, K. Detection of amplified oncogenes by genome DNA microarrays in human primary esophageal squamous cell carcinoma: Comparison with conventional comparative genomic hybridization analysis. Cancer Genet. Cytogenet. 2003, 146, 16–21. [Google Scholar] [CrossRef]
  60. Zhao, J.; Roth, J.; Bode-Lesniewska, B.; Pfaltz, M.; Heitz, P.U.; Komminoth, P. Combined comparative genomic hybridization and genomic microarray for detection of gene amplifications in pulmonary artery intimal sarcomas and adrenocortical tumors. Genes Chromosomes Cancer 2002, 34, 48–57. [Google Scholar] [CrossRef] [PubMed]
  61. Ahmad, I.; Iwata, T.; Leung, H.Y. Mechanisms of FGFR-mediated carcinogenesis. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2012, 1823, 850–860. [Google Scholar] [CrossRef] [PubMed]
  62. Jacquemier, J.; Adelaide, J.; Parc, P.; Penault-Llorca, F.; Planche, J.; Delapeyriere, O.; Birnbaum, D. Expression of the FGFR1 gene in human breast-carcinoma cells. Int. J. Cancer 1994, 59, 373–378. [Google Scholar] [CrossRef] [PubMed]
  63. Meyer, K.B.; Maia, A.T.; O’Reilly, M.; Teschendorff, A.E.; Chin, S.F.; Caldas, C.; Ponder, B.A. Allele-specific up-regulation of FGFR2 increases susceptibility to breast cancer. PLoS Biol. 2008, 6, e108. [Google Scholar] [CrossRef] [PubMed]
  64. Chin, K.; DeVries, S.; Fridlyand, J.; Spellman, P.T.; Roydasgupta, R.; Kuo, W.L.; Lapuk, A.; Neve, R.M.; Qian, Z.; Ryder, T.; et al. Genomic and transcriptional aberrations linked to breast cancer pathophysiologies. Cancer Cell 2006, 10, 529–541. [Google Scholar] [CrossRef] [PubMed]
  65. Gelsi-Boyer, V.; Orsetti, B.; Cervera, N.; Finetti, P.; Sircoulomb, F.; Rougé, C.; Lasorsa, L.; Letessier, A.; Ginestier, C.; Monville, F.; et al. Comprehensive profiling of 8p11-12 amplification in breast cancer. Mol. Cancer Res. 2005, 3, 655–667. [Google Scholar] [CrossRef] [PubMed]
  66. Letessier, A.; Sircoulomb, F.; Ginestier, C.; Cervera, N.; Monville, F.; Gelsi-Boyer, V.; Esterni, B.; Geneix, J.; Finetti, P.; Zemmour, C.; et al. Frequency, prognostic impact, and subtype association of 8p12, 8q24, 11q13, 12p13, 17q12, and 20q13 amplifications in breast cancers. BMC Cancer 2006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Bernard-Pierrot, I.; Gruel, N.; Stransky, N.; Vincent-Salomon, A.; Reyal, F.; Raynal, V.; Vallot, C.; Pierron, G.; Radvanyi, F.; Delattre, O. Characterization of the recurrent 8p11-12 amplicon identifies PPAPDC1B, a phosphatase protein, as a new therapeutic target in breast cancer. Cancer Res. 2008, 68, 7165–7175. [Google Scholar] [CrossRef] [PubMed]
  68. Garcia, M.J.; Pole, J.C.; Chin, S.F.; Teschendorff, A.; Naderi, A.; Ozdag, H.; Vias, M.; Kranjac, T.; Subkhankulova, T.; Paish, C.; et al. A 1 Mb minimal amplicon at 8p11-12 in breast cancer identifies new candidate oncogenes. Oncogene 2005, 24, 5235–5245. [Google Scholar] [CrossRef] [PubMed]
  69. Reis-Filho, J.S.; Simpson, P.T.; Turner, N.C.; Lambros, M.B.; Jones, C.; Mackay, A.; Grigoriadis, A.; Sarrio, D.; Savage, K.; Dexter, T.; et al. FGFR1 emerges as a potential therapeutic target for lobular breast carcinomas. Clin. Cancer Res. 2006, 12, 6652–6662. [Google Scholar] [CrossRef] [PubMed]
  70. Hunter, D.J.; Kraft, P.; Jacobs, K.B.; Cox, D.G.; Yeager, M.; Hankinson, S.E.; Wacholder, S.; Wang, Z.; Welch, R.; Hutchinson, A.; et al. A genome-wide association study identifies alleles in FGFR2 associated with risk of sporadic postmenopausal breast cancer. Nat. Genet. 2007, 39, 870–874. [Google Scholar] [CrossRef] [PubMed]
  71. Jang, J.H.; Shin, K.H.; Park, J.G. Mutations in fibroblast growth factor receptor 2 and fibroblast growth factor receptor 3 genes associated with human gastric and colorectal cancers. Cancer Res. 2001, 61, 3541–3543. [Google Scholar] [PubMed]
  72. Kunii, K.; Davis, L.; Gorenstein, J.; Hatch, H.; Yashiro, M.; di Bacco, A.; Elbi, C.; Lutterbach, B. FGFR2-amplified gastric cancer cell lines require FGFR2 and Erbb3 signaling for growth and survival. Cancer Res. 2008, 68, 2340–2348. [Google Scholar] [CrossRef] [PubMed]
  73. Nord, H.; Segersten, U.; Sandgren, J.; Wester, K.; Busch, C.; Menzel, U.; Komorowski, J.; Dumanski, J.P.; Malmström, P.U.; de Ståhl, T.D. Focal amplifications are associated with high grade and recurrences in stage Ta bladder carcinoma. Int. J. Cancer 2010, 126, 1390–1402. [Google Scholar] [CrossRef] [PubMed]
  74. Freier, K.; Schwaenen, C.; Sticht, C.; Flechtenmacher, C.; Mühling, J.; Hofele, C.; Radlwimmer, B.; Lichter, P.; Joos, S. Recurrent FGFR1 amplification and high FGFR1 protein expression in oral squamous cell carcinoma (OSCC). Oral Oncol. 2007, 43, 60–66. [Google Scholar] [CrossRef] [PubMed]
  75. Gorringe, K.L.; Jacobs, S.; Thompson, E.R.; Sridhar, A.; Qiu, W.; Choong, D.Y.; Campbell, I.G. High-resolution single nucleotide polymorphism array analysis of epithelial ovarian cancer reveals numerous microdeletions and amplifications. Clin. Cancer Res. 2007, 13, 4731–4739. [Google Scholar] [CrossRef] [PubMed]
  76. Rand, V.; Huang, J.; Stockwell, T.; Ferriera, S.; Buzko, O.; Levy, S.; Busam, D.; Li, K.; Edwards, J.B.; Eberhart, C.; et al. Sequence survey of receptor tyrosine kinases reveals mutations in glioblastomas. Proc. Natl. Acad. Sci. USA 2005, 102, 14344–14349. [Google Scholar] [CrossRef] [PubMed]
  77. Van Rhijn, B.W.; van Tilborg, A.A.; Lurkin, I.; Bonaventure, J.; de Vries, A.; Thiery, J.P.; van der Kwast, T.H.; Zwarthoff, E.C.; Radvanyi, F. Novel fibroblast growth factor receptor 3 (FGFR3) mutations in bladder cancer previously identified in non-lethal skeletal disorders. Eur. J. Hum. Genet. 2002, 10, 819–824. [Google Scholar] [CrossRef] [PubMed]
  78. Greenman, C.; Stephens, P.; Smith, R.; Dalgliesh, G.L.; Hunter, C.; Bignell, G.; Davies, H.; Teague, J.; Butler, A.; Stevens, C.; et al. Patterns of somatic mutation in human cancer genomes. Nature 2007, 446, 153–158. [Google Scholar] [CrossRef] [PubMed]
  79. Hernández, S.; de Muga, S.; Agell, L.; Juanpere, N.; Esgueva, R.; Lorente, J.A.; Mojal, S.; Serrano, S.; Lloreta, J. FGFR3 mutations in prostate cancer: Association with low-grade tumors. Mod. Pathol. 2009, 22, 848–856. [Google Scholar] [CrossRef] [PubMed]
  80. Dutt, A.; Salvesen, H.B.; Chen, T.H.; Ramos, A.H.; Onofrio, R.C.; Hatton, C.; Nicoletti, R.; Winckler, W.; Grewal, R.; Hanna, M.; et al. Drug-sensitive FGFR2 mutations in endometrial carcinoma. Proc. Natl. Acad. Sci. USA 2008, 105, 8713–8717. [Google Scholar] [CrossRef] [PubMed]
  81. Chesi, M.; Brents, L.A.; Ely, S.A.; Bais, C.; Robbiani, D.F.; Mesri, E.A.; Kuehl, W.M.; Bergsagel, P.L. Activated fibroblast growth factor receptor 3 is an oncogene that contributes to tumor progression in multiple myeloma. Blood 2001, 97, 729–736. [Google Scholar] [CrossRef] [PubMed]
  82. Chou, A.; Dekker, N.; Jordan, R.C. Identification of novel fibroblast growth factor receptor 3 gene mutations in actinic cheilitis and squamous cell carcinoma of the lip. Oral Surg. Oral Med. Oral Pathol. Oral Radiol. Endodontol. 2009, 107, 535–541. [Google Scholar] [CrossRef] [PubMed]
  83. Davies, H.; Hunter, C.; Smith, R.; Stephens, P.; Greenman, C.; Bignell, G.; Teague, J.; Butler, A.; Edkins, S.; Stevens, C.; et al. Somatic mutations of the protein kinase gene family in human lung cancer. Cancer Res. 2005, 65, 7591–7595. [Google Scholar] [PubMed]
  84. Pollock, P.M.; Gartside, M.G.; Dejeza, L.C.; Powell, M.A.; Mallon, M.A.; Davies, H.; Mohammadi, M.; Futreal, P.A.; Stratton, M.R.; Trent, J.M.; et al. Frequent activating FGFR2 mutations in endometrial carcinomas parallel germline mutations associated with craniosynostosis and skeletal dysplasia syndromes. Oncogene 2007, 26, 7158–7162. [Google Scholar] [CrossRef] [PubMed]
  85. Cappellen, D.; de Oliveira, C.; Ricol, D.; de Medina, S.; Bourdin, J.; Sastre-Garau, X.; Chopin, D.; Thiery, J.P.; Radvanyi, F. Frequent activating mutations of FGFR3 in human bladder and cervix carcinomas. Nat. Genet. 1999, 23, 18–20. [Google Scholar] [PubMed]
  86. Cheng, L.; Zhang, S.; Davidson, D.D.; MacLennan, G.T.; Koch, M.O.; Montironi, R.; Lopez-Beltran, A. Molecular determinants of tumor recurrence in the urinary bladder. Future Oncol. 2009, 5, 843–857. [Google Scholar] [CrossRef] [PubMed]
  87. Jackson, C.C.; Medeiros, L.J.; Miranda, R.N. 8p11 myeloproliferative syndrome: A review. Hum. Pathol. 2010, 41, 461–476. [Google Scholar] [CrossRef] [PubMed]
  88. Yagasaki, F.; Wakao, D.; Yokoyama, Y.; Uchida, Y.; Murohashi, I.; Kayano, H.; Taniwaki, M.; Matsuda, A.; Bessho, M. Fusion of ETV6 to fibroblast growth factor receptor 3 in peripheral T-cell lymphoma with at (4; 12)(p16; p13) chromosomal translocation. Cancer Res. 2001, 61, 8371–8374. [Google Scholar] [PubMed]
  89. Roumiantsev, S.; Krause, D.S.; Neumann, C.A.; Dimitri, C.A.; Asiedu, F.; Cross, N.C.; van Etten, R.A. Distinct stem cell myeloproliferative/T lymphoma syndromes induced by ZNF198-FGFR1 and BCR-FGFR1 fusion genes from 8p11 translocations. Cancer Cell 2004, 5, 287–298. [Google Scholar] [CrossRef]
  90. Xiao, S.; Nalabolu, S.R.; Aster, J.C.; Ma, J.; Abruzzo, L.; Jaffe, E.S.; Stone, R.; Weissman, S.M.; Hudson, T.J.; Fletcher, J.A. FGFR1 is fused with a novel zinc-finger gene, ZNF198, in the t (8; 13) leukaemia/lymphoma syndrome. Nat. Genet. 1998, 18, 84–87. [Google Scholar] [CrossRef] [PubMed]
  91. Demiroglu, A.; Steer, E.J.; Heath, C.; Taylor, K.; Bentley, M.; Allen, S.L.; Koduru, P.; Brody, J.P.; Hawson, G.; Rodwell, R.; et al. The t (8; 22) in chronic myeloid leukemia fuses BCR to FGFR1: Transforming activity and specific inhibition of FGFR1 fusion proteins. Blood 2001, 98, 3778–3783. [Google Scholar] [CrossRef] [PubMed]
  92. Ren, M.; Li, X.; Cowell, J.K. Genetic fingerprinting of the development and progression of T-cell lymphoma in a murine model of atypical myeloproliferative disorder initiated by the ZNF198–fibroblast growth factor receptor-1 chimeric tyrosine kinase. Blood 2009, 114, 1576–1584. [Google Scholar] [CrossRef] [PubMed]
  93. Knights, V.; Cook, S.J. De-regulated FGF receptors as therapeutic targets in cancer. Pharmacol. Ther. 2010, 125, 105–117. [Google Scholar] [CrossRef] [PubMed]
  94. Hilberg, F.; Roth, G.J.; Krssak, M.; Kautschitsch, S.; Sommergruber, W.; Tontsch-Grunt, U.; Garin-Chesa, P.; Bader, G.; Zoephel, A.; Quant, J.; et al. BIBF 1120: Triple angiokinase inhibitor with sustained receptor blockade and good antitumor efficacy. Cancer Res. 2008, 68, 4774–4782. [Google Scholar] [CrossRef] [PubMed]
  95. Hahn, K.; Oglivie, G.; Rusk, T.; Devauchelle, P.; Leblanc, A.; Legendre, A.; Powers, B.; Leventhal, P.S.; Kinet, J.P.; Palmerini, F.; et al. Masitinib is safe and effective for the treatment of canine mast cell tumors. J. Vet. Intern. Med. 2008, 22, 1301–1309. [Google Scholar] [CrossRef] [PubMed]
  96. Fabbro, D.; Manley, P.W. Su-6668. SUGEN. Curr. Opin. Investig. Drugs (Lond. Engl.) 2001, 2, 1142–1148. [Google Scholar]
  97. Kumar, R.; Knick, V.B.; Rudolph, S.K.; Johnson, J.H.; Crosby, R.M.; Crouthamel, M.C.; Hopper, T.M.; Miller, C.G.; Harrington, L.E.; Onori, J.A.; et al. Pharmacokinetic-pharmacodynamic correlation from mouse to human with pazopanib, a multikinase angiogenesis inhibitor with potent antitumor and antiangiogenic activity. Mol. Cancer Ther. 2007, 6, 2012–2021. [Google Scholar] [CrossRef] [PubMed]
  98. Marek, L.; Ware, K.E.; Fritzsche, A.; Hercule, P.; Helton, W.R.; Smith, J.E.; McDermott, L.A.; Coldren, C.D.; Nemenoff, R.A.; Merrick, D.T.; et al. Fibroblast growth factor (FGF) and FGF receptor-mediated autocrine signaling in non-small-cell lung cancer cells. Mol. Pharmacol. 2009, 75, 196–207. [Google Scholar] [CrossRef] [PubMed]
  99. McDermott, L.A.; Simcox, M.; Higgins, B.; Nevins, T.; Kolinsky, K.; Smith, M.; Yang, H.; Li, J.K.; Chen, Y.; Ke, J.; et al. RO4383596, an orally active KDR, FGFR, and PDGFR inhibitor: Synthesis and biological evaluation. Bioorg. Med. Chem. 2005, 13, 4835–4841. [Google Scholar] [CrossRef] [PubMed]
  100. Ma, P.C.; Maulik, G.; Christensen, J.; Salgia, R. c-Met: Structure, functions and potential for therapeutic inhibition. Cancer Metastasis Rev. 2003, 22, 309–325. [Google Scholar] [CrossRef] [PubMed]
  101. Peruzzi, B.; Bottaro, D.P. Targeting the c-Met signaling pathway in cancer. Clin. Cancer Res. 2006, 12, 3657–3660. [Google Scholar] [CrossRef] [PubMed]
  102. 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]
  103. Furge, K.A.; Zhang, Y.W.; vande Woude, W.G. Met receptor tyrosine kinase: Enhanced signaling through adapter proteins. Oncogene 2000, 19, 5582–5589. [Google Scholar] [CrossRef] [PubMed]
  104. 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]
  105. 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]
  106. Miller, C.T.; Lin, L.; Casper, A.M.; Lim, J.; Thomas, D.G.; Orringer, M.B.; Chang, A.C.; Chambers, A.F.; Giordano, T.J.; Glover, T.W.; et al. Genomic amplification of MET with boundaries within fragile site FRA7G and upregulation of MET pathways in esophageal adenocarcinoma. Oncogene 2006, 25, 409–418. [Google Scholar] [CrossRef] [PubMed]
  107. Beroukhim, R.; Getz, G.; Nghiemphu, L.; Barretina, J.; Hsueh, T.; Linhart, D.; Vivanco, I.; Lee, J.C.; Huang, J.H.; Alexander, S.; et al. Assessing the significance of chromosomal aberrations in cancer: Methodology and application to glioma. Proc. Natl. Acad. Sci. USA 2007, 104, 20007–20012. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Hecht, M.; Papoutsi, M.; Tran, H.D.; Wilting, J.; Schweigerer, L. Hepatocyte growth factor/c-Met signaling promotes the progression of experimental human neuroblastomas. Cancer Res. 2004, 64, 6109–6118. [Google Scholar] [CrossRef] [PubMed]
  110. Choi, Y.L.; Tsukasaki, K.; O’neill, M.C.; Yamada, Y.; Onimaru, Y.; Matsumoto, K.; Ohashi, J.; Yamashita, Y.; Tsutsumi, S.; Kaneda, R.; et al. A genomic analysis of adult T-cell leukemia. Oncogene 2007, 26, 1245–1255. [Google Scholar] [CrossRef] [PubMed]
  111. Kang, J.; Dolled-Filhart, M.; Ocal, I.T.; Singh, B.; Lin, C.Y.; Dickson, R.B.; Rimm, D.L.; Camp, R.L. Tissue microarray analysis of hepatocyte growth factor/Met pathway components reveals a role for Met, matriptase, and hepatocyte growth factor activator inhibitor 1 in the progression of node-negative breast cancer. Cancer Res. 2003, 63, 1101–1105. [Google Scholar] [PubMed]
  112. Lengyel, E.; Prechtel, D.; Resau, J.H.; Gauger, K.; Welk, A.; Lindemann, K.; Salanti, G.; Richter, T.; Knudsen, B.; Vande Woude, G.F.; et al. C-Met overexpression in node-positive breast cancer identifies patients with poor clinical outcome independent of Her2/neu. Int. J. Cancer 2005, 113, 678–682. [Google Scholar] [CrossRef] [PubMed]
  113. Schiering, N.; Knapp, S.; Marconi, M.; Flocco, M.M.; Cui, J.; Perego, R.; Rusconi, L.; Cristiani, C. Crystal structure of the tyrosine kinase domain of the hepatocyte growth factor receptor c-Met and its complex with the microbial alkaloid K-252a. Proc. Natl. Acad. Sci. USA 2003, 100, 12654–12659. [Google Scholar] [CrossRef] [PubMed]
  114. Underiner, T.L.; Herbertz, T.; Miknyoczki, S.J. Discovery of small molecule c-Met inhibitors: Evolution and profiles of clinical candidates. Anti-Cancer Agents Med. Chem. 2010, 10, 7–27. [Google Scholar] [CrossRef]
  115. Comoglio, P.M.; Giordano, S.; Trusolino, L. Drug development of MET inhibitors: Targeting oncogene addiction and expedience. Nat. Rev. Drug Discov. 2008, 7, 504–516. [Google Scholar] [CrossRef] [PubMed]
  116. Previdi, S.; Abbadessa, G.; Dalò, F.; France, D.S.; Broggini, M. Breast cancer-derived bone metastasis can be effectively reduced through specific c-MET inhibitor tivantinib (ARQ 197) and shRNA c-MET knockdown. Mol. Cancer Ther. 2012, 11, 214–223. [Google Scholar] [CrossRef] [PubMed]
  117. Bagai, R.; Fan, W.; Ma, P.C. ARQ-197, an oral small-molecule inhibitor of c-Met for the treatment of solid tumors. IDrugs Investig. Drugs J. 2010, 13, 404–414. [Google Scholar]
  118. Lennartsson, J.; Rönnstrand, L. Stem cell factor receptor/c-Kit: From basic science to clinical implications. Physiol. Rev. 2012, 92, 1619–1649. [Google Scholar] [CrossRef] [PubMed]
  119. Eder, J.P.; vande Woude, G.F.; Boerner, S.A.; LoRusso, P.M. Novel therapeutic inhibitors of the c-Met signaling pathway in cancer. Clin. Cancer Res. 2009, 15, 2207–2214. [Google Scholar] [CrossRef] [PubMed]
  120. Morrison, D.K. MAP kinase pathways. Cold Spring Harb. Perspect. Biol. 2012, 4. [Google Scholar] [CrossRef] [PubMed]
  121. Chong, H.; Guan, K.L. Regulation of Raf through phosphorylation and N terminus-C terminus interaction. J. Biol. Chem. 2003, 278, 36269–36276. [Google Scholar] [CrossRef] [PubMed]
  122. Lochhead, P.; Gilley, R.; Cook, S. ERK5 and its role in tumour development. Biochem. Soc. Trans. 2012, 40, 251–256. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, X.; Tournier, C. Regulation of cellular functions by the ERK5 signalling pathway. Cell. Signal. 2006, 18, 753–760. [Google Scholar] [CrossRef] [PubMed]
  124. Kato, Y.; Tapping, R.I.; Huang, S.; Watson, M.H.; Ulevitch, R.J.; Lee, J.D. Bmk1/Erk5 is required for cell proliferation induced by epidermal growth factor. Nature 1998, 395, 713–716. [Google Scholar] [PubMed]
  125. Wang, X.; Merritt, A.J.; Seyfried, J.; Guo, C.; Papadakis, E.S.; Finegan, K.G.; Kayahara, M.; Dixon, J.; Boot-Handford, R.P.; Cartwright, E.J.; et al. Targeted deletion of mek5causes early embryonic death and defects in the extracellular signal-regulated kinase 5/myocyte enhancer factor 2 cell survival pathway. Mol. Cell. Biol. 2005, 25, 336–345. [Google Scholar] [CrossRef] [PubMed]
  126. Hayashi, M.; Fearns, C.; Eliceiri, B.; Yang, Y.; Lee, J.D. Big mitogen-activated protein kinase 1/extracellular signal-regulated kinase 5 signaling pathway is essential for tumor-associated angiogenesis. Cancer Res. 2005, 65, 7699–7706. [Google Scholar] [PubMed]
  127. Mehta, P.B.; Jenkins, B.L.; McCarthy, L.; Thilak, L.; Robson, C.N.; Neal, D.E.; Leung, H.Y. MEK5 overexpression is associated with metastatic prostate cancer, and stimulates proliferation, MMP-9expression and invasion. Oncogene 2003, 22, 1381–1389. [Google Scholar] [CrossRef] [PubMed]
  128. Wang, X.; Pesakhov, S.; Harrison, J.S.; Kafka, M.; Danilenko, M.; Studzinski, G.P. The MAPK ERK5, but not ERK1/2, inhibits the progression of monocytic phenotype to the functioning macrophage. Exp. Cell Res. 2015, 330, 199–211. [Google Scholar] [CrossRef] [PubMed]
  129. Wagner, E.F.; Nebreda, A.R. Signal integration by JNK and p38 MAPK pathways in cancer development. Nat. Rev. Cancer 2009, 9, 537–549. [Google Scholar] [CrossRef] [PubMed]
  130. Sabapathy, K.; Hochedlinger, K.; Nam, S.Y.; Bauer, A.; Karin, M.; Wagner, E.F. Distinct roles for JNK1 and JNK2 in regulating JNK activity and c-Jun-dependent cell proliferation. Mol. Cell 2004, 15, 713–725. [Google Scholar] [CrossRef] [PubMed]
  131. Toshiharu, S.; Maeda, S.; Chang, L.; Karin, M. Loss of hepatic NF-κB activity enhances chemical hepatocarcinogenesis through sustained c-Jun N-terminal kinase 1 activation. Proc. Natl. Acad. Sci. USA 2006, 103, 10544–10551. [Google Scholar]
  132. Davies, C.; Tournier, C. Exploring the function of the JNK (c-Jun N-terminal kinase) signalling pathway in physiological and pathological processes to design novel therapeutic strategies. Biochem. Soc. Trans. 2012, 40, 85–89. [Google Scholar] [CrossRef] [PubMed]
  133. Lijian, H.; Zatloukal, K.; Scheuch, H.; Stepniak, E.; Wagner, E.F. Proliferation of human HCC cells and chemically induced mouse liver cancers requires JNK1-dependent p21 downregulation. J. Clin. Investig. 2008, 118, 3943–3953. [Google Scholar]
  134. Chen, N.; Nomura, M.; She, Q.B.; Ma, W.Y.; Bode, A.M.; Wang, L.; Flavell, R.A.; Dong, Z. Suppression of skin tumorigenesis in c-Jun NH2-terminal kinase-2-deficient mice. Cancer Res. 2001, 61, 3908–3912. [Google Scholar] [PubMed]
  135. She, Q.B.; Chen, N.; Bode, A.M.; Flavell, R.A.; Dong, Z. Deficiency of c-Jun-NH2-terminal kinase-1 in mice enhances skin tumor development by 12-O-tetradecanoylphorbol-13-acetate. Cancer Res. 2002, 62, 1343–1348. [Google Scholar] [PubMed]
  136. Qingshan, C.; Zhang, Y.; Beezhold, K.J.; Bhatia, D.; Zhao, H.; Chen, J.; Castranova, V.; Shi, X.; Chen, F. Sustained JNK1 activation is associated with altered histone H3 methylations in human liver cancer. J. Hepatol. 2009, 50, 323–333. [Google Scholar]
  137. Cuenda, A.; Rousseau, S. p38 MAP-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta 2007, 1773, 1358–1375. [Google Scholar] [CrossRef] [PubMed]
  138. Cuadrado, A.; Nebreda, A.R. Mechanisms and functions of p38 MAPK signalling. Biochem. J. 2010, 429, 403–417. [Google Scholar] [CrossRef] [PubMed]
  139. Dhillon, A.S.; Hagan, S.; Rath, O.; Kolch, W. MAP kinase signalling pathways in cancer. Oncogene 2007, 26, 3279–3290. [Google Scholar] [CrossRef] [PubMed]
  140. Koul, H.K.; Pal, M.; Koul, S. Role of p38 MAP kinase signal transduction in solid tumors. Genes Cancer 2013, 4, 342–359. [Google Scholar] [CrossRef] [PubMed]
  141. Bhowmick, N.A.; Zent, R.; Ghiassi, M.; McDonnell, M.; Moses, H.L. Integrin β1 signaling is necessary for transforming growth factor-β activation of p38MAPK and epithelial plasticity. J. Biol. Chem. 2001, 276, 46707–46713. [Google Scholar] [CrossRef] [PubMed]
  142. Cheng, T.L.; Symons, M.; Jou, T.S. Regulation of anoikis by Cdc42 and Rac1. Exp. Cell Res. 2004, 295, 497–511. [Google Scholar] [CrossRef] [PubMed]
  143. Boutros, T.; Chevet, E.; Metrakos, P. Mitogen-activated protein (MAP) kinase/MAP kinase phosphatase regulation: Roles in cell growth, death, and cancer. Pharmacol. Rev. 2008, 60, 261–310. [Google Scholar] [CrossRef] [PubMed]
  144. Santarpia, L.; Lippman, S.M.; El-Naggar, A.K. Targeting the MAPK-RAS-RAF signaling pathway in cancer therapy. Expert Opin. Ther. Targets 2012, 16, 103–119. [Google Scholar] [CrossRef] [PubMed]
  145. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef] [PubMed]
  146. Poulikakos, P.I.; Rosen, N. Mutant BRAF melanomas—Dependence and resistance. Cancer Cell 2011, 19, 11–15. [Google Scholar] [CrossRef] [PubMed]
  147. Xing, M. BRAF mutation in papillary thyroid cancer: Pathogenic role, molecular bases, and clinical implications. Endocr. Rev. 2007, 28, 742–762. [Google Scholar] [CrossRef] [PubMed]
  148. Alessi, D.R.; James, S.R.; Downes, C.P.; Holmes, A.B.; Gaffney, P.R.; Reese, C.B.; Cohen, P. Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Bα. Curr. Biol. 1997, 7, 261–269. [Google Scholar] [CrossRef]
  149. Campbell, R.M.; Anderson, B.D.; Brooks, N.A.; Brooks, H.B.; Chan, E.M.; de Dios, A.; Gilmour, R.; Graff, J.R.; Jambrina, E.; Mader, M.; et al. Characterization of LY2228820 dimesylate, a potent and selective inhibitor of p38 MAPK with antitumor activity. Mol. Cancer Ther. 2014, 13, 364–374. [Google Scholar] [CrossRef] [PubMed]
  150. Hemmings, B.A.; Restuccia, D.F. Pi3k-pkb/akt pathway. Cold Spring Harbor Perspect. Biol. 2012, 4, a011189. [Google Scholar] [CrossRef] [PubMed]
  151. Zaytseva, Y.Y.; Valentino, J.D.; Gulhati, P.; Evers, B.M. mTOR inhibitors in cancer therapy. Cancer Lett. 2012, 319, 1–7. [Google Scholar] [CrossRef] [PubMed]
  152. GarciaEcheverria, C. Blocking the mTOR pathway: A drug discovery perspective. Biochem. Soc. Trans. 2011, 39, 451–455. [Google Scholar] [CrossRef] [PubMed]
  153. Wander, S.A.; Hennessy, B.T.; Slingerland, J.M. Next-generation mTOR inhibitors in clinical oncology: How pathway complexity informs therapeutic strategy. J. Clin. Investig. 2011, 121, 1231–1241. [Google Scholar] [CrossRef] [PubMed]
  154. Howes, A.L.; Chiang, G.G.; Lang, E.S.; Ho, C.B.; Powis, G.; Vuori, K.; Abraham, R.T. The phosphatidylinositol 3-kinase inhibitor, PX-866, is a potent inhibitor of cancer cell motility and growth in three-dimensional cultures. Mol. Cancer Ther. 2007, 6, 2505–2514. [Google Scholar] [CrossRef] [PubMed]
  155. Shapiro, G.I.; Rodon, J.; Bedell, C.; Kwak, E.L.; Baselga, J.; Braña, I.; Pandya, S.S.; Scheffold, C.; Laird, A.D.; Nguyen, L.T.; et al. Phase I safety, pharmacokinetic, and pharmacodynamic study of SAR245408 (XL147), an oral pan-class I PI3K inhibitor, in patients with advanced solid tumors. Clin. Cancer Res. 2014, 20, 233–245. [Google Scholar] [CrossRef] [PubMed]
  156. Bowles, D.W.; Ma, W.W.; Senzer, N.; Brahmer, J.R.; Adjei, A.A.; Davies, M.; Lazar, A.J.; Vo, A.; Peterson, S.; Walker, L.; et al. A multicenter phase 1 study of PX-866 in combination with docetaxel in patients with advanced solid tumours. Br. J. Cancer 2013, 109, 1085–1092. [Google Scholar] [CrossRef] [PubMed]
  157. Munugalavadla, V.; Mariathasan, S.; Slaga, D.; Du, C.; Berry, L.; del Rosario, G.; Yan, Y.; Boe, M.; Sun, L.; Friedman, L.S.; et al. The PI3K inhibitor GDC-0941 combines with existing clinical regimens for superior activity in multiple myeloma. Oncogene 2014, 33, 316–325. [Google Scholar] [CrossRef] [PubMed]
  158. Bendell, J.C.; Rodon, J.; Burris, H.A.; de Jonge, M.; Verweij, J.; Birle, D.; Demanse, D.; de Buck, S.S.; Ru, Q.C.; Peters, M.; et al. Phase I, dose-escalation study of BKM120, an oral pan-Class I PI3K inhibitor, in patients with advanced solid tumors. J. Clin. Oncol. 2011, 30, 282–290. [Google Scholar] [CrossRef] [PubMed]
  159. Hoellenriegel, J.; Meadows, S.A.; Sivina, M.; Wierda, W.G.; Kantarjian, H.; Keating, M.J.; Giese, N.; O’Brien, S.; Yu, A.; Miller, L.L.; et al. The phosphoinositide 3′-kinase delta inhibitor, CAL-101, inhibits B-cell receptor signaling and chemokine networks in chronic lymphocytic leukemia. Blood 2011, 118, 3603–3612. [Google Scholar] [CrossRef] [PubMed]
  160. Richardson, P.G.; Eng, C.; Kolesar, J.; Hideshima, T.; Anderson, K.C. Perifosine, an oral, anti-cancer agent and inhibitor of the Akt pathway: Mechanistic actions, pharmacodynamics, pharmacokinetics, and clinical activity. Expert Opin. Drug Metab. Toxicol. 2012, 8, 623–633. [Google Scholar] [CrossRef] [PubMed]
  161. Rhodes, N.; Heerding, D.A.; Duckett, D.R.; Eberwein, D.J.; Knick, V.B.; Lansing, T.J.; McConnell, R.T.; Gilmer, T.M.; Zhang, S.Y.; Robell, K.; et al. Characterization of an Akt kinase inhibitor with potent pharmacodynamic and antitumor activity. Cancer Res. 2008, 68, 2366–2374. [Google Scholar] [CrossRef] [PubMed]
  162. Agarwal, E.; Chaudhuri, A.; Leiphrakpam, P.D.; Haferbier, K.L.; Brattain, M.G.; Chowdhury, S. Akt inhibitor MK-2206 promotes anti-tumor activity and cell death by modulation of AIF and Ezrin in colorectal cancer. BMC Cancer 2014, 14, 145. [Google Scholar] [CrossRef] [PubMed]
  163. Bhagwat, S.V.; Gokhale, P.C.; Crew, A.P.; Cooke, A.; Yao, Y.; Mantis, C.; Kahler, J.; Workman, J.; Bittner, M.; Dudkin, L.; et al. Preclinical characterization of OSI-027, a potent and selective inhibitor of mTORC1 and mTORC2: Distinct from rapamycin. Mol. Cancer Ther. 2011, 10, 1394–1406. [Google Scholar] [CrossRef] [PubMed]
  164. Chresta, C.M.; Davies, B.R.; Hickson, I.; Harding, T.; Cosulich, S.; Critchlow, S.E.; Vincent, J.P.; Ellston, R.; Jones, D.; Sini, P.; et al. AZD8055 is a potent, selective, and orally bioavailable ATP-competitive mammalian target of rapamycin kinase inhibitor with in vitro and in vivo antitumor activity. Cancer Res. 2010, 70, 288–298. [Google Scholar] [CrossRef] [PubMed]
  165. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef] [PubMed]
  166. Engelman, J.A. Targeting PI3K signalling in cancer: Opportunities, challenges and limitations. Nat. Rev. Cancer 2009, 9, 550–562. [Google Scholar] [CrossRef] [PubMed]
  167. Carpten, J.D.; Faber, A.L.; Horn, C.; Donoho, G.P.; Briggs, S.L.; Robbins, C.M.; Hostetter, G.; Boguslawski, S.; Moses, T.Y.; Savage, S.; et al. A transforming mutation in the pleckstrin homology domain of AKT1 in cancer. Nature 2007, 448, 439–444. [Google Scholar] [CrossRef] [PubMed]
  168. Kim, M.S.; Jeong, E.G.; Yoo, N.J.; Lee, S.H. Mutational analysis of oncogenic AKT E17K mutation in common solid cancers and acute leukaemias. Br. J. Cancer 2008, 98, 1533–1535. [Google Scholar] [CrossRef] [PubMed]
  169. Malanga, D.; Scrima, M.; de Marco, C.; Fabiani, F.; de Rosa, N.; de Gisi, S.; Malara, N.; Savino, R.; Rocco, G.; Chiappetta, G.; et al. Activating E17K mutation in the gene encoding the protein kinase AKT in a subset of squamous cell carcinoma of the lung. Cell Cycle 2008, 7, 665–669. [Google Scholar] [CrossRef] [PubMed]
  170. Davies, M.A.; Stemke-Hale, K.; Tellez, C.; Calderone, T.L.; Deng, W.; Prieto, V.G.; Lazar, A.J.; Gershenwald, J.E.; Mills, G.B. A novel AKT3 mutation in melanoma tumours and cell lines. Br. J. Cancer 2008, 99, 1265–1268. [Google Scholar] [CrossRef] [PubMed]
  171. Cancer Genome Atlas Research Network. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 2008, 455, 1061–1068. [Google Scholar]
  172. Philp, A.J.; Campbell, I.G.; Leet, C.; Vincan, E.; Rockman, S.P.; Whitehead, R.H.; Thomas, R.J.; Phillips, W.A. The phosphatidylinositol 3′-kinase p85α gene is an oncogene in human ovarian and colon tumors. Cancer Res. 2001, 61, 7426–7429. [Google Scholar] [PubMed]
  173. Pedrero, J.M.G.; Carracedo, D.G.; Pinto, C.M.; Zapatero, A.H.; Rodrigo, J.P.; Nieto, C.S.; Gonzalez, M.V. Frequent genetic and biochemical alterations of the PI 3-K/AKT/PTEN pathway in head and neck squamous cell carcinoma. Int. J. Cancer 2005, 114, 242–248. [Google Scholar] [CrossRef] [PubMed]
  174. Woenckhaus, J.; Steger, K.; Werner, E.; Fenic, I.; Gamerdinger, U.; Dreyer, T.; Stahl, U. Genomic gain of PIK3CA and increased expression of p110alpha are associated with progression of dysplasia into invasive squamous cell carcinoma. J. Pathol. 2002, 198, 335–342. [Google Scholar] [CrossRef] [PubMed]
  175. Massion, P.P.; Kuo, W.L.; Stokoe, D.; Olshen, A.B.; Treseler, P.A.; Chin, K.; Chen, C.; Polikoff, D.; Jain, A.N.; Pinkel, D.; et al. Genomic copy number analysis of non-small cell lung cancer using array comparative genomic hybridization implications of the phosphatidylinositol 3-kinase pathway. Cancer Res. 2002, 62, 3636–3640. [Google Scholar] [PubMed]
  176. Staal, S.P. Molecular cloning of the akt oncogene and its human homologues AKT1 and AKT2: Amplification of AKT1 in a primary human gastric adenocarcinoma. Proc. Natl. Acad. Sci. USA 1987, 84, 5034–5037. [Google Scholar] [CrossRef] [PubMed]
  177. Cheng, J.Q.; Ruggeri, B.; Klein, W.M.; Sonoda, G.; Altomare, D.A.; Watson, D.K.; Testa, J.R. Amplification of AKT2 in human pancreatic cells and inhibition of AKT2 expression and tumorigenicity by antisense RNA. Proc. Natl. Acad. Sci. USA 1996, 93, 3636–3641. [Google Scholar] [CrossRef] [PubMed]
  178. Ruggeri, B.A.; Huang, L.; Wood, M.; Cheng, J.Q.; Testa, J.R. Amplification and overexpression of the AKT2 oncogene in a subset of human pancreatic ductal adenocarcinomas. Mol. Carcinog. 1998, 21, 81–86. [Google Scholar] [CrossRef]
  179. Bellacosa, A.; de Feo, D.; Godwin, A.K.; Bell, D.W.; Cheng, J.Q.; Altomare, D.A.; Wan, M.; Dubeau, L.; Scambia, G.; Masciullo, V.; et al. Molecular alterations of the AKT2 oncogene in ovarian and breast carcinomas. Int. J. Cancer 1995, 64, 280–285. [Google Scholar] [CrossRef] [PubMed]
  180. Regad, T. Molecular and cellular pathogenesis of melanoma initiation and progression. Cell. Mol. Life Sci. 2013, 70, 4055–4065. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Regad, T. Targeting RTK Signaling Pathways in Cancer. Cancers 2015, 7, 1758-1784. https://doi.org/10.3390/cancers7030860

AMA Style

Regad T. Targeting RTK Signaling Pathways in Cancer. Cancers. 2015; 7(3):1758-1784. https://doi.org/10.3390/cancers7030860

Chicago/Turabian Style

Regad, Tarik. 2015. "Targeting RTK Signaling Pathways in Cancer" Cancers 7, no. 3: 1758-1784. https://doi.org/10.3390/cancers7030860

Article Metrics

Back to TopTop