Next Article in Journal
A Controlled Study of Major Depressive Episodes in Long-Term Childhood, Adolescence, and Young Adult Cancer Survivors (The NOR-CAYACS Study)
Next Article in Special Issue
Metabolic Plasticity in Melanoma Progression and Response to Oncogene Targeted Therapies
Previous Article in Journal
FGF/FGFR-Dependent Molecular Mechanisms Underlying Anti-Cancer Drug Resistance
Previous Article in Special Issue
Oxidative Stress and Autophagy as Key Targets in Melanoma Cell Fate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tyrosine Kinase Inhibitors in Adult Glioblastoma: An (Un)Closed Chapter?

1
Cancer Signaling Unit, Navarrabiomed, Hospital Universitario de Navarra (HUN), Universidad Pública de Navarra (UPNA), 31008 Pamplona, Spain
2
Health Research Institute of Navarre (IdiSNA), 31008 Pamplona, Spain
*
Authors to whom correspondence should be addressed.
Cancers 2021, 13(22), 5799; https://doi.org/10.3390/cancers13225799
Submission received: 29 October 2021 / Revised: 12 November 2021 / Accepted: 17 November 2021 / Published: 18 November 2021
(This article belongs to the Special Issue Feature Paper from Journal Reviewers)

Abstract

:

Simple Summary

Glioblastoma multiforme (GBM) is the most common type of malignant brain tumor. GBM patients face a dire future, as they rarely survive longer than 15 months after diagnosis. Typically, patients undergo surgery to remove the tumor followed by combined radiotherapy and chemotherapy. However, these therapies usually extend survival only for several months, since tumors invariably regrow, which is called recurrence. Targeted therapies against specific genes that control GBM tumor-growth have been tested in clinical trials for years without success. In this article, we describe the main scientific findings leading to the testing of these targeted therapies in GBM, and discuss the potential causes behind the failure of these clinical trials. We highlight the importance of performing molecular analyses in tumors to determine the presence of those genes controlling GBM growth, before administering drugs specifically blocking their activity. In doing so clinicians could identify patients that could potentially benefit the most. Furthermore, we discuss the reasons to test these drugs in newly diagnosed patients rather than in patients under recurrence. In summary, the aim of this review is to propose alternative approaches for the design of clinical trials testing targeted therapies in GBM patients based on available scientific evidence.

Abstract

Glioblastoma (GBM) is the most common and lethal form of malignant brain tumor. GBM patients normally undergo surgery plus adjuvant radiotherapy followed by chemotherapy. Numerous studies into the molecular events driving GBM highlight the central role played by the Epidermal Growth Factor Receptor (EGFR), as well as the Platelet-derived Growth Factor Receptors PDGFRA and PDGFRB in tumor initiation and progression. Despite strong preclinical evidence for the therapeutic potential of tyrosine kinase inhibitors (TKIs) that target EGFR, PDGFRs, and other tyrosine kinases, clinical trials performed during the last 20 years have not led to the desired therapeutic breakthrough for GBM patients. While clinical trials are still ongoing, in the medical community there is the perception of TKIs as a lost opportunity in the fight against GBM. In this article, we review the scientific rationale for the use of TKIs targeting glioma drivers. We critically analyze the potential causes for the failure of TKIs in the treatment of GBM, and we propose alternative approaches to the clinical evaluation of TKIs in GBM patients.

1. Introduction

Glioblastoma multiforme (GBM) is the most common and malignant adult primary tumor, with an annual incidence rate of 3.23 per 100,000, accounting for almost 50% of all malignant brain tumors [1]. There are 20,000 new cases diagnosed in the United States of America every year; nearly half of the patients are over 65, although GBM can occur at any age [2]. GBM remains among the cancers with poorest prognosis, with a 15-month median overall survival (OS) after diagnosis and a 5-year survival rate below 7% [1]. Conventional management of GBM consists of surgical removal of the tumor, followed by combined radiotherapy and chemotherapy, and usually six monthly cycles of adjuvant temozolomide (TMZ) [3]. However, adjuvant therapy protocols have low success rates, extending survival for just three months, as patients invariably relapse [4]. Recurrence rates reach around 90%, and, in these cases, GBM recurs with a poorer prognosis (i.e., median PFS of 1.5–6 months and median OS of 2–9 months) [5,6]. Recurrences are mostly local, within 2 cm of the initial tumor margin, usually are not accessible to surgery [5], and are less sensitive to therapy than the original tumor; indeed there is no standard treatment for recurrent GBM.
The World Health Organization (WHO) classifies gliomas into four degrees of malignancy, based on histopathological and clinical criteria [7]. GBMs are categorized as grade IV gliomas, characterized by high mitotic activity, microvascular proliferation, necrosis, resistance to apoptosis, invasion to adjacent brain tissue, and accumulation of genomic aberrations [7,8,9]. From a clinical point of view, GBMs have been historically differentiated into two groups: primary and secondary.
Primary or de novo cases occur without clinical evidence of a lower-grade glioma (LGG), and form the majority (90%) of cases, while secondary cases derive from the progression of LGGs [10]. Recently, however, the WHO classification has been updated based on advanced molecular analyses in order to improve the precision of the diagnosis of tumors. This new classification takes into account both the molecular and histological characteristics: it integrates the genotypic and phenotypic parameters and classifies GBMs based on their isocitrate dehydrogenase (IDH) mutation status as IDH-wild type, IDH-mutant, or as GBM with an inconclusive or unavailable IDH mutation status, such as GBM NOS (not otherwise specified) [11]. IDH-wild type GBM displays mutations in the telomerase reverse transcriptase (TERT) promoter (72%), in TP53 (27%) and in the phosphatase and tensin homolog (PTEN) (24%), as well as amplifications (35%) of the epidermal growth factor receptor (EGFR) gene. IDH-mutant GBM also harbors mutations in TP53 (81%), but is characterized by a high degree (71%) in mutations in ATRX [11].
Over time, further molecular alterations typical for GBM have been identified and high-throughput molecular analyses have helped to improve a molecular classification of these tumors.

2. Molecular Classification of GBM

In the last few decades, ambitious molecular studies have revealed the remarkable degree of inter-tumor heterogeneity present in GBM [12,13,14]. Results from The Cancer Genome Atlas (TCGA) program revealed that the most frequent alterations in GBM predominantly affect three cell-signaling pathways: the p53 pathway, the retinoblastoma (Rb) pathway, and the receptor tyrosine kinase (RTK) signaling pathways [13]. Indeed, copy number alterations of RTKs, such as the EGFR or platelet-derived growth factor receptor α (PDGFRA), in GBM had been reported already more than 30 years ago [15,16]. EGFR expression is common in GBM, and tumors characterized by overexpression of the EGFR can predict poor patient survival [17]. Following the detection of EGFR amplifications, gene signatures associated with EGFR overexpression and other GBM clinical features have been defined [17,18,19]. These signatures also helped to identify molecular subclasses such as proneural, mesenchymal, and proliferative [20]. In the latter study, EGFR amplifications were predominantly found in tumors of either proliferative or mesenchymal subclasses.
Further characterization by the TCGA led to the definition of four molecular GBM subtypes—mesenchymal, classical, neural, and proneural—taking into account genetic and epigenetic alterations, response to treatments, and prognosis [21]. Thereby, the mesenchymal subtype is linked to mutation or deletion of the two well-known tumor suppressor genes neurofibromin 1 (NF1) and PTEN. NF1 encodes a Ras-GTPase activating protein that blocks RAS signaling, while the product of PTEN suppresses the activation of the PI3K pathway downstream of RAS [22]. The classical subtype is associated with amplification and mutation of EGFR, deletion of the cyclin-dependent kinase inhibitors CDKN2A and 2B, and gain of chromosome 7 concomitant with loss of chromosome 10. The neural subtype also displays EGFR amplification and PTEN deletions, and the proneural subtype is associated with PDGFRA gene mutation and/or amplification, as well as TP53 and IDH1 mutations.
Importantly, distinct cells in the same tumor often recapitulate programs from distinct subtypes, and these subtypes can co-exist in different regions of the same tumor, leading to intratumor heterogeneity [14,23,24]. Nevertheless, subtypes can change during tumor development and also through therapy, whereby disease progression and therapy resistance is predominantly linked to the mesenchymal phenotype [20,25].
Recently, an integrative unified model of cellular states, genetics, and plasticity has been described [26]. Single-cell lineage tracing and expression analysis described that malignant cells in GBM exist in four main cellular states that are related to distinct neural cell types. The individual cellular states exhibit plasticity, and their establishment is influenced by signals from the tumor microenvironment. The distinct states have been described as neural progenitor-like (NPC-like), oligodendrocyte-progenitor-like (OPC-like), astrocyte-like (AC-like) and, finally, mesenchymal-like (MES-like) states. The relative frequency of cells in each state is influenced by copy number amplifications of the CDK4, EGFR and PDGFRA loci, and by mutations in the NF1 locus [26]. For instance, TCGA tumors with high-level genetic amplification of EGFR are associated with higher AC-like bulk scores. High-level amplifications of PDGFRA and CDK4 are associated with OPC-like and NPC regulators, and NF1 alterations were correlated with MES-high tumors [26].
Apart from gene amplification, GBM presents EGFR deletions and point mutations. There are several variants of EGFR, defined by the deletion of different exons: EGFRvI (deletion of exons 1–13) [27]; EGFRvII (exons 14 and 15 deletion) [28]; EGFRvIII (exon 2 to 7); EGFRvIV (exons 25–27); and EGFRvV (exons 25–28) [29]. The protein product of the EGFRvIII variant lacks a sequence of 267 amino acids in the extracellular ligand-binding domain, leading to a constitutive activation of the EGFR pathway. This variant is the most common, and can be found in 20–50% of GBMs that carry EGFR amplification [13]. Furthermore, EGFR point mutations have been identified in almost 25% of GBM samples. Such mutations include the substitution of arginine by lysine at position 108 (R108K), the substitution of an alanine at residue 289 by valine, aspartic acid or threonine (A289V/D/T), and the substitution of glycine by aspartic acid at position 598 (G598D). These mutations lead to a constitutively active kinase activity [30].
Regarding PDGFRA, it has been shown to be altered by overexpression, amplification, mutation, or rearrangements. PDGFRA gene rearrangements, such as fusion between kinase insert domain receptor (KDR) (VEGFRII) and PDGFRA gene, have been found in PDGFRA-amplified GBM. Intragenic deletion rearrangements such as PDGFRAΔ8, 9, which is formed by an in-frame deletion of 243 bp in exons 8 and 9 of the extracellular domain, were observed in 40% of the GBM with PDGFRA amplification [31]. Different studies have also identified PDGFRA point mutations localized in different domains of the receptor, mainly in the Ig-like extracellular domain, such as C235Y and W349C, potentially disrupting ligand interaction [21]. More mutations have been reported in the extracellular domain, such as Y288H and the activating mutations Y288C and P345S. In addition, activating mutations like V561D in the juxtamembrane and D842V in kinase domain have been encountered, In addition to R1037K mutation, which is found in the intracellular kinase domain, and s frequent in 1 of 316 samples [32]. Some of those mutations, such as V536E, which can be found in the transmembrane domain, have been functionally characterized, and have reported a gain of function by stimulating cell growth and signaling via ERK and STAT5 in the absence of ligand [33].
Although currently not considered for GBM tumor subtyping, in 50% of cases with EGFR amplifications, tumors also carry a truncated form of EGFR that possesses constitutive kinase activity, which is named EGFRvIII, or otherwise EGFR type III, de2–7, ΔEGFR) [34]. Increased EGFRvIII signaling has been correlated with glioma progression and poor prognosis. Finally the RTK MET (Mesenchymal–Epithelial Transition factor) is also amplified in a significant proportion of gliomas, and appears to be co-amplified with EGFRvIII [35]. Apart from MET amplification and overexpression, a mutant form of MET, METex14 is present in 14% of secondary GBM and is characterized by the deletion of exon 14 in the intracellular domain, which harbors the binding site for Cbl. As a consequence, METex14 shows constitutive activation and decreased protein degradation [36].
Altered expression of fibroblast growth factors receptors (FGFRs) has also been described in GBM. For instance, FGFR1 and FGFR2 gene amplification, abnormal activation, or single nucleotide polymorphisms (SNPs) [37]. In addition, oncogenic fusions of FGFR3 and FGFR1 to the transforming acidic coiled-coil (TACC) proteins, which generate oncogenic forms, have been reported [38]. In fact, FGFR3-TACC3 fusion gene is present in 3% of GBMs [38]. Recently, a comparative integrated analysis found four new clinically actionable alterations in FGFR2 and FGFR3 that promote an aggressive phenotype [39].

3. RTK as Drivers in GBM

3.1. RTK Signalling

EGFR, also known as ErbB1/HER1, is one of four members of the EGFR/ErbB family in humans. The other members are ErbB2/HER2, ErbB3/HER3, which have an impaired kinase domain, and ErbB4/HER4. The receptors can form homo- and hetero dimers, depending on the ligands they interact with. Ligand binding induces dimerization, which results in a conformational change, and kinase domain activation followed by auto/trans-phosphorylation events at specific tyrosine residues that fully activate the individual receptors [40]. Ligands that induce EGFR homodimers are EGF itself, transforming growth factor α (TGF-α), amphiregulin (ARG), and epigen (EGN) [40]. After ligand binding, the fully activated EGFR can bind, and phosphorylate a wide range of effectors, including kinases of the SRC family, PLC gamma, the RAC guanine-nucleotide exchange factor VAV, transcription factors such as STAT5, and, most importantly, adaptor proteins that trigger the activation of the RAS GTPases and subsequent stimulation of the RAF/MAPK and PI3K-AKT pathways. EGFRvIII, the truncated form of EGFR, has a deletion of exons 2–7, affecting the extracellular domain. As a consequence, EGFRvIII cannot bind EGF, but shows constitutive activity due to decreased receptor internalization and protein degradation [41].
Receptor dimerization also plays a role in PDGFR signaling, whereby PDGFRa and PDGFRb can homo- or heterodimerize. These dimers are activated by binding of platelet-derived growth factor a, b, c, and d (PDGFA, B, C or D), which requires prior dimerization of the ligands [42]. Specifically, PDGFAA, BB, CC and AB ligand dimers activate PDGFRAA homodimers, while the PDGFRAB heterodimer is mostly activated PDGFBB, PDGFAB dimers, and PDGFRBB dimers are activated by PDGFBB and PDGFDD [42]. The binding of PDGFs and activation of PDGFRA/B can occur in a exocrine, paracrine, and autocrine manner [43]. Similarly to the EGFR, ligand-bound PDGFRs auto/trans-phosphorylate specific tyrosine residues, which leads to the binding and activation of many of the abovementioned EGFR effectors, including the RAF/MAK and PI3K/AKT pathways. The RTK MET is a receptor for the hepatocyte growth factor (HGF), and it is synthesized as a single-chain precursor that is cleaved by furin yielding an N-terminal α-chain and a C-terminal β-chain [44]. Active, cleaved HGF binds the alpha and the beta chain to induce receptor dimerization and subsequent activation [45]. Activated, autophoshorylated MET recruits and activates downstream effectors that include SRC kinases and, via RAS, the RAF/MAPK and PI3K/AKT pathways [46].
The duration and intensity of RTK activation is crucial for the establishment of distinct transcriptional programs that impact on cell proliferation, differentiation, survival, migration, invasion etc. Furthermore, such responses are also dependent on cell type specific traits and the context of the tissue in which RTK activation and signaling takes place [47].
Increased knowledge of the structure, activating mechanisms of RTKs and downstream signaling modules have substantially improved our understanding of the cellular machinery that mediates gliomagenesis and maintains the malignant phenotype of transformed glia.

3.2. Pre-Clinical Evidence of RTK Involvement in GBM

Preclinical studies on RTK signaling in brain tumors involved established glioma cell lines, xenograft tumors (cell line-based and patient derived) and genetically engineered mouse models (GEMM). Over the years, these models have led to a better understanding of the role of EGFR, PDGFRs, and other RTKs in gliomagenesis as well as glioma progression, and have been crucial for testing new, potentially active therapeutic agents.
The use of GEMMs has demonstrated that deregulated PDGFR or EGFR signaling in an adequate genetic background promotes gliomagenesis [48,49,50]. In order to understand EGFR signaling in GBM and to better predict the efficacy of targeted therapeutics, a variety of preclinical models of GBM based on overexpression of EGFR and EGFRvIII have been developed [51]. The majority of glioma mouse models have employed cre/LoxP technology to create genetic alterations. A modification of this technology incorporates a Cre-mediated multifluorescent protein expressing system, which allows for the dissection of developmental processes of gliomagenesis and detecting morphologically heterogeneous tumor populations in gliomas [52]. Alternative systems to generate targeted mutations in a tissue-specific manner are the RCAS (replication-competent ASLV long terminal repeat with a splice acceptor) [53] and the Sleeping Beauty (SB) transposon/transposase system [54]. Studies using these systems showed that expression of EGFR in adult brain tissues is not a transforming event, but that loss of p16lnk4a, p19Arf, and PTEN cooperates with EGFR in gliomagenesis. They also showed that EGFR signals through its canonical pathways, whereas tumors arising from expression of mutant EGFRvIII do not use these same pathways [51]. A murine PDGFB-driven glioblastoma model based on the RCAS/Tva system [49] has been used to study the effects of corticosteroids within the tumor microenvironment and their negative impact on radiotherapy [55]. A PDGFRα-driven mouse model based on autocrine receptor stimulation revealed that the tubulin-binding protein Stathmin1 is a PDGFRα phospho-regulated target, whose misregulation confers sensitivity to vinblastine (VB) cytotoxicity [56]. In addition, plenty of studies have been made in order to study the role of PDGF ligands in gliomagenesis [57]. For instance, transgenic mice expressing PDGFB on a Tp53 null background develop brain tumors resembling human GBMs [58]. Also injection of RCAS-PDFGA in Pten of Cdkn2a null mice [59]. For instance an animal model of ATRX-deficient GBM was created to show that loss of ATRX accelerated tumor growth rate and reduced median survival [60].
Early genetic approaches utilizing RNA interference suggested that EGFR depletion from glioma cells could induce a partial cell cycle arrest in G2M [61], while later studies showed that treatment with EGFR specific siRNAs had no inhibitory effect on cell proliferation, migration, and activation status of EGFR-coupled signaling cascades [62]. Moreover, sole pharmacological inhibition of EGFR by the tyrosine kinase inhibitor erlotinib displayed no activity in 2D clonogenic survival assays, nor in 3D GBM spheroids [63]. These reports indicate that specific down-regulation or inhibition of EGFR is not sufficient as a single agent therapeutic approach.
However, combining the downregulation of EGFR by siRNA with the up-regulation of PTEN expression in PTEN-deficient U251 cells resulted in cell cycle arrest, suppression of proliferation, reduction in invasion and promotion of apoptosis, and growth reduction of U251 subcutaneous xenografts [64]. Anti-proliferative effects were also seen when cells were treated with the EGFR inhibitor AG1478 in combination with the GSI-X Notch signaling inhibitor [65]. Moreover, a combination of the EGFR inhibitor afatinib with temozolomide significantly decreased xenograft growth and progression of intracranially injected U87EGFRvIII GBM xenografts [66]. Combining an anti-EGFRvIII monoclonal antibody with rapamycin resulted in an inhibition in the growth of U87-EGFRvIII and U251- EGFRvIII cells when injected subcutaneously into nude mice [67]. Furthermore, D2C7-IT, a novel immunotoxin targeting wild-type EGFR as well as EGFRvIII proteins, when combined with checkpoint inhibitors, improved survival in intracranial glioma models [68].
Targeting the PDGFRA with a neutralizing human monoclonal antibody inhibits the growth of tumor xenografts [69], but, notably, there is more preclinical evidence linking PDGFR-β to glioma proliferation and survival than PDGFRA. A PDGFR-β-specific shRNA can reduce viability and enhance GBM radiosensitivity [70], and silencing PDGFRB by RNAi was shown to enhance the radiosensitivity of C6 glioma cells in vivo [70]. The PDGFRB inhibitor AG1433 induces cytotoxicity in high grade glioma cell lines [71]. The combination of the PDGFRB inhibitor JNJ-10198409 and the IGF-1R inhibitor PPP/CAS 477-47-4 reduced Akt and Erk1/2 phosphorylation, and diminished cell proliferation, through a G2/M blockade of the cell cycle [72]. In addition, the use of anti-PDGF antibodies also resulted in a reduction of cell viability and induction of autophagy in glioma cells [73].

4. Factors Limiting the Effectiveness of TKIs in GBM

RTK inhibitors of the family of TKIs can be grouped into reversible and irreversible inhibitors. Irreversible TKIs bind their target via covalent bounds, whereas reversible inhibitors are based on non-covalent binding. The latter are subdivided between ATP-competitive inhibitors (Type I) that occupy the ATP-binding pocket, and molecules that bind the inactive form of the RTK adjacently to the ATP pocket (Type II). The ATP-binding domain is structurally well conserved among certain RTKs, which limits the binding specificity of these drugs [74]. Most covalent/irreversible TKI tested in GBMs target EGFR (afatinib, neratininb, osimertinib). As evidenced in Table 1, apart from EGFR inhibitors, the vast majority of drugs trialed so far targeting PDGFR signaling in GBM also block several other RTKs. Indeed, some of them, such as ponatinib, vandetanib, dasatinib, and cabozantinib, can affect the activity of more than eight different kinases [75,76] (Table 1). As a consequence of this lack of specificity, the inhibition of driver kinases is diluted and the probability of off target effects leading to the activation of compensation mechanisms is increased. Furthermore such unspecificity can lead to increased systemic toxicity/adverse effects that limit the treatment duration and, therefore, efficacy [77,78].

4.1. Blood Brain Barrier (BBB) and Drug Accumulation

The blood–brain barrier (BBB) is a very selective membrane that limits the entry of drugs, biomolecules, and cells to the central nervous system (CNS). The BBB is bordered by the basal lamina, a glycoprotein-rich extracellular matrix (ECM), also formed of endothelial cells, pericytes, and astrocytes [79]. There are a variety of mechanisms for substances to cross the BBB, such as transmembrane diffusion and active transport among others [80]. Passive diffusion of small molecules through the BBB depends on their lipophilicity; small hydrophobic molecules diffuse transcellularly, whereas small hydrophilic compounds can enter the brain via the paracellular route [79]. Active transport across the BBB endothelium is regulated by ATP-binding cassette transporters (ABC transporters) located within vessel walls. ABC transporters regulate efflux from the endothelium into the luminal compartment. However, these ABC transporters are often responsible of decreasing the uptake rate of potential drugs crossing de BBB, since most anti-neoplastic low molecular weight drugs are substrates for ABC proteins [79]. The most common ABC transporters are P-glycoprotein (P-gp; also known as ABCB1), and breast cancer resistance protein (BCRP; also known as ABCG2) [79]. For instance, brain accumulation of TKIs such as regorafenib, gefitinib, and tivozanib is restricted by P-gp and BCRP [81,82,83]. Some other studies showed that oral administration of Imatinib resulted only in a marginal flux across the blood-brain barrier [84,85]. There are some other drugs, such as axitinib and tesevatinib, that were detected in the brain of the animals, and were able to permeabilize an in vitro BBB model, which strongly suggests that it could efficiently reach human brain tumors [86,87]. Sunitinib, dasatinib, and sorafenib are drugs capable of entering the brain [88,89,90]. Studies assessing the pharmacokinetic of the multikinase inhibitor ponatinib demonstrated that P-gp and BCRP restricted ponatinib brain accumulation [91,92], although a comparative analysis showed that ponatinib has better BBB penetration, and achieves higher brain plasma concentrations than dasatinib [93].
Nevertheless, there is evidence suggesting that the BBB is by no means intact, and that therefore the failure of TKI in GBM cannot be entirely blamed on the BBB blocking drug access to the brain. As shown in Table 2 and Table 3, most clinical trials with TKIs have traditionally recruited patients under recurrence. These patients have undergone surgery and radiotherapy prior to chemotherapy, with both treatments debilitating BBB integrity [94]. Furthermore, tumor-induced/associated neovasculature (the main characteristic of GBM) is leaky and more disorganized than the normal/physiological BBB. However, there is ample evidence that, generally speaking, TKIs do not reach high-enough intra-tumoral therapeutic concentrations. In this regard, the formulation of drugs and the incorporation of nanocarriers and other drug delivery systems might provide renewed hope to chemotherapy in brain tumors. Similarly, local administration of drugs directly into the tumor resection cavity is another strategy being considered in order to bypass the restrictions imposed by the BBB [94,95,96].

4.2. Patient Selection

At present, the inclusion criteria used in clinical trials testing TKIs against GBM RTKs are based on the patients status [97]. Most trials consider diverse parameters, such as number of leucocytes and thrombocytes; absence of intracerebral inflammation; adequate hepatic, renal and bone marrow function; and previous treatment received (e.g., surgery undergone), among other factors. The vast majority of trials using small molecules inhibitors targeting RTKs in GBM have not taken into account tumor subtype. In other words, the expression of either EGFR or PDGFRs, drivers of glioma progression, is rarely included when assessing the efficacy of a drug. Only around 15% of all those clinical trials have taken into account expression of RTKs, and 9 out of 10 of them are focused on recurrent GBM rather than newly diagnosed patients [97]. For instance, in an open-label trial of imatinib mesylate with patients with unresectable, recurrent GBM expressing PDGFR (NCT00171938, 2004), immunohistochemical documentation of expression of PDGFR was required for inclusion [97]. In 2012, another multicenter study (NCT01520870) assessed the efficacy and safety of the multi-kinase dacomitinib in patients with recurrent GBM with EGFR gene amplification and/or EGFRvIII mutation, which was determined by in situ hybridization fluorescent (FISH) and/or PCR respectively in the primary tumor [97]. On the other hand, and as the authors acknowledge, no proof of the stability of EGFR amplification in recurrent tumors was provided [97]. More recently, a proof of concept trial investigating crenolanib monotherapy (NCT02626364) in patients with recurrent/refractory GBM included patients with PDGFRA gene amplification, as determined by FISH, at the time of diagnosis or time of recurrence [97]. In 2020, the trial NCT04424966 enrolled participants with recurrent high-grade glioma with FGFR1 K656E or FGFR3 K650E mutations or FGFR3-TACC3 translocation (demonstrated by NGS sequencing, IHC and/or RT- PCR) for the clinical assessment of infigratinib [97].
Historically the majority of clinical trials have focused on recurrent patients compared to newly diagnosed patients (Table 4). However, the proportion of ongoing trials focused on newly diagnosed GBM patients seems to be increasing (Table 4). Notably, less than 25% of the ongoing trials incorporated RTK target expression as inclusion criteria. The NCT03231501 trial started in 2018 to evaluate the EGFR inhibitor epitinib in newly diagnosed patients with EGFR gene amplification [97]. Nevertheless, it is still fairly obvious that the molecular landscape of GBM patients has rarely been used as inclusion criteria.
Certainly, patient stratification based on target/RTK expression is relevant in newly diagnosed patients. These patients are normally subjected to surgery for tumor removal prior to conventional adjuvant therapy. In principle, target expression studies could be carried out, either prospectively or retrospectively. This approach could provide useful insight into the efficacy of TKIs and its correlation with target expression. Unfortunately, such an approach is more difficult to realize in clinical trials recruiting patients with recurrent GBM. These patients are less likely to undergo surgery, and it has become more evident that the genetic landscape of recurrent tumors does not correlate with that of the primary tumor of origin [98,99]. Indeed the mutagenic characteristic of the DNA alkylating agent temozolomide can drive the evolution of recurrent GBM [100]. Similarly, ionizing radiation can change the mutational profile of primary gliomas via the induction of double-strand breaks. Analysis of post-radiation occurring high-grade astrocytomas showed that, compared to spontaneous high-grade gliomas, radiated tumors had an increased prevalence of genomic aneuploidy. This was accompanied by a significant increase in the frequency of PDGFRA, MET, BRAF, and RRAS2 amplifications [101]. Thus there is ample evidence of the potential of current adjuvant therapies to change the genetic landscape of first diagnosed tumor compared to recurrent tumors [102]. Consequently, analysis of primary tumors cannot guide the trial of TKIs on recurrent GBM.
There is clearly a need to develop clinical trials where TKIs are tested as post-surgical adjuvant therapy, whether in newly diagnosed patients or in recurrence. That said, there is lack of information regarding the response of newly diagnosed patients to TKIs. This subset of patients can be easily stratified upon for their tumor molecular profile. Furthermore, newly diagnosed patients are more likely to present a better fitness profile that patients under recurrence [103,104,105].

4.3. Tumor Heterogeneity

As mentioned above, distinct cells in the same tumor frequently represent different subtypes, which can co-exist within the tumor, leading to intratumor heterogeneity. GBM tumor subtyping is based on gene amplification of actionable key glioma drivers (50% EGFR amplification in classical or 10–15% PDGFRA amplification in proneural subtypes). However, intratumor heterogeneity leads to an overlap in the expression of these characteristic markers, leading to a mosaic pattern of receptor expression [23]. For instance, mRNA and protein expression studies have shown that high expression and activation of these ‘marker’ RTKs is more widespread than indicated by copy number analyses. Indeed, the PDGFRA protein has been detected in between 25–75% of GBM tumors [106,107], and PDGFRB protein (able to activate similar pathways than PDGFRA via dimerization) may be expressed in up to 60% of cases [106]. Moreover, the ligand PDGFA, which can activate PDGFRA as well as PDGFRB is expressed in up to 80% of tumors [104]. Finally, PDGFR expression is not only dependent on the genetic traits of tumor cells, but can also be regulated by surrounding stromal cell populations such as microglial cells [108].
Using mass spectrometry, Schaff et al. were able to detect and quantify EGFR protein in 48 out of 51 GBM samples, including 22 cases with no EGFR amplification [109]. Verhaak et al. showed that in many tumors of the proneural type EGFR was co-expressed with PDGFRA and vice versa, ‘classical’ tumor types are often positive for PDGFRA. Szerlip et al. described the co-amplification and activation of different RTKs; indeed 43% of GBM with PDGFRA amplification displayed co-amplification of EGFR or MET [110].
This intratumor heterogeneity in RTK expression might reflect what has recently been proposed, namely that GBM tumors represent a dynamic system based on cellular states [26], which adds even more complexity.
Apart from the heterogeneity found amongst individual tumor cells, one must also consider that the tumor cells interact with, and thus shape the tumor microenvironment. This microenvironment comprises astrocytes, neurons, pericytes, endothelial cells, and fibroblasts, as well as immune cells, including macrophages and microglia [111]. Macrophages and the microglia appear to be enriched in tumors with NF1 deficiency, but how this is linked to the expression or activation of particular RTKs is not known [14].
Cells of the tumor microenvironment can be compromised and exploited by the tumor cells. These cells establish a particular extracellular matrix environment and secrete an array of cytokines and growth factors, such as EGF, PDGF, and VEGF, that impact on tissue remodeling and angiogenesis [112], but can potentially also activate RTKs in glioma cells, as it has been established, for instance, in melanoma [113]. Indeed, microglia can stimulate the invasion of glioma cells, and this is partially dependent on EGFR activation [114]. Endothelial cells have been shown to support the propagation of brain tumor stem cells [115], which pose a source for therapy resistance. Importantly however, TKIs such as sunitinib and ponatinib are able to suppress the self-renewing capacity of glioma stem cells [106]. Overall, while the impact of microenvironment on the efficacy of immunotherapies is currently studied, not much is known yet about how it will interfere with the action of TKIs. Nevertheless, it is conceivable that intratumor heterogeneity allows unresponsive tumor cells to escape drug treatment, and that a heterogeneous tumor microenvironment could support drug treated cells, which together poses a crucial challenge to TKI-based therapies in GBM.

4.4. Mechanisms of Resistance to TKIs

There are several mechanisms described by which tumor cells withstand RTK inhibition. These involve mutations and target gene amplification, as well as autocrine re-activation of the receptor and mutations in down-stream signaling components.
Due to its high frequency of mutation and overexpression in GBM, EGFR targeting TKIs are widely used in pre-clinical and clinical studies, and have so far provided some inside into resistance mechanisms. A common phenomenon in acquired resistance to EGFRi is the appearance of inhibitor resistant mutations such as the T790M mutation, which hinders first generation ATP-competitive inhibitors to bind the kinase domain [116]. Third-generation irreversible inhibitors such as osimertinib, which is currently being trialed in GBM (Table 1), are thought to overcome such a challenge [117]. However, targeting EGFR in glioblastoma is particularly challenged by the amplification of extrachromosomal DNA, which contributes to significant variability in the expression of EGFRvIII [118]. Epigenetic down-regulation of EGFRvIII expression as response to drug treatment has been demonstrated with erlotinib in pre-clinical GBM models, leading to so called ‘target independence’ [119]. Compensatory activation of alternative RTKs can also be achieved through epigenetic mechanisms. For instance, PDGFRB expression is induced by EGFR inhibitors, leading to EGFRi resistance [120]. Alternative reactivation of downstream signaling pathways, such as PI3K signaling by PTEN loss, is another resistance mechanism found with EGFR TKis [121]. There is also sufficient preclinical evidence showing that activation of MET, IGF1R, ERBB2, or ERBB3 confers resistance to EGFR TKis [122,123].
Antitumor efficacy to PDGFR-targeting drugs seems to depend on similar compensatory signaling mechanisms, as the co-expression of ERBB3, IGF1R and TGFBR2 in PDGFR expressing glioblastoma cells contributes to PDGFR inhibitor resistance [124]. In addition, insulin can promote resistance to PDGFR inhibition in gliomas driven by PDGFB [125]. Thus, intriguingly the individual mechanisms to escape the inhibitory effects of TKIs share common signaling components, which can be informative for the design of more effective combination therapies.

4.5. A Role for Corticoids in the Design of RTKi Trials?

Neuro-inflammation and peritumoral edema is one of the main factors affecting the wellbeing of GBM patients with regard to neurologic symptoms, such as blurred vision, dizziness, nausea, aphasia, and headaches [126]. These comorbidities have great impact in the course of the disease. Glioblastoma patients are very often administered glucocorticoids (mainly dexamethasone) to manage brain swelling, and increasing evidence is mounting suggesting a negative correlation between dexamethasone administration and overall survival [106,127,128,129]. As has been described in many cell and tumor types, glucocorticoid receptor activation leads to wide transcriptomic rewiring. Importantly, in GBM dexamethasone promotes the hyperactivation of the PDGFR pathway, establishing a transcriptional program that promotes radio-resistance through bypassing the mitotic checkpoint via modulation of the Spindle Assembly Complex [106]. However, treatment with multikinase inhibitors targeting PDGFR overcame the radioprotective activity of dexamethasone. Moreover combining dexamethasone with TKIs produced a synergistic inhibitory effect on tumor cell growth in vitro and in vivo [106]. These findings suggest a complex interplay between dexamethasone and TKIs. However, glucocorticoid administration has not been taken into account when designing, developing, and analyzing the data of clinical trials assessing TKIs.

5. Conclusions

Despite global efforts to incorporate targeted therapy into GBM management, the Stupp regime continues to be the standard of care for GBM patients. Inhibitors of RTKs that drive GBM progression have so far failed to produce significant clinical results. In the face of the dire prognosis that GBM patients endure, this situation is clearly discouraging. However, considering the weaknesses in the design of previous trials, it appears there is room for improvement, and TKIs should not yet be discarded as potentially therapeutic opportunities in GBM. In this review, we have aimed to highlight the factors that might need to be taken into account in order to better conceive future clinical trials testing TKIs in GBM patients. There is compelling evidence to include molecular testing for TKI target expression in order to enable patient stratification. Moreover, such information will allow for the meaningful analysis of patient data from a completed trial. Notably, we propose that analysis of protein expression levels, together with gene amplification/mutation, should be implemented to provide a more accurate view of the molecular landscape for each patient. For instance, immunohistochemical analyses could inform about the impact of tumor heterogeneity on patients’ response to treatment. Since both genetic and protein expression analysis can be easily performed, clinicians have the opportunity to use either of them as inclusion criteria, or for posterior correlation analyses between clinical response and molecular profile. Given the increased evidence demonstrating that the mutational and/or molecular landscape of recurrent tumors can differ from that of newly diagnosed tumor tissue, it would be important to, at least, obtain biopsies of the recurrent tumor for analyses if resection is not an option. Alternatively, the option of assessing TKIs in newly diagnosed GBM patients would be the most informative option. In this regard, the combination of present adjuvant therapies (radiotherapy and/or temozolomide) together with TKIs appears to be an understudied therapeutic opportunity.

Author Contributions

Conceptualization, investigation: P.A. and I.A. writing—original draft preparation: P.A. and I.A. Writing—review and editing: P.A. and I.A. All authors have read and agreed to the published version of the manuscript.

Funding

P.A. is a recipient of the Navarrabiomed Postdoctoral Fellowship. I.A. also acknowledges support through a Miguel Servet II fellowship program (Ref: CPII20/00011) from the Instituto de Salud Carlos III-FEDER.

Acknowledgments

We are indebted to Claudia Wellbrock for helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ostrom, Q.T.; Patil, N.; Cioffi, G.; Waite, K.; Kruchko, C.; Barnholtz-Sloan, J.S. CBTRUS statistical report: Primary brain and other central nervous system tumors diagnosed in the United States in 2013–2017. Neuro-Oncology 2020, 22, IV1–IV96. [Google Scholar] [CrossRef]
  2. Ostrom, Q.T.; Bauchet, L.; Davis, F.G.; Deltour, I.; Fisher, J.L.; Langer, C.E.; Pekmezci, M.; Schwartzbaum, J.A.; Turner, M.C.; Walsh, K.M.; et al. The epidemiology of glioma in adults: A state of the science review. Neuro-Oncology 2014, 16, 896–913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Stupp, R.; Mason, W.P.; van den Bent, M.J.; Weller, M.; Fisher, B.; Taphoorn, M.J.B.; Belanger, K.; Brandes, A.A.; Marosi, C.; Bogdahn, U.; et al. Radiotherapy plus Concomitant and Adjuvant Temozolomide for Glioblastoma. N. Engl. J. Med. 2005, 352, 987–996. [Google Scholar] [CrossRef]
  4. Stupp, R.; Hegi, M.E.; Mason, W.P.; van den Bent, M.J.; Taphoorn, M.J.; Janzer, R.C.; Ludwin, S.K.; Allgeier, A.; Fisher, B.; Belanger, K.; et al. Effects of radiotherapy with concomitant and adjuvant temozolomide versus radiotherapy alone on survival in glioblastoma in a randomised phase III study: 5-year analysis of the EORTC-NCIC trial. Lancet Oncol. 2009, 10, 459–466. [Google Scholar] [CrossRef]
  5. Weller, M.; Cloughesy, T.; Perry, J.R.; Wick, W. Standards of care for treatment of recurrent glioblastoma-are we there yet? Neuro-Oncology 2013, 15, 4–27. [Google Scholar] [CrossRef] [Green Version]
  6. Audureau, E.; Chivet, A.; Ursu, R.; Corns, R.; Metellus, P.; Noel, G.; Zouaoui, S.; Guyotat, J.; Le Reste, P.J.; Faillot, T.; et al. Prognostic factors for survival in adult patients with recurrent glioblastoma: A decision-tree-based model. J. Neurooncol. 2018, 136, 565–576. [Google Scholar] [CrossRef] [PubMed]
  7. Louis, D.N.; Ohgaki, H.; Wiestler, O.D.; Cavenee, W.K.; Burger, P.C.; Jouvet, A.; Scheithauer, B.W.; Kleihues, P. The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol. 2007, 114, 97–109. [Google Scholar] [CrossRef] [Green Version]
  8. Furnari, F.B.; Fenton, T.; Bachoo, R.M.; Mukasa, A.; Stommel, J.M.; Stegh, A.; Hahn, W.C.; Ligon, K.L.; Louis, D.N.; Brennan, C.; et al. Malignant astrocytic glioma: Genetics, biology, and paths to treatment. Genes Dev. 2007, 21, 2683–2710. [Google Scholar] [CrossRef] [Green Version]
  9. Parsons, D.W.; Jones, S.; Zhang, X.; Lin, J.C.-H.; Leary, R.J.; Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I.-M.; Gallia, G.L.; et al. An integrated genomic analysis of human glioblastoma multiforme. Science 2008, 321, 1807–1812. [Google Scholar] [CrossRef] [Green Version]
  10. Ohgaki, H.; Kleihues, P. Genetic Pathways to Primary and Secondary Glioblastoma. Am. J. Pathol. 2007, 170, 1445–1453. [Google Scholar] [CrossRef] [Green Version]
  11. Louis, D.N.; Perry, A.; Reifenberger, G.; von Deimling, A.; Figarella-Branger, D.; Cavenee, W.K.; Ohgaki, H.; Wiestler, O.D.; Kleihues, P.; Ellison, D.W. The 2016 World Health Organization Classification of Tumors of the Central Nervous System: A summary. Acta Neuropathol. 2016, 131, 803–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. McLendon, R.; Friedman, A.; Bigner, D.; Van Meir, E.G.; Brat, D.J.; Mastrogianakis, G.M.; Olson, J.J.; Mikkelsen, T.; Lehman, N.; Aldape, K.; et al. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 2008, 455, 1061–1068. [Google Scholar] [CrossRef]
  13. Brennan, C.W.; Verhaak, R.G.W.; McKenna, A.; Campos, B.; Noushmehr, H.; Salama, S.R.; Zheng, S.; Chakravarty, D.; Sanborn, J.Z.; Berman, S.H.; et al. The somatic genomic landscape of glioblastoma. Cell 2013, 155, 462–477. [Google Scholar] [CrossRef]
  14. Wang, Q.; Hu, B.; Hu, X.; Kim, H.; Squatrito, M.; Scarpace, L.; deCarvalho, A.C.; Lyu, S.; Li, P.; Li, Y.; et al. Tumor Evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with Immunological Changes in the Microenvironment. Cancer Cell 2017, 32, 42–56.e6. [Google Scholar] [CrossRef] [Green Version]
  15. Libermann, T.A.; Nusbaum, H.R.; Razon, N.; Kris, R.; Lax, I.; Soreq, H.; Whittle, N.; Waterfield, M.D.; Ullrich, A.; Schlessinger, J. Amplification, enhanced expression and possible rearrangement of EGF receptor gene in primary human brain tumours of glial origin. Nature 1985, 313, 144–147. [Google Scholar] [CrossRef]
  16. Muleris, M.; Almeida, A.; Dutrillaux, A.M.; Pruchon, E.; Vega, F.; Delattre, J.Y.; Poisson, M.; Malfoy, B.; Dutrillaux, B. Oncogene amplification in human gliomas: A molecular cytogenetic analysis. Oncogene 1994, 9, 2717–2722. [Google Scholar]
  17. Freije, W.A.; Castro-Vargas, F.E.; Fang, Z.; Horvath, S.; Cloughesy, T.; Liau, L.M.; Mischel, P.S.; Nelson, S.F. Gene expression profiling of gliomas strongly predicts survival. Cancer Res. 2004, 64, 6503–6510. [Google Scholar] [CrossRef] [Green Version]
  18. Mischel, P.S.; Shai, R.; Shi, T.; Horvath, S.; Lu, K.V.; Choe, G.; Seligson, D.; Kremen, T.J.; Palotie, A.; Liau, L.M.; et al. Identification of molecular subtypes of glioblastoma by gene expression profiling. Oncogene 2003, 22, 2361–2373. [Google Scholar] [CrossRef] [Green Version]
  19. Shai, R.; Shi, T.; Kremen, T.J.; Horvath, S.; Liau, L.M.; Cloughesy, T.F.; Mischel, P.S.; Nelson, S.F. Gene expression profiling identifies molecular subtypes of gliomas. Oncogene 2003, 22, 4918–4923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Phillips, H.S.; Kharbanda, S.; Chen, R.; Forrest, W.F.; Soriano, R.H.; Wu, T.D.; Misra, A.; Nigro, J.M.; Colman, H.; Soroceanu, L.; et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 2006, 9, 157–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Verhaak, R.G.W.; Hoadley, K.A.; Purdom, E.; Wang, V.; Qi, Y.; Wilkerson, M.D.; Miller, C.R.; Ding, L.; Golub, T.; Mesirov, J.P.; et al. Integrated Genomic Analysis Identifies Clinically Relevant Subtypes of Glioblastoma Characterized by Abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 2010, 17, 98–110. [Google Scholar] [CrossRef] [Green Version]
  22. Arozarena, I.; Calvo, F.; Crespo, P. Ras, an actor on many stages: Posttranslational modifications, localization, and site-specified events. Genes Cancer 2011, 2, 182–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Patel, A.P.; Tirosh, I.; Trombetta, J.J.; Shalek, A.K.; Gillespie, S.M.; Wakimoto, H.; Cahill, D.P.; Nahed, B.V.; Curry, W.T.; Martuza, R.L.; et al. Single-cell RNA-seq highlights intratumoral heterogeneity in primary glioblastoma. Science 2014, 344, 1396–1401. [Google Scholar] [CrossRef] [Green Version]
  24. Sottoriva, A.; Spiteri, I.; Piccirillo, S.G.M.; Touloumis, A.; Collins, V.P.; Marioni, J.C.; Curtis, C.; Watts, C.; Tavaré, S. Intratumor heterogeneity in human glioblastoma reflects cancer evolutionary dynamics. Proc. Natl. Acad. Sci. USA 2013, 110, 4009–4014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fedele, M.; Cerchia, L.; Pegoraro, S.; Sgarra, R.; Manfioletti, G. Proneural-mesenchymal transition: Phenotypic plasticity to acquire multitherapy resistance in glioblastoma. Int. J. Mol. Sci. 2019, 20, 2746. [Google Scholar] [CrossRef] [Green Version]
  26. Neftel, C.; Laffy, J.; Filbin, M.G.; Hara, T.; Shore, M.E.; Rahme, G.J.; Richman, A.R.; Silverbush, D.; Shaw, M.L.; Hebert, C.M.; et al. An Integrative Model of Cellular States, Plasticity, and Genetics for Glioblastoma. Cell 2019, 178, 835–849.e21. [Google Scholar] [CrossRef] [PubMed]
  27. Wong, A.J.; Ruppert, J.M.; Bigner, S.H.; Grzeschik, C.H.; Humphrey, P.A.; Bigner, D.S.; Vogelstein, B. Structural alterations of the epidermal growth factor receptor gene in human gliomas. Proc. Natl. Acad. Sci. USA 1992, 89, 2965–2969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Humphrey, P.A.; Gangarosa, L.M.; Wong, A.J.; Archer, G.E.; Lund-Johansen, M.; Bjerkvig, R.; Laerum, O.-D.; Friedman, H.S.; Bigner, D.D. Deletion-mutant epidermal growth factor receptor in human gliomas: Effect of type II mutation on receptor function. Biochem. Biophys. Res. Commun. 1991, 178, 1413–1420. [Google Scholar] [CrossRef]
  29. Ekstrand, A.J.; Sugawa, N.; James, C.D.; Collins, V.P. Amplified and rearranged epidermal growth factor receptor genes in human glioblastomas reveal deletions of sequences encoding portions of the N- and/or C-terminal tails. Proc. Natl. Acad. Sci. USA 1992, 89, 4309–4313. [Google Scholar] [CrossRef] [Green Version]
  30. Furnari, F.B.; Cloughesy, T.F.; Cavenee, W.K.; Mischel, P.S. Heterogeneity of epidermal growth factor receptor signalling networks in glioblastoma. Nat. Rev. Cancer 2015, 15, 302–310. [Google Scholar] [CrossRef] [Green Version]
  31. Ozawa, T.; Brennan, C.W.; Wang, L.; Squatrito, M.; Sasayama, T.; Nakada, M.; Huse, J.T.; Pedraza, A.; Utsuki, S.; Yasui, Y.; et al. PDGFRA gene rearrangements are frequent genetic events in PDGFRA-amplified glioblastomas. Genes Dev. 2010, 24, 2205–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Ip, C.K.M.; Ng, P.K.S.; Jeong, K.J.; Shao, S.H.; Ju, Z.; Leonard, P.G.; Hua, X.; Vellano, C.P.; Woessner, R.; Sahni, N.; et al. Neomorphic PDGFRA extracellular domain driver mutations are resistant to PDGFRA targeted therapies. Nat. Commun. 2018, 9, 4583. [Google Scholar] [CrossRef] [PubMed]
  33. Velghe, A.I.; Van Cauwenberghe, S.; Polyansky, A.A.; Chand, D.; Montano-Almendras, C.P.; Charni, S.; Hallberg, B.; Essaghir, A.; Demoulin, J.-B. PDGFRA alterations in cancer: Characterization of a gain-of-function V536E transmembrane mutant as well as loss-of-function and passenger mutations. Oncogene 2014, 33, 2568–2576. [Google Scholar] [CrossRef] [Green Version]
  34. Felsberg, J.; Hentschel, B.; Kaulich, K.; Gramatzki, D.; Zacher, A.; Malzkorn, B.; Kamp, M.; Sabel, M.; Simon, M.; Westphal, M.; et al. Epidermal growth factor receptor variant III (EGFRvIII) positivity in EGFR-amplified glioblastomas: Prognostic role and comparison between primary and recurrent tumors. Clin. Cancer Res. 2017, 23, 6846–6855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Paulsson, J.; Lindh, M.B.; Jarvius, M.; Puputti, M.; Nistér, M.; Nupponen, N.N.; Paulus, W.; Söderberg, O.; Dresemann, G.; Von Deimling, A.; et al. Prognostic but not predictive role of platelet-derived growth factor receptors in patients with recurrent glioblastoma. Int. J. Cancer 2011, 128, 1981–1988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Hu, H.; Mu, Q.; Bao, Z.; Chen, Y.; Liu, Y.; Chen, J.; Wang, K.; Wang, Z.; Nam, Y.; Jiang, B.; et al. Mutational Landscape of Secondary Glioblastoma Guides MET-Targeted Trial in Brain Tumor. Cell 2018, 175, 1665–1678.e18. [Google Scholar] [CrossRef] [Green Version]
  37. Lasorella, A.; Sanson, M.; Iavarone, A. FGFR-TACC gene fusions in human glioma. Neuro-Oncology 2017, 19, 475–483. [Google Scholar] [CrossRef] [Green Version]
  38. Singh, D.; Chan, J.M.; Zoppoli, P.; Niola, F.; Sullivan, R.; Castano, A.; Liu, E.M.; Reichel, J.; Porrati, P.; Pellegatta, S.; et al. Transforming Fusions of FGFR and TACC Genes in Human Glioblastoma. Science 2012, 337, 1231–1235. [Google Scholar] [CrossRef] [Green Version]
  39. Georgescu, M.-M.; Islam, M.Z.; Li, Y.; Traylor, J.; Nanda, A. Novel targetable FGFR2 and FGFR3 alterations in glioblastoma associate with aggressive phenotype and distinct gene expression programs. Acta Neuropathol. Commun. 2021, 9, 69. [Google Scholar] [CrossRef]
  40. Lemmon, M.A.; Schlessinger, J.; Ferguson, K.M. The EGFR family: Not so prototypical receptor tyrosine kinases. Cold Spring Harb. Perspect. Biol. 2014, 6, a020768. [Google Scholar] [CrossRef]
  41. 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] [Green Version]
  42. Heldin, C.-H.; Lennartsson, J. Structural and functional properties of platelet-derived growth factor and stem cell factor receptors. Cold Spring Harb. Perspect. Biol. 2013, 5, a009100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Lokker, N.A.; Sullivan, C.M.; Hollenbach, S.J.; Israel, M.A.; Giese, N.A. Platelet-derived growth factor (PDGF) autocrine signaling regulates survival and mitogenic pathways in glioblastoma cells: Evidence that the novel PDGF-C and PDGF-D ligands may play a role in the development of brain tumors. Cancer Res. 2002, 62, 3729–3735. [Google Scholar] [PubMed]
  44. Gherardi, E.; Youles, M.E.; Miguel, R.N.; Blundell, T.L.; Iamele, L.; Gough, J.; Bandyopadhyay, A.; Hartmann, G.; Butler, P.J.G. Functional map and domain structure of MET, the product of the c-met protooncogene and receptor for hepatocyte growth factor/scatter factor. Proc. Natl. Acad. Sci. USA 2003, 100, 12039–12044. [Google Scholar] [CrossRef] [Green Version]
  45. Gherardi, E.; Sandin, S.; Petoukhov, M.V.; Finch, J.; Youles, M.E.; Ofverstedt, L.-G.; Miguel, R.N.; Blundell, T.L.; Vande Woude, G.F.; Skoglund, U.; et al. Structural basis of hepatocyte growth factor/scatter factor and MET signalling. Proc. Natl. Acad. Sci. USA. 2006, 103, 4046–4051. [Google Scholar] [CrossRef] [Green Version]
  46. Mulcahy, E.Q.X.; Colόn, R.R.; Abounader, R. HGF/MET Signaling in Malignant Brain Tumors. Int. J. Mol. Sci. 2020, 21, 7546. [Google Scholar] [CrossRef]
  47. Vasudevan, H.N.; Mazot, P.; He, F.; Soriano, P. Receptor tyrosine kinases modulate distinct transcriptional programs by differential usage of intracellular pathways. Elife 2015, 4. [Google Scholar] [CrossRef]
  48. Dai, C.; Celestino, J.C.; Okada, Y.; Louis, D.N.; Fuller, G.N.; Holland, E.C. PDGF autocrine stimulation dedifferentiates cultured astrocytes and induces oligodendrogliomas from and oligoastrocytomas neural progenitors and astrocytes in vivo. Genes Dev. 2001, 15, 1913–1925. [Google Scholar] [CrossRef] [Green Version]
  49. Hambardzumyan, D.; Amankulor, N.M.; Helmy, K.Y.; Becher, O.J.; Holland, E.C. Modeling adult gliomas using RCAS/t-va technology. Transl. Oncol. 2009, 2, 89–95. [Google Scholar] [CrossRef] [Green Version]
  50. Holland, E.C.; Hively, W.P.; DePinho, R.A.; Varmus, H.E. A constitutively active epidermal growth factor receptor cooperates with disruption of G1 cell-cycle arrest pathways to induce glioma-like lesions in mice. Genes Dev. 1998, 12, 3675–3685. [Google Scholar] [CrossRef] [Green Version]
  51. Zhu, H.; Acquaviva, J.; Ramachandran, P.; Boskovitz, A.; Woolfenden, S.; Pfannl, R.; Bronson, R.T.; Chen, J.W.; Weissleder, R.; Housman, D.E.; et al. Oncogenic EGFR signaling cooperates with loss of tumor suppressor gene functions in gliomagenesis. Proc. Natl. Acad. Sci. USA 2009, 106, 2712–2716. [Google Scholar] [CrossRef] [Green Version]
  52. Hara, T.; Verma, I.M. Modeling Gliomas Using Two Recombinases. Cancer Res. 2019, 79, 3983–3991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Huse, J.T.; Holland, E.C. Genetically engineered mouse models of brain cancer and the promise of preclinical testing. Brain Pathol. 2009, 19, 132–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wiesner, S.M.; Decker, S.A.; Larson, J.D.; Ericson, K.; Forster, C.; Gallardo, J.L.; Long, C.; Demorest, Z.L.; Zamora, E.A.; Low, W.C.; et al. De novo induction of genetically engineered brain tumors in mice using plasmid DNA. Cancer Res. 2009, 69, 431–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Pitter, K.L.; Tamagno, I.; Alikhanyan, K.; Hosni-Ahmed, A.; Pattwell, S.S.; Donnola, S.; Dai, C.; Ozawa, T.; Chang, M.; Chan, T.A.; et al. Corticosteroids compromise survival in glioblastoma. Brain 2016, 139, 1458–1471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Jun, H.J.; Appleman, V.A.; Wu, H.J.; Rose, C.M.; Pineda, J.J.; Yeo, A.T.; Delcuze, B.; Lee, C.; Gyuris, A.; Zhu, H.; et al. A PDGFRα-driven mouse model of glioblastoma reveals a stathmin1-mediated mechanism of sensitivity to vinblastine. Nat. Commun. 2018, 9. [Google Scholar] [CrossRef] [PubMed]
  57. Heldin, C.H.; Lennartsson, J.; Westermark, B. Involvement of platelet-derived growth factor ligands and receptors in tumorigenesis. J. Intern. Med. 2018, 283, 16–44. [Google Scholar] [CrossRef] [Green Version]
  58. Hede, S.-M.; Hansson, I.; Afink, G.B.; Eriksson, A.; Nazarenko, I.; Andrae, J.; Genove, G.; Westermark, B.; Nistér, M. GFAP promoter driven transgenic expression of PDGFB in the mouse brain leads to glioblastoma in a Trp53 null background. Glia 2009, 57, 1143–1153. [Google Scholar] [CrossRef]
  59. Ozawa, T.; Riester, M.; Cheng, Y.-K.; Huse, J.T.; Squatrito, M.; Helmy, K.; Charles, N.; Michor, F.; Holland, E.C. Most human non-GCIMP glioblastoma subtypes evolve from a common proneural-like precursor glioma. Cancer Cell 2014, 26, 288–300. [Google Scholar] [CrossRef] [Green Version]
  60. Koschmann, C.; Calinescu, A.-A.; Nunez, F.J.; Mackay, A.; Fazal-Salom, J.; Thomas, D.; Mendez, F.; Kamran, N.; Dzaman, M.; Mulpuri, L.; et al. ATRX loss promotes tumor growth and impairs nonhomologous end joining DNA repair in glioma. Sci. Transl. Med. 2016, 8, 328ra28. [Google Scholar] [CrossRef] [Green Version]
  61. Fan, Q.-W.; Weiss, W.A. RNA interference against a glioma-derived allele of EGFR induces blockade at G2M. Oncogene 2005, 24, 829–837. [Google Scholar] [CrossRef] [Green Version]
  62. Vollmann, A.; Vornlocher, H.-P.; Stempfl, T.; Brockhoff, G.; Apfel, R.; Bogdahn, U. Effective silencing of EGFR with RNAi demonstrates non-EGFR dependent proliferation of glioma cells. Int. J. Oncol. 2006, 28, 1531–1542. [Google Scholar] [CrossRef] [Green Version]
  63. Gomez-Roman, N.; Stevenson, K.; Gilmour, L.; Hamilton, G.; Chalmers, A.J. A novel 3D human glioblastoma cell culture system for modeling drug and radiation responses. Neuro-Oncology 2017, 19, 229–241. [Google Scholar] [CrossRef] [Green Version]
  64. Han, L.; Zhang, A.; Xu, P.; Yue, X.; Yang, Y.; Wang, G.; Jia, Z.; Pu, P.; Kang, C. Combination gene therapy with PTEN and EGFR siRNA suppresses U251 malignant glioma cell growth in vitro and in vivo. Med. Oncol. 2010, 27, 843–852. [Google Scholar] [CrossRef]
  65. Cenciarelli, C.; Marei, H.E.S.; Zonfrillo, M.; Pierimarchi, P.; Paldino, E.; Casalbore, P.; Felsani, A.; Vescovi, A.L.; Maira, G.; Mangiola, A. PDGF receptor alpha inhibition induces apoptosis in glioblastoma cancer stem cells refractory to anti-Notch and anti-EGFR treatment. Mol. Cancer 2014, 13, 247. [Google Scholar] [CrossRef] [Green Version]
  66. Vengoji, R.; Macha, M.A.; Nimmakayala, R.K.; Rachagani, S.; Siddiqui, J.A.; Mallya, K.; Gorantla, S.; Jain, M.; Ponnusamy, M.P.; Batra, S.K.; et al. Afatinib and Temozolomide combination inhibits tumorigenesis by targeting EGFRvIII-cMet signaling in glioblastoma cells. J. Exp. Clin. Cancer Res. 2019, 38, 266. [Google Scholar] [CrossRef]
  67. Xu, W.; Bi, Y.; Kong, J.; Zhang, J.; Wang, B.; Li, K.; Tian, M.; Pan, X.; Shi, B.; Gu, J.; et al. Combination of an anti-EGFRvIII antibody CH12 with Rapamycin synergistically inhibits the growth of EGFRvIII+PTEN-glioblastoma in vivo. Oncotarget 2016, 7, 24752–24765. [Google Scholar] [CrossRef] [PubMed]
  68. Chandramohan, V.; Bao, X.; Yu, X.; Parker, S.; McDowall, C.; Yu, Y.-R.; Healy, P.; Desjardins, A.; Gunn, M.D.; Gromeier, M.; et al. Improved efficacy against malignant brain tumors with EGFRwt/EGFRvIII targeting immunotoxin and checkpoint inhibitor combinations. J. Immunother. Cancer 2019, 7, 142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Loizos, N.; Xu, Y.; Huber, J.; Liu, M.; Lu, D.; Finnerty, B.; Rolser, R.; Malikzay, A.; Persaud, A.; Corcoran, E.; et al. Targeting the platelet-derived growth factor receptor alpha with a neutralizing human monoclonal antibody inhibits the growth of tumor xenografts: Implications as a potential therapeutic target. Mol. Cancer Ther. 2005, 4, 369–379. [Google Scholar] [CrossRef] [PubMed]
  70. Hong, J.; Wang, X.; Peng, Y.; Peng, J.; Wang, J.; Dong, Y.; He, D.; Peng, Z.; Tu, Q.; Sheng, L.; et al. Silencing platelet-derived growth factor receptor-β enhances the radiosensitivity of C6 glioma cells in vitro and in vivo. Oncol. Lett. 2017, 14, 329–336. [Google Scholar] [CrossRef]
  71. Alexandru, O.; Sevastre, A.-S.; Castro, J.; Artene, S.-A.; Tache, D.E.; Purcaru, O.S.; Sfredel, V.; Tataranu, L.G.; Dricu, A. Platelet-Derived Growth Factor Receptor and Ionizing Radiation in High Grade Glioma Cell Lines. Int. J. Mol. Sci. 2019, 20, 4663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Carrasco-Garcia, E.; Martinez-Lacaci, I.; Mayor-López, L.; Tristante, E.; Carballo-Santana, M.; García-Morales, P.; Ventero Martin, M.P.; Fuentes-Baile, M.; Rodriguez-Lescure, Á.; Saceda, M. PDGFR and IGF-1R Inhibitors Induce a G2/M Arrest and Subsequent Cell Death in Human Glioblastoma Cell Lines. Cells 2018, 7, 131. [Google Scholar] [CrossRef] [Green Version]
  73. Takeuchi, H.; Kanzawa, T.; Kondo, Y.; Kondo, S. Inhibition of platelet-derived growth factor signalling induces autophagy in malignant glioma cells. Br. J. Cancer 2004, 90, 1069–1075. [Google Scholar] [CrossRef] [PubMed]
  74. Ciardiello, F.; Tortora, G. EGFR antagonists in cancer treatment. N. Engl. J. Med. 2008, 358, 1160–1174. [Google Scholar] [CrossRef] [Green Version]
  75. Open Targets Platform. Available online: https://platform.opentargets.org/ (accessed on 11 October 2021).
  76. Drug Bank. Available online: https://go.drugbank.com/ (accessed on 11 October 2021).
  77. Shah, D.R.; Shah, R.R.; Morganroth, J. Tyrosine kinase inhibitors: Their on-target toxicities as potential indicators of efficacy. Drug Saf. 2013, 36, 413–426. [Google Scholar] [CrossRef]
  78. Hartmann, J.T.; Haap, M.; Kopp, H.-G.; Lipp, H.-P. Tyrosine kinase inhibitors—A review on pharmacology, metabolism and side effects. Curr. Drug Metab. 2009, 10, 470–481. [Google Scholar] [CrossRef]
  79. Arvanitis, C.D.; Ferraro, G.B.; Jain, R.K. The blood-brain barrier and blood-tumour barrier in brain tumours and metastases. Nat. Rev. Cancer 2020, 20, 26–41. [Google Scholar] [CrossRef]
  80. Banks, W.A. Characteristics of compounds that cross the blood-brain barrier. BMC Neurol. 2009, 9, S3. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, J.; Bruin, M.A.C.; Gan, C.; Lebre, M.C.; Rosing, H.; Beijnen, J.H.; Schinkel, A.H. Brain accumulation of tivozanib is restricted by ABCB1 (P-glycoprotein) and ABCG2 (breast cancer resistance protein) in mice. Int. J. Pharm. 2020, 581, 119277. [Google Scholar] [CrossRef] [PubMed]
  82. Agarwal, S.; Sane, R.; Gallardo, J.L.; Ohlfest, J.R.; Elmquist, W.F. Distribution of gefitinib to the brain is limited by P-glycoprotein (ABCB1) and breast cancer resistance protein (ABCG2)-mediated active efflux. J. Pharmacol. Exp. Ther. 2010, 334, 147–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kort, A.; Durmus, S.; Sparidans, R.W.; Wagenaar, E.; Beijnen, J.H.; Schinkel, A.H. Brain and Testis Accumulation of Regorafenib is Restricted by Breast Cancer Resistance Protein (BCRP/ABCG2) and P-glycoprotein (P-GP/ABCB1). Pharm. Res. 2015, 32, 2205–2216. [Google Scholar] [CrossRef] [Green Version]
  84. Holdhoff, M.; Supko, J.G.; Gallia, G.L.; Hann, C.L.; Bonekamp, D.; Ye, X.; Cao, B.; Olivi, A.; Grossman, S.A. Intratumoral concentrations of imatinib after oral administration in patients with glioblastoma multiforme. J. Neurooncol. 2010, 97, 241–245. [Google Scholar] [CrossRef] [Green Version]
  85. Senior, K. Gleevec does not cross blood-brain barrier. Lancet Oncol. 2003, 4, 198. [Google Scholar] [CrossRef]
  86. Pagnuzzi-Boncompagni, M.; Picco, V.; Vial, V.; Planas-Bielsa, V.; Vandenberghe, A.; Grépin, R.; Durivault, J.; Montemagno, C.; Martial, S.; Doyen, J.; et al. The anti-angiogenic compound axitinib demonstrates low toxicity and anti-tumoral effects against medulloblastoma. bioRxiv 2020, hal-02991404. [Google Scholar] [CrossRef]
  87. Tonra, J.R.; Poyurovsky, M.; Liu, K.G.; Patel, J.; Rao, N.; Tilton, R.; Ryan, J.L.; Berger, M.S.; Witte, L.; Kim, J.-I.; et al. Abstract 2590: KD019: Blood brain barrier penetrant HER2/neu, Src, and EGFR inhibitor. Cancer Res. 2015, 75, 2590. [Google Scholar] [CrossRef]
  88. Oberoi, R.K.; Mittapalli, R.K.; Elmquist, W.F. Pharmacokinetic assessment of efflux transport in sunitinib distribution to the brain. J. Pharmacol. Exp. Ther. 2013, 347, 755–764. [Google Scholar] [CrossRef] [Green Version]
  89. Porkka, K.; Koskenvesa, P.; Lundán, T.; Rimpiläinen, J.; Mustjoki, S.; Smykla, R.; Wild, R.; Luo, R.; Arnan, M.; Brethon, B.; et al. Dasatinib crosses the blood-brain barrier and is an efficient therapy for central nervous system Philadelphia chromosome-positive leukemia. Blood 2008, 112, 1005–1012. [Google Scholar] [CrossRef] [Green Version]
  90. Wei, J.; Wang, Z.; Wang, W.; Liu, X.; Wan, J.; Yuan, Y.; Li, X.; Ma, L.; Liu, X. Oxidative Stress Activated by Sorafenib Alters the Temozolomide Sensitivity of Human Glioma Cells Through Autophagy and JAK2/STAT3-AIF Axis. Front. Cell Dev. Biol. 2021, 9, 660005. [Google Scholar] [CrossRef] [PubMed]
  91. Laramy, J.K.; Kim, M.; Parrish, K.E.; Sarkaria, J.N.; Elmquist, W.F. Pharmacokinetic Assessment of Cooperative Efflux of the Multitargeted Kinase Inhibitor Ponatinib Across the Blood-Brain Barrier. J. Pharmacol. Exp. Ther. 2018, 365, 249–261. [Google Scholar] [CrossRef]
  92. Kort, A.; van Hoppe, S.; Sparidans, R.W.; Wagenaar, E.; Beijnen, J.H.; Schinkel, A.H. Brain Accumulation of Ponatinib and Its Active Metabolite, N-Desmethyl Ponatinib, Is Limited by P-Glycoprotein (P-GP/ABCB1) and Breast Cancer Resistance Protein (BCRP/ABCG2). Mol. Pharm. 2017, 14, 3258–3268. [Google Scholar] [CrossRef]
  93. Ravi, K.; Franson, A.; Homan, M.J.; Roberts, H.; Pai, M.P.; Miklja, Z.; He, M.; Wen, B.; Benitez, L.L.; Perissinotti, A.J.; et al. Comparative pharmacokinetic analysis of the blood-brain barrier penetration of dasatinib and ponatinib in mice. Leuk. Lymphoma 2021, 62, 1990–1994. [Google Scholar] [CrossRef] [PubMed]
  94. Lundy, D.J.; Nguyễn, H.; Hsieh, P.C.H. Emerging Nano-Carrier Strategies for Brain Tumor Drug Delivery and Considerations for Clinical Translation. Pharmaceutics 2021, 13, 1193. [Google Scholar] [CrossRef] [PubMed]
  95. Alghamdi, M.; Gumbleton, M.; Newland, B. Local delivery to malignant brain tumors: Potential biomaterial-based therapeutic/adjuvant strategies. Biomater. Sci. 2021, 9, 6037–6051. [Google Scholar] [CrossRef] [PubMed]
  96. Miller, J.J.; Wen, P.Y. Emerging targeted therapies for glioma. Expert Opin. Emerg. Drugs 2016, 21, 441–452. [Google Scholar] [CrossRef]
  97. Clinical Trials.gov. Available online: https://clinicaltrials.gov/. (accessed on 3 September 2021).
  98. Wang, J.; Cazzato, E.; Ladewig, E.; Frattini, V.; Rosenbloom, D.I.S.; Zairis, S.; Abate, F.; Liu, Z.; Elliott, O.; Shin, Y.-J.; et al. Clonal evolution of glioblastoma under therapy. Nat. Genet. 2016, 48, 768–776. [Google Scholar] [CrossRef] [Green Version]
  99. Campos, B.; Olsen, L.R.; Urup, T.; Poulsen, H.S. A comprehensive profile of recurrent glioblastoma. Oncogene 2016, 35, 5819–5825. [Google Scholar] [CrossRef]
  100. Johnson, B.E.; Mazor, T.; Hong, C.; Barnes, M.; Aihara, K.; McLean, C.Y.; Fouse, S.D.; Yamamoto, S.; Ueda, H.; Tatsuno, K.; et al. Mutational analysis reveals the origin and therapy-driven evolution of recurrent glioma. Science 2014, 343, 189–193. [Google Scholar] [CrossRef] [Green Version]
  101. López, G.Y.; Van Ziffle, J.; Onodera, C.; Grenert, J.P.; Yeh, I.; Bastian, B.C.; Clarke, J.; Oberheim Bush, N.A.; Taylor, J.; Chang, S.; et al. The genetic landscape of gliomas arising after therapeutic radiation. Acta Neuropathol. 2019, 137, 139–150. [Google Scholar] [CrossRef] [Green Version]
  102. Schäfer, N.; Gielen, G.H.; Rauschenbach, L.; Kebir, S.; Till, A.; Reinartz, R.; Simon, M.; Niehusmann, P.; Kleinschnitz, C.; Herrlinger, U.; et al. Longitudinal heterogeneity in glioblastoma: Moving targets in recurrent versus primary tumors. J. Transl. Med. 2019, 17, 96. [Google Scholar] [CrossRef] [Green Version]
  103. van Linde, M.E.; Brahm, C.G.; de Witt Hamer, P.C.; Reijneveld, J.C.; Bruynzeel, A.M.E.; Vandertop, W.P.; van de Ven, P.M.; Wagemakers, M.; van der Weide, H.L.; Enting, R.H.; et al. Treatment outcome of patients with recurrent glioblastoma multiforme: A retrospective multicenter analysis. J. Neurooncol. 2017, 135, 183–192. [Google Scholar] [CrossRef] [Green Version]
  104. Carson, K.A.; Grossman, S.A.; Fisher, J.D.; Shaw, E.G. Prognostic factors for survival in adult patients with recurrent glioma enrolled onto the new approaches to brain tumor therapy CNS consortium phase I and II clinical trials. J. Clin. Oncol. 2007, 25, 2601–2606. [Google Scholar] [CrossRef]
  105. Mineo, J.-F.; Bordron, A.; Baroncini, M.; Ramirez, C.; Maurage, C.-A.; Blond, S.; Dam-Hieu, P. Prognosis factors of survival time in patients with glioblastoma multiforme: A multivariate analysis of 340 patients. Acta Neurochir. 2007, 149, 243–245. [Google Scholar] [CrossRef]
  106. Aldaz, P.; Auzmendi-iriarte, J.; Durántez, M.; Lasheras-otero, I.; Carrasco-garcia, E.; Victoria Zelaya, M.; Bragado, L.; Olías-arjona, A.; Egaña, L.; Samprón, N.; et al. Identification of a dexamethasone mediated radioprotection mechanism reveals new therapeutic vulnerabilities in glioblastoma. Cancers 2021, 13, 361. [Google Scholar] [CrossRef]
  107. Martinho, O.; Longatto-Filho, A.; Lambros, M.B.K.; Martins, A.; Pinheiro, C.; Silva, A.; Pardal, F.; Amorim, J.; Mackay, A.; Milanezi, F.; et al. Expression, mutation and copy number analysis of platelet-derived growth factor receptor A (PDGFRA) and its ligand PDGFA in gliomas. Br. J. Cancer 2009, 101, 973–982. [Google Scholar] [CrossRef] [Green Version]
  108. Wallmann, T.; Zhang, X.-M.; Wallerius, M.; Bolin, S.; Joly, A.-L.; Sobocki, C.; Leiss, L.; Jiang, Y.; Bergh, J.; Holland, E.C.; et al. Microglia Induce PDGFRB Expression in Glioma Cells to Enhance Their Migratory Capacity. iScience 2018, 9, 71–83. [Google Scholar] [CrossRef] [Green Version]
  109. Schaff, L.R.; Yan, D.; Thyparambil, S.; Tian, Y.; Cecchi, F.; Rosenblum, M.; Reiner, A.S.; Panageas, K.S.; Hembrough, T.; Lin, A.L. Characterization of MGMT and EGFR protein expression in glioblastoma and association with survival. J. Neurooncol. 2020, 146, 163–170. [Google Scholar] [CrossRef]
  110. Szerlip, N.J.; Pedraza, A.; Chakravarty, D.; Azim, M.; McGuire, J.; Fang, Y.; Ozawa, T.; Holland, E.C.; Huse, J.T.; Jhanwar, S.; et al. Intratumoral heterogeneity of receptor tyrosine kinases EGFR and PDGFRA amplification in glioblastoma defines subpopulations with distinct growth factor response. Proc. Natl. Acad. Sci. USA 2012, 109, 3041–3046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Fanelli, G.N.; Grassini, D.; Ortenzi, V.; Pasqualetti, F.; Montemurro, N.; Perrini, P.; Naccarato, A.G.; Scatena, C. Decipher the Glioblastoma Microenvironment: The First Milestone for New Groundbreaking Therapeutic Strategies. Genes 2021, 12, 445. [Google Scholar] [CrossRef]
  112. Broekman, M.L.; Maas, S.L.N.; Abels, E.R.; Mempel, T.R.; Krichevsky, A.M.; Breakefield, X.O. Multidimensional communication in the microenvirons of glioblastoma. Nat. Rev. Neurol. 2018, 14, 482–495. [Google Scholar] [CrossRef] [PubMed]
  113. Wellbrock, C.; Arozarena, I. The Complexity of the ERK/MAP-Kinase Pathway and the Treatment of Melanoma Skin Cancer. Front. Cell Dev. Biol. 2016, 4, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Coniglio, S.J.; Eugenin, E.; Dobrenis, K.; Stanley, E.R.; West, B.L.; Symons, M.H.; Segall, J.E. Microglial Stimulation of Glioblastoma Invasion Involves Epidermal Growth Factor Receptor (EGFR) and Colony Stimulating Factor 1 Receptor (CSF-1R) Signaling. Mol. Med. 2012, 18, 519–527. [Google Scholar] [CrossRef]
  115. Calabrese, C.; Poppleton, H.; Kocak, M.; Hogg, T.L.; Fuller, C.; Hamner, B.; Oh, E.Y.; Gaber, M.W.; Finklestein, D.; Allen, M.; et al. A perivascular niche for brain tumor stem cells. Cancer Cell 2007, 11, 69–82. [Google Scholar] [CrossRef] [Green Version]
  116. Zhao, Y.; Wang, H.; He, C. Drug resistance of targeted therapy for advanced non-small cell lung cancer harbored EGFR mutation: From mechanism analysis to clinical strategy. J. Cancer Res. Clin. Oncol. 2021, 147, 3653–3664. [Google Scholar] [CrossRef]
  117. Pan, P.C.; Magge, R.S. Mechanisms of EGFR Resistance in Glioblastoma. Int. J. Mol. Sci. 2020, 21, 8471. [Google Scholar] [CrossRef]
  118. Francis, J.M.; Zhang, C.-Z.; Maire, C.L.; Jung, J.; Manzo, V.E.; Adalsteinsson, V.A.; Homer, H.; Haidar, S.; Blumenstiel, B.; Pedamallu, C.S.; et al. EGFR variant heterogeneity in glioblastoma resolved through single-nucleus sequencing. Cancer Discov. 2014, 4, 956–971. [Google Scholar] [CrossRef] [Green Version]
  119. Nathanson, D.A.; Gini, B.; Mottahedeh, J.; Visnyei, K.; Koga, T.; Gomez, G.; Eskin, A.; Hwang, K.; Wang, J.; Masui, K.; et al. Targeted therapy resistance mediated by dynamic regulation of extrachromosomal mutant EGFR DNA. Science 2014, 343, 72–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Akhavan, D.; Pourzia, A.L.; Nourian, A.A.; Williams, K.J.; Nathanson, D.; Babic, I.; Villa, G.R.; Tanaka, K.; Nael, A.; Yang, H.; et al. De-repression of PDGFRβ transcription promotes acquired resistance to EGFR tyrosine kinase inhibitors in glioblastoma patients. Cancer Discov. 2013, 3, 534–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Mellinghoff, I.K.; Wang, M.Y.; Vivanco, I.; Haas-Kogan, D.A.; Zhu, S.; Dia, E.Q.; Lu, K.V.; Yoshimoto, K.; Huang, J.H.Y.; Chute, D.J.; et al. Molecular determinants of the response of glioblastomas to EGFR kinase inhibitors. N. Engl. J. Med. 2005, 353, 2012–2024. [Google Scholar] [CrossRef] [Green Version]
  122. Saleem, H.; Kulsoom Abdul, U.; Küçükosmanoglu, A.; Houweling, M.; Cornelissen, F.M.G.; Heiland, D.H.; Hegi, M.E.; Kouwenhoven, M.C.M.; Bailey, D.; Würdinger, T.; et al. The TICking clock of EGFR therapy resistance in glioblastoma: Target Independence or target Compensation. Drug Resist. Updates 2019, 43, 29–37. [Google Scholar] [CrossRef] [PubMed]
  123. Ma, Y.; Tang, N.; Thompson, R.C.; Mobley, B.C.; Clark, S.W.; Sarkaria, J.N.; Wang, J. InsR/IGF1R Pathway Mediates Resistance to EGFR Inhibitors in Glioblastoma. Clin. Cancer Res. 2016, 22, 1767–1776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Song, K.; Yuan, Y.; Lin, Y.; Wang, Y.-X.; Zhou, J.; Gai, Q.-J.; Zhang, L.; Mao, M.; Yao, X.-X.; Qin, Y.; et al. ERBB3, IGF1R, and TGFBR2 expression correlate with PDGFR expression in glioblastoma and participate in PDGFR inhibitor resistance of glioblastoma cells. Am. J. Cancer Res. 2018, 8, 792–809. [Google Scholar]
  125. Almiron Bonnin, D.A.; Ran, C.; Havrda, M.C.; Liu, H.; Hitoshi, Y.; Zhang, Z.; Cheng, C.; Ung, M.; Israel, M.A. Insulin-Mediated Signaling Facilitates Resistance to PDGFR Inhibition in Proneural hPDGFB-Driven Gliomas. Mol. Cancer Ther. 2017, 16, 705–716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Omuro, A.; DeAngelis, L.M. Glioblastoma and Other Malignant Gliomas: A Clinical Review. JAMA 2013, 310, 1842–1850. [Google Scholar] [CrossRef] [PubMed]
  127. Schloss, M.H.; Freidberg, S.R.; Heatley, G.J.; Lo, T.C.M. Glucocorticoid Dependency as A Prognostic Factor in Radiotherapy for Cerebral Gliomas. Acta Oncol. 1989, 28, 51–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Shields, L.B.; Shelton, B.J.; Shearer, A.J.; Chen, L.; Sun, D.A.; Parsons, S.; David Bourne, T.; LaRocca, R.; Spalding, A.C. Dexamethasone administration during definitive radiation and temozolomide renders a poor prognosis in a retrospective analysis of newly diagnosed glioblastoma patients. Radiat. Oncol. 2015, 4–11. [Google Scholar] [CrossRef] [Green Version]
  129. Watne, K.; Hannisdal, E.; Nome, O.; Hager, B.; Hirschberg, H. Prognostic Factors in Malignant Gliomas with Special Reference to Intra-Arterial Chemotherapy. Acta Oncol. 1993, 32, 307–310. [Google Scholar] [CrossRef] [PubMed]
Table 1. Receptor Tyrosine Kinase inhibitors targeting drivers of GBM trialed in glioblastoma patients.
Table 1. Receptor Tyrosine Kinase inhibitors targeting drivers of GBM trialed in glioblastoma patients.
DrugTrade NameHuman TargetsFirst FDA ApprovalClinical Use in Cancer
IMATINIBGLEEVECKIT, ABL1, PDGFRB2001CMLCML, GIST
GEFITINIBIRESSAEGFR2003NSCLCNSCLC
ERLOTINIBTARCEVAEGFR2004NSCLCNSCLC, PCa
SORAFENIBNEXAVARBRAF, PDGFRB, FLT1, FLT4, KDR, FLT3, RAF1, RET, KIT2005RCCRCC
SUNITINIBSUTENTPDGFRB, PDGFRA, KIT, FLT3, CSF1R, FLT1, FLT4, KDR2006RCCRCC
DASATINIBSPRYCELFYN, SRC, LCK, YES1, BLK, HCK, LYN, FRK, FGR, SRMS, EPHA2, PDGFRB, ABL1, KIT2006CMLCML
LAPATINIBTYVERBEGFR, ERBB22007BCaBCa
PAZOPANIBVOTRIENTPDGFRB, PDGFRA, FGFR3, FGFR1, FLT1, FLT4, KDR, ITK, CSF1R, KIT, LCK2009RCCRCC, SARCOMA
VANDETANIBCAPRELSA, ZACTIMAERBB2,3,4, EGFR, FLT1, FLT4, KDR, PTK6, EPHA (1–7,10) PHA8, EPHB (1–4,6) RET, SRC,TEK2011TCTC
REGORAFENIBSTIVARGATEK, RAF1, FGFR1, DDR2, BRAF, RET, MAPK11, FLT1,4, KDR, FGFR2, FRK, PDGFRB, PDGFRA, ABL1, KIT2012CRC, GISTCRC, GIST, HCC
CABOZANTINIBCABOMETYX, COMETRIQKDR, MET2012TCTC
AXITINIBINLYTAFLT1, FLT4, KDR2012RCCRCC, SARCOMA
PONATINIBICLUSIGBCR, ABL1, PDGFR, FGFR, EPHR, KIT, SRC RET, FLT32012CMLCML, Ph+ALL
AFATINIBGIOTRIFERBB2, ERBB4, EGFR2013NSCLCNSCLC
NINTEDANIBOFEV, VARGATEFFLT1, FLT4, KDR, PDGFRB, PDGFRA, FGFR3, FGFR1, FGFR4, FGFR22014PFNSCLC
OSIMERTINIBTAGRISSOEGFR2015NSCLCNSCLC
NERATINIB EGFR, ERBB4, ERBB22017BCaBCa
DACOMITINIB ERBB2, EGFR, ERBB42018NSCLCNSCLC
INFIGRATINIB FGFR3, FGFR1, FGFR4, FGFR22021CCACCA
NSCLC: non-small cell lung cancer; TC: thyroid cancer, CRC: Colorectal Cancer; GIST: gastrointestinal stromal tumor; HCC: hepatocellular carcinoma; RCC: renal cell carcinoma; CML: chronic myeloid leukemia; Ph+ALL: Philadelphia chromosome-positive acute lymphoblastic leukemia; PF: pulmonary fibrosis BCa: breast cancer; CCA: metastatic cholangiocarcinoma.
Table 2. Clinical trials of Receptor Tyrosine Kinase Inhibitors in Newly diagnosed Glioblastoma.
Table 2. Clinical trials of Receptor Tyrosine Kinase Inhibitors in Newly diagnosed Glioblastoma.
DrugNTC NumberPhaseStatusStart Date
GEFITINIBNCT002387972Completed1 February 2003
VATALANIBNCT003858531Completed1 September 2006
SORAFENIBNCT005448172Completed1 April 2007
DASATINIBNCT008959602Withdrawn1 May 2009
SUNITINIBNCT011001772Completed1 June 2009
TANDUTINIBNCT009048521Withdrawn1 June 2009
AFATINIBNCT009774311Completed17 September 2009
AXITINIBNCT015081172Terminated1 August 2011
SUNITINIBNCT029285752Unknown status1 August 2012
LAPATINIBNCT015915772Active, not recruiting7 December 2012
PAZOPANIBNCT023314981Active, not recruiting1 June 2015
NERATINIBNCT029777802Recruiting9 February 2017
EPITINIBNCT032315011Recruiting26 January 2018
ANLOTINIBNCT041196741Recruiting15 January 2019
REGORAFENIBNCT039704472Recruiting30 July 2019
ANLOTINIBNCT041574782Not yet recruiting1 January 2020
ANLOTINIBNCT047252142Recruiting15 January 2021
Table 3. Clinical trials of receptor tyrosine kinase inhibitors in recurrent or progressive glioblastoma.
Table 3. Clinical trials of receptor tyrosine kinase inhibitors in recurrent or progressive glioblastoma.
DrugNTC NumberPhaseStatusStart date
ERLOTINIBNCT003378832Completed1 July 2003
IMATINIBNCT001719382Terminated1 April 2004
IMATINIBNCT001543753Completed1 October 2004
GEFITINIBNCT002508872Completed1 July 2005
IMATINIBNCT002907712Terminated1 February 2006
ERLOTINIBNCT003014181Completed1 March 2006
SUNITINIBNCT006060082Completed1 March 2007
SUNITINIBNCT008648640Completed1 May 2007
CEDIRANIBNCT005032041Completed1 September 2007
SORAFENIBNCT005974932Completed1 September 2007
SUNITINIBNCT005353792Unknown status1 October 2007
CABOZANTINIBNCT007042882Completed1 May 2008
SUNITINIBNCT009231172Terminated1 June 2008
CEDIRANIBNCT007771533Completed1 October 2008
VANDETANIBNCT008210801Completed1 October 2008
DASATINIBNCT008921772Completed1 October 2009
DASATINIBNCT009483891Terminated1 October 2009
DACOMITINIBNCT011125272Completed1 April 2010
NINTEDANIBNCT012514842Completed1 January 2011
CEDIRANIBNCT013108552Terminated1 May 2011
GEFITINIBNCT013108552Terminated1 May 2011
ERLOTINIBNCT011108761Terminated1 June 2011
DACOMITINIBNCT015208702Completed1 February 2012
NINTEDANIBNCT016666001Terminated1 August 2012
SORAFENIBNCT018177512Active, not recruiting11 April 2013
DOVITINIBNCT019727501Unknown status1 October 2013
INFIGRATINIBNCT019757012Completed9 December 2013
REGORAFENIBNCT029262222Active, not recruiting1 November 2015
CRENOLANIBNCT026263642Completed1 April 2016
SORAFENIBNCT014346021Active, not recruiting11 April 2016
TESEVATINIBNCT028444392Completed1 June 2016
AXITINIBNCT032913142Completed3 May 2017
CEDIRANIBNCT029746212Active, not recruiting15 September 2017
SUNITINIBNCT030258932Recruiting31 August 2018
OSIMERTINIBNCT037323522Active, not recruiting28 November 2018
REGORAFENIBNCT039704472Recruiting30 July 2019
REGORAFENIBNCT040516062Recruiting31 July 2019
INFIGRATINIBNCT044249660Recruiting21 July 2020
ANLOTINIBNCT045478552Recruiting11 September 2020
Table 4. Inclusion criteria (%) in GBM TKIs clinical trials.
Table 4. Inclusion criteria (%) in GBM TKIs clinical trials.
Clinical TrialsInclusion CriteriaNewly DiagnosedRecurrentTotal
AllOthers30.91%69.09%100%
TKR expression or mutation5.88%18.42%14.54%
Currently ongoingOthers47.06%52.94%100%
TKR expression or mutation12.50%33.33%23.53%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aldaz, P.; Arozarena, I. Tyrosine Kinase Inhibitors in Adult Glioblastoma: An (Un)Closed Chapter? Cancers 2021, 13, 5799. https://doi.org/10.3390/cancers13225799

AMA Style

Aldaz P, Arozarena I. Tyrosine Kinase Inhibitors in Adult Glioblastoma: An (Un)Closed Chapter? Cancers. 2021; 13(22):5799. https://doi.org/10.3390/cancers13225799

Chicago/Turabian Style

Aldaz, Paula, and Imanol Arozarena. 2021. "Tyrosine Kinase Inhibitors in Adult Glioblastoma: An (Un)Closed Chapter?" Cancers 13, no. 22: 5799. https://doi.org/10.3390/cancers13225799

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