Next Article in Journal
Enhanced Preprocessing Approach Using Ensemble Machine Learning Algorithms for Detecting Liver Disease
Next Article in Special Issue
An Artificial Placenta Experimental System in Sheep: Critical Issues for Successful Transition and Survival up to One Week
Previous Article in Journal
N-Glycoprofiling of SLC35A2-CDG: Patient with a Novel Hemizygous Variant
Previous Article in Special Issue
Autopsies Revealed Pathological Features of COVID-19 in Unvaccinated vs. Vaccinated Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antitumor Potential of Antiepileptic Drugs in Human Glioblastoma: Pharmacological Targets and Clinical Benefits

by
Manuela Stella
1,2,†,
Giammarco Baiardi
1,2,†,
Stefano Pasquariello
1,2,
Fabio Sacco
1,2,
Irene Dellacasagrande
1,
Alessandro Corsaro
1,
Francesca Mattioli
1,2,‡ and
Federica Barbieri
1,*,‡
1
Pharmacology and Toxicology Unit, Department of Internal Medicine, University of Genova, 16132 Genova, Italy
2
Clinical Pharmacology Unit, EO Ospedali Galliera, 16128 Genova, Italy
*
Author to whom correspondence should be addressed.
Co-first authors.
Co-last authors.
Submission received: 13 January 2023 / Revised: 7 February 2023 / Accepted: 10 February 2023 / Published: 16 February 2023
(This article belongs to the Special Issue 10th Anniversary of Biomedicines—Novel Targets for Cranial Tumors)

Abstract

:
Glioblastoma (GBM) is characterized by fast-growing cells, genetic and phenotypic heterogeneity, and radio-chemo-therapy resistance, contributing to its dismal prognosis. Various medical comorbidities are associated with the natural history of GBM. The most disabling and greatly affecting patients’ quality of life are neurodegeneration, cognitive impairment, and GBM-related epilepsy (GRE). Hallmarks of GBM include molecular intrinsic mediators and pathways, but emerging evidence supports the key role of non-malignant cells within the tumor microenvironment in GBM aggressive behavior. In this context, hyper-excitability of neurons, mediated by glutamatergic and GABAergic imbalance, contributing to GBM growth strengthens the cancer-nervous system crosstalk. Pathogenic mechanisms, clinical features, and pharmacological management of GRE with antiepileptic drugs (AEDs) and their interactions are poorly explored, yet it is a potentially promising field of research in cancer neuroscience. The present review summarizes emerging cooperative mechanisms in oncogenesis and epileptogenesis, focusing on the neuron-to-glioma interface. The main effects and efficacy of selected AEDs used in the management of GRE are discussed in this paper, as well as their potential beneficial activity as antitumor treatment. Overall, although still many unclear processes overlapping in GBM growth and seizure onset need to be elucidated, this review focuses on the intriguing targeting of GBM-neuron mutual interactions to improve the outcome of the so challenging to treat GBM.

1. Introduction

Glioblastoma (GBM) is the most common primary malignant CNS tumor (48%) [1]. In the United States, the average annual age-adjusted incidence of GBM is 3.2/100,000 [1], while in Europe, it ranges from 3.3/100,000 in France [2] to 4.17/100,000 in Spain [3]. GBM occurs mainly in older adults (>65 years), and the male:female ratio is 1.6:1 [4,5], and the global incidence of GBM increases with age [6]. GBM is an astrocytic tumor characterized by high cellular heterogeneity, invasiveness, and microvascular proliferation. All these characteristics concur in patients’ short median survival and frequent recurrence, also favored by an immunosuppressive tumor-induced microenvironment fostering the self-renewal of highly infiltrative and drug-resistant GBM stem cells (GSCs) in the peritumoral region [7].
In 2016, the World Health Organization (WHO) newly classified CNS tumors by increasing the importance of their molecular profile in addition to histopathologic features; GBM is defined as a WHO grade IV glioma [8]. The 2021 WHO classification further expanded the relevance of molecular diagnostics, adding isocitrate dehydrogenase (IDH) mutational status and 1p/19q co-deletion to better define glioma subtypes; currently, a diffuse astrocytic glioma with microvascular proliferation, necrosis, and/or specific gene alterations (i.e., TERT mutation, epidermal growth factor receptor (EGFR) amplification, and chromosome rearrangement, IDHwt without 1p/19q co-deletion) is diagnosed as GBM WHO grade 4 [9].
Despite continuous advances in our understanding of GBM pathogenesis, from self-renewal of GSCs [10] to pro-tumorigenic microenvironment [11] or epigenetic phenomena [12] driving tumor progression, the prognosis of affected patients remains poor. Since 2005, standard of care (SOC) therapy consists of surgical resection, concurrent radiotherapy (RT, total 60 Gy) with the alkylating agent temozolomide (TMZ, 75 mg/m2/day for 6 wk), and further adjuvant TMZ (150–200 mg/m2/day for 5 days for six 28-day cycles) [13], ensuring an overall survival (OS) of approximately 15 months in newly diagnosed GBM [14]. Tumor-treating fields can also be given concurrently with adjuvant TMZ [15].
Methylation of the O6-methylguanine-DNA methyltransferase (MGMT) promoter is the strongest predictor for benefit from TMZ; thus, this drug could be withheld in selected patients with unmethylated MGMT tumors [13]. Neither the extension of TMZ treatment nor the addition of bevacizumab, a humanized anti-VEGF monoclonal antibody, yielded further survival benefits for newly diagnosed GBM [16,17].
Recurrence is nearly inevitable after a median interval of ~7 months [18], and agreement on SOC is lacking [19]. Since the efficacy of current treatment for the recurrent disease remains poor, the National Comprehensive Cancer Network (NCCN) recommends enrollment in clinical trials as the preferred therapeutic approach [20]. Surgery may have a role for symptomatic lesions, while other treatment options include re-irradiation, TMZ rechallenge, PCV (procarbazine, CCNU [lomustine], vincristine), bevacizumab, nitrosourea, or target therapies [20]. TMZ rechallenge may be a reasonable option for MGMT methylated GBMs [21,22], while bevacizumab, often effective in reducing clinical symptoms related to peritumoral edema, failed to demonstrate survival benefit [23]. Despite progress made in developing targeted therapies [24], none has been shown to prolong survival in randomized clinical trials [14,19]. Immunotherapies (vaccination, immune checkpoint blockade, oncolytic viral therapies, chimeric antigen receptor T-cell therapies) are a growing perspective, despite the fact that their potential efficacy needs to be demonstrated [25].

2. Glioblastoma-Related Epilepsy (GRE)

Glioma patients’ quality of life largely depends on the neurological decline due to the tumor itself or treatment-related toxicity and epilepsy (for definition, see [26]), which negatively affects neurocognitive functions and outcome [27].
In GBM patients, the diagnosis of epilepsy is usually made after one seizure episode. An inverse relationship between the degree of malignancy and the frequency of seizures has been reported: patients with low-grade glioma are affected by seizures in 65–95% of cases, while they occur in 30–50% of GBM patients [28]. GBM often originates in less epileptogenic areas, and patients may not live long enough to develop epilepsy, thus explaining the relative lower incidence of epilepsy in these tumors [29,30]. Multiple factors concur to GRE onset, such as increased intracranial pressure and edema, alterations of the peritumoral cortex, inflammation, and vascular insufficiency, likely triggering excitotoxicity and hyperexcitability of peritumoral neurons [31]. Underlying molecular mechanisms involve disruption of inhibitory/excitatory transmission balance mainly due to alterations of GABAergic and glutamatergic regulation [31], whose functions connecting epileptogenicity and GBM progression are detailed in the following paragraphs. The involvement of glioma-derived thrombospondin, a regulator of synaptogenesis, in excitatory synapse formation within the peritumoral cortex of a glioma-cell-implanted rat model has been described [32]. Furthermore, the association between connexin 43, a multifunction protein that forms gap junction channels, and GRE has been suggested and discussed [33,34].
Based on the relevance of alterations in GBM surrounding tissue in GRE, the pathophysiology of epilepsy may have roots in the tumor microenvironment; indeed, in the peritumoral area from a rat model of glioma, the density of GABAergic neurons was significantly decreased, and transcriptomic analysis revealed that 5 of 19 genes differentially expressed were associated with epilepsy and neurodevelopmental disorders [35].
The presence of GRE is considered one of the major risk factors for long-term disability since antiepileptic drugs often have heavy side effects and interactions with chemotherapy and supportive therapy; therefore, further investigation on neurons within the GBM microenvironment will identify targetable mechanisms in a translational perspective.

3. Antiepileptic Drugs in GRE: Clinical Management

Seizure control in GBM patients can be hardly accomplished by surgery alone; however, the extent of resection may determine better seizure outcome [36], while a beneficial effect of radio-chemo-therapy has not yet been defined: it seems effective for low-grade glioma [37] but not in GBM elderly patients [38]. Therefore, the best treatment strategy of GRE with AEDs can guarantee epilepsy control with a great impact on the patient’s quality of life. Management of GRE with a specific AED is challenging and controversial. Considering the risk of AED adverse events, physical disability, and neurocognitive impairments related to the tumor site, a multidisciplinary approach is needed.
To date, no universal guidelines are available for GRE therapy, and the antiepileptic efficacy of drugs in non-tumor epilepsy is not directly translatable to GBM patients due to drug–drug interactions and neurological conditions. In GRE, no superior efficacy of one AED over another has been demonstrated [39,40], and we are waiting for results from ongoing trials (NCT03048084, NCT03636958, NCT04497142). Current recommendations on AEDs use in newly diagnosed GBM suggest not treating patients who have never experienced a seizure episode, while it should be prescribed for patients who had at least one comitial episode to prevent the high risk of recurrent seizures [40,41]. Nevertheless, this issue is still discussed [42], and currently, refractory seizures may be treated with the combination of AEDs with a different mechanism of action [39].
The non-enzyme inducing AEDs, levetiracetam (LEV), lamotrigine (LMG), topiramate (TPM), lacosamide (LCM), pregabalin (PRG), and valproic acid (VPA) are preferred as monotherapy, as they have fewer adverse effects and interactions with other concomitant therapies in cancer patients. VPA acts as an inhibitor of glucuronidation; TPM is a weak inducer of CYP3A4 and might also inhibit CYP2C19. Perampanel (PER) and brivaracetam (BRV) are potentially broad-spectrum antiepileptic drugs that may be useful in add-on and do not appear to act as potent inducers of cytochrome P450 isoenzymes [43]. However, VPA, TPM, and PER can induce enzyme inhibition, increasing the toxicity of antitumor substrates. [44,45].
The use of enzyme-inducing AEDs (phenytoin, PHT, phenobarbital, and carbamazepine) is generally discouraged as they are potent inducers of several CytP450 isoenzymes, leading to clinically relevant alteration of antitumor agent pharmacokinetics (e.g., lomustine, vincristine), possibly increasing their side effects and reducing efficacy [40]. When a drug interaction is suspected, AEDs serum concentrations should be monitored [46]. However, taking into account that the metabolism of TMZ is not affected by old AEDs, the main reason for their limited use is the higher safety, tolerability, and effectiveness of newer AEDs [47].
A systematic review demonstrates that monotherapy with LEV, PHT, and PRG had higher efficacy in GRE, with LEV showing a lower failure rate [47], although variability among studies and patient populations prevents a definite statement. LEV shows a more favorable efficacy profile compared to VPA, with similar toxicity [47,48].
Drug resistance to first-line monotherapy in patients with GRE occurs in approximately 30% of patients and requires an add-on. However, no evidence of the superiority of a drug over another in resistant patients has been reported [40]. Among add-on treatments, PER efficacy and safety in GRE have been described [49,50], as well as for BRV [51], an analog of LEV, currently under evaluation (NCT05029960), although larger prospective studies are lacking.
Given the emerging interest in cannabinoids (i.e., cannabidiol, CBD) as promising antiepileptic agents in patients with refractory seizures [52], CBD-enriched products might help control seizures also in glioma patients. CBD showed good safety and tolerability [53], while clinical efficacy in GRE needs to be tested.
Therapeutic criteria for GRE are still based on studies on general epilepsy; thus, a satisfactory GRE therapy is still challenging. In this scenario, further insights into specific pathophysiological mechanisms underpinning GRE and drug resistance might improve AED choice and patient outcome.

4. Molecules at the Crossroads of GBM Development and Epilepsy

Many mechanisms cooperate and overlap in tumorigenesis and epileptogenesis, which mainly impact GRE pathophysiology. Indeed, GRE shows a definite clinical profile but implies several causes related to both the patient’s (age, genetic factors, therapy, physical status) and tumor (subtype, location, infiltration in the brain parenchyma, edema, microenvironment, neoangiogenesis, inflammation, infiltrating microglia) features.
At the molecular level, the epileptogenic-related process includes IDH mutation status, mTOR signaling pathway, and neurotransmitter (GABA, glutamate) dysregulation—all shared relevant mechanisms in GBM development.

4.1. Intrinsic Molecular Features of GBM Progression and Their Role in GRE

IDH enzymes participate in the metabolic process and cellular homeostasis catalyzing the oxidative decarboxylation of isocitrate. Mutations in IDHs are frequent in a variety of human malignancies, including gliomas, which correlate with seizure risk [54]. Preoperative epilepsy is more frequent in IDHwt low-grade glioma patients than in IDHwt GBM [55].
IDH mutation likely leads to the formation of D-2HG (D-2-hydroxyglutarate), an oncometabolite structurally similar to glutamate, able to enhance the activity of tumor-adjacent cortical neurons interacting with NMDA receptors and, therefore, boosting seizures [56]. In IDHmut tumors, high D-2HG concentration leads to metabolic disruptions in surrounding cortical neurons that consequently promote seizures, increasing preoperative seizure risk. Further evidence suggests that D-2HG supports a metabolic change (hypermetabolic phenotype) in GBM surrounding neurons, characterized by lactic dehydrogenase-A increase, defects in the tricarboxylic acid cycle [56,57], as well as upregulation of mTOR (mechanistic target of rapamycin) activity [56,58], already known as pro-epileptogenic stimuli. Indeed, mTOR signaling is responsible for cell growth and survival, and alterations are related to multiple brain disorders, including neoplasms and epilepsy [59]. Interestingly, in GBM, the PI3K/AKT/mTOR pathway is almost uniformly activated due to activating mutations in PIK3CA or PIK3R1 or loss of phosphatase and TENsin homolog (PTEN) [60,61]. Dysfunction of the mTOR pathway also modifies GABA and glutamate signaling, two main actors in epileptogenesis [62].
Furthermore, mTOR, the most intrinsic properties of GBM, depicting its specific molecular profile, directly influences the development of epilepsy. Mutations in tumor suppressor genes, such as tumor protein 53 (TP53), PTEN, and neurofibromatosis type 1 (NF1), whose downstream altered pathways drive GBM development, are also closely related to peritumoral hyperexcitability, as observed in epileptogenic GBMs developed in a PTEN, NF1, and p53 CRISPR-based animal model [63,64,65]. Frequent TP53 mutations in GBM lead to upregulation of the protein, which in turn triggers MET and EGFR activity, enhancing a tumor invasive behavior and matching with TP53 overexpression associated with preoperative seizures [66]. Similarly, PTEN mutations causing dysregulation of the mTOR pathway result in epileptic disorders [67,68].
Pro-epileptogenic effects of NF1 mutations are evident in patients bearing this multisystem disorder, commonly leading to brain tumor predisposition and neurological manifestations, in which seizures occur in 4–7% of patients [69]. In GBM, neurofibromin, the NF1 gene product, negatively regulates mTOR signaling through downregulation of the Ras/MAPK pathway, hyperactivation of mTOR, and increased cell proliferation, favoring tumor progression [70]. Notably, alterations of the RTK/RAS/PI3K pathway, found in 90% of GBM, are also driven by mutations in the PI3K catalytic subunit PIK3CA [60], as described in patient-derived xenograft mice models harboring mutant PIK3CA, in which gliomagenesis and neuronal hyperactivity showed a reciprocal influence, likely through a secreted factor (e.g., glypican 3, GPC3), selectively expressed in these tumors [71]. A complementary activity of secreted molecules involved in neuronal activity, such as neuroligin-3 or glutamate, has been associated with GBM growth [72].
Overall, the intrinsic molecular features of GBM share most of their activities with epilepsy, affecting both tumor and neuronal behavior.

4.2. Neurotransmitter Synaptic Inputs to GBM Progression and GRE

Cancer cells may co-opt neuronal signaling pathway, and reciprocally, neuronal activity may drive GBM malignant behavior by either autocrine or paracrine (non-synaptic) mechanisms. GRE is often linked to unbalanced inhibitory and excitatory synaptic signals leading to excessive excitability in GBM surrounding tissues [73]. Therefore, the involvement of GABAergic and glutamatergic systems, which maintain a finely-tuned excitation-inhibition balance in normal tissue, deserves particular interest in oncogenesis and epileptogenesis (Figure 1).

4.3. GABAergic Signaling

The inhibitory GABA transmission is mediated by ligand-gated chloride-permeable channel GABAA receptors (GABAARs), and the response type (depolarization vs. hyperpolarization) is mainly driven by intracellular Cl concentrations in post-synaptic neurons, regulated by a neuronal-specific solute carrier (SLC) family protein 12 (SLC12A), the member of the K+-Cl cotransporter 2, KCC2 [74] (Figure 1).
Surrounding neocortical cells in epilepsy-associated lesions show downregulation of GABAARs and reduced synapses, likely leading to unbalance between excitatory and inhibitory transmission, which, in turn, may contribute to the pathogenesis of focal seizures [75]. As in other human focal epilepsies, pathologic changes in Cl homeostasis can switch GABAergic signaling from hyperpolarizing to depolarizing. Excitatory effects of GABA in the human peritumoral neocortex contribute to epileptogenic glioma [76].
Epileptogenicity of GABAergic neurotransmission in GBM seems to be linked to dysregulation of two cation-chloride cotransporters: (i) downregulation of KCC2, which is usually expressed at high levels in cerebral tissue and extrude Cl; and (ii) upregulation of Na–K–2Cl cotransporter 1 (NKCC1), which is normally expressed at low levels actively uptakes Cl into cells [76,77]. This process leads to an increase in neuronal intracellular Cl concentration, which, contrarily to what is observed in a normal brain (hyperpolarization), shifts towards a depolarizing GABAergic excitatory response [62,78], contributing to GRE (Figure 1). KCC2 is downregulated in the peritumoral region while NKCC1 is highly expressed [79], and gliomas show perturbation of Cl homeostasis since GBM cells accumulate Cl through the activity of NKCC1 transporter, supporting their proliferative and migratory ability [80] (Figure 1).

4.4. Glutamatergic Signaling

In addition to GABA signal disruption, glutamatergic mechanisms, influencing intracellular Cl concentration, and imbalance of the excitatory neurotransmitter glutamate in neurons, contribute to cell excitability, the pathophysiology of seizures, and epileptogenicity in the tumor-surrounding tissue (Figure 1). Ionotropic glutamate AMPA and NMDA receptors are predominantly involved in seizure onset [81]. Interestingly, in the glioma microenvironment, in addition to neuron-released glutamate, GBM cells, contrarily to normal astrocytes which sequester the neurotransmitter, release glutamate, likely exerting a pro-epileptic activity in the peritumoral brain (Figure 2). Insights into glutamatergic interactions in glioma are increasing: GBM cells express high levels of SLC7A11solute carrier family 7 member 11or xCT, a subunit of xc-cystine-glutamate transporter system, which acts as a cystine/glutamate antiporter across the plasma membrane. Increased exchange of extracellular cystine and intracellular glutamate induces glutathione synthesis, concurring either endogenous antioxidant activity sustaining GBM cell survival or glutamate release (Figure 2). SLC7A11 upregulation correlates with tumor invasion and outcome in GBM patients as well as with the onset of seizures, as described in animal models of human GBM xenografts implanted into the mouse brain [82,83], although the seizures in these experimental animals could not be completely representative of rhythmic and periodic patterns of seizure in humans. However, Buckingham et al. [83] describe a novel GBM-related pathophysiological mechanism of seizure onset, in which glutamate released by the GBM cells through SLC7A11 causes hyper-excitability of surrounding neurons. Surgical removal of GBM mass might improve GRE by reducing glutamate-releasing GBM cells.
GBMs also show the downregulation of another transporter of the SLC family, the excitatory amino acid transporter 2 (EAAT2), which normally rapidly retrieves glutamate to protect neurons from excitotoxicity, limiting glutamate re-uptake from the extracellular space, thus favoring seizure onset [84]. In addition, the upregulation of branched-chain amino acid transaminase 1 (BCAT1), which forms glutamate by transferring an α-amino group BCAAs to α-ketoglutarate, increases glutamate amount and release in IDHwt GBMs [85].
GBM cells highly express AMPA receptors (AMPAR, tetrameric assemblies of four GluR1–GluR4 subunits) but are often deficient (or low-expressing) in the Ca2+ impermeable GluR2 subunit [86]. Indeed, when the glutamatergic signal is blocked, survival, migration, and proliferation of glioma cells are impaired [87]. Glutamate exerts its tumorigenic effects via autocrine (activating glutamate receptors on GBM cells themselves) and paracrine (on adjacent astrocytes and neurons) stimuli and through the activation of AMPAR lacking GluR2, increasing Ca2+ influx and triggering growth-related MAPK and Akt pathways (Figure 2) [88,89]. AMPARs are Ca2+-permeable if they lack or contain the unedited GluR2, since in the brain, when this subunit is subjected to RNA editing (a post-transcriptional modification which converts a glutamine codon in the GluA2 gene in arginine codon in the mRNA), by the adenine deaminase RNA specific (ADAR2) enzyme, AMPARs are Ca2+-impermeable [90]. Thus, GluR2 RNA editing tunes glutamatergic neurotransmission and might play a key role in normal and pathologic brain, including glioma [91]. Interestingly, low ADAR2 protein in GBM cells promotes cell proliferation and migration and is associated with a patient’s shorter survival [92,93].
Overlapping epileptogenic and oncogenic mechanisms have been recently corroborated by the identification of functional synapses between GBM cells and neurons [94], forming tripartite glutamatergic synapses of cancer cells, and pre- and postsynaptic neurons [95]. These neuron–glioma networks are regulated by glutamate via AMPAR-dependent transmission and cancer-cell migration and proliferation, thus boosting GBM progression (Figure 1) [94,95].
Consequently, there is an emerging view based on glutamatergic signaling in neuron–GBM cell bidirectional interaction, which connects seizures and brain tumor progression and offers potential common therapeutic targets to treat both diseases.

4.5. Endogenous Cannabinoids

Endocannabinoid (eCB) system is composed of two major GPCR cannabinoid-specific receptors, namely CB1 and CB2, activated by eCBs (2-arachidoyl glycerol, 2-AG, and anandamide, AEA) and regulated by anabolic and catabolic enzymes. CB1 receptors are abundantly expressed in the CNS and on neurons, while CB2 receptors are principally present in the immune cells (e.g., leukocytes, microglia) [96] and other cell types, including cancer cells. Neuromodulatory signaling of eCBs is involved in a variety of brain physiological processes, and its alteration is relevant to psychiatric and neurological disorders, including epilepsy [97]. In animal models, eCB level is increased in mice with seizure-induced neurotoxicity (acute neuronal depolarization as by kainic acid, KA) [98,99], and deletion of CB1 in hippocampal glutamatergic excitatory neurons enhances seizure severity in KA-induced mice [100,101]. In epileptic patients, CB1 receptors and eCBs (i.e., AEA) are downregulated [102,103]. As previously highlighted, hyper-activation of glutamatergic transmission is a key pathogenic event for seizure generation; therefore, mechanisms regulating excitatory neurotransmission, such as the eCB system, might elucidate the pathophysiology of epilepsy and suggest novel therapeutic strategies.
As far as GBM, eCBs are still largely unexplored and contradictory: high 2-AG levels were detected in GBM [104,105], while AEA was found higher [106] or lower [107] in gliomas than in non-tumor tissues, and it could exert an in vitro and in vivo antitumor activity, as reported in GBM cell lines [104]. CB1 and CB2 receptors are variably present in human tumors; downregulation of CB1 levels in GBM, compared to a healthy brain, has been reported [108], as well as upregulation [105,108]. Concerning CB2, it is highly expressed in GBM [105], and antitumor effects of agonists have been reported in human GBM xenografts [109] and GBM stem cells [110].
Overall, evidence suggests a key role for eCB signaling in seizure onset and possibly as an add-on in resistant epilepsy [111], while its potential activity and mechanisms leading to GBM proliferation are largely unexplored and uncertain.

5. Repurposing Antiepileptic Drugs for the Treatment of Glioblastoma: Pharmacologic Targets

The scenario so far described underlines critical point in GBM prognosis and management: (i) GBM has a fatal clinical course despite aggressive and multimodal treatments; (ii) epilepsy frequently develops in GBM patients; (iii) the two diseases show shared pathogenic processes.
Drug repurposing, identifying new therapeutic uses for already-available drugs, is a promising tool that speeds up drug discovery time. In GBM, several known compounds have been explored for new use as antitumor activity [112,113,114]. Therefore, repositioning AEDs can be a promising option able to improve patient survival and control both seizures and GBM growth and recurrence.
LEV and VPA are the most studied drugs, possibly having a dual function as antiepileptic and antineoplastic, and are largely used in patients with GRE. Studies have also been carried out on other AEDs generally used as a second choice or add-on in case of therapeutic failure on the basis of the mechanism of action potentially impacting tumor growth (e.g., PER and CBD).

5.1. Antitumor Efficacy of AEDs Used in GRE: Preclinical and Clinical Perspectives

5.1.1. Levetiracetam

LEV exerts its antiepileptic effects through binding to the synaptic vesicle glycoprotein 2A (SV2A), which regulates neurotransmitter vesicular dynamics in neurons, reducing Ca2+ release and acting as a negative allosteric modulator of GABA- and glycin-gated currents, thus supporting GABA release [115,116].
Antitumor efficacy of LEV, analyzed in GBM cell lines, has been attributed to the inhibitory action on MGMT, overcoming this GBM resistance mechanism and enhancing TMZ effects [117] by decreasing MGMT expression and activating apoptotic pathway [118]. Interestingly, LEV shows antitumor effects at concentrations in the serum therapeutic range for seizure prophylaxis [119]. In addition, LEV, combined with IFN-α, enhanced the anti-tumor activity of TMZ in MGMT-positive GBM cells [120].
In the clinical setting, the survival benefit of LEV has been evaluated in GBM patient treated with current SOC.
Retrospective analyses show that LEV improves patients’ PFS and OS [121,122,123]. However, a pooled analysis of four different clinical trials (NCT00943826, NCT00689221, NCT00884741, NCT00813943) on a large series of cases treated with LEV at the start of chemo-radio-therapy failed to find an association with patients’ survival [124]. This controversial association between LEV and clinical outcome was further investigated in an observational study on IDHwt GBMs, reporting a possible OS benefit when LEV is used during the whole standard SOC duration [125].
Overall, the observational retrospective studies contain information and selection bias, while prospective randomized controlled trials and studies including specific molecular profiles will help address the actual efficacy of LEV in GBM. Currently, an open-label single-arm phase 2 study, with external and historical control, enrolling 73 patients (NCT02815410), reported minimal survival benefits over the radio-chemo-therapy [126]. A study protocol for a double-blind randomized clinical trial focusing on the clinical benefits of LEV + TMZ in the treatment of GBM has been planned [127].

5.1.2. Valproic Acid

VPA is a potent anticonvulsant acting through multiple mechanisms, such as interaction with GABA transaminase, succinate semialdehyde dehydrogenase, postsynaptic GABA and glutamate receptors, and ion channels [128]. Antitumor effects of VPA have been extensively explored in both the pre-clinical and clinical settings [129] being a histone deacetylase (HDAC) inhibitor (epigenetic drug) [130], changing not only in histone acetylation but also in DNA methylation in glioma cell lines [131].
Besides HDAC, other targets have been associated with VPA antitumor activity in GBM cells, such as the upregulation of brain-derived neurotrophic factor (BDNF) [132] and signaling pathways, such as ERK/Akt [133], Akt/mTOR [134], and Wnt [135]. In glioma cells, VPA promotes apoptosis [133,134] and autophagy [134,136] and impairs cell proliferation and invasiveness [135,137]. However, the effective dose (millimolar) of VPA is far above that used for the treatment of epilepsy [138]. At a concentration of up to 100 µM, VPA failed to inhibit glioma cell growth and metastasis in vitro [139]. Beyond antitumor effects as a single agent, VPA sensitizes GBM cells to several anticancer drugs, such as nitrosoureas [140], TMZ [141,142,143], gefitinib [144], etoposide [145], and radiation therapy [143,146].
Therefore, based on in vitro and in vivo preclinical results, VPA may manage both epilepsy and glioma as a potential adjuvant drug to enhance patients’ response to SOC.
Several studies show that VPA increases the median survival of patients affected by GBM [147,148,149]. Retrospective studies suggest that adding VPA to TMZ-based chemoradiotherapy in newly diagnosed GBMs slightly prolongs survival at the expense of increased thrombocytopenia and leukopenia [148,149,150]. In addition, VPA used during radiotherapy decreases side effects and prolongs OS and PFS [151].
A recent update of an open-label phase 2 study (NCI-06-C-0112) on 37 patients confirms previous data reporting that the addition of VPA improves their PFS and OS [152]. A meta-analysis confirmed the prolonged survival with VPA combined treatment [153]; conversely, analysis of pooled randomized trials enrolling 1869 patients did not associate with extended PFS or OS [124].
As a whole, preclinical evidence supports VPA antitumor and TMZ-sensitizing efficacy, while in the clinical setting, further investigation is needed to justify VPA as a potential adjuvant in current GBM therapy.

5.1.3. Perampanel

PER is a selective non-competitive AMPA receptor antagonist. The shared mechanism of AMPA-activation connecting seizure activity and GBM growth support the rationale for AMPA-receptor blocker evaluation of antitumor activity, first investigated with talampanel, a PER analog, which is still not on the market due to its poor pharmacokinetics and short half-life [154], without reaching concordant significant results [155,156,157].
The antiproliferative mechanism of PER is not yet clear and preliminary data are obtained in a wide range of concentrations and different established glioma cell lines. In a study comparing antitumor effects in GBM cell lines of various AEDs, including LEV and VPA, only PER showed inhibitory effects on cell proliferation, migration, and invasion, without induction of apoptosis and reduction of extracellular glutamate levels [139,158]. Whereas, in other works, PER antineoplastic activity is mediated by apoptosis, possibly due to increased GluR2/3 expression synergizing with TMZ [159,160].
In in vivo experiments in C6 rat glioma xenografts, PER did not affect either tumor growth or animal survival, while it blocked epileptiform discharges in organotypic glioma slices and reduced glucose uptake in C6 glioma cells in vitro [161]. Similarly, in another orthotopic rat model of glioma (F98), which promotes an epileptiform phenotype, the therapeutic efficacy of PER, as an adjuvant to standard radiochemotherapy, failed to impact tumor progression and animal survival, while tumor-associated epilepsy was abolished, maintaining the glutamatergic network activity on healthy peritumoral tissue of treated animals [162].
Interestingly, in the scenario of neuro-glioma synapses which produce postsynaptic currents mediated by AMPA receptors, the 14-day treatment of GBM xenografted mice with PER exerts significant antiproliferative activity on GBM cells [95] and in mice bearing pediatric glioma [94].
At the clinical level, only one retrospective study evaluated PER impact on seizures and tumor progression in 12 GBM patients with refractory epilepsy, detecting by MRI a reduction of tumor volume [163]. Currently, while in the management of GRE, PER in add-on could be a valid therapeutic option, and a study is ongoing to confirm its safety and efficacy (NCT04650204); trials evaluating PER efficacy on GBM progression and survival are lacking.
However, the connection between increased neuronal activity and glioma progression mediated by glutamatergic synapses warrants further investigation in which AMPA receptor inhibition by PER may be exploited both at the preclinical and clinical levels.

5.1.4. Cannabidiol

Cannabidiol (CBD) is one of the major molecules found in cannabis and hemp, and pharmaceutical-grade CBD has FDA and ENMA approval for seizures in Dravet syndrome and Lennox–Gastaut syndrome, as well as for tuberous complex [164].
Although the mechanism of action is not yet fully understood, CBD has numerous targets; in particular, its anticonvulsant activity is due to the interaction with three receptors—the transient receptor potential vanilloid-1 (TRPV1), the orphan G protein-coupled receptor-55 (GPR55) and the equilibrative nucleoside transporter 1 (ENT-1), implicated in the regulation of neuronal excitability [165].
The in vitro and in vivo anti-glioma activity of different CB1/CB2 agonists (cannabinoids) and endocannabinoids has been described (for a review, see [97]).
Phytocannabinoids, namely CBD, in preclinical studies, demonstrated antiproliferative and pro-apoptotic effects and inhibition of the migration of GBM [166], acting as a sensitizer to chemotherapeutic agents [167]. In human primary GBM stem-like, CBD exerts cytotoxic activity likely through modulation of a key transcription factor in GBM, the nuclear factor kappa B (NF-κB), by promoting DNA binding and preventing posttranslational modification of the NF-κB subunit RELA/p65 [168].
In a retrospective trial, CBD seems to demonstrate an increase in the survival of patients with GBM [169].
Results from a phase 1b randomized trial of cannabinoid oromucosal spray (nabiximols, containing delta-9-tetrahydrocannabinol-THC, CBD, and additional cannabinoid and non-cannabinoid components) with TMZ in patients with recurrent GBM showed satisfactory safety and tolerability, no drug–drug interaction, and survival at 1 year of 83% for nabiximols and 44% for placebo arm [170].
Currently, there is an ongoing phase 1b trial (NCT03529448) assessing the safety and antitumor activity of the THC + CBD combination with standard therapy in newly-diagnosed GBM.
Further studies will show whether compounds that exert promising effects on tumor cells, in addition to the psychoactive THC, such as CBD and inhibitors of endocannabinoid degradation, could be potential combination partners for established chemotherapeutic agents in GBM treatment.

5.2. Potential Repositioning of Other AEDs Used in GRE

Few in vitro investigations report on the potential antitumor effects of other approved AEDs used in GRE.
Among them, LCM has been shown to be ineffective [171] or to exert cytotoxicity and anti-migratory effects in the micromolar range [172] in GBM cells lines, likely via HDAC inhibition, as observed in a breast cancer model [173]. Curiously, the inactive enantiomer of this AED, S-lacosamide (S-LCM), is able to decrease GBM cell proliferation and the growth of orthotopic tumors in a murine GBM model [174].
BRV, with a 15- to 30-fold higher affinity for SV2A than LEV, shares the above-described antitumor effects with LCM [172].
LMG fails to reach significant cytotoxicity at therapeutic concentrations in vitro [138]; however, P3K/AKT signaling might represent its additional target, also blocking voltage-dependent calcium and sodium channels, to exploit antitumor effects as observed in breast cancer models [175].
PGB, a GABA analog, exhibits anti-neuroinflammatory effects, preventing substance P (SP)-mediated IL-6 and IL-8 production in the U373 GBM cell line through the inhibition of p38 MAPK and NF-κB signaling molecules [176].
Similarly, a slight cytotoxic activity in GBM cells has been described for TPM [138].
Stiripentol, an agonist of GABAAR, approved for the treatment of seizures associated with Dravet syndrome [177], showed selective cytotoxic and anti-migratory activity in GBM cells and additive or synergistic effects with TMZ [178]. Interestingly, stiripentol decreases GBM invasion and growth in xenografted mice [179], likely via lactate dehydrogenase (LDH) block, an enzyme involved in both neuron hyperpolarization [180] and highly glycolytic metabolism of GBM cells, which catalyzes lactate production and correlates with high proliferation and invasion [181].
Comprehensive antitumor mechanisms of the above AEDs remain unclear, as well as their efficacy in vivo; however, several mechanisms might represent a novel therapeutic approach for targeting GBM (Table 1).

6. Conclusions

GBM is a lethal, highly proliferating, and invasive cancer that tightly interconnects with brain tissue, coopting astrocytes, macrophages/microglia, and endothelial cells in the tumor microenvironment, regulated by both brain activity and neuron–glioma synaptic interaction (Figure 3). Communication may occur directly between neuron and glioma cells or by regulation of other cell types within the microenvironment through neurotransmitter signaling. Neuronal mechanisms can trigger GBM cell growth and invasion and sustain intratumoral cellular heterogeneity. Concomitantly, cellular and molecular alterations in GBM cells and peritumoral tissues may generate epileptic symptoms in patients with GBM via shared signals mediated by neurotransmitters (glutamate, GABA).
Further studies are issued to clarify the anti-tumor efficacy of AEDs based on specific biomarkers of GBM so that two diseases, GRE and GBM, could be targeted with significant clinical benefit simultaneously (Figure 3). The emerging research field of cancer neuroscience will help identify specific and mutual pathophysiology and vulnerabilities of GBM progression and seizure onset, potentially promising for future bidirectional therapeutic targeting of common potential drivers, such as glutamatergic signaling, as studies with different AEDs, although presently not conclusive, suggest.

Author Contributions

Research and analysis of the cited articles, M.S., G.B., S.P., F.S., I.D., A.C., F.M. and F.B.; writing—original draft preparation: M.S. and G.B.; writing—review and editing: M.S., G.B., S.P., F.S., I.D., A.C., F.M. and F.B.; conceptualization and supervision, F.M. and F.B. All authors have read and agreed to the published version of the manuscript.

Funding

PNRR-MUR-M4C2 PE0000006 Research Program “MNESYS”—A multiscale integrated approach to the study of the nervous system in health and disease.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ostrom, Q.T.; Gittleman, H.; Truitt, G.; Boscia, A.; Kruchko, C.; Barnholtz-Sloan, J.S.; Duncan, D.L. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2011–2015 Introduction. Neuro. Oncol. 2018, 20, 1–86. [Google Scholar] [CrossRef] [Green Version]
  2. Fabbro-Peray, P.; Zouaoui, S.; Darlix, A.; Fabbro, M.; Pallud, J.; Rigau, V.; Mathieu-Daude, H.; Bessaoud, F.; Fabienne Bauchet, L.; Riondel, A.; et al. Association of Patterns of Care, Prognostic Factors, and Use of Radiotherapy-Temozolomide Therapy with Survival in Patients with Newly Diagnosed Glioblastoma: A French National Population-Based Study. J. Neurooncol. 2019, 142, 91–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tejada Solís, S.; Plans Ahicart, G.; Iglesias Lozano, I.; De Quintana Schmidt, C.; Fernández Coello, A.; Hostalot Panisello, C.; Ley Urzaiz, L.; García Romero, J.C.; Díez Valle, R.; González Sánchez, J.; et al. Glioblastoma Treatment Guidelines: Consensus by the Spanish Society of Neurosurgery Tumor Section. Neurocirugía 2020, 31, 289–297. [Google Scholar] [CrossRef] [PubMed]
  4. Low, J.T.; Ostrom, Q.T.; Cioffi, G.; Neff, C.; Waite, K.A.; Kruchko, C.; Barnholtz-Sloan, J.S. Primary Brain and Other Central Nervous System Tumors in the United States (2014–2018): A Summary of the CBTRUS Statistical Report for Clinicians. Neurooncol. Pract. 2022, 9, 165–182. [Google Scholar] [CrossRef]
  5. Carrano, A.; Jose Juarez, J.; Incontri, D.; Ibarra, A.; Cazares, H.G.; Minchiotti, G.; Fico, A.; Castresana, J.S. Sex-Specific Differences in Glioblastoma. Cells 2021, 10, 1783. [Google Scholar] [CrossRef]
  6. Leece, R.; Xu, J.; Ostrom, Q.T.; Chen, Y.; Kruchko, C.; Barnholtz-Sloan, J.S. Global Incidence of Malignant Brain and Other Central Nervous System Tumors by Histology, 2003–2007. Neuro. Oncol. 2017, 19, 1553–1564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Gieryng, A.; Pszczolkowska, D.; Walentynowicz, K.A.; Rajan, W.D.; Kaminska, B. Immune Microenvironment of Gliomas. Lab. Investig. 2017, 97, 498–518. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Louis, D.N.; Perry, A.; Wesseling, P.; Brat, D.J.; Cree, I.A.; Figarella-Branger, D.; Hawkins, C.; Ng, H.K.; Pfister, S.M.; Reifenberger, G.; et al. The 2021 WHO Classification of Tumors of the Central Nervous System: A Summary. Neuro. Oncol. 2021, 23, 1231–1251. [Google Scholar] [CrossRef]
  10. Mattei, V.; Santilli, F.; Martellucci, S.; Monache, S.D.; Fabrizi, J.; Colapietro, A.; Angelucci, A.; Festuccia, C. The Importance of Tumor Stem Cells in Glioblastoma Resistance to Therapy. Int. J. Mol. Sci. 2021, 22, 3863. [Google Scholar] [CrossRef]
  11. Andersen, R.S.; Anand, A.; Harwood, D.S.L.; Kristensen, B.W. Tumor-Associated Microglia and Macrophages in the Glioblastoma Microenvironment and Their Implications for Therapy. Cancers 2021, 13, 4255. [Google Scholar] [CrossRef]
  12. Uddin, M.S.; Mamun, A.; Alghamdi, B.S.; Tewari, D.; Jeandet, P.; Sarwar, M.S.; Ashraf, G.M. Epigenetics of Glioblastoma Multiforme: From Molecular Mechanisms to Therapeutic Approaches. Semin. Cancer Biol. 2022, 83, 100–120. [Google Scholar] [CrossRef] [PubMed]
  13. 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] [PubMed]
  14. Wen, P.Y.; Weller, M.; Lee, E.Q.; Alexander, B.M.; Barnholtz-Sloan, J.S.; Barthel, F.P.; Batchelor, T.T.; Bindra, R.S.; Chang, S.M.; Antonio Chiocca, E.; et al. Glioblastoma in Adults: A Society for Neuro-Oncology (SNO) and European Society of Neuro-Oncology (EANO) Consensus Review on Current Management and Future Directions. Neuro. Oncol. 2020, 22, 1073–1113. [Google Scholar] [CrossRef] [PubMed]
  15. Stupp, R.; Taillibert, S.; Kanner, A.; Read, W.; Steinberg, D.M.; Lhermitte, B.; Toms, S.; Idbaih, A.; Ahluwalia, M.S.; Fink, K.; et al. Effect of Tumor-Treating Fields Plus Maintenance Temozolomide vs. Maintenance Temozolomide Alone on Survival in Patients With Glioblastoma: A Randomized Clinical Trial. JAMA 2017, 318, 2306–2316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Gramatzki, D.; Kickingereder, P.; Hentschel, B.; Felsberg, J.; Herrlinger, U.; Schackert, G.; Tonn, J.C.; Westphal, M.; Sabel, M.; Schlegel, U.; et al. Limited Role for Extended Maintenance Temozolomide for Newly Diagnosed Glioblastoma. Neurology 2017, 88, 1422–1430. [Google Scholar] [CrossRef]
  17. Chinot, O.L.; Wick, W.; Mason, W.; Henriksson, R.; Saran, F.; Nishikawa, R.; Carpentier, A.F.; Hoang-Xuan, K.; Kavan, P.; Cernea, D.; et al. Bevacizumab plus Radiotherapy-Temozolomide for Newly Diagnosed Glioblastoma. N. Engl. J. Med. 2014, 370, 709–722. [Google Scholar] [CrossRef] [Green Version]
  18. Lamborn, K.R.; Yung, W.K.A.; Chang, S.M.; Wen, P.Y.; Cloughesy, T.F.; DeAngelis, L.M.; Robins, H.I.; Lieberman, F.S.; Fine, H.A.; Fink, K.L.; et al. Progression-Free Survival: An Important End Point in Evaluating Therapy for Recurrent High-Grade Gliomas. Neuro. Oncol. 2008, 10, 162–170. [Google Scholar] [CrossRef]
  19. Vargas López, A.J. Glioblastoma in Adults: A Society for Neuro-Oncology (SNO) and European Society of Neuro-Oncology (EANO) Consensus Review on Current Management and Future Directions. Neuro. Oncol. 2021, 23, 502–503. [Google Scholar] [CrossRef]
  20. Nabors, L.B.; Portnow, J.; Ahluwalia, M.; Baehring, J.; Brem, H.; Brem, S.; Butowski, N.; Campian, J.L.; Clark, S.W.; Fabiano, A.J.; et al. Central Nervous System Cancers, Version 3.2020, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2020, 18, 1537–1570. [Google Scholar] [CrossRef]
  21. Weller, M.; Tabatabai, G.; Kästner, B.; Felsberg, J.; Steinbach, J.P.; Wick, A.; Schnell, O.; Hau, P.; Herrlinger, U.; Sabel, M.C.; et al. MGMT Promoter Methylation Is a Strong Prognostic Biomarker for Benefit from Dose-Intensified Temozolomide Rechallenge in Progressive Glioblastoma: The DIRECTOR Trial. Clin. Cancer Res. 2015, 21, 2057–2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wick, W.; Gorlia, T.; Bendszus, M.; Taphoorn, M.; Sahm, F.; Harting, I.; Brandes, A.A.; Taal, W.; Domont, J.; Idbaih, A.; et al. Lomustine and Bevacizumab in Progressive Glioblastoma. N. Engl. J. Med. 2017, 377, 1954–1963. [Google Scholar] [CrossRef] [PubMed]
  23. Gramatzki, D.; Roth, P.; Rushing, E.J.; Weller, J.; Andratschke, N.; Hofer, S.; Korol, D.; Regli, L.; Pangalu, A.; Pless, M.; et al. Bevacizumab May Improve Quality of Life, but Not Overall Survival in Glioblastoma: An Epidemiological Study. Ann. Oncol. 2018, 29, 1431–1436. [Google Scholar] [CrossRef] [PubMed]
  24. Le Rhun, E.; Preusser, M.; Roth, P.; Reardon, D.A.; Van Den Bent, M.; Wen, P.; Reifenberger, G.; Weller, M. Molecular Targeted Therapy of Glioblastoma. Cancer Treat Rev. 2019, 80, 101896. [Google Scholar] [CrossRef]
  25. Andersen, B.M.; Reardon, D.A. Immunotherapy Approaches for Adult Glioma: Knowledge Gained from Recent Clinical Trials. Curr. Opin. Neurol. 2022, 35, 803–813. [Google Scholar] [CrossRef]
  26. Fisher, R.S.; Van Emde Boas, W.; Blume, W.; Elger, C.; Genton, P.; Lee, P.; Engel, J. Epileptic Seizures and Epilepsy: Definitions Proposed by the International League Against Epilepsy (ILAE) and the International Bureau for Epilepsy (IBE). Epilepsia 2005, 46, 470–472. [Google Scholar] [CrossRef]
  27. Liang, S.; Fan, X.; Zhao, M.; Shan, X.; Li, W.; Ding, P.; You, G.; Hong, Z.; Yang, X.; Luan, G.; et al. Clinical Practice Guidelines for the Diagnosis and Treatment of Adult Diffuse Glioma-Related Epilepsy. Cancer Med 2019, 8, 4527–4535. [Google Scholar] [CrossRef] [Green Version]
  28. Herman, S.T. Epilepsy after Brain Insult: Targeting Epileptogenesis. Neurology 2002, 59, S21–S26. [Google Scholar] [CrossRef]
  29. Englot, D.J.; Chang, E.F.; Vecht, C.J. Epilepsy and Brain Tumors. Handb. Clin. Neurol. 2016, 134, 267–285. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, D.Y.; Chen, C.C.; Crawford, J.R.; Wang, S.G. Tumor-Related Epilepsy: Epidemiology, Pathogenesis and Management. J. Neurooncol. 2018, 139, 13–21. [Google Scholar] [CrossRef]
  31. Pallud, J.; Capelle, L.; Huberfeld, G. Tumoral Epileptogenicity: How Does It Happen? Epilepsia 2013, 54 (Suppl. 9), 30–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wang, Y.H.; Huang, T.L.; Chen, X.; Yu, S.X.; Li, W.; Chen, T.; Li, Y.; Kuang, Y.Q.; Shu, H.F. Glioma-Derived TSP2 Promotes Excitatory Synapse Formation and Results in Hyperexcitability in the Peritumoral Cortex of Glioma. J. Neuropathol. Exp. Neurol. 2021, 80, 137–149. [Google Scholar] [CrossRef] [PubMed]
  33. Aronica, E.; Gorter, J.A.; Jansen, G.H.; Leenstra, S.; Yankaya, B.; Troost, D. Expression of Connexin 43 and Connexin 32 Gap-Junction Proteins in Epilepsy-Associated Brain Tumors and in the Perilesional Epileptic Cortex. Acta Neuropathol. 2001, 101, 449–459. [Google Scholar] [CrossRef]
  34. Dong, H.; Zhou, X.W.; Wang, X.; Yang, Y.; Luo, J.W.; Liu, Y.H.; Mao, Q. Complex Role of Connexin 43 in Astrocytic Tumors and Possible Promotion of Glioma-associated Epileptic Discharge (Review). Mol. Med. Rep. 2017, 16, 7890–7900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Komiyama, K.; Iijima, K.; Kawabata-Iwakawa, R.; Fujihara, K.; Kakizaki, T.; Yanagawa, Y.; Yoshimoto, Y.; Miyata, S. Glioma Facilitates the Epileptic and Tumor-Suppressive Gene Expressions in the Surrounding Region. Sci. Rep. 2022, 12, 6805. [Google Scholar] [CrossRef]
  36. Li, L.; Fang, S.; Li, G.; Zhang, K.; Huang, R.; Wang, Y.; Zhang, C.; Li, Y.; Zhang, W.; Zhang, Z.; et al. Glioma-Related Epilepsy in Patients with Diffuse High-Grade Glioma after the 2016 WHO Update: Seizure Characteristics, Risk Factors, and Clinical Outcomes. J. Neurosurg. 2021, 136, 67–75. [Google Scholar] [CrossRef] [PubMed]
  37. Koekkoek, J.A.F.; Kerkhof, M.; Dirven, L.; Heimans, J.J.; Reijneveld, J.C.; Taphoorn, M.J.B. Seizure Outcome after Radiotherapy and Chemotherapy in Low-Grade Glioma Patients: A Systematic Review. Neuro. Oncol. 2015, 17, 924–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Climans, S.A.; Brandes, A.A.; Cairncross, J.G.; Ding, K.; Fay, M.; Laperriere, N.; Menten, J.; Nishikawa, R.; O’Callaghan, C.J.; Perry, J.R.; et al. Temozolomide and Seizure Outcomes in a Randomized Clinical Trial of Elderly Glioblastoma Patients. J. Neurooncol. 2020, 149, 65–71. [Google Scholar] [CrossRef]
  39. Van Der Meer, P.B.; Taphoorn, M.J.B.; Koekkoek, J.A.F. C URRENT OPINION Management of Epilepsy in Brain Tumor Patients. Curr. Opin. Oncol. 2022, 34, 685–690. [Google Scholar] [CrossRef]
  40. Maschio, M.; Aguglia, U.; Avanzini, G.; Banfi, P.; Buttinelli, C.; Capovilla, G.; Casazza, M.M.L.; Colicchio, G.; Coppola, A.; Costa, C.; et al. Management of Epilepsy in Brain Tumors. Neurol. Sci. 2019, 40, 2217–2234. [Google Scholar] [CrossRef]
  41. Walbert, T.; Harrison, R.A.; Schiff, D.; Avila, E.K.; Chen, M.; Kandula, P.; Lee, J.W.; le Rhun, E.; Stevens, G.H.J.; Vogelbaum, M.A.; et al. SNO and EANO Practice Guideline Update: Anticonvulsant Prophylaxis in Patients with Newly Diagnosed Brain Tumors. Neuro. Oncol. 2021, 23, 1835–1844. [Google Scholar] [CrossRef]
  42. Stocksdale, B.; Nagpal, S.; Hixson, J.D.; Johnson, D.R.; Rai, P.; Shivaprasad, A.; Tremont-Lukats, I.W. Neuro-Oncology Practice Clinical Debate: Long-Term Antiepileptic Drug Prophylaxis in Patients with Glioma. Neurooncol. Pract. 2020, 7, 583–588. [Google Scholar] [CrossRef] [PubMed]
  43. Medicines Complete. Martindale: The Complete Drug Reference. 2020. Available online: https://about.medicinescomplete.com/publication/martindale-the-complete-drug-reference/ (accessed on 10 January 2023).
  44. Bénit, C.P.; Vecht, C.J. Seizures and Cancer: Drug Interactions of Anticonvulsants with Chemotherapeutic Agents, Tyrosine Kinase Inhibitors and Glucocorticoids. Neurooncol. Pract. 2016, 3, 245–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bourg, V.; Lebrun, C.; Chichmanian, R.M.; Thomas, P.; Frenay, M. Nitroso-Urea-Cisplatin-Based Chemotherapy Associated with Valproate: Increase of Haematologic Toxicity. Ann. Oncol. 2001, 12, 217–219. [Google Scholar] [CrossRef] [PubMed]
  46. Armstrong, T.S.; Grant, R.; Gilbert, M.R.; Lee, J.W.; Norden, A.D. Epilepsy in Glioma Patients: Mechanisms, Management, and Impact of Anticonvulsant Therapy. Neuro. Oncol. 2016, 18, 779–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. De Bruin, M.E.; Van Der Meer, P.B.; Dirven, L.; Taphoorn, M.J.B.; Koekkoek, J.A.F. Efficacy of Antiepileptic Drugs in Glioma Patients with Epilepsy: A Systematic Review. Neurooncol. Pract. 2021, 8, 501–517. [Google Scholar] [CrossRef] [PubMed]
  48. Van Der Meer, P.B.; Dirven, L.; Fiocco, M.; Vos, M.J.; Kouwenhoven, M.C.M.; Van Den Bent, M.J.; Taphoorn, M.J.B.; Koekkoek, J.A.F. First-Line Antiepileptic Drug Treatment in Glioma Patients with Epilepsy: Levetiracetam vs Valproic Acid. Epilepsia 2021, 62, 1119–1129. [Google Scholar] [CrossRef]
  49. Maschio, M.; Zarabla, A.; Maialetti, A.; Giannarelli, D.; Koudriavtseva, T.; Villani, V.; Zannino, S. Perampanel in Brain Tumor-Related Epilepsy: Observational Pilot Study. Brain Behav. 2020, 10, e01612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Coppola, A.; Zarabla, A.; Maialetti, A.; Villani, V.; Koudriavtseva, T.; Russo, E.; Nozzolillo, A.; Sueri, C.; Belcastro, V.; Balestrini, S.; et al. Perampanel Confirms to Be Effective and Well-Tolerated as an Add-On Treatment in Patients With Brain Tumor-Related Epilepsy (PERADET Study). Front. Neurol. 2020, 11, 592. [Google Scholar] [CrossRef]
  51. Maschio, M.; Maialetti, A.; Mocellini, C.; Domina, E.; Pauletto, G.; Costa, C.; Mascia, A.; Romoli, M.; Giannarelli, D. Effect of Brivaracetam on Efficacy and Tolerability in Patients With Brain Tumor-Related Epilepsy: A Retrospective Multicenter Study. Front. Neurol. 2020, 11, 813. [Google Scholar] [CrossRef]
  52. Golub, V.; Reddy, D.S. Cannabidiol Therapy for Refractory Epilepsy and Seizure Disorders. Adv. Exp. Med. Biol. 2021, 1264, 93–110. [Google Scholar] [CrossRef] [PubMed]
  53. Souza, J.D.R.; Pacheco, J.C.; Rossi, G.N.; De-Paulo, B.O.; Zuardi, A.W.; Guimarães, F.S.; Hallak, J.E.C.; Crippa, J.A.; Dos Santos, R.G. Adverse Effects of Oral Cannabidiol: An Updated Systematic Review of Randomized Controlled Trials (2020–2022). Pharmaceutics 2022, 14, 2598. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, H.; Judkins, J.; Thomas, C.; Wu, M.; Khoury, L.; Benjamin, C.G.; Pacione, D.; Golfinos, J.G.; Kumthekar, P.; Ghamsari, F.; et al. Mutant IDH1 and Seizures in Patients with Glioma. Neurology 2017, 88, 1805–1813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Van Opijnen, M.P.; Tesileanu, C.M.S.; Dirven, L.; Van Der Meer, P.B.; Wijnenga, M.M.J.; Vincent, A.J.P.E.; Broekman, M.L.D.; Dubbink, H.J.; Kros, J.M.; van Duinen, S.G.; et al. IDH1/2 Wildtype Gliomas Grade 2 and 3 with Molecular Glioblastoma-like Profile Have a Distinct Course of Epilepsy Compared to IDH1/2 Wildtype Glioblastomas. Neuro. Oncol. 2022; Epub ahead of print. [Google Scholar] [CrossRef]
  56. Mortazavi, A.; Fayed, I.; Bachani, M.; Dowdy, T.; Jahanipour, J.; Khan, A.; Owotade, J.; Walbridge, S.; Inati, S.K.; Steiner, J.; et al. IDH-Mutated Gliomas Promote Epileptogenesis through d-2-Hydroxyglutarate-Dependent MTOR Hyperactivation. Neuro. Oncol. 2022, 24, 1423–1435. [Google Scholar] [CrossRef] [PubMed]
  57. Garrett, M.; Sperry, J.; Braas, D.; Yan, W.; Le, T.M.; Mottahedeh, J.; Ludwig, K.; Eskin, A.; Qin, Y.; Levy, R.; et al. Metabolic Characterization of Isocitrate Dehydrogenase (IDH) Mutant and IDH Wildtype Gliomaspheres Uncovers Cell Type-Specific Vulnerabilities. Cancer Metab. 2018, 6, 4. [Google Scholar] [CrossRef] [Green Version]
  58. Carbonneau, M.; Gagné, L.M.; Lalonde, M.-E.; Germain, M.-A.; Motorina, A.; Guiot, M.-C.; Secco, B.; Vincent, E.E.; Tumber, A.; Hulea, L.; et al. The Oncometabolite 2-Hydroxyglutarate Activates the MTOR Signalling Pathway. Nat. Commun. 2016, 7, 12700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Crino, P.B. The MTOR Signalling Cascade: Paving New Roads to Cure Neurological Disease. Nat. Rev. Neurol. 2016, 12, 379–392. [Google Scholar] [CrossRef]
  60. 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. [Google Scholar] [CrossRef]
  61. Venkatesan, S.; Hoogstraat, M.; Caljouw, E.; Pierson, T.; Spoor, J.K.H.; Zeneyedpour, L.; Dubbink, H.J.; Dekker, L.J.; Van Der Kaaij, M.; Kloezeman, J.; et al. TP53 Mutated Glioblastoma Stem-like Cell Cultures Are Sensitive to Dual MTORC1/2 Inhibition While Resistance in TP53 Wild Type Cultures Can Be Overcome by Combined Inhibition of MTORC1/2 and Bcl-2. Oncotarget 2016, 7, 58435–58444. [Google Scholar] [CrossRef]
  62. Huberfeld, G.; Vecht, C.J. Seizures and Gliomas--towards a Single Therapeutic Approach. Nat. Rev. Neurol. 2016, 12, 204–216. [Google Scholar] [CrossRef]
  63. Hatcher, A.; Yu, K.; Meyer, J.; Aiba, I.; Deneen, B.; Noebels, J.L. Pathogenesis of Peritumoral Hyperexcitability in an Immunocompetent CRISPR-Based Glioblastoma Model. J. Clin. Investig. 2020, 130, 2286–2300. [Google Scholar] [CrossRef] [Green Version]
  64. Winter, R.; Pembrey, M. Interpretation of the Heterogeneity in the Linkage Relationships of DNA Markers around the Fragile X Locus. Hum. Genet. 1987, 77, 297–298. [Google Scholar] [CrossRef]
  65. Hu, S.; Kao, H.-Y.; Yang, T.; Wang, Y. Early and Bi-Hemispheric Seizure Onset in a Rat Glioblastoma Multiforme Model. Neurosci. Lett. 2022, 766, 136351. [Google Scholar] [CrossRef]
  66. Toledo, M.; Sarria-Estrada, S.; Quintana, M.; Maldonado, X.; Martinez-Ricarte, F.; Rodon, J.; Auger, C.; Aizpurua, M.; Salas-Puig, J.; Santamarina, E.; et al. Epileptic Features and Survival in Glioblastomas Presenting with Seizures. Epilepsy Res. 2017, 130, 1–6. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, J.K.; Lee, J.H. Mechanistic Target of Rapamycin Pathway in Epileptic Disorders. J. Korean Neurosurg. Soc. 2019, 62, 272–287. [Google Scholar] [CrossRef]
  68. Nguyen, L.H.; Bordey, A. Convergent and Divergent Mechanisms of Epileptogenesis in MTORopathies. Front. Neuroanat. 2021, 15, 664695. [Google Scholar] [CrossRef]
  69. Sorrentino, U.; Bellonzi, S.; Mozzato, C.; Brasson, V.; Toldo, I.; Parrozzani, R.; Clementi, M.; Cassina, M.; Trevisson, E. Epilepsy in NF1: Epidemiologic, Genetic, and Clinical Features. A Monocentric Retrospective Study in a Cohort of 784 Patients. Cancers 2021, 13, 6336. [Google Scholar] [CrossRef]
  70. Johannessen, C.M.; Reczek, E.E.; James, M.F.; Brems, H.; Legius, E.; Cichowski, K. The NF1 Tumor Suppressor Critically Regulates TSC2 and MTOR. Proc. Natl. Acad. Sci. USA 2005, 102, 8573–8578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Yu, K.; Lin, C.C.J.; Hatcher, A.; Lozzi, B.; Kong, K.; Huang-Hobbs, E.; Cheng, Y.T.; Beechar, V.B.; Zhu, W.; Zhang, Y.; et al. PIK3CA Variants Selectively Initiate Brain Hyperactivity during Gliomagenesis. Nature 2020, 578, 166–171. [Google Scholar] [CrossRef] [PubMed]
  72. Venkatesh, H.S.; Johung, T.B.; Caretti, V.; Noll, A.; Tang, Y.; Nagaraja, S.; Gibson, E.M.; Mount, C.W.; Polepalli, J.; Mitra, S.S.; et al. Neuronal Activity Promotes Glioma Growth through Neuroligin-3 Secretion. Cell 2015, 161, 803–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Cucchiara, F.; Pasqualetti, F.; Giorgi, F.S.; Danesi, R.; Bocci, G. Epileptogenesis and Oncogenesis: An Antineoplastic Role for Antiepileptic Drugs in Brain Tumours? Pharmacol. Res. 2020, 156, 104786. [Google Scholar] [CrossRef] [PubMed]
  74. Kahle, K.T.; Rinehart, J.; Lifton, R.P. Phosphoregulation of the Na-K-2Cl and K-Cl Cotransporters by the WNK Kinases. Biochim. Biophys. Acta 2010, 1802, 1150–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wolf, H.K.; Roos, D.; Blümcke, I.; Pietsch, T.; Wiestler, O.D. Perilesional Neurochemical Changes in Focal Epilepsies. Acta Neuropathol. 1996, 91, 376–384. [Google Scholar] [CrossRef] [PubMed]
  76. Pallud, J.; Le Van Quyen, M.; Bielle, F.; Pellegrino, C.; Varlet, P.; Cresto, N.; Baulac, M.; Duyckaerts, C.; Kourdougli, N.; Chazal, G.; et al. Cortical GABAergic Excitation Contributes to Epileptic Activities around Human Glioma. Sci. Transl. Med. 2014, 6, 244ra89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Campbell, S.L.; Robel, S.; Cuddapah, V.A.; Robert, S.; Buckingham, S.C.; Kahle, K.T.; Sontheimer, H. GABAergic Disinhibition and Impaired KCC2 Cotransporter Activity Underlie Tumor-Associated Epilepsy. Glia 2015, 63, 23–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Huberfeld, G.; Wittner, L.; Clemenceau, S.; Baulac, M.; Kaila, K.; Miles, R.; Rivera, C. Perturbed Chloride Homeostasis and GABAergic Signaling in Human Temporal Lobe Epilepsy. J. Neurosci. 2007, 27, 9866–9873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. MacKenzie, G.; O’Toole, K.K.; Maguire, J. Compromised GABAergic Inhibition Contributes to Tumor-Associated Epilepsy. Epilepsy Res. 2016, 126, 185–196. [Google Scholar] [CrossRef] [Green Version]
  80. Habela, C.W.; Ernest, N.J.; Swindall, A.F.; Sontheimer, H. Chloride Accumulation Drives Volume Dynamics Underlying Cell Proliferation and Migration. J. Neurophysiol. 2009, 101, 750–757. [Google Scholar] [CrossRef] [Green Version]
  81. Hanada, T. Ionotropic Glutamate Receptors in Epilepsy: A Review Focusing on AMPA and NMDA Receptors. Biomolecules 2020, 10, 464. [Google Scholar] [CrossRef] [Green Version]
  82. Robert, S.M.; Buckingham, S.C.; Campbell, S.L.; Robel, S.; Holt, K.T.; Ogunrinu-Babarinde, T.; Warren, P.P.; White, D.M.; Reid, M.A.; Eschbacher, J.M.; et al. SLC7A1 Expression Is Associated with Seizures, Predicts Poor Survival in Patients with Malignant Glioma. Sci. Transl. Med. 2015, 7, 289ra86. [Google Scholar] [CrossRef] [Green Version]
  83. Sørensen, M.F.; Heimisdóttir, S.B.; Sørensen, M.D.; Mellegaard, C.S.; Wohlleben, H.; Kristensen, B.W.; Beier, C.P. High Expression of Cystine–Glutamate Antiporter XCT (SLC7A11) Is an Independent Biomarker for Epileptic Seizures at Diagnosis in Glioma. J. Neurooncol. 2018, 138, 49–53. [Google Scholar] [CrossRef] [PubMed]
  84. Buccoliero, A.M.; Caporalini, C.; Scagnet, M.; Mussa, F.; Giordano, F.; Sardi, I.; Migliastro, I.; Moscardi, S.; Conti, V.; Barba, C.; et al. Angiocentric Glioma-Associated Seizures: The Possible Role of EATT2, Pyruvate Carboxylase and Glutamine Synthetase. Seizure 2021, 86, 152–154. [Google Scholar] [CrossRef] [PubMed]
  85. Tönjes, M.; Barbus, S.; Park, Y.J.; Wang, W.; Schlotter, M.; Lindroth, A.M.; Pleier, S.V.; Bai, A.H.C.; Karra, D.; Piro, R.M.; et al. BCAT1 Promotes Cell Proliferation through Amino Acid Catabolism in Gliomas Carrying Wild-Type IDH1. Nat. Med. 2013, 19, 901–908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Buckingham, S.C.; Campbell, S.L.; Haas, B.R.; Montana, V.; Robel, S.; Ogunrinu, T.; Sontheimer, H. Glutamate Release by Primary Brain Tumors Induces Epileptic Activity. Nat. Med. 2011, 17, 1269–1274. [Google Scholar] [CrossRef] [PubMed]
  87. Rzeski, W.; Ikonomidou, C.; Turski, L. Glutamate Antagonists Limit Tumor Growth. Biochem. Pharmacol. 2002, 64, 1195–1200. [Google Scholar] [CrossRef]
  88. Schenk, U.; Menna, E.; Kim, T.; Passafaro, M.; Chang, S.; de Camilli, P.; Matteoli, M. A Novel Pathway for Presynaptic Mitogen-Activated Kinase Activation via AMPA Receptors. J. Neurosci. 2005, 25, 1654–1663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Ishiuchi, S.; Yoshida, Y.; Sugawara, K.; Aihara, M.; Ohtani, T.; Watanabe, T.; Saito, N.; Tsuzuki, K.; Okado, H.; Miwa, A.; et al. Ca2+-Permeable AMPA Receptors Regulate Growth of Human Glioblastoma via Akt Activation. J. Neurosci. 2007, 27, 7987–8001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Wright, A.; Vissel, B. The Essential Role of AMPA Receptor GluR2 Subunit RNA Editing in the Normal and Diseased Brain. Front. Mol. Neurosci. 2012, 5, 34. [Google Scholar] [CrossRef] [Green Version]
  91. Maas, S.; Patt, S.; Schrey, M.; Rich, A. Underediting of Glutamate Receptor GluR-B MRNA in Malignant Gliomas. Proc. Natl. Acad. Sci. USA 2001, 98, 14687–14692. [Google Scholar] [CrossRef] [Green Version]
  92. Cesarini, V.; Silvestris, D.A.; Galeano, F.; Tassinari, V.; Martini, M.; Locatelli, F.; Gallo, A. ADAR2 Protein Is Associated with Overall Survival in GBM Patients and Its Decrease Triggers the Anchorage-Independent Cell Growth Signature. Biomolecules 2022, 12, 1142. [Google Scholar] [CrossRef]
  93. Colman, H.; Zhang, L.; Sulman, E.P.; McDonald, J.M.; Shooshtari, N.L.; Rivera, A.; Popoff, S.; Nutt, C.L.; Louis, D.N.; Cairncross, J.G.; et al. A Multigene Predictor of Outcome in Glioblastoma. Neuro. Oncol. 2010, 12, 49–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Venkatesh, H.S.; Morishita, W.; Geraghty, A.C.; Silverbush, D.; Gillespie, S.M.; Arzt, M.; Tam, L.T.; Espenel, C.; Ponnuswami, A.; Ni, L.; et al. Electrical and Synaptic Integration of Glioma into Neural Circuits. Nature 2019, 573, 539–545. [Google Scholar] [CrossRef] [PubMed]
  95. Venkataramani, V.; Tanev, D.I.; Strahle, C.; Studier-Fischer, A.; Fankhauser, L.; Kessler, T.; Körber, C.; Kardorff, M.; Ratliff, M.; Xie, R.; et al. Glutamatergic Synaptic Input to Glioma Cells Drives Brain Tumour Progression. Nature 2019, 573, 532–538. [Google Scholar] [CrossRef] [PubMed]
  96. Shahbazi, F.; Grandi, V.; Banerjee, A.; Trant, J.F. Cannabinoids and Cannabinoid Receptors: The Story so Far. iScience 2020, 23, 101301. [Google Scholar] [CrossRef] [PubMed]
  97. Cristino, L.; Bisogno, T.; Di Marzo, V. Cannabinoids and the Expanded Endocannabinoid System in Neurological Disorders. Nat. Rev. Neurol. 2020, 16, 9–29. [Google Scholar] [CrossRef]
  98. Schlicker, E.; Kathmann, M. Modulation of Transmitter Release via Presynaptic Cannabinoid Receptors. Trends Pharm. Sci. 2001, 22, 565–572. [Google Scholar] [CrossRef]
  99. Marsicano, G.; Goodenough, S.; Monory, K.; Hermann, H.; Eder, M.; Cannich, A.; Azad, S.C.; Cascio, M.G.; Ortega-Gutiérrez, S.; van der Stelt, M.; et al. CB1 Cannabinoid Receptors and On-Demand Defense against Excitotoxicity. Science 2003, 302, 84–88. [Google Scholar] [CrossRef] [Green Version]
  100. Naidoo, V.; Karanian, D.A.; Vadivel, S.K.; Locklear, J.R.; Wood, J.A.T.; Nasr, M.; Quizon, P.M.P.; Graves, E.E.; Shukla, V.; Makriyannis, A.; et al. Equipotent Inhibition of Fatty Acid Amide Hydrolase and Monoacylglycerol Lipase—Dual Targets of the Endocannabinoid System to Protect against Seizure Pathology. Neurotherapeutics 2012, 9, 801–813. [Google Scholar] [CrossRef] [Green Version]
  101. Monory, K.; Massa, F.; Egertová, M.; Eder, M.; Blaudzun, H.; Westenbroek, R.; Kelsch, W.; Jacob, W.; Marsch, R.; Ekker, M.; et al. The Endocannabinoid System Controls Key Epileptogenic Circuits in the Hippocampus. Neuron 2006, 51, 455–466. [Google Scholar] [CrossRef] [Green Version]
  102. Ludányi, A.; Eross, L.; Czirják, S.; Vajda, J.; Halász, P.; Watanabe, M.; Palkovits, M.; Maglóczky, Z.; Freund, T.F.; Katona, I. Downregulation of the CB1 Cannabinoid Receptor and Related Molecular Elements of the Endocannabinoid System in Epileptic Human Hippocampus. J. Neurosci. 2008, 28, 2976–2990. [Google Scholar] [CrossRef] [Green Version]
  103. Romigi, A.; Bari, M.; Placidi, F.; Grazia Marciani, M.; Malaponti, M.; Torelli, F.; Izzi, F.; Prosperetti, C.; Zannino, S.; Corte, F.; et al. Cerebrospinal Fluid Levels of the Endocannabinoid Anandamide Are Reduced in Patients with Untreated Newly Diagnosed Temporal Lobe Epilepsy. Epilepsia 2010, 51, 768–772. [Google Scholar] [CrossRef] [PubMed]
  104. Maccarrone, M.; Attinà, M.; Cartoni, A.; Bari, M.; Finazzi-Agrò, A. Gas Chromatography-Mass Spectrometry Analysis of Endogenous Cannabinoids in Healthy and Tumoral Human Brain and Human Cells in Culture. J. Neurochem. 2001, 76, 594–601. [Google Scholar] [CrossRef] [PubMed]
  105. Wu, X.; Han, L.; Zhang, X.; Li, L.; Jiang, C.; Qiu, Y.; Huang, R.; Xie, B.; Lin, Z.; Ren, J.; et al. Alteration of Endocannabinoid System in Human Gliomas. J. Neurochem. 2012, 120, 842–849. [Google Scholar] [CrossRef] [PubMed]
  106. Petersen, G.; Moesgaard, B.; Schmid, P.C.; Schmid, H.H.O.; Broholm, H.; Kosteljanetz, M.; Hansen, H.S. Endocannabinoid Metabolism in Human Glioblastomas and Meningiomas Compared to Human Non-Tumour Brain Tissue. J. Neurochem. 2005, 93, 299–309. [Google Scholar] [CrossRef]
  107. De Jesús, M.L.; Hostalot, C.; Garibi, J.M.; Sallés, J.; Meana, J.J.; Callado, L.F. Opposite Changes in Cannabinoid CB1 and CB2 Receptor Expression in Human Gliomas. Neurochem. Int. 2010, 56, 829–833. [Google Scholar] [CrossRef] [PubMed]
  108. Ellert-Miklaszewska, A.; Ciechomska, I.A.; Kaminska, B. Cannabinoid Signaling in Glioma Cells. Adv. Exp. Med. Biol. 2020, 1202, 223–241. [Google Scholar] [CrossRef]
  109. Esposito, G.; Ligresti, A.; Izzo, A.A.; Bisogno, T.; Ruvo, M.; di Rosa, M.; di Marzo, V.; Iuvone, T. The Endocannabinoid System Protects Rat Glioma Cells against HIV-1 Tat Protein-Induced Cytotoxicity: Mechanism and Regulation. J. Biol. Chem. 2002, 277, 50348–50354. [Google Scholar] [CrossRef] [Green Version]
  110. Aguado, T.; Carracedo, A.; Julien, B.; Velasco, G.; Milman, G.; Mechoulamluis, R.; Alvarez, L.; Guzmán, M.; Galve-Roperh, I. Cannabinoids Induce Glioma Stem-like Cell Differentiation and Inhibit Gliomagenesis. J. Biol. Chem. 2007, 282, 6854–6862. [Google Scholar] [CrossRef] [Green Version]
  111. Patel, S.; Grinspoon, R.; Fleming, B.; Skirvin, L.A.; Wade, C.; Wolper, E.; Bruno, P.L.; Thiele, E.A. The Long-Term Efficacy of Cannabidiol in the Treatment of Refractory Epilepsy. Epilepsia 2021, 62, 1594–1603. [Google Scholar] [CrossRef]
  112. Ntafoulis, I.; Koolen, S.L.W.; Leenstra, S.; Lamfers, M.L.M. Drug Repurposing, a Fast-Track Approach to Develop Effective Treatments for Glioblastoma. Cancers 2022, 14, 3705. [Google Scholar] [CrossRef]
  113. Barbieri, F.; Verduci, I.; Carlini, V.; Zona, G.; Pagano, A.; Mazzanti, M.; Florio, T. Repurposed Biguanide Drugs in Glioblastoma Exert Antiproliferative Effects via the Inhibition of Intracellular Chloride Channel 1 Activity. Front. Oncol. 2019, 9, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Würth, R.; Barbieri, F.; Florio, T. New Molecules and Old Drugs as Emerging Approaches to Selectively Target Human Glioblastoma Cancer Stem Cells. Biomed. Res. Int. 2014, 126586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Lynch, B.A.; Lambeng, N.; Nocka, K.; Kensel-Hammes, P.; Bajjalieh, S.M.; Matagne, A.; Fuks, B. The Synaptic Vesicle Protein SV2A Is the Binding Site for the Antiepileptic Drug Levetiracetam. Proc. Natl. Acad. Sci. USA 2004, 101, 9861–9866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Rigo, J.M.; Hans, G.; Nguyen, L.; Rocher, V.; Belachew, S.; Malgrange, B.; Leprince, P.; Moonen, G.; Selak, I.; Matagne, A.; et al. The Anti-Epileptic Drug Levetiracetam Reverses the Inhibition by Negative Allosteric Modulators of Neuronal GABA- and Glycine-Gated Currents. Br. J. Pharmacol. 2002, 136, 659–672. [Google Scholar] [CrossRef]
  117. Marutani, A.; Nakamura, M.; Nishimura, F.; Nakazawa, T.; Matsuda, R.; Hironaka, Y.; Nakagawa, I.; Tamura, K.; Takeshima, Y.; Motoyama, Y.; et al. Tumor-Inhibition Effect of Levetiracetam in Combination with Temozolomide in Glioblastoma Cells. Neurochem. J. 2017, 11, 43–49. [Google Scholar] [CrossRef]
  118. Bobustuc, G.C.; Baker, C.H.; Limaye, A.; Jenkins, W.D.; Pearl, G.; Avgeropoulos, N.G.; Konduri, S.D. Levetiracetam Enhances P53-Mediated MGMT Inhibition and Sensitizes Glioblastoma Cells to Temozolomide. Neuro. Oncol. 2010, 12, 917–927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Scicchitano, B.M.; Sorrentino, S.; Proietti, G.; Lama, G.; Dobrowolny, G.; Catizone, A.; Binda, E.; Larocca, L.M.; Sica, G. Levetiracetam Enhances the Temozolomide Effect on Glioblastoma Stem Cell Proliferation and Apoptosis. Cancer Cell Int. 2018, 18, 136. [Google Scholar] [CrossRef]
  120. Ni, X.R.; Guo, C.C.; Yu, Y.J.; Yu, Z.H.; Cai, H.P.; Wu, W.C.; Ma, J.X.; Chen, F.R.; Wang, J.; Chen, Z.P. Combination of Levetiracetam and IFN-α Increased Temozolomide Efficacy in MGMT-Positive Glioma. Cancer Chemother. Pharmacol. 2020, 86, 773–782. [Google Scholar] [CrossRef] [PubMed]
  121. Kim, Y.H.; Kim, T.; Joo, J.D.; Han, J.H.; Kim, Y.J.; Kim, I.A.; Yun, C.H.; Kim, C.Y. Survival Benefit of Levetiracetam in Patients Treated with Concomitant Chemoradiotherapy and Adjuvant Chemotherapy with Temozolomide for Glioblastoma Multiforme. Cancer 2015, 121, 2926–2932. [Google Scholar] [CrossRef]
  122. Roh, T.H.; Moon, J.H.; Park, H.H.; Kim, E.H.; Hong, C.K.; Kim, S.H.; Kang, S.G.; Chang, J.H. Association between Survival and Levetiracetam Use in Glioblastoma Patients Treated with Temozolomide Chemoradiotherapy. Sci. Rep. 2020, 10, 10783. [Google Scholar] [CrossRef]
  123. Cardona, A.F.; Rojas, L.; Wills, B.; Bernal, L.; Ruiz-Patiño, A.; Arrieta, O.; Hakim, E.J.; Hakim, F.; Mejía, J.A.; Useche, N.; et al. Efficacy and Safety of Levetiracetam vs. Other Antiepileptic Drugs in Hispanic Patients with Glioblastoma. J. Neurooncol. 2018, 136, 363–371. [Google Scholar] [CrossRef]
  124. Happold, C.; Gorlia, T.; Chinot, O.; Gilbert, M.R.; Nabors, L.B.; Wick, W.; Pugh, S.L.; Hegi, M.; Cloughesy, T.; Roth, P.; et al. Does Valproic Acid or Levetiracetam Improve Survival in Glioblastoma? A Pooled Analysis of Prospective Clinical Trials in Newly Diagnosed Glioblastoma. J. Clin. Oncol. 2016, 34, 731–739. [Google Scholar] [CrossRef]
  125. Pallud, J.; Huberfeld, G.; Dezamis, E.; Peeters, S.; Moiraghi, A.; Gavaret, M.; Guinard, E.; Dhermain, F.; Varlet, P.; Oppenheim, C.; et al. Effect of Levetiracetam Use Duration on Overall Survival of Isocitrate Dehydrogenase Wild-Type Glioblastoma in Adults: An Observational Study. Neurology 2022, 98, E125–E140. [Google Scholar] [CrossRef]
  126. Hwang, K.; Kim, J.; Kang, S.G.; Jung, T.Y.; Kim, J.H.; Kim, S.H.; Kang, S.H.; Hong, Y.K.; Kim, T.M.; Kim, Y.J.; et al. Levetiracetam as a Sensitizer of Concurrent Chemoradiotherapy in Newly Diagnosed Glioblastoma: An Open-Label Phase 2 Study. Cancer Med. 2022, 11, 371–379. [Google Scholar] [CrossRef] [PubMed]
  127. Sun, M.; Huang, N.; Tao, Y.; Wen, R.; Zhao, G.; Zhang, X.; Xie, Z.; Cheng, Y.; Mao, J.; Liu, G. The Efficacy of Temozolomide Combined with Levetiracetam for Glioblastoma (GBM) after Surgery: A Study Protocol for a Double-Blinded and Randomized Controlled Trial. Trials 2022, 23, 234. [Google Scholar] [CrossRef] [PubMed]
  128. Romoli, M.; Mazzocchetti, P.; D’Alonzo, R.; Siliquini, S.; Rinaldi, V.E.; Verrotti, A.; Calabresi, P.; Costa, C. Valproic Acid and Epilepsy: From Molecular Mechanisms to Clinical Evidences. Curr. Neuropharmacol. 2019, 17, 926–946. [Google Scholar] [CrossRef] [PubMed]
  129. Duenas-Gonzalez, A.; Candelaria, M.; Perez-Plascencia, C.; Perez-Cardenas, E.; de la Cruz-Hernandez, E.; Herrera, L.A. Valproic Acid as Epigenetic Cancer Drug: Preclinical, Clinical and Transcriptional Effects on Solid Tumors. Cancer Treat. Rev. 2008, 34, 206–222. [Google Scholar] [CrossRef] [PubMed]
  130. Phiel, C.J.; Zhang, F.; Huang, E.Y.; Guenther, M.G.; Lazar, M.A.; Klein, P.S. Histone Deacetylase Is a Direct Target of Valproic Acid, a Potent Anticonvulsant, Mood Stabilizer, and Teratogen. J. Biol. Chem. 2001, 276, 36734–36741. [Google Scholar] [CrossRef] [Green Version]
  131. Barciszewska, A.M.; Belter, A.; Gawrońska, I.; Giel-Pietraszuk, M.; Naskręt-Barciszewska, M.Z. Cross-Reactivity between Histone Demethylase Inhibitor Valproic Acid and DNA Methylation in Glioblastoma Cell Lines. Front. Oncol. 2022, 12, 1033035. [Google Scholar] [CrossRef]
  132. Castro, L.M.R.; Gallant, M.; Niles, L.P. Novel Targets for Valproic Acid: Up-Regulation of Melatonin Receptors and Neurotrophic Factors in C6 Glioma Cells. J. Neurochem. 2005, 95, 1227–1236. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, C.; Liu, S.; Yuan, X.; Hu, Z.; Li, H.; Wu, M.; Yuan, J.; Zhao, Z.; Su, J.; Wang, X.; et al. Valproic Acid Promotes Human Glioma U87 Cells Apoptosis and Inhibits Glycogen Synthase Kinase-3β Through ERK/Akt Signaling. Cell. Physiol. Biochem. 2016, 39, 2173–2185. [Google Scholar] [CrossRef]
  134. Han, W.; Yu, F.; Cao, J.; Dong, B.; Guan, W.; Shi, J. Valproic Acid Enhanced Apoptosis by Promoting Autophagy Via Akt/MTOR Signaling in Glioma. Cell Transpl. 2020, 29. [Google Scholar] [CrossRef]
  135. Riva, G.; Cilibrasi, C.; Bazzoni, R.; Cadamuro, M.; Negroni, C.; Butta, V.; Strazzabosco, M.; Dalprà, L.; Lavitrano, M.; Bentivegna, A. Valproic Acid Inhibits Proliferation and Reduces Invasiveness in Glioma Stem Cells Through Wnt/β Catenin Signalling Activation. Genes 2018, 9, 522. [Google Scholar] [CrossRef] [Green Version]
  136. Fu, J.; Shao, C.J.; Chen, F.R.; Ng, H.K.; Chen, Z.P. Autophagy Induced by Valproic Acid Is Associated with Oxidative Stress in Glioma Cell Lines. Neuro. Oncol. 2010, 12, 328–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Chen, Y.; Tsai, Y.H.; Tseng, S.H. Valproic Acid Affected the Survival and Invasiveness of Human Glioma Cells through Diverse Mechanisms. J. Neurooncol. 2012, 109, 23–33. [Google Scholar] [CrossRef] [PubMed]
  138. Lee, C.Y.; Lai, H.Y.; Chiu, A.; Chan, S.H.; Hsiao, L.P.; Lee, S.T. The Effects of Antiepileptic Drugs on the Growth of Glioblastoma Cell Lines. J. Neurooncol. 2016, 127, 445–453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Lange, F.; Weßlau, K.; Porath, K.; Hörnschemeyer, J.; Bergner, C.; Krause, B.J.; Mullins, C.S.; Linnebacher, M.; Köhling, R.; Kirschstein, T. AMPA Receptor Antagonist Perampanel Affects Glioblastoma Cell Growth and Glutamate Release in Vitro. PLoS ONE 2019, 14, e0211644. [Google Scholar] [CrossRef] [Green Version]
  140. Ciusani, E.; Balzarotti, M.; Calatozzolo, C.; De Grazia, U.; Boiardi, A.; Salmaggi, A.; Croci, D. Valproic Acid Increases the in Vitro Effects of Nitrosureas on Human Glioma Cell Lines. Oncol. Res. 2007, 16, 453–463. [Google Scholar] [CrossRef] [PubMed]
  141. Li, Z.; Xia, Y.; Bu, X.; Yang, D.; Yuan, Y.; Guo, X.; Zhang, G.; Wang, Z.; Jiao, J. Effects of Valproic Acid on the Susceptibility of Human Glioma Stem Cells for TMZ and ACNU. Oncol. Lett. 2018, 15, 9877–9883. [Google Scholar] [CrossRef]
  142. Ryu, C.H.; Yoon, W.S.; Park, K.Y.; Kim, S.M.; Lim, J.Y.; Woo, J.S.; Jeong, C.H.; Hou, Y.; Jeun, S.S. Valproic Acid Downregulates the Expression of MGMT and Sensitizes Temozolomide-Resistant Glioma Cells. J. Biomed. Biotechnol. 2012, 987495. [Google Scholar] [CrossRef] [Green Version]
  143. Van Nifterik, K.A.; Van Den Berg, J.; Slotman, B.J.; Lafleur, M.V.M.; Sminia, P.; Stalpers, L.J.A. Valproic Acid Sensitizes Human Glioma Cells for Temozolomide and γ-Radiation. J. Neurooncol. 2012, 107, 61–67. [Google Scholar] [CrossRef] [PubMed]
  144. Chang, C.Y.; Li, J.R.; Wu, C.C.; Ou, Y.C.; Chen, W.Y.; Kuan, Y.H.; Wang, W.Y.; Chen, C.J. Valproic Acid Sensitizes Human Glioma Cells to Gefitinib-Induced Autophagy. IUBMB Life 2015, 67, 869–879. [Google Scholar] [CrossRef] [Green Version]
  145. Das, C.M.; Aguilera, D.; Vasquez, H.; Prasad, P.; Zhang, M.; Wolff, J.E.; Gopalakrishnan, V. Valproic Acid Induces P21 and Topoisomerase-II (Alpha/Beta) Expression and Synergistically Enhances Etoposide Cytotoxicity in Human Glioblastoma Cell Lines. J. Neurooncol. 2007, 85, 159–170. [Google Scholar] [CrossRef] [PubMed]
  146. Zhou, Y.; Xu, Y.; Wang, H.; Niu, J.; Hou, H.; Jiang, Y. Histone Deacetylase Inhibitor, Valproic Acid, Radiosensitizes the C6 Glioma Cell Line in Vitro. Oncol. Lett. 2014, 7, 203–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Kuo, Y.J.; Yang, Y.H.; Lee, I.Y.; Chen, P.C.; Yang, J.T.; Wang, T.C.; Lin, M.H.C.; Yang, W.H.; Cheng, C.Y.; Chen, K.T.; et al. Effect of Valproic Acid on Overall Survival in Patients with High-Grade Gliomas Undergoing Temozolomide: A Nationwide Population-Based Cohort Study in Taiwan. Medicine 2020, 99, e21147. [Google Scholar] [CrossRef]
  148. Redjal, N.; Reinshagen, C.; Le, A.; Walcott, B.P.; McDonnell, E.; Dietrich, J.; Nahed, B.V. Valproic Acid, Compared to Other Antiepileptic Drugs, Is Associated with Improved Overall and Progression-Free Survival in Glioblastoma but Worse Outcome in Grade II/III Gliomas Treated with Temozolomide. J. Neurooncol. 2016, 127, 505–514. [Google Scholar] [CrossRef]
  149. Kerkhof, M.; Dielemans, J.C.M.; Van Breemen, M.S.; Zwinkels, H.; Walchenbach, R.; Taphoorn, M.J.; Vecht, C.J. Effect of Valproic Acid on Seizure Control and on Survival in Patients with Glioblastoma Multiforme. Neuro. Oncol. 2013, 15, 961–967. [Google Scholar] [CrossRef] [Green Version]
  150. Weller, M.; Gorlia, T.; Cairncross, J.G.; van den Bent, M.J.; Mason, W.; Belanger, K.; Brandes, A.A.; Bogdahn, U.; Macdonald, D.R.; Forsyth, P.; et al. Prolonged Survival with Valproic Acid Use in the EORTC/NCIC Temozolomide Trial for Glioblastoma. Neurology 2011, 77, 1156–1164. [Google Scholar] [CrossRef] [Green Version]
  151. Watanabe, S.; Kuwabara, Y.; Suehiro, S.; Yamashita, D.; Tanaka, M.; Tanaka, A.; Ohue, S.; Araki, H. Valproic Acid Reduces Hair Loss and Improves Survival in Patients Receiving Temozolomide-Based Radiation Therapy for High-Grade Glioma. Eur. J. Clin. Pharmacol. 2017, 73, 357–363. [Google Scholar] [CrossRef]
  152. Krauze, A.V.; Megan, M.; Theresa, C.Z.; Peter, M.; Shih, J.H.; Tofilon, P.J.; Rowe, L.; Gilbert, M.; Camphausen, K. The Addition of Valproic Acid to Concurrent Radiation Therapy and Temozolomide Improves Patient Outcome: A Correlative Analysis of RTOG 0525, SEER and a Phase II NCI Trial. Cancer Stud. Ther. 2020, 5, 1–15. [Google Scholar] [CrossRef]
  153. Yuan, Y.; Xiang, W.; Qing, M.; Yanhui, L.; Jiewen, L.; Yunhe, M. Survival Analysis for Valproic Acid Use in Adult Glioblastoma Multiforme: A Meta-Analysis of Individual Patient Data and a Systematic Review. Seizure 2014, 23, 830–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Langan, Y.M.; Lucas, R.; Jewell, H.; Toublanc, N.; Schaefer, H.; Sander, J.W.A.S.; Patsalos, P.N. Talampanel, a New Antiepileptic Drug: Single- and Multiple-Dose Pharmacokinetics and Initial 1-Week Experience in Patients with Chronic Intractable Epilepsy. Epilepsia 2003, 44, 46–53. [Google Scholar] [CrossRef] [PubMed]
  155. Howes, J.F.; Bell, C. Talampanel. Neurotherapeutics 2007, 4, 126–129. [Google Scholar] [CrossRef] [Green Version]
  156. Iwamoto, F.M.; Kreisl, T.N.; Kim, L.; Duic, J.P.; Butman, J.A.; Albert, P.S.; Fine, H.A. Phase 2 Trial of Talampanel, a Glutamate Receptor Inhibitor, for Adults with Recurrent Malignant Gliomas. Cancer 2010, 116, 1776–1782. [Google Scholar] [CrossRef] [PubMed]
  157. Grossman, S.A.; Ye, X.; Chamberlain, M.; Mikkelsen, T.; Batchelor, T.; Desideri, S.; Piantadosi, S.; Fisher, J.; Fine, H.A. Talampanel with Standard Radiation and Temozolomide in Patients with Newly Diagnosed Glioblastoma: A Multicenter Phase II Trial. J. Clin. Oncol. 2009, 27, 4155–4161. [Google Scholar] [CrossRef] [Green Version]
  158. Yagi, C.; Tatsuoka, J.; Sano, E.; Hanashima, Y.; Ozawa, Y.; Yoshimura, S.; Yamamuro, S.; Sumi, K.; Hara, H.; Katayama, Y.; et al. Anti-tumor Effects of Anti-epileptic Drugs in Malignant Glioma Cells. Oncol. Rep. 2022, 48, 216. [Google Scholar] [CrossRef]
  159. Salmaggi, A.; Corno, C.; Maschio, M.; Donzelli, S.; D’urso, A.; Perego, P.; Ciusani, E. Synergistic Effect of Perampanel and Temozolomide in Human Glioma Cell Lines. J. Pers. Med. 2021, 11, 390. [Google Scholar] [CrossRef]
  160. Tatsuoka, J.; Sano, E.; Hanashima, Y.; Yagi, C.; Yamamuro, S.; Sumi, K.; Hara, H.; Takada, K.; Kanemaru, K.; Komine-Aizawa, S.; et al. Anti-Tumor Effects of Perampanel in Malignant Glioma Cells. Oncol. Lett. 2022, 24, 1–9. [Google Scholar] [CrossRef]
  161. Mayer, J.; Kirschstein, T.; Resch, T.; Porath, K.; Krause, B.J.; Köhling, R.; Lange, F. Perampanel Attenuates Epileptiform Phenotype in C6 Glioma. Neurosci. Lett. 2020, 715, 134629. [Google Scholar] [CrossRef]
  162. Lange, F.; Hartung, J.; Liebelt, C.; Boisserée, J.; Resch, T.; Porath, K.; Hörnschemeyer, J.; Reichart, G.; Sellmann, T.; Neubert, V.; et al. Perampanel Add-on to Standard Radiochemotherapy in Vivo Promotes Neuroprotection in a Rodent F98 Glioma Model. Front. Neurosci. 2020, 14, 598266. [Google Scholar] [CrossRef]
  163. Izumoto, S.; Miyauchi, M.; Tasaki, T.; Okuda, T.; Nakagawa, N.; Nakano, N.; Kato, A.; Fujita, M. Seizures and Tumor Progression in Glioma Patients with Uncontrollable Epilepsy Treated with Perampanel. Anticancer Res. 2018, 38, 4361–4366. [Google Scholar] [CrossRef] [PubMed]
  164. Gherzi, M.; Milano, G.; Fucile, C.; Calevo, M.G.; Mancardi, M.M.; Nobili, L.; Astuni, P.; Marini, V.; Barco, S.; Cangemi, G.; et al. Safety and Pharmacokinetics of Medical Cannabis Preparation in a Monocentric Series of Young Patients with Drug Resistant Epilepsy. Complement. Ther. Med. 2020, 51, 102402. [Google Scholar] [CrossRef] [PubMed]
  165. Gray, R.A.; Whalley, B.J. The Proposed Mechanisms of Action of CBD in Epilepsy. Epileptic Disord. 2020, 22, 10–15. [Google Scholar] [CrossRef]
  166. Vaccani, A.; Massi, P.; Colombo, A.; Rubino, T.; Parolaro, D. Cannabidiol Inhibits Human Glioma Cell Migration through a Cannabinoid Receptor-Independent Mechanism. Br. J. Pharmacol. 2005, 144, 1032–1036. [Google Scholar] [CrossRef] [Green Version]
  167. López-Valero, I.; Saiz-Ladera, C.; Torres, S.; Hernández-Tiedra, S.; García-Taboada, E.; Rodríguez-Fornés, F.; Barba, M.; Dávila, D.; Salvador-Tormo, N.; Guzmán, M.; et al. Targeting Glioma Initiating Cells with A Combined Therapy of Cannabinoids and Temozolomide. Biochem. Pharmacol. 2018, 157, 266–274. [Google Scholar] [CrossRef] [PubMed]
  168. Volmar, M.N.M.; Cheng, J.; Alenezi, H.; Richter, S.; Haug, A.; Hassan, Z.; Goldberg, M.; Li, Y.; Hou, M.; Herold-Mende, C.; et al. Cannabidiol Converts NF-ΚB into a Tumor Suppressor in Glioblastoma with Defined Antioxidative Properties. Neuro. Oncol. 2021, 23, 1898. [Google Scholar] [CrossRef] [PubMed]
  169. Likar, R.; Koestenberger, M.; Stutschnig, M.; Nahler, G. Cannabidiol Μay Prolong Survival in Patients with Glioblastoma Multiforme. Cancer Diagn. Progn. 2021, 1, 77–82. [Google Scholar] [CrossRef] [PubMed]
  170. Twelves, C.; Sabel, M.; Checketts, D.; Miller, S.; Tayo, B.; Jove, M.; Brazil, L.; Short, S.C.; McBain, C.; Haylock, B.; et al. A Phase 1b Randomised, Placebo-Controlled Trial of Nabiximols Cannabinoid Oromucosal Spray with Temozolomide in Patients with Recurrent Glioblastoma. Br. J. Cancer 2021, 124, 1379–1387. [Google Scholar] [CrossRef]
  171. Doello, K.; Mesas, C.; Quiñonero, F.; Rama, A.R.; Vélez, C.; Perazzoli, G.; Ortiz, R. Antitumor Effect of Traditional Drugs for Neurological Disorders: Preliminary Studies in Neural Tumor Cell Lines. Neurotox. Res. 2022, 40, 1645–1652. [Google Scholar] [CrossRef]
  172. Rizzo, A.; Donzelli, S.; Girgenti, V.; Sacconi, A.; Vasco, C.; Salmaggi, A.; Blandino, G.; Maschio, M.; Ciusani, E. In Vitro Antineoplastic Effects of Brivaracetam and Lacosamide on Human Glioma Cells. J. Exp. Clin. Cancer Res. 2017, 36, 76. [Google Scholar] [CrossRef] [Green Version]
  173. Bang, S.R.; Ambavade, S.D.; Jagdale, P.G.; Adkar, P.P.; Waghmare, A.B.; Ambavade, P.D. Lacosamide Reduces HDAC Levels in the Brain and Improves Memory: Potential for Treatment of Alzheimer’s Disease. Pharm. Biochem. Behav. 2015, 134, 65–69. [Google Scholar] [CrossRef] [PubMed]
  174. Moutal, A.; Villa, L.S.; Yeon, S.K.; Householder, K.T.; Park, K.D.; Sirianni, R.W.; Khanna, R. CRMP2 Phosphorylation Drives Glioblastoma Cell Proliferation. Mol. Neurobiol. 2018, 55, 4403–4416. [Google Scholar] [CrossRef] [PubMed]
  175. Pellegrino, M.; Rizza, P.; Nigro, A.; Ceraldi, R.; Ricci, E.; Perrotta, I.; Aquila, S.; Lanzino, M.; Andò, S.; Morelli, C.; et al. FoxO3a Mediates the Inhibitory Effects of the Antiepileptic Drug Lamotrigine on Breast Cancer Growth. Mol. Cancer Res. 2018, 16, 923–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Yamaguchi, K.; Kumakura, S.; Someya, A.; Iseki, M.; Inada, E.; Nagaoka, I. Anti-inflammatory Actions of Gabapentin and Pregabalin on the Substance P-induced Mitogen-activated Protein Kinase Activation in U373 MG Human Glioblastoma Astrocytoma Cells. Mol. Med. Rep. 2017, 16, 6109–6115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Frampton, J.E. Stiripentol: A Review in Dravet Syndrome. Drugs 2019, 79, 1785–1796. [Google Scholar] [CrossRef]
  178. Yadav, A.; Alnakhli, A.; Vemana, H.P.; Bhutkar, S.; Muth, A.; Dukhande, V.V. Repurposing an Antiepileptic Drug for the Treatment of Glioblastoma. Pharm. Res. 2022, 39, 2871–2883. [Google Scholar] [CrossRef]
  179. Guyon, J.; Fernandez-Moncada, I.; Larrieu, C.M.; Bouchez, C.L.; Pagano Zottola, A.C.; Galvis, J.; Chouleur, T.; Burban, A.; Joseph, K.; Ravi, V.M.; et al. Lactate Dehydrogenases Promote Glioblastoma Growth and Invasion via a Metabolic Symbiosis. EMBO Mol. Med. 2022, 14, e15343. [Google Scholar] [CrossRef]
  180. Sada, N.; Lee, S.; Katsu, T.; Otsuki, T.; Inoue, T. Epilepsy Treatment. Targeting LDH Enzymes with a Stiripentol Analog to Treat Epilepsy. Science 2015, 347, 1362–1367. [Google Scholar] [CrossRef]
  181. Colen, C.B.; Shen, Y.; Ghoddoussi, F.; Yu, P.; Francis, T.B.; Koch, B.J.; Monterey, M.D.; Galloway, M.P.; Sloan, A.E.; Mathupala, S.P. Metabolic Targeting of Lactate Efflux by Malignant Glioma Inhibits Invasiveness and Induces Necrosis: An in Vivo Study. Neoplasia 2011, 13, 620–632. [Google Scholar] [CrossRef] [Green Version]
  182. Ryu, J.Y.; Min, K.L.; Chang, M.J. Effect of Anti-Epileptic Drugs on the Survival of Patients with Glioblastoma Multiforme: A Retrospective, Single-Center Study. PLoS ONE 2019, 14, e0225599. [Google Scholar] [CrossRef] [Green Version]
  183. Krauze, A.V.; Myrehaug, S.D.; Chang, M.G.; Holdford, D.J.; Smith, S.; Shih, J.; Tofilon, P.J.; Fine, H.A.; Camphausen, K. A Phase 2 Study of Concurrent Radiation Therapy, Temozolomide, and the Histone Deacetylase Inhibitor Valproic Acid for Patients with Glioblastoma. Int. J. Radiat. Oncol. Biol. Phys. 2015, 92, 986–992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Neuro-glioma interplay: role in glioblastoma (GBM) progression and in GBM-related epilepsy. (Details of the molecular processes are described in the text). The schematic diagram depicts the interactions between neurons and GBM cells, leading to seizure onset and GBM growth, focusing on the role of neurotransmitters (GABA and glutamate) imbalance. In neurons of epileptogenic tissue, upregulation of NKCC1 (Na–K–2Cl cotransporter (1), and the concomitant low expression of KCC2 (K–2Cl symporter (2), which normally control the low intracellular Cl concentration required for GABAAR-mediated inhibition, is significantly decreased. Consequently, intracellular Cl concentration increases and triggers a depolarizing GABAergic excitatory response firing seizure onset. In parallel, GBM shows perturbation of Cl homeostasis accumulating high concentrations of Cl. The neuro-gliomal synapses lead to interactions mediated by glutamate receptors (AMPAR) and overlapping mechanisms involved in both oncogenesis and epileptogenesis. GBM cell express calcium-permeable AMPAR, which are stimulated by glutamate in an autocrine (GBM cells release a high amount of glutamate) and paracrine (neuronal) manner promoting GBM cell proliferation and invasion. This mechanism is mainly mediated by high upregulation of the cystine/glutamate antiporter SLC7A11 (Solute Carrier Family 7 Member 11) on the GBM cell synaptic membrane, which associates with seizure onset.
Figure 1. Neuro-glioma interplay: role in glioblastoma (GBM) progression and in GBM-related epilepsy. (Details of the molecular processes are described in the text). The schematic diagram depicts the interactions between neurons and GBM cells, leading to seizure onset and GBM growth, focusing on the role of neurotransmitters (GABA and glutamate) imbalance. In neurons of epileptogenic tissue, upregulation of NKCC1 (Na–K–2Cl cotransporter (1), and the concomitant low expression of KCC2 (K–2Cl symporter (2), which normally control the low intracellular Cl concentration required for GABAAR-mediated inhibition, is significantly decreased. Consequently, intracellular Cl concentration increases and triggers a depolarizing GABAergic excitatory response firing seizure onset. In parallel, GBM shows perturbation of Cl homeostasis accumulating high concentrations of Cl. The neuro-gliomal synapses lead to interactions mediated by glutamate receptors (AMPAR) and overlapping mechanisms involved in both oncogenesis and epileptogenesis. GBM cell express calcium-permeable AMPAR, which are stimulated by glutamate in an autocrine (GBM cells release a high amount of glutamate) and paracrine (neuronal) manner promoting GBM cell proliferation and invasion. This mechanism is mainly mediated by high upregulation of the cystine/glutamate antiporter SLC7A11 (Solute Carrier Family 7 Member 11) on the GBM cell synaptic membrane, which associates with seizure onset.
Biomedicines 11 00582 g001
Figure 2. Main glutamatergic signaling in glioblastoma cells and peritumoral neurons leads to tumor progression and epileptic onset. SLC7A11 (cystine/glutamate antiporter solute carrier family 7 member 11) is often upregulated in glioblastoma (GBM) and associated with seizures. GBM cells release glutamate via SLC7A11 in exchange for cysteine (precursor glutathione which mitigates oxidative stress and sustains GBM cell survival) uptake. The abundant release of glutamate concurs to enhance the excitotoxicity of GBM surrounding neurons and contributes to tumor growth in a paracrine manner. The upregulated glutamatergic activity in neurons stimulates glioma growth and signaling via ionotropic glutamate AMPA and NMDA receptors (AMPAR and NMDAR), which are also involved in seizure onset. GluR2 subunit-lacking AMPARs are highly Ca2+ permeable leading to increased Ca2+ influx in GBM cells that (a) enhances glutamate release to surrounding cells, (b) triggers growth- and invasion-related pathways in GBM cells, (c) impacts neuro-gliomal glutamatergic synapses. NMDARs are highly expressed in synapses but also in glia and GBM cells. NMDARs concur to increase intracellular Ca2+ levels via activation of mGluRs glutamatergic transmission in both neurons and glia, and glioma. Metabotropic glutamate receptors (mGluRs) are G-protein-coupled receptors that affect ion exchange when activated by glutamate and regulate major pro-survival and proliferative signaling pathways such as MEK/ERK (mitogen-activated protein kinase kinase 1, MEK; extracellular signal-regulated kinase ERK), and PI3K/Akt (phosphatidylinositol-3 kinase, PI3K; serine/threonine kinase, Akt).
Figure 2. Main glutamatergic signaling in glioblastoma cells and peritumoral neurons leads to tumor progression and epileptic onset. SLC7A11 (cystine/glutamate antiporter solute carrier family 7 member 11) is often upregulated in glioblastoma (GBM) and associated with seizures. GBM cells release glutamate via SLC7A11 in exchange for cysteine (precursor glutathione which mitigates oxidative stress and sustains GBM cell survival) uptake. The abundant release of glutamate concurs to enhance the excitotoxicity of GBM surrounding neurons and contributes to tumor growth in a paracrine manner. The upregulated glutamatergic activity in neurons stimulates glioma growth and signaling via ionotropic glutamate AMPA and NMDA receptors (AMPAR and NMDAR), which are also involved in seizure onset. GluR2 subunit-lacking AMPARs are highly Ca2+ permeable leading to increased Ca2+ influx in GBM cells that (a) enhances glutamate release to surrounding cells, (b) triggers growth- and invasion-related pathways in GBM cells, (c) impacts neuro-gliomal glutamatergic synapses. NMDARs are highly expressed in synapses but also in glia and GBM cells. NMDARs concur to increase intracellular Ca2+ levels via activation of mGluRs glutamatergic transmission in both neurons and glia, and glioma. Metabotropic glutamate receptors (mGluRs) are G-protein-coupled receptors that affect ion exchange when activated by glutamate and regulate major pro-survival and proliferative signaling pathways such as MEK/ERK (mitogen-activated protein kinase kinase 1, MEK; extracellular signal-regulated kinase ERK), and PI3K/Akt (phosphatidylinositol-3 kinase, PI3K; serine/threonine kinase, Akt).
Biomedicines 11 00582 g002
Figure 3. Antiepileptic drugs in human glioblastoma—a quick overview of current knowledge and future perspectives.
Figure 3. Antiepileptic drugs in human glioblastoma—a quick overview of current knowledge and future perspectives.
Biomedicines 11 00582 g003
Table 1. Suggested targets and antitumor efficacy of antiepileptic drugs in GBM.
Table 1. Suggested targets and antitumor efficacy of antiepileptic drugs in GBM.
DrugAntiepileptic Targets Antitumor
Targets
Preclinical Studies Clinical
Studies
#Clinical
Trials *
Brivaracetam Na+ channels
SV2A
GABA
MGMT [172][51] -
Cannabidiol/cannabinoids TRPV1
GPR55
ENT-1
NF-κb [166,167,168][169,170] NCT03529448 NR
NCT03607643 NR
NCT03529448 NR
NCT0181260 C
NCT01812616 C
Carbamazepine Na+ channels u.k.[158] - -
Lamotrigine Na+ channels
Ca2+ channels
PTEN
PI3K/Akt
[138] - -
LacosamideNa+ channels HDAC
CRMP2
[172,174] - -
LevetiracetamSV2A
Ca2+ channels
GABA
MGMT
HDAC
[117,118,119,120,171][121,122,123,125,126,182]NCT03048084 R
NCT00629889 C
NCT02815410
Perampanel AMPARAMPA/glutamate[139,160,159][163]NCT04650204 R
Pregabalinα2δ subunit
Ca2+ channels
p38 MAPK
NF-κB
[176] - NCT00629889 C
StiripentolGABAAR LDH[178,179] - -
TopiramateNa+ channels
GABAAR
AMPA/kainateR
Ca2+ channels
u.k.[138] - -
Valproic AcidGABA
Na+ channels
NMDAR
Ca2+ channels
HDAC
BDNF
ERK/Akt
Akt/mTOR
Wnt
[133,134,135,136,137,138,140,141,142,143,144,145,146,171][147,148,149,150,151,182,183]NCT01817751 NR
NCT03243461 R
NCT03048084 R
NCT00879437 C
NCT00302159 C
NCT02648633 C
* Clinicaltrials.gov (accessed on 10 January 2023); NR, not recruiting; R, recruiting; C, completed. Abbreviations: u.k. = unknown; SV2A = synaptic vesicle glycoprotein 2A; GABA = γ-Aminobutyric acid; TRPV1 = transient receptor potential vanilloid-1; GPR55 = orphan G protein-coupled receptor-55; ENT-1 = equilibrative nucleoside transporter 1; AMPA = α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; α2δ = alpha2-delta protein, an subunit of voltage-gated Ca2+ channels; NMDA = N-methyl-D-aspartate; MGMT = O(6)-methylguanine-DNA methyltransferase; NF-κb = nuclear factor kappa B; HDAC = histone deacetylase; PTEN = phosphatase and TENsin homolog; PI3K = phosphatidylinositol-3 kinase; AKT = Protein kinase B; CRMP2 = collapsin response mediator protein 2; p38 MAPK = 38-kDa mitogen-activated protein; LDH = lactate dehydrogenase; BDNF = brain-derived neurotrophic factor); ERK = extracellular signal-regulated kinase; mTOR = Mammalian target of rapamycin; Wnt = Wingless/Integrated pathway.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stella, M.; Baiardi, G.; Pasquariello, S.; Sacco, F.; Dellacasagrande, I.; Corsaro, A.; Mattioli, F.; Barbieri, F. Antitumor Potential of Antiepileptic Drugs in Human Glioblastoma: Pharmacological Targets and Clinical Benefits. Biomedicines 2023, 11, 582. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines11020582

AMA Style

Stella M, Baiardi G, Pasquariello S, Sacco F, Dellacasagrande I, Corsaro A, Mattioli F, Barbieri F. Antitumor Potential of Antiepileptic Drugs in Human Glioblastoma: Pharmacological Targets and Clinical Benefits. Biomedicines. 2023; 11(2):582. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines11020582

Chicago/Turabian Style

Stella, Manuela, Giammarco Baiardi, Stefano Pasquariello, Fabio Sacco, Irene Dellacasagrande, Alessandro Corsaro, Francesca Mattioli, and Federica Barbieri. 2023. "Antitumor Potential of Antiepileptic Drugs in Human Glioblastoma: Pharmacological Targets and Clinical Benefits" Biomedicines 11, no. 2: 582. https://0-doi-org.brum.beds.ac.uk/10.3390/biomedicines11020582

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