Next Article in Journal
Removal of Pb2+, CrT, and Hg2+ Ions from Aqueous Solutions Using Amino-Functionalized Magnetic Nanoparticles
Next Article in Special Issue
Antiseizure Effects of Scoparone, Borneol and Their Impact on the Anticonvulsant Potency of Four Classic Antiseizure Medications in the Mouse MES Model—An Isobolographic Transformation
Previous Article in Journal
Acylcarnitines in Ophthalmology: Promising Emerging Biomarkers
Previous Article in Special Issue
A Potent Antagonist of Smoothened in Hedgehog Signaling for Epilepsy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances and Challenges of Cannabidiol as an Anti-Seizure Strategy: Preclinical Evidence

by
Cecilia Zavala-Tecuapetla
1,*,
Hiram Luna-Munguia
2,†,
María-Leonor López-Meraz
3,† and
Manola Cuellar-Herrera
4,†
1
Laboratory of Physiology of Reticular Formation, National Institute of Neurology and Neurosurgery, Insurgentes Sur 3877, La Fama, Mexico City 14269, Mexico
2
Departamento de Neurobiologia Conductual y Cognitiva, Instituto de Neurobiologia, Universidad Nacional Autonoma de Mexico, Campus UNAM-Juriquilla, Queretaro 76230, Mexico
3
Instituto de Investigaciones Cerebrales, Universidad Veracruzana, Luis Castelazo Ayala s/n, Col. Industrial Ánimas, Xalapa 91190, Mexico
4
Epilepsy Clinic, Hospital General de México Dr. Eduardo Liceaga, Dr. Balmis 148, Doctores, Mexico City 06720, Mexico
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(24), 16181; https://doi.org/10.3390/ijms232416181
Submission received: 29 October 2022 / Revised: 24 November 2022 / Accepted: 16 December 2022 / Published: 19 December 2022
(This article belongs to the Special Issue Advances in Epilepsy and Antiepileptic Drugs 2.0)

Abstract

:
The use of Cannabis for medicinal purposes has been documented since ancient times, where one of its principal cannabinoids extracted from Cannabis sativa, cannabidiol (CBD), has emerged over the last few years as a promising molecule with anti-seizure potential. Here, we present an overview of recent literature pointing out CBD’s pharmacological profile (solubility, metabolism, drug-drug interactions, etc.,), CBD’s interactions with multiple molecular targets as well as advances in preclinical research concerning its anti-seizure effect on both acute seizure models and chronic models of epilepsy. We also highlight the recent attention that has been given to other natural cannabinoids and to synthetic derivatives of CBD as possible compounds with therapeutic anti-seizure potential. All the scientific research reviewed here encourages to continue to investigate the probable therapeutic efficacy of CBD and its related compounds not only in epilepsy but also and specially in drug-resistant epilepsy, since there is a dire need for new and effective drugs to treat this disease.

Graphical Abstract

1. Introduction

Cannabis is a genus of plants that belongs to the Cannabaceae family [1]. Among its species, C. sativa L. is a chemically complex plant due to the at least 545 different chemical compounds identified since the 1940s [2], which include more than 100 phytocannabinoids [3,4].
Being the second most abundant phytocannabinoid and due to its multiple effects resulting from its actions on several molecular targets, cannabidiol (CBD) has gained attention due to its attractive therapeutic potential to treat cancer [5], pain [6], Huntington’s disease [7], several psychiatric disorders [8], epilepsy [9], Alzheimer’s disease [10], Parkinson’s disease [11], autoimmune diseases such as rheumatoid arthritis (RA) [12,13], and lately, COVID-19 [14,15], among others.
Currently, new clinical trials on the efficacy of CBD not only for RA (NCT04911127) but also for COVID-19 (NCT04686539, NCT04467918) and lately also for post-COVID syndrome (Long-COVID, NCT04997395) are being conducted (consult ClinicalTrials.gov).

2. Cannabidiol (CBD)

CBD is a highly lipophilic terpenophenol with a wide range of pharmacological effects but without psychoactive consequences [16]. CBD has showed anti-inflammatory [17,18], immunomodulatory [19], and neuroprotective [20] effects. Regarding the latter, CBD's anti-oxidant properties [17,21,22] seems to reduce apoptosis and favors neuroprotective effects and cell viability [23,24], effects that can be therapeutically explored.
Among its pharmacological characteristics, CBD shows poor oral bioavailability (around 13–19%) [25]. Recent evidence points out the existing need to really understand how the co-administration of food and oil content in excipients impacts or not CBD’s pharmacokinetics behavior and dosing schemes, which is crucial information for its therapeutic success and for the development of new innovative formulations [26,27]. Moreover, CBD undergoes hepatic conjugative [28] and oxidative metabolism by specific cytochrome P450 enzymes such as CYP3A4, CYP2C19, CYP1A1, CYP1A2, CYP2C9, CYP2D6, and CYP3A5 [29]. In this regard, the possible interactions between CBD, other drugs, and these enzymes should be considered in order to avoid altering CBD's bioavailability.
Finally, sex differences in CBD's actions should not be ruled out [30,31], and although the idea that CBD has low toxicity has prevailed, recent studies indicate increasing levels of liver enzymes (alanine aminotransferase (ALT); aspartate aminotransferase (AST)) due to CBD treatment [32,33,34]. CBD's effects caused by prolonged use on patients with chronic diseases should be kept in mind, since main adverse effects include: headache, drowsiness, decreased appetite, diarrhea, skin rashes, fatigue, and sleep disorders [9,34].

2.1. Cannabidiol: A Multitarget Molecule

The pharmacodynamic profile of CBD is complex due to its interaction with a wide variety of molecular targets including enzymes, transporters, receptors, and ion channels, among others (Figure 1) [6,35]. Remarkably, these interactions trigger specific mechanisms and signaling pathways culminating in very diverse effects, with high and broad therapeutic potential for diverse diseases.

2.1.1. Cannabinoid Receptors

Cannabinoid receptors (cannabinoid receptor type 1, CB1; cannabinoid receptor type 2, CB2) activate Gi/o proteins which leads to membrane hyperpolarization [36] and to a decrease in neurotransmitter release from pre-synaptic terminals [37].
CBD has showed a poor affinity for both CB1 and CB2 receptors [38,39]. However, it acts as a non-competitive negative allosteric modulator of CB1 receptors [40,41], while with CB2 receptors, CBD acts as an inverse agonist but only at high concentrations [42]. Furthermore, it has been reported that CBD inhibits the reuptake of anandamide and its degradation by the fatty acid amide hydrolase (FAAH), which increases the endogenous cannabinoid tone; in this way CBD indirectly activates CB1 receptors [42,43,44].

2.1.2. Adenosine Receptors

Adenosine is an endogenous purine nucleoside that regulates several physiological functions via four cell-surface G-protein-coupled receptors (A1, A2A, A2B, and A3) [45]. Anti-inflammatory and neuroprotective effects of CBD can be due to either acting as an agonist of A2A receptors, or as a competitive inhibitor at the equilibrative nucleoside transporter 1 (ENT1), which mediates the reuptake of adenosine and therefore its blockade leads to an increase in adenosine extracellular levels [46,47,48,49,50,51].

2.1.3. Serotonin Receptors

Serotonin (5-HT), an endogenous monoamine neurotransmitter, acts on fourteen different receptors classified into seven different classes (5-HT1 to 5-HT7) [52]. In particular, the activation of 5-HT1A receptors causes hyperpolarization via Gi/o-protein-coupled potassium channels [52]. CBD is a low-affinity agonist of the human 5-HT1A receptors [53]. However, it has been proposed that the 5-HT1A-mediated effects of CBD are more attributable to allosteric interactions with the receptor binding site and/or interference with certain intracellular pathways [54].

2.1.4. Dopamine Receptors

Dopamine, other monoamine neurotransmitter, can act on two types of specific G protein-coupled receptors: D1-like and D2-like dopamine receptors, which activation can elicit different intracellular responses [55]. In this regard, cannabinoids seem to modulate dopaminergic signaling [46,56,57,58,59], and specifically, CBD has shown to have a partial agonistic action on dopamine D2-like receptors [60,61], action that may block the D2 receptors when they are overreactive, or activate them when they are underactive [61].

2.1.5. G-Protein-Coupled Receptor 55 (GPR55)

GPR55 is a seven-transmembrane orphan G-protein-coupled receptor that can be activated by endocannabinoids and by lipid transmitters [62,63]. CBD acts as a suppressor of GPR55 activation, which facilitates an increase in neurotransmitter release [62,64], and has been associated with CBD’s anti-inflammatory effects [65].

2.1.6. Peroxisome Proliferator-Activated Receptors (PPARs)

These receptors belong to the type II nuclear hormone receptor superfamily and three isoforms have been identified: PPARα, PPARβ/ɗ, and PPARү [66]. Specifically, CBD is an agonist of PPARү, and this agonist action is able to induce neuroprotective and anti-inflammatory effects [17,18,67].

2.1.7. GABA Receptors

GABA (main inhibitory neurotransmitter) binds to GABAA receptors promoting a rapid influx of chloride ions and resulting in hyperpolarization and inhibition of the cell [68,69]. Interestingly, GABAA receptors represent a target for CBD, either enhancing inhibitory GABAA receptor activation [70], potentiating GABA-mediated inhibitory currents [71], or enhancing the amplitude of the GABA-evoked currents [72].

2.1.8. Opioid Receptors

The opioid system is integrated by four associated G-protein-coupled receptors [µ (MOR), ɗ (DOR), ҡ (KOR), and the non-opioid, nociceptin/orphanin FQ (NOP)] and diverse opioid neuropeptides [73,74]. CBD behaves as a negative allosteric modulator of MOR and DOR [75] and decreases MOR gene expression [76].

2.1.9. Transient Receptor Potential Vanilloid (TRPV) Channels

TRPV channels are direct regulators of intracellular calcium [77]. TRPV1 is a non-selective cation channel [78] which can be activated by CBD [79,80,81]. This agonist action on TRPV1 channels [43], increases entry of extracellular calcium ions through the channel [82], activating the enzyme calcineurin and dephosphorylating TRPV1 channel, to finally desensitize it [83]. Moreover, CBD can also activate TRPV2 channels [42], which under basal conditions, are located in the endoplasmic reticulum, and once activated, they translocate to the plasma membrane [84]. More recently, a study focused on the role of TRPV2 channels in human brain endothelial cells, pointed out that CBD could be considered as a pharmacological tool to regulate the tightness of the blood brain barrier [85].

2.1.10. Other Ion Channels

Some ion channels being proposed as additional CBD targets include: (a) voltage-gated sodium channels, where CBD decreases the sodium current, although the exact mechanism of interaction is still unknown [86,87]; (b) low voltage-activated T-type calcium channels, inhibited by CBD [88]; and (c) potassium channels that generate M-current, where CBD showed recently to enhance neuronal M-current [89].

3. Cannabidiol's Actions for Seizures, Epilepsy, and Drug-Resistant Epilepsy: Preclinical evidence

3.1. Mechanisms Underlying Seizures and Epilepsy

Epilepsy is a very complex neurological disease [90]. The lack of a full effective pharmacological treatment despite the availability of a large number of anti-seizure medications (ASMs) and the risk of developing drug-resistant epilepsy are relevant aspects of this disease.
In terms of etiology, epilepsy can have a structural, genetic, infectious, metabolic, immune, or unknown origin [91], although the origin of epileptic seizures is a multifactorial phenomenon.
From a neurochemical perspective, it has been identified that epileptic seizures occur due to an “epileptic discharge”, i.e., the loss in the dynamic balance between inhibitory (mainly by GABA) and excitatory (mainly by glutamate) neurotransmitter systems. Thus, an alteration between these systems promotes the neuronal hyperexcitability and hypersynchronicity distinctive of epilepsy [92,93]. However, the mechanisms involved in the fine modulation of this synaptic dysfunction are multiple and not unique to these two neurotransmitter systems previously mentioned. Changes in the synthesis and release of neurotransmitters, their receptors, molecules involved in their transductional mechanism, as well as alterations in diverse ionic channels have been implicated in epilepsy [93]. Additionally, changes in neuronal connections and circuits (axonal and dendritic plasticity), neuronal cell loss, aberrant neurogenesis, and neuroinflammatory pathways activation, also contribute to the epileptic etiopathogenesis process [94].

3.2. Mechanisms Underlying the Anti-Seizure Effects of CBD

As we previously pointed out, CBD is a multitarget molecule and therefore can modulate neuronal excitability. A recent study reported that from CBD's interactions with transmembrane receptors, almost 74% inhibit cell activity, whereas only 26% stimulate it [6]. It is important to highlight that from CBD's interactions and the resulting inhibitory effect on cell activity, CBD’s efficacy as a seizure-controlling drug could be explained.
First, although CBD does not promote direct activation of CB1R, an indirect activation of this receptor results from increasing levels of anandamide due to CBD's actions (FAAH blockage and anandamide cellular uptake inhibition) [42,43]. This effect could suppress neuronal hyperactivity, protecting against seizures [95,96,97]. Moreover, CBD seems to prevent CB1R alterations on limbic brain structures, favoring the control of neuronal activity [98].
CBD can also increase extracellular levels of adenosine by ENT1 blockage [46]. This effect allows activation of A1 receptors by adenosine, with anti-seizure effects as a result [99].
On the other hand, although the serotonergic signaling is known to be involved in epilepsy [100] and CBD shows affinity for 5-HT1A receptors [53], the role of these receptors is uncertain in the anti-seizure effects of CBD [101].
In relation to dopamine neurotransmission, since CBD has been proposed as a partial agonist of the D2 dopamine receptor [60,61], it is possible that CBD's anti-seizure effects are due in part to its actions on the dopaminergic system since alterations of this system have been associated as a possible pathophysiological mechanism of epilepsy [102,103,104,105,106].
Another important target is the GPR55 receptor that regulates glutamate release through a calcium-dependent mechanism, modulating neuronal excitability [107]. Therefore, antagonism of GPR55 receptors by CBD can reduce glutamate release and have anti-seizure efficacy [108,109].
Additionally, abnormalities of GABAergic transmission have been associated with epilepsy development [110,111]. CBD has been reported as a positive allosteric modulator of GABAA receptors [71,112], possibly enhancing GABAergic inhibitory transmission and explaining additional CBD’s anti-seizure effects.
With respect to TRPV1 channels, their activation increases neuronal excitability [113] and, in consequence, facilitates seizures [114,115,116,117,118,119]. Although CBD has agonistic actions on TRPV1 channels, it induces a rapid desensitization of these channels [42] that could indirectly decrease the neuronal excitability and explain CBD's anti-seizure effects [120,121].
Finally, the T-type calcium channels and voltage-gated sodium channels are involved in the regulation of cell excitability and therefore they have a role in epilepsy [122,123]. It has been demonstrated that CBD causes the inhibition of these channels [88,124], effects that could mediate the anti-seizure actions of CBD [87,125]. Additionally, CBD has been shown to enhance neuronal M-current, a potassium current whose activation could also underlie CBD's anti-seizure efficacy [89].

3.3. CBD and Seizures

In the search of new therapeutic options, CBD has managed to position itself as a promising molecule to treat seizures [126].
In acute seizure models (as chemoconvulsants [3-mercaptopropionic acid, 3-MA; pentylenetetrazol, PTZ; kainic acid, KA; pilocarpine; N-methyl-D-aspartate, NMDA; soman]; electrical stimulation [maximal electroshock, MES; 6 Hz model]; acoustic stimulation, high temperatures or virus infection), CBD has demonstrated its anti-seizure capacity at variable doses (see Table S1).
The anti-seizure effect of CBD has been tested in rodents of different ages, which gives additional information regarding the effect of this cannabinoid in different stages of neurodevelopment. Most of the research performed in adult animals used males, and only one study analyzed the effect of CBD considering the female estrous cycle. The preclinical evidence about the effectivity of CBD in combination with other ASMs is limited. Only few studies have evaluated CBD's effects on combination with classical (phenytoin, clonazepam, midazolam) and other more recent ASMs (tiagabine, lacosamide, oxcarbazepine, gabapentin, levetiracetam, topiramate, pregabalin, lamotrigine). Most results support a synergic effect when CBD and ASMs are combined, but the information is not conclusive. Table S1 briefly summarizes the main findings obtained from these studies that we have consulted and included for this review.
CBD's anti-seizure potential in the MES model is observed at an ED50 of 80–83.5 mg/kg in adult mice and 53.2–68.78 mg/kg in adult rats, avoiding generalized tonic-clonic seizures (GTCS) and increasing the seizure threshold (the electrical current that produces seizures) [127,128,129,130,131]. Additionally, in that model, CBD (100 mg/kg) increases the anti-seizure effect of some ASMs (topiramate, oxcarbazepine, pregabalin) when it is administered in combination with them [130]. Other observed effects with CBD were the attenuation of the anti-seizure efficacy of levetiracetam (100 mg/kg CBD), and lack of effect on the anti-seizure efficacy of lamotrigine or lacosamide [130].
CBD has a protective effect against psychomotor seizures in adult mice at doses of 50–360 mg/kg (ED50 = 144–173 mg/kg) [128,129,130,132,133]. The same effect is observed against hyperthermia-induced seizures in developing and adult mice [109,134,135,136]. For example, Yu et al. [136] found that CBD (3–30 mg/kg) increased seizure latency, reduced seizure severity and the proportion of fourteen postnatal (P) days old (P14) mice that achieved GTCS. A similar effect was observed in P14–16 Scn1a+/− mice (a model of Dravet Syndrome, DS) where CBD (100 mg/kg) managed to increase the temperature threshold for hyperthermia-induced seizures [134].
In the acute audiogenic seizure model of genetic origin, acute administration of CBD (100 mg/kg i.p.,) reduced generalized clonic convulsions, but without entirely abolishing seizures [137].
Chronic CBD treatment (100 mg/kg i.p. bid, 2 weeks) had no significant effects on seizure behavior [137]. Moreover, coadministration of valproic acid (VPA) plus CBD did not alter the therapeutic outcome of VPA as a monotherapy [137].
Another experimental approach extensively used to characterize the effect of CBD on seizure control is the PTZ model in mice and rats [31,108,121,128,129,133,138]. In this model, CBD was able to protect mice (80, 83.5 mg/kg) and rats (88.9 mg/kg) from tonic-clonic seizures in a dose-dependent manner [128,129]. Vilela et al. [121] evaluated CBD pretreatment in different protocols with PTZ. In all experiments, the authors found that CBD (60 mg/kg; i.p.) decreased both seizure frequency and duration. The results by Uttl and colleagues [138] also showed that CBD administration (60 mg/kg) reduced seizure severity and completely blocked the tonic phase in mice. We noticed that actually there is only one study in which female rats were tested and CBD (50 mg/kg) was administered during the proestrus-estrus transition phase [31]. In this study, CBD reduces the duration and severity of PTZ induced seizures, although the latency, incidence of seizures, and mortality rate remained unchanged [31].
Only a couple of reports have evaluated the anti-seizure effect of CBD in models with a primary glutamatergic mechanism: the NMDA-induced seizure model, where CBD's pre-treatment (60 mg/kg) did not protect P12 male rats from seizures [138]; and the intrahippocampal administration of CBD has been reported to be effective to inhibit KA-induced seizures in P20 rats [139].
In pilocarpine-induced behavioral seizures, CBD's pretreatment (1–100 mg/kg) promoted an anticonvulsant effect, increasing the latency and reducing the severity of seizures [140,141].
Finally, in the model of recurrent generalized seizures induced by repetitive application of 3-MA in rats, CBD (200 mg/kg) did not modify the expression of the 3-MA-induced seizures but reduced the incidence of status epilepticus (SE). The combination of CBD with phenobarbital reduced the prevalence of SE [142].
All the results previously mentioned obtained from different acute preclinical models support CBD's anti-seizure effect, in both the developing and the adult brain, where acute administration is prevalent and the doses of CBD employed vary between the studies mentioned above. However, there is less information on the effect of combining CBD with ASMs, and it is worth noting that although results suggest mostly a potentiation effect, some studies also show that CBD does not enhance ASMs efficacy or may even promote a pro-seizure effect [130,135,137,142,143]. Further preclinical research is required to evaluate the probable interactions between CBD and currently used ASMs.

3.4. CBD, Epilepsy, and Drug-Resistant Epilepsy

The preclinical evidence regarding the effect of CBD in chronic models of epilepsy is still limited (Table S1).
For example, in the pilocarpine epilepsy model, CBD (10 mg/kg i.p.) demonstrated anticonvulsant effects, either by single or multiple administrations, decreasing both SE severity and mortality rate [144], while acute CBD (10 mg/kg i.v.) pretreatment was found to attenuate the maximum seizure severity [129].
In the mouse corneal kindling model, acute administration of CBD (220 mg/kg; i.p.) reduced the frequency of limbic seizures in a dose-dependent manner on kindled mice [128]. Similar results were observed by Patra et al. [129] in the same animal model. CBD also produced partial suppression of both generalized seizures (ED50 = 283 mg/kg) and focal seizures (ED40 = 320 mg/kg) in the amygdala-kindling model [145].
After chronic administration of CBD (50 mg/kg; i.p./daily/28 days/prior to PTZ stimulus) during the PTZ-kindling development, CBD managed to delay the kindling progression although it did not avoid the kindled state [146]. In this same model, another dose of CBD (60 mg/kg; i.p.) was also able to delay the progression of kindling [121]. In another study, daily administration of CBD for 2 weeks (100 mg/kg, i.p., once a day) during the post-kindling phase, did not protect against flurothyl induced seizures, suggesting that CBD could not prevent the pro-epileptogenic process developed in this mouse model of Angelman Syndrome [147]. However, recently, Lazarini-Lopez et al. [98] suggested that CBD could indeed have antiepileptogenic effects, since the chronic administration of CBD (25 mg/kg i.p.; bid, for 10 days) prevented the development of limbic seizures in the audiogenic kindling model, as well as the increase in neuronal activity and the changes in CB1 receptor expression [98].
Finally, CBD's effect in drug-resistant epilepsy has not been thoroughly evaluated (Table S1). For example, in the 6-Hz corneal stimulation model of “psychomotor seizures”, considered as a drug-resistant epilepsy model [148], CBD (164 and 173 mg/kg, i.p.) protected against forelimb seizures and facial clonus [128,129]. On the contrary, in the lamotrigine-resistant amygdala kindling model, CBD (100 and 300 mg/kg, i.p.) did not have any type of protection against epileptic seizures in kindled rats [128].
All the preceding studies performed in chronic models of epilepsy support a partial suppression of seizures by CBD through acute and sometimes chronic protocols of administration. It is necessary to thoroughly research CBD as a disease-modifying agent (by its anti-inflammatory, anti-oxidant, or neuroprotective actions), and its possible coadministration with ASMs. The fasting or fed state of the experimental animal could influence both the brain levels and the anti-seizure effects observed with CBD, a factor that is still undetermined. More preclinical research is required before any conclusions can be made about the therapeutic potential of CBD in drug-resistant epilepsy treatment.

4. Other Natural Cannabinoids with Anti-Seizure Potential

Despite the great attention that CBD has received in the search for new therapeutic options for the control of seizures and epilepsy, new interest has emerged in other phytocannabinoids from C. sativa, that have begun to be evaluated as therapeutic options due to their anti-seizure properties observed in preclinical models (see Table S2).
Cannabichromene (CBC), a phytocannabinoid frequently detected in artisanal cannabis oils, and other CBC-related derivatives (5-fluoro-CBC, cannabichromenic acid CBCA, cannabichromevarinic acid CBCVA) have recently shown anti-seizure effects against hyperthermia-induced seizures in Scn1a+/− mouse model of DS [149].
A similar anti-seizure potential was observed with other phytocannabinoids (also present in artisanal cannabis oils), such as cannabigerolic acid (CBGA), cannabigerovarinic acid (CBGVA), and cannabidivarinic acid (CBDVA) (Figure 2), against seizure activity induced by MES test and against hyperthermia-induced seizures in Scn1a+/− mouse model of DS [150]. It was noticed that CBGA shows superior potency against hyperthermia-induced seizures than CBD, but its anti-seizure effects may depend on the type of preclinical seizure model tested [150] (Table S2). Since CBGA serves as a biosynthetic precursor for CBD [151], it should be taken into account if CBGA's anti-seizure effects are due to CBGA's metabolism or due to the direct effect of CBD.
Moreover, cannabidiolic acid (CBDA) (Figure 2), a phytocannabinoid acid and a precursor of CBD, has recently been shown to also raise the thermogenic seizures threshold in the Scn1aRX/+ mouse model of DS [152]. An interesting finding from that study was that CBDA has poor brain penetration when dissolved in vegetable oil (95% canola oil, 5% sunflower oil), but when the drug was dissolved in Tween 80-based vehicle (ethanol-Tween 80-saline; 1:1:18 ratio) its brain concentration increased [152].
These data emphasize the importance of selecting adequate vehicle preparations for cannabinoids. However, because exposure to heat or light decarboxylates CBDA into CBD [153], magnesium ions have been used to stabilize CBDA-enriched hemp extracts (Chylobinoid and Mg-CBDA) [154] (Table S2). Both extracts demonstrated anti-seizure effects at the MES test in rats [154]. Another study using a CBD/CBDA-rich hemp extract oil showed a reduction in the frequency of canine refractory epileptic seizures during three months of treatment [155].
Cannabidivarin (CBDV) (Figure 2), a propyl analog of CBD, has been evaluated as a purified drug or as a cannabis extract rich in CBDV/CBD, due to its anti-seizure properties observed at the preclinical level [151,156,157]. In a model of acute seizures, CBDV (400 mg/kg oral gavage) decreased PTZ-induced seizure severity and increased onset latency, effects that were associated with changes in the expression levels of some epilepsy-related genes [156]. CBDV in vitro, suppressed epileptiform activity in brain slices, while in vivo had anti-seizure effects on several preclinical models [151,157].
Moreover, CBDV suppressed PTZ-induced seizures in P10 rats, while in P20 rats, CBDV suppressed seizures in several models, including the PTZ, MES, and DMCM (methyl-6,7-dimethoxy-4-ethyl-beta-carboline-3-carboxylate) models. Due to the above, it is suggested that CBDV could have therapeutic value as a novel treatment for childhood epilepsy [158] (Table S2).
As a major metabolite of CBD, 7-OH-CBD has also exerted anti-seizure effects in the MES model in mice [159].
Besides CBD, other compounds such as cannabinol (CBN) and linalool (LN, a cannabis's common terpene) have reduced seizure behavior in the Scn1Lab−/− mutant zebrafish, a DS model [160] (Table S2). Still, further research of their anti-seizure potential is necessary.
Other examples of natural cannabinoids are two olive oil hemp extracts (K1, without volatile compounds; K2; extract rich in terpenes) derived from Cannabis sativa L. (highly enriched in CBD) that reduced the occurrence of convulsive seizures in the 6-Hz corneal stimulation mouse model, where it is suggested that terpenic components could enhance the anti-seizure activity of cannabinoids contained in the K2 oil hemp extract [161] (Table S2). These findings suggest not only that both cannabinoids and terpenes found in oil extracts could influence the anti-seizure effects observed, but also that oil extracts could have potential therapeutic effects against seizures, epilepsy, or drug-resistant epilepsy.
Some other CBD analogues [162] awaiting pharmacological evaluation for their therapeutic benefits include: cannabinodiol (CBND-C5) and cannabinodivarin (CBND-C3), two aromatic CBD's analogs isolated from Lebanese hashish [163]; and cannabielsoin (CBE), a cannabinoid metabolite identified in plants as a product of photo-oxidation from CBD and CBDA [164].
In relation to clinical trials, there are two recent studies that assessed CBDV's efficacy [165,166]. In one study, add-on CBDV (400 to 800 mg b.i.d./2 w + 800 mg b.i.d./6 w) treatment in adult patients with focal seizures, there was no change in seizure frequency between the CBDV and placebo treatment groups, after 8 weeks of treatment [165]. The other study included five girls with Rett Syndrome, where CBDV (oral solution, 50 mg/mL; 10 mg/kg/day) induced a trend toward reduction in mean monthly seizure frequency [166].
Further pharmacological characterization of the biological effects at the preclinical level of the phytocannabinoids reviewed here as well as of the cannabis extracts is necessary, in order to assess whether or not these compounds could be considered in the future as new therapeutic options to improve seizures, epilepsy, or drug-resistant epilepsy outcomes, besides the already known ASMs.

5. Synthetic CBD Derivatives

Other pure molecules, with therapeutic potential, are the so-called synthetic CBD derivatives, which have major potency, efficacy, or pharmacokinetic profile (Figure 2) [162]. Structural modifications of the CBD's molecule include: the alkyl side-chain, the propylcyclohexane, substitution of the phenolic hydroxyl groups, and quinone based synthetic analogs [162].
The pharmacological activity of synthetic CBD derivatives has recently been started to be investigated in a few acute seizure models.
For example, the 1,1-dimethylheptyl (DMH)-CBD derivatives as the pinene dimethoxy-DMH-CBD derivative (HU-308), or the O-1602 (another synthetic cannabinoid), have not shown anti-seizure effects against clonic seizures in the DMCM model (Figure 2) [167].
On the other hand, the 8,9-dihydrocannabidiol (H2CBD), a fully synthetic CBD's analogue, exhibited a dose-dependent anti-seizure action in acute PTZ-induced generalized seizures in rats, with a protective effect comparable to CBD (Figure 2) [168]. Because H2CBD is fully synthetic, it has been suggested that this molecule could not be considered a controlled substance and therefore could circumvent legal issues surrounding cannabis-based therapies [168].
There is only one clinical study that for the first time reported a reduction of seizure frequency in drug-resistant epileptic patients due to treatment with a pharmaceutical grade, synthetically derived CBD [169].
Further preclinical research (in acute and chronic animal models) regarding these synthetic molecules for more precise characterization is required in order to establish whether they have or not anti-seizure or antiepileptogenic effectivity, a better or comparable effectiveness to natural derived CBD, and how they could have or not therapeutic potential in drug-resistant epilepsy.

6. Future Perspectives

6.1. Mitochondrial Dysfunction, Epilepsy, and CBD

Mitochondria has critical cellular functions as the generation of energy (ATP), calcium homeostasis, and control of cell death. Moreover, it is the primary site of reactive oxygen species (ROS) production [170].
In general, mitochondrial dysfunction includes depletion of ATP, the blockade of enzymatic activity in the electron transport chain, generation of ROS, decrease of mitochondrial DNA (mtDNA), calcium imbalance mechanisms and alteration of apoptotic pathways [171]. All these events triggered by mitochondrial dysfunction increase neuronal excitability and promote cellular damage [172]. Therefore, mitochondrial dysfunction has been proposed as an active contributor to the onset and progression of epilepsy [172,173,174,175,176,177,178,179,180,181,182].
In this regard, CBD seems to have the capacity to reduce the intracellular calcium concentration when it is abnormally elevated as in excitotoxic circumstances, via direct interaction with the mitochondrial Na+/Ca2+ exchanger (NCX) [183]. The effect of CBD over the intracellular calcium concentration could prove to have anti-seizure properties by restoring the calcium balance and therefore decreasing neuronal excitability, especially relevant during mitochondrial dysfunction associated with epilepsy [171,179]. CBD also has a direct interaction with the outer-mitochondrial membrane protein, the voltage-dependent anion channel 1 (VDAC1), interaction that in this case increases intracellular calcium concentrations in BV-2 microglial cells [184]. The above suggests that CBD's mitochondrial regulation of the intracellular calcium concentration could be a cell type-dependent effect [185].
On the other hand, it has been demonstrated that not only CB1 receptors are present in brain outer mitochondrial membranes [186,187], but also it has been corroborated their participation in the inhibition of brain mitochondrial respiration [188]. Specifically, CBD has been reported as a full inhibitor of mitochondrial respiration through inhibition of electron transport chain complexes I, II/III, and IV [188,189]. However, it has been suggested a biphasic effect of CBD on mitochondrial respiration [185], where low concentrations could activate mitochondrial respiration [190] and high concentrations could inhibit it [188,189,191]. This is quite relevant since the modulation of mitochondrial respiration could lead to either neuroprotective or neurotoxic effects, affecting both brain mitochondrial metabolism and neuronal integrity, which may contribute to the onset and progression of epilepsy.
Mitophagy, a selective autophagy of mitochondria, is an important mitochondrial mechanism that eliminates dysfunctional or damaged mitochondria to maintain mitochondrial homeostasis and cell survival [192]. Interestingly, defects in mitophagy due to mitochondrial dysfunction have begun to be studied as a pathophysiological mechanism of epilepsy [193,194,195,196].
Recently, it has been reported that CBD activates the PINK1/Parkin pathway in a dose-dependent manner [197]. This pathway initiates not only the mitophagy process, but also the production of mitochondrial derived vesicles (MDV; to transport damaged proteins and lipids to lysosomes) [198,199]. How these CBD's actions over the PINK1/Parkin pathway can have therapeutic value during mitochondrial dysfunction and consequently in epilepsy progression, are questions that remain to be clarified.
Further research is needed that focus on the effects of CBD over specific cell types (e.g., neurons, microglia, astrocytes), as well as an evaluation of the concentrations and doses required for the desired therapeutic effects of CBD, with respect to mitochondrial dysfunction.
If further research proves that CBD can indeed modulate mitochondrial dysfunction, then this would be a different mechanism via which CBD could be a promising tool to treat epilepsy.

6.2. Mammalian Target of Rapamycin, Epilepsy, and CBD

The mammalian target of rapamycin (mTOR) can affect neuronal signaling and excitability, neurotransmitter receptor expression, and synaptic plasticity [200]. This signaling pathway has been proposed as a pathophysiological mechanism in epilepsy [200], since hyperactivated signaling has been reported in both preclinical models as well as in human tissue resected from epileptic patients [201,202,203,204,205,206,207,208,209].
New findings have found that the genetic deletion of PI3Kγ abolishes not only CBD's anti-seizure effects but also CBD's neuroprotection against pilocarpine-induced seizures in mice [141]. The inhibition of mTOR, which is an important component of the PI3K pathway, also blocked CBD's anti-seizure effects [141]. These results suggest that some CBD's actions could be mediated by the PI3K/mTOR pathway [141]. Further studies are required to establish the therapeutic implications of this new mechanism of CBD, in both epilepsy and drug-resistant epilepsy.

7. Cannabidiol: Clinical Evidence

Nowadays, the only brand version of CBD approved by the US Food and Drug Administration (FDA) and European Medicines Agency (EMA), and indicated for the treatment of severe epilepsy syndromes (Dravet Syndrome, DS; Lennox-Gastaut Syndrome, LGS) and for tuberous sclerosis complex (TSC) [32,34,210,211], is the highly purified CBD formulation Epidiolex®/Epidyolex® (oral solution in sesame oil) [212].
Since then, diverse clinical trials have been conducted to evaluate the safety, tolerability, and efficacy of CBD in humans, mainly in patients with severe epilepsy syndromes and TSC, diseases associated with drug resistant seizures in children and young adults [32,34,211,213,214,215,216,217,218,219].
In these trials, CBD has demonstrated an improvement in the patient’s overall condition as adjunctive treatment, by observing a decrease in the seizure frequency. Patel et al. [220] recently reported that CBD does not lose its efficacy in controlling seizures over a treatment period of up to 60 months, in patients with drug-resistant epilepsy of various underlying etiologies (children, young adults, adults). Moreover, other study showed that CBD's efficacy lasts over two years of treatment, in adults and children with drug-resistant epilepsy [221].
On the other hand, CBD treatment has been well tolerated, with adverse events mostly mild or moderate in severity, across different populations including children, adolescents, and younger adults, during the treatment of drug resistant seizures associated with these severe forms of epilepsy. A recent study reported the first systematic review and meta-analysis of the adverse effects of CBD across a range of medical indications [222]. These clinical trials data were mainly obtained from childhood epilepsy syndromes, and suggest that CBD is well tolerated and has few adverse effects, despite the high doses used of CBD [222]. Serious adverse events related to abnormal liver function tests (increased liver transaminases, a sign of hepatotoxicity) or pneumonia were reported. Moreover, other adverse events reported were diarrhea, reduced appetite, and somnolence [222]. It has been estimated that the time to onset of CBD treatment effect (seizure reduction and adverse events), may occur within the first week of treatment and most adverse events resolved within 14-week period (e.g., in patients with LGS) [223].
More information can be consulted through the US National Institutes of Health Clinical Trials Registry (http://www.clinicaltrials.gov, accessed on; searching Epilepsy + CBD), or it can consult some recent reviews of this topic [224,225,226,227,228].

8. Conclusions

The cannabidiol compound has gained popularity in the search for new therapeutic options to treat seizures, epilepsy, or drug-resistant epilepsy. However, CBD's medical use has been restricted since it is still a controlled substance in many countries worldwide.
Evidence obtained from preclinical studies highly support CBD’s anti-seizure effect, which involves neuronal excitability modulation throughout different mechanisms.
Our review summarizes new current data that should be taken into account in future research to investigate the anti-seizure potential of CBD, and lately, other natural and synthetic cannabinoids, from mechanisms of action and pharmacological profiles to innovative formulations (e.g., the molecule's protection from heat or light by using magnesium ions).
Of relevance, due to the fact that CBD is a poor water-soluble molecule, its dissolution in different vehicle preparations, oil-based, surfactants or non-ionic surfactants, may promote differences in the final concentration of CBD in both plasma and brain, highlighting the importance of vehicle selection. In addition, other factors that may affect CBD’s final brain levels include molecule's purity, used dosage, route of administration, fed state, and sex. Potential differences in CBD efficacy could be influenced by the animal species and seizure or epilepsy model chosen. CBD's drug interactions with ASMs should be the main goal for further research in order to avoid toxic or counterproductive effects after chronic treatment with CBD.
Regarding the clinical application of CBD, both its efficacy and optimal doses for the treatment of other epileptic conditions such as temporal lobe epilepsy (a known type of drug-resistant epilepsy) or status epilepticus, in populations of adult patients, remain to be investigated. Many questions still remain unsolved and further research is necessary before considering CBD and other cannabinoids as new therapeutic choices (e.g., monotherapy, adjunctive therapy) with potential to treat not only seizures or epilepsy, but hopefully drug-resistant epilepsy as well in adult patients.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232416181/s1.

Author Contributions

Conceptualization, writing—review and editing, C.Z.-T.; writing—original draft preparation. H.L.-M., M.-L.L.-M. and M.C.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Armando Israel Gutierrez-Lerma for English review and improvement. H.L.-M. is grateful for the support of UNAM-DGAPA-PAPIIT (project IA200621).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sommano, S.R.; Chittasupho, C.; Ruksiriwanich, W.; Jantrawut, P. The Cannabis Terpenes. Molecules 2020, 25, 5792. [Google Scholar] [CrossRef] [PubMed]
  2. ElSohly, M.; Gul, W. Constituents of Cannabis Sativa. In Handbook of Cannabis; Pertwee, R., Ed.; Oxford University Press: Oxford, UK, 2014. [Google Scholar] [CrossRef]
  3. Dayanandan, P.; Kaufman, P.B. Trichomes of Cannabis sativa L. (Cannabaceae). Am. J. Bot. 1976, 63, 578–591. [Google Scholar] [CrossRef] [Green Version]
  4. Hanus, L.O.; Meyer, S.M.; Muñoz, E.; Taglialatela-Scafati, O.; Appendino, G. Phytocannabinoids: A unified critical inventory. Nat. Prod. Rep. 2016, 33, 1357–1392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Seltzer, E.S.; Watters, A.K.; MacKenzie, D., Jr.; Granat, L.M.; Zhang, D. Cannabidiol (CBD) as a Promising Anti-Cancer Drug. Cancers 2020, 12, 3203. [Google Scholar] [CrossRef] [PubMed]
  6. Mlost, J.; Bryk, M.; Starowicz, K. Cannabidiol for Pain Treatment: Focus on Pharmacology and Mechanism of Action. Int. J. Mol. Sci. 2020, 21, 8870. [Google Scholar] [CrossRef]
  7. Sagredo, O.; Pazos, M.R.; Satta, V.; Ramos, J.A.; Pertwee, R.G.; Fernandez-Ruiz, J. Neuroprotective effects of phytocannabinoid-based medicines in experimental models of Huntington’s disease. J. Neurosci. Res. 2011, 89, 1509–1518. [Google Scholar] [CrossRef]
  8. García-Gutiérrez, M.S.; Navarrete, F.; Gasparyan, A.; Austrich-Olivares, A.; Sala, F.; Manzanares, J. Cannabidiol: A Potential New Alternative for the Treatment of Anxiety, Depression, and Psychotic Disorders. Biomolecules 2020, 10, 1575. [Google Scholar] [CrossRef]
  9. Franco, V.; Perucca, E. Pharmacological and Therapeutic Properties of Cannabidiol for Epilepsy. Drugs 2019, 79, 1435–1454. [Google Scholar] [CrossRef]
  10. Watt, G.; Karl, T. In vivo Evidence for Therapeutic Properties of Cannabidiol (CBD) for Alzheimer’s Disease. Front. Pharmacol. 2017, 8, 20. [Google Scholar] [CrossRef]
  11. Patricio, F.; Morales-Andrade, A.A.; Patricio-Martínez, A.; Limón, I.D. Cannabidiol as a Therapeutic Target: Evidence of its Neuroprotective and Neuromodulatory Function in Parkinson’s Disease. Front. Pharmacol. 2020, 11, 595635. [Google Scholar] [CrossRef]
  12. Lowin, T.; Tingting, R.; Zurmahr, J.; Classen, T.; Schneider, M.; Pongratz, G. Cannabidiol (CBD): A killer for inflammatory rheumatoid arthritis synovial fibroblasts. Cell Death Dis. 2020, 11, 714. [Google Scholar] [CrossRef]
  13. Rodríguez-Mesa, X.M.; Moreno-Vergara, A.F.; Contreras-Bolaños, L.A.; Guevara-Moriones, N.; Mejía-Piñeros, A.L.; Santander-González, S.P. Therapeutic Prospects of Cannabinoids in the Immunomodulation of Prevalent Autoimmune Diseases. Cannabis Cannabinoid Res. 2021, 6, 196–210. [Google Scholar] [CrossRef]
  14. Esposito, G.; Pesce, M.; Seguella, L.; Sanseverino, W.; Lu, J.; Corpetti, C.; Sarnelli, G. The potential of cannabidiol in the COVID-19 pandemic. Br. J. Pharmacol. 2020, 177, 4967–4970. [Google Scholar] [CrossRef]
  15. Khalsa, J.H.; Bunt, G.; Maggirwar, S.B.; Kottilil, S. COVID-19 and Cannabidiol (CBD). J. Addict. Med. 2021, 15, 355–356. [Google Scholar] [CrossRef]
  16. Aguillon, A.R.; Leao, R.A.C.; Miranda, L.S.M.; de Souza, R.O.M.A. Cannabidiol discovery and synthesis-a target-oriented analysis in drug production processes. Chemistry 2021, 27, 5577–5600. [Google Scholar] [CrossRef]
  17. Esposito, G.; Scuderi, C.; Valenza, M.; Togna, G.I.; Latina, V.; De Filippis, D.; Cipriano, M.; Carratu, M.R.; Iuvone, T.; Steardo, L. Cannabidiol reduces AB-induced neuroinflammation and promotes hippocampal neurogenesis through PPARү involvement. PLoS ONE 2011, 6, e28668. [Google Scholar] [CrossRef]
  18. Atalay, S.; Jarocka-Karpowicz, I.; Skrzydlewska, E. Antioxidative and Anti-Inflammatory Properties of Cannabidiol. Antioxidants 2019, 9, 21. [Google Scholar] [CrossRef] [Green Version]
  19. Kozela, E.; Pietr, M.; Juknat, A.; Rimmerman, N.; Levy, R.; Vogel, Z. Cannabinoids delta(9)-tetrahydrocannabinol and cannabidiol differentially inhibit the lipopolysaccharide-activated NF-kappaB and interferon-beta/STAT proinflammatory pathways in BV-2 microglial cells. J. Biol. Chem. 2010, 285, 1616–1626. [Google Scholar] [CrossRef] [Green Version]
  20. Schiavon, A.P.; Soares, L.M.; Bonato, J.M.; Milani, H.; Guimarães, F.S.; Weffort de Oliveira, R.M. Protective effects of cannabidiol against hippocampal cell death and cognitive impairment induced by bilateral common carotid artery occlusion in mice. Neurotox. Res. 2014, 26, 307–316. [Google Scholar] [CrossRef]
  21. Hampson, A.J.; Grimaldi, M.; Lolic, M.; Wink, D.; Rosenthal, R.; Axelrod, J. Neuroprotective antioxidants from marijuana. Ann. N. Y. Acad. Sci. 2000, 899, 274–282. [Google Scholar] [CrossRef]
  22. Hacke, A.C.M.; Lima, D.; de Costa, F.; Deshmukh, K.; Li, N.; Chow, A.M.; Marques, J.A.; Pereira, R.P.; Kerman, K. Probing the antioxidant activity of Δ9-tetrahydrocannabinol and cannabidiol in Cannabis sativa extracts. Analyst 2019, 144, 4952–4961. [Google Scholar] [CrossRef] [PubMed]
  23. El-Remessy, A.B.; Khalil, I.E.; Matragoon, S.; Abou-Mohamed, G.; Tsai, N.J.; Roon, P.; Caldwell, R.B.; Caldwell, R.W.; Green, K.; Liou, G.I. Neuroprotective effect of (-) delta9-tetrahydrocannabinol and cannabidiol in N-methyl-D-aspartate-induced retinal neurotoxicity: Involvement of peroxynitrite. Am. J. Pathol. 2003, 163, 1997–2008. [Google Scholar] [CrossRef] [PubMed]
  24. El-Remessy, A.B.; Al-Shabrawey, M.; Khalifa, Y.; Tsai, N.T.; Caldwell, R.B.; Liou, G.I. Neuroprotective and blood-retinal barrier-preserving effects of cannabidiol in experimental diabetes. Am. J. Pathol. 2006, 168, 235–244. [Google Scholar] [CrossRef] [Green Version]
  25. Mechoulam, R.; Parker, L.A.; Gallily, R. Cannabidiol: An overview of some pharmacological aspects. J. Clin. Pharmacol. 2002, 42, 11S–19S. [Google Scholar] [CrossRef] [PubMed]
  26. Deiana, S.; Watanabe, A.; Yamasaki, Y.; Amada, N.; Arthur, M.; Fleming, S.; Woodcock, H.; Dorward, P.; Pigliacampo, B.; Close, S.; et al. Plasma and brain pharmacokinetic profile of cannabidiol (CBD), cannabidivarine (CBDV), ▲9-tetrahydrocannabivarin (THCV) and cannabigerol (CBG) in rats and mice following oral and intraperitoneal administration and CBD action on obsessive-compulsive behaviour. Psychopharmacology 2012, 219, 859–873. [Google Scholar] [CrossRef]
  27. Silmore, L.H.; Willmer, A.R.; Capparelli, E.V.; Rosania, G.R. Food effects on the formulation, dosing, and administration of cannabidiol (CBD) in humans: A systematic review of clinical studies. Pharmacotherapy 2021, 41, 405–420. [Google Scholar] [CrossRef]
  28. Mazur, A.; Lichti, C.F.; Prather, P.L.; Zielinska, A.K.; Bratton, S.M.; Gallus-Zawada, A.; Finel, M.; Miller, G.P.; Radominska-Pandya, A.; Moran, J.H. Characterization of human hepatic and extrahepatic UDP-glucuronosyltransferase enzymes involved in the metabolism of classic cannabinoids. Drug Metab. Dispos. 2009, 37, 1496–1504. [Google Scholar] [CrossRef] [Green Version]
  29. Jiang, R.; Yamaori, S.; Takeda, S.; Yamamoto, I.; Watanabe, K. Identification of cytochrome P450 enzymes responsible for metabolism of cannabidiol by human liver microsomes. Life Sci. 2011, 89, 165–170. [Google Scholar] [CrossRef]
  30. Fattore, L.; Fratta, W. How important are sex differences in cannabinoid action? Br. J. Pharmacol. 2010, 160, 544–548. [Google Scholar] [CrossRef] [Green Version]
  31. Janisset, N.R.L.L.; Romariz, S.A.A.; Hashiguchi, D.; Quintella, M.L.; Gimenes, C.; Yokoyama, T.; Filev, R.; Carlini, E.; Barbosa da Silva, R.; Faber, J.; et al. Partial protective effects of cannabidiol against PTZ-induced acute seizures in female rats during the proestrus-estrus transition. Epilepsy Behav. 2022, 129, 108615. [Google Scholar] [CrossRef]
  32. Devinsky, O.; Patel, A.D.; Cross, J.H.; Villanueva, V.; Wirrell, E.C.; Privitera, M.; Greenwood, S.M.; Roberts, C.; Checketts, D.; VanLandingham, K.E.; et al. GWPCARE3 Study Group. Effect of Cannabidiol on Drop Seizures in the Lennox-Gastaut Syndrome. N. Engl. J. Med. 2018, 378, 1888–1897. [Google Scholar] [CrossRef] [Green Version]
  33. Ewing, L.E.; Skinner, C.M.; Quick, C.M.; Kennon-McGill, S.; McGill, M.R.; Walker, L.A.; ElSohly, M.A.; Gurley, B.J.; Koturbash, I. Hepatotoxicity of a Cannabidiol-Rich Cannabis Extract in the Mouse Model. Molecules 2019, 24, 1694. [Google Scholar] [CrossRef] [Green Version]
  34. Thiele, E.A.; Bebin, E.M.; Filloux, F.; Kwan, P.; Loftus, R.; Sahebkar, F.; Sparagana, S.; Wheless, J. Long-term cannabidiol treatment for seizures in patients with tuberous sclerosis complex: An open-label extension trial. Epilepsia 2022, 63, 426–439. [Google Scholar] [CrossRef]
  35. De Almeida, D.L.; Devi, L.A. Diversity of molecular targets and signaling pathways for CBD. Pharmacol. Res. Perspect. 2020, 8, e00682. [Google Scholar] [CrossRef]
  36. Szabo, B.; Schlicker, E. Effects of cannabinoids on neurotransmission. Handb. Exp. Pharmacol. 2005, 168, 327–365. [Google Scholar] [CrossRef]
  37. Kreitzer, A.C. Neurotransmission: Emerging roles of endocannabinoids. Curr. Biol. 2005, 15, R549–R551. [Google Scholar] [CrossRef] [Green Version]
  38. Mechoulam, R.; Peters, M.; Murillo-Rodriguez, E.; Hanus, L.O. Cannabidiol-recent advances. Chem. Biodivers. 2007, 4, 1678–1692. [Google Scholar] [CrossRef]
  39. McPartland, J.M.; Duncan, M.; Di Marzo, V.; Pertwee, R.G. Are cannabidiol and ▲(9)-tetrahydrocannabivarin negative modulators of the endocannabinoid system? A systematic review. Br. J. Pharmacol. 2015, 172, 737–753. [Google Scholar] [CrossRef] [Green Version]
  40. Laprairie, R.B.; Bagher, A.M.; Kelly, M.E.M.; Denovan-Wright, E.M. Cannabidiol is a negative allosteric modulator of the cannabinoid CB1 receptor. Br. J. Pharmacol. 2015, 172, 4790–4805. [Google Scholar] [CrossRef]
  41. Tham, M.; Yilmaz, O.; Alaverdashvili, M.; Kelly, M.E.M.; Denovan-Wright, E.M.; Laprairie, R.B. Allosteric and orthosteric pharmacology of cannabidiol and cannabidiol-dimethylheptyl at the type 1 and type 2 cannabinoid receptors. Br. J. Pharmacol. 2019, 176, 1455–1469. [Google Scholar] [CrossRef] [Green Version]
  42. De Petrocellis, L.; Ligresti, A.; Moriello, A.S.; Allara, M.; Bisogno, T.; Petrosino, S.; Stott, C.G.; Di Marzo, V. Effects of cannabinoids and cannabinoid-enriched Cannabis extracts on TRP channels and endocannabinoid metabolic enzymes. Br. J. Pharmacol. 2011, 163, 1479–1494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Bisogno, T.; Hanus, L.; De Petrocellis, L.; Tchilibon, S.; Ponde, D.E.; Brandi, I.; Moriello, A.S.; Davis, J.B.; Mechoulam, R.; Di Marzo, V. Molecular targets for cannabidiol and its synthetic analogues: Effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of anandamide. Br. J. Pharmacol. 2001, 134, 845–852. [Google Scholar] [CrossRef] [PubMed]
  44. Leweke, F.M.; Piomelli, D.; Pahlisch, F.; Muhl, D.; Gerth, C.W.; Hoyer, C.; Klosterkotter, J.; Hellmich, M.; Koethe, D. Cannabidiol enhances anandamide signaling and alleviates psychotic symptoms of schizophrenia. Transl. Psychiatry 2012, 2, e94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Fredholm, B.B.; Arslan, G.; Halldner, L.; Kull, B.; Schulte, G.; Wasserman, W. Structure and function of adenosine receptors and their genes. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 362, 364–374. [Google Scholar] [CrossRef] [PubMed]
  46. Pandolfo, P.; Silveirinha, V.; dos Santos-Rodrigues, A.; Venance, L.; Ledent, C.; Takahashi, R.N.; Cunha, R.A.; Köfalvi, A. Cannabinoids inhibit the synaptic uptake of adenosine and dopamine in the rat and mouse striatum. Eur. J. Pharmacol. 2011, 655, 38–45. [Google Scholar] [CrossRef]
  47. Carrier, E.J.; Auchampach, J.A.; Hillard, C.J. Inhibition of an equilibrative nucleoside transporter by cannabidiol: A mechanism of cannabinoid immunosuppression. Proc. Natl. Acad. Sci. USA 2006, 103, 7895–7900. [Google Scholar] [CrossRef] [Green Version]
  48. Liou, G.I.; Auchampach, J.A.; Hillard, C.J.; Zhu, G.; Yousufzai, B.; Mian, S.; Khan, S.; Khalifa, Y. Mediation of cannabidiol anti-inflammation in the retina by equilibrative nucleoside transporter and A2A adenosine receptor. Investig. Ophthalmol. Vis. Sci. 2008, 49, 5526–5531. [Google Scholar] [CrossRef] [Green Version]
  49. Mecha, M.; Feliu, A.; Iñigo, P.M.; Mestre, L.; Carrillo-Salinas, F.J.; Guaza, C. Cannabidiol provides long-lasting protection against the deleterious effects of inflammation in a viral model of multiple sclerosis: A role for A2A receptors. Neurobiol. Dis. 2013, 59, 141–150. [Google Scholar] [CrossRef]
  50. Castillo, A.; Tolon, M.R.; Fernandez-Ruiz, J.; Romero, J.; Martinez-Orgado, J. The neuroprotective effect of cannabidiol in an in vitro model of newborn hypoxic-ischemic brain damage in mice is mediated by CB(2) and adenosine receptors. Neurobiol. Dis. 2010, 37, 434–440. [Google Scholar] [CrossRef]
  51. Martin-Moreno, A.M.; Reigada, D.; Ramirez, B.G.; Mechoulam, R.; Innamorato, N.; Cuadrado, A.; de Ceballos, M.L. Cannabidiol and other cannabinoids reduce microglial activation in vitro and in vivo: Relevance to Alzheimer’s disease. Mol. Pharmacol. 2011, 79, 964–973. [Google Scholar] [CrossRef] [Green Version]
  52. Hoyer, D.; Hannon, J.P.; Martin, G.R. Molecular, pharmacological and functional diversity of 5-HT receptors. Pharmacol. Biochem. Behav. 2002, 71, 533–554. [Google Scholar] [CrossRef]
  53. Russo, E.B.; Burnett, A.; Hall, B.; Parker, K.K. Agonistic properties of cannabidiol at 5-HT1a receptors. Neurochem. Res. 2005, 30, 1037–1043. [Google Scholar] [CrossRef]
  54. Rock, E.M.; Bolognini, D.; Limebeer, C.L.; Cascio, M.G.; Anavi-Goffer, S.; Fletcher, P.J.; Mechoulam, R.; Pertwee, R.G.; Parker, L.A. Cannabidiol, a non-psychotropic component of cannabis, attenuates vomiting and nausea-like behaviour via indirect agonism of 5-HT(1A) somatodendritic autoreceptors in the dorsal raphe nucleus. Br. J. Pharmacol. 2012, 165, 2620–2634. [Google Scholar] [CrossRef] [Green Version]
  55. Klein, M.O.; Battagello, D.S.; Cardoso, A.R.; Hauser, D.N.; Bittencourt, J.C.; Correa, R.G. Dopamine: Functions, Signaling, and Association with Neurological Diseases. Cell. Mol. Neurobiol. 2019, 39, 31–59. [Google Scholar] [CrossRef]
  56. Cheer, J.F.; Wassum, K.M.; Heien, M.L.; Phillips, P.E.; Wightman, R.M. Cannabinoids enhance subsecond dopamine release in the nucleus accumbens of awake rats. J. Neurosci. 2004, 24, 4393–4400. [Google Scholar] [CrossRef] [Green Version]
  57. Terzian, A.L.; Drago, F.; Wotjak, C.T.; Micale, V. The Dopamine and Cannabinoid Interaction in the Modulation of Emotions and Cognition: Assessing the Role of Cannabinoid CB1 Receptor in Neurons Expressing Dopamine D1 Receptors. Front. Behav. Neurosci. 2011, 5, 49. [Google Scholar] [CrossRef] [Green Version]
  58. Hudson, R.; Rushlow, W.; Laviolette, S.R. Phytocannabinoids modulate emotional memory processing through interactions with the ventral hippocampus and mesolimbic dopamine system: Implications for neuropsychiatric pathology. Psychopharmacology 2018, 235, 447–458. [Google Scholar] [CrossRef]
  59. Oliveira da Cruz, J.F.; Busquets-Garcia, A.; Zhao, Z.; Varilh, M.; Lavanco, G.; Bellocchio, L.; Robin, L.; Cannich, A.; Julio-Kalajzić, F.; Lesté-Lasserre, T.; et al. Specific Hippocampal Interneurons Shape Consolidation of Recognition Memory. Cell Rep. 2020, 32, 108046. [Google Scholar] [CrossRef]
  60. Seeman, P. Cannabidiol is a partial agonist at dopamine D2High receptors, predicting its antipsychotic clinical dose. Transl. Psychiatry 2016, 6, e920. [Google Scholar] [CrossRef]
  61. Shrader, S.H.; Tong, Y.G.; Duff, M.B.; Freedman, J.H.; Song, Z.H. Involvement of dopamine receptor in the actions of non-psychoactive phytocannabinoids. Biochem. Biophys. Res. Commun. 2020, 533, 1366–1370. [Google Scholar] [CrossRef]
  62. Ryberg, E.; Larsson, N.; Sjogren, S.; Hjorth, S.; Hermansson, N.O.; Leonova, J.; Elebring, T.; Nilsson, K.; Drmota, T.; Greasley, P.J. The orphan receptor GPR55 is a novel cannabinoid receptor. Br. J. Pharmacol. 2007, 152, 1092–1101. [Google Scholar] [CrossRef] [PubMed]
  63. Oka, S.; Nakajima, K.; Yamashita, A.; Kishimoto, S.; Sugiura, T. Identification of GPR55 as a lysophosphatidylinositol receptor. Biochem. Biophys. Res. Commun. 2007, 362, 928–934. [Google Scholar] [CrossRef] [PubMed]
  64. Chiurchiu, V.; Lanuti, M.; De Bardi, M.; Battistini, L.; Maccarrone, M. The differential characterization of GPR55 receptor in human peripheral blood reveals a distinctive expression in monocytes and NK cells and a proinflammatory role in these innate cells. Int. Immunol. 2015, 27, 153–160. [Google Scholar] [CrossRef] [Green Version]
  65. Gonzalez-Garcia, C.; Moreno-Torres, I.; Garcia-Hernandez, R.; Campos-Ruiz, L.; Rodriguez Esparragoza, L.; Coronado, M.J.; Garcia-Grande, A.; Garcia-Merino, A.; Sanchez-Lopez, A.J. Mechanisms of action of cannabidiol in adoptively transferred experimental autoimmune encephalomyelitis. Exp. Neurol. 2017, 298, 57–67. [Google Scholar] [CrossRef] [PubMed]
  66. Alexander, S.; Cidlowski, J.A.; Kelly, E.; Marrion, N.; Peters, J.A.; Benson, H.E.; Faccenda, E.; Pawson, A.J.; Sharman, J.L.; Southan, C.; et al. CGTP Collaborators. The concise guide to PHARMACOLOGY 2015/2016: Nuclear hormone receptors. Br. J. Pharmacol. 2015, 172, 5956–5978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Sonego, A.B.; Prado, D.S.; Vale, G.T.; Sepulveda-Diaz, J.E.; Cunha, T.M.; Tirapelli, C.R.; Del Bel, E.A.; Raisman-Vozari, R.; Guimaraes, F.S. Cannabidiol prevents haloperidol-induced vacuos chewing movements and inflammatory changes in mice via PPARү receptors. Brain Behav. Immun. 2018, 74, 241–251. [Google Scholar] [CrossRef]
  68. Johnston, G.A. GABAA receptor pharmacology. Pharmacol. Ther. 1996, 69, 173–198. [Google Scholar] [CrossRef]
  69. Olsen, R.W.; Sieghart, W. GABA A receptors: Subtypes provide diversity of function and pharmacology. Neuropharmacology 2009, 56, 141–148. [Google Scholar] [CrossRef] [Green Version]
  70. Aso, E.; Andres-Benito, P.; Ferrer, I. Delineating the efficacy of a cannabis-based medicine at advanced stages of dementia in a murine model. J. Alzheimers Dis. 2016, 54, 903–912. [Google Scholar] [CrossRef]
  71. Bakas, T.; van Nieuwenhuijzen, P.S.; Devenish, S.O.; McGregor, I.S.; Arnold, J.C.; Chebib, M. The direct actions of cannabidiol and 2-arachidonoyl glycerol at GABAA receptors. Pharmacol. Res. 2017, 119, 358–370. [Google Scholar] [CrossRef]
  72. Ruffolo, G.; Cifelli, P.; Roseti, C.; Thom, M.; van Vliet, E.A.; Limatola, C.; Aronica, E.; Palma, E. A novel GABAergic dysfunction in human Dravet syndrome. Epilepsia 2018, 59, 2106–2117. [Google Scholar] [CrossRef] [Green Version]
  73. Waldhoer, M.; Bartlett, S.E.; Whistler, J.L. Opioid receptors. Annu. Rev. Biochem. 2004, 73, 953–990. [Google Scholar] [CrossRef] [Green Version]
  74. Benarroch, E.E. Endogenous opioid systems: Current concepts and clinical correlations. Neurology 2012, 79, 807–814. [Google Scholar] [CrossRef]
  75. Kathmann, M.; Flau, K.; Redmer, A.; Trankle, C.; Schlicker, E. Cannabidiol is an allosteric modulator at mu- and delta-opioid receptors. Naunyn Schmiedebergs Arch. Pharmacol. 2006, 372, 354–361. [Google Scholar] [CrossRef]
  76. Viudez-Martinez, A.; Garcia-Gutierrez, M.S.; Navarron, C.M.; Morales-Calero, M.I.; Navarrete, F.; Torres-Suarez, A.I.; Manzanares, J. Cannabidiol reduces ethanol consumption, motivation and relapse in mice. Addict. Biol. 2018, 23, 154–164. [Google Scholar] [CrossRef]
  77. Wu, L.J.; Sweet, T.B.; Clapham, D.E. International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family. Pharmacol. Rev. 2010, 62, 381–404. [Google Scholar] [CrossRef]
  78. Geppetti, P.; Materazzi, S.; Nicoletti, P. The transient receptor potential vanilloid 1: Role in airway inflammation and disease. Eur. J. Pharmacol. 2006, 533, 207–214. [Google Scholar] [CrossRef]
  79. Zygmunt, P.M.; Petersson, J.; Andersson, D.A.; Chuang, H.; Sorgard, M.; Di Marzo, V.; Julius, D.; Hogestatt, E.D. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature 1999, 400, 452–457. [Google Scholar] [CrossRef]
  80. Correll, C.C.; Phelps, P.T.; Anthes, J.C.; Umland, S.; Greenfeder, S. Cloning and pharmacological characterization of mouse TRPV1. Neurosci. Lett. 2004, 370, 55–60. [Google Scholar] [CrossRef]
  81. McNamara, F.N.; Randall, A.; Gunthorpe, M.J. Effects of piperine, the pungent component of black pepper, at the human vanilloid recpetor (TRPV1). Br. J. Pharmacol. 2005, 144, 781–790. [Google Scholar] [CrossRef] [Green Version]
  82. Caterina, M.J.; Schumacher, M.A.; Tominaga, M.; Rosen, T.A.; Levine, J.D.; Julius, D. The capsaicin receptor: A heat-activated ion channel in the pain pathway. Nature 1997, 389, 816–824. [Google Scholar] [CrossRef] [PubMed]
  83. Novakova-Tousova, K.; Vyklicky, L.; Susankova, K.; Benedikt, J.; Samad, A.; Teisinger, J.; Vlachova, V. Functional changes in the vanilloid receptor subtype 1 channel during and after acute desensitization. Neuroscience 2007, 149, 144–154. [Google Scholar] [CrossRef] [PubMed]
  84. Peralvarez-Marin, A.; Doñate-Macian, P.; Gaudet, R. What do we know about the transient receptor potential vanilloid 2 (TRPV2) ion channel? FEBS J. 2013, 280, 5471–5487. [Google Scholar] [CrossRef] [Green Version]
  85. Luo, H.; Rossi, E.; Saubamea, B.; Chasseigneaux, S.; Cochois, V.; Choublier, N.; Smirnova, M.; Glacial, F.; Perriere, N.; Bourdoulous, S.; et al. Cannabidiol increases proliferation, migration, tubulogenesis, and integrity of human brain endothelial cells through TRPV2 activation. Mol. Pharm. 2019, 16, 1312–1326. [Google Scholar] [CrossRef] [PubMed]
  86. Huang, C.W.; Lin, P.C.; Chen, J.L.; Lee, M.J. Cannabidiol Selectively Binds to the Voltage-Gated Sodium Channel Nav1.4 in Its Slow-Inactivated State and Inhibits Sodium Current. Biomedicines 2021, 9, 1141. [Google Scholar] [CrossRef]
  87. Ghovanloo, M.R.; Ruben, P.C. Cannabidiol and Sodium Channel Pharmacology: General Overview, Mechanism, and Clinical Implications. Neuroscientist 2022, 28, 318–334. [Google Scholar] [CrossRef]
  88. Ross, H.R.; Napier, I.; Connor, M. Inhibition of recombinant human T-type calcium channels by Delta9-tetrahydrocannabinol and cannabidiol. J. Biol. Chem. 2008, 283, 16124–16134. [Google Scholar] [CrossRef] [Green Version]
  89. Zhang, H.B.; Heckman, L.; Niday, Z.; Jo, S.; Fujita, A.; Shim, J.; Pandey, R.; Al Jandal, H.; Jayakar, S.; Barrett, L.B.; et al. Cannabidiol activates neuronal Kv7 channels. Elife 2022, 11, e73246. [Google Scholar] [CrossRef]
  90. Berg, A.T.; Berkovic, S.F.; Brodie, M.J.; Buchhalter, J.; Cross, J.H.; van Emde Boas, W.; Engel, J.; French, J.; Glauser, T.A.; Mathern, G.W.; et al. Revised terminology and concepts for organization of seizures and epilepsies: Report of the ILAE Commission on Classification and Terminology, 2005–2009. Epilepsia 2010, 51, 676–685. [Google Scholar] [CrossRef]
  91. Scheffer, I.E.; Berkovic, S.; Capovilla, G.; Connolly, M.B.; French, J.; Guilhoto, L.; Hirsch, E.; Jain, S.; Mathern, G.W.; Moshé, S.L.; et al. ILAE classification of the epilepsies: Position paper of the ILAE Commission for Classification and Terminology. Epilepsia 2017, 58, 512–521. [Google Scholar] [CrossRef] [Green Version]
  92. Staley, K. Molecular mechanisms of epilepsy. Nat. Neurosci. 2015, 18, 367–372. [Google Scholar] [CrossRef] [Green Version]
  93. Sills, G.J.; Rogawski, M.A. Mechanisms of action of currently used antiseizure drugs. Neuropharmacology 2020, 168, 107966. [Google Scholar] [CrossRef]
  94. Pitkänen, A.; Ekolle Ndode-Ekane, X.; Lapinlampi, N.; Puhakka, N. Epilepsy biomarkers—Toward etiology and pathology specificity. Neurobiol. Dis. 2019, 123, 42–58. [Google Scholar] [CrossRef]
  95. Naidoo, V.; Nikas, S.P.; Karanian, D.A.; Hwang, J.; Zhao, J.; Wood, J.T.; Alapafuja, S.O.; Vadivel, S.K.; Butler, D.; Makriyannis, A.; et al. A new generation fatty acid amide hydrolase inhibitor protects against kainate-induced excitotoxicity. J. Mol. Neurosci. 2011, 43, 493–502. [Google Scholar] [CrossRef] [Green Version]
  96. Naidoo, V.; Karanian, D.A.; Vadivel, S.K.; Locklear, J.R.; Wood, J.T.; Nasr, M.; Quizon, P.M.; 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]
  97. Vilela, L.R.; Medeiros, D.C.; Rezende, G.H.; de Oliveira, A.C.; Moraes, M.F.; Moreira, F.A. Effects of cannabinoids and endocannabinoid hydrolysis inhibition on pentylenetetrazole-induced seizure and electroencephalographic activity in rats. Epilepsy Res. 2013, 104, 195–202. [Google Scholar] [CrossRef]
  98. Lazarini-Lopes, W.; Do Val-da Silva, R.A.; da Silva-Júnior, R.M.P.; Silva-Cardoso, G.K.; Leite-Panissi, C.R.A.; Leite, J.P.; Garcia-Cairasco, N. Chronic cannabidiol (CBD) administration induces anticonvulsant and antiepileptogenic effects in a genetic model of epilepsy. Epilepsy Behav. 2021, 119, 107962. [Google Scholar] [CrossRef]
  99. Weltha, L.; Reemmer, J.; Boison, D. The role of adenosine in epilepsy. Brain Res. Bull. 2019, 151, 46–54. [Google Scholar] [CrossRef]
  100. Petrucci, A.N.; Joyal, K.G.; Purnell, B.S.; Buchanan, G.F. Serotonin and sudden unexpected death in epilepsy. Exp. Neurol. 2020, 325, 113145. [Google Scholar] [CrossRef]
  101. Pelz, M.C.; Schoolcraft, K.D.; Larson, C.; Spring, M.G.; López, H.H. Assessing the role of serotonergic receptors in cannabidiol’s anticonvulsant efficacy. Epilepsy Behav. 2017, 73, 111–118. [Google Scholar] [CrossRef]
  102. Pacia, S.V.; Doyle, W.K.; Broderick, P.A. Biogenic amines in the human neocortex in patients with neocortical and mesial temporal lobe epilepsy: Identification with in situ microvoltammetry. Brain Res. 2001, 899, 106–111. [Google Scholar] [CrossRef] [PubMed]
  103. Werhahn, K.J.; Landvogt, C.; Klimpe, S.; Buchholz, H.G.; Yakushev, I.; Siessmeier, T.; Müller-Forell, W.; Piel, M.; Rösch, F.; Glaser, M.; et al. Decreased dopamine D2/D3-receptor binding in temporal lobe epilepsy: An [18F]fallypride PET study. Epilepsia 2006, 47, 1392–1396. [Google Scholar] [CrossRef] [PubMed]
  104. Rocha, L.; Alonso-Vanegas, M.; Villeda-Hernández, J.; Mújica, M.; Cisneros-Franco, J.M.; López-Gómez, M.; Zavala-Tecuapetla, C.; Frías-Soria, C.L.; Segovia-Vila, J.; Borsodi, A. Dopamine abnormalities in the neocortex of patients with temporal lobe epilepsy. Neurobiol. Dis. 2012, 45, 499–507. [Google Scholar] [CrossRef] [PubMed]
  105. Alcantara-Gonzalez, D.; Floran, B.; Escartin, E.; Rocha, L. Changes on D2-like receptor induced Gi protein activation and hippocampal dopamine release in kindled rats. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 40, 246–251. [Google Scholar] [CrossRef]
  106. Bameri, B.; Shaki, F.; Ahangar, N.; Ataee, R.; Samadi, M.; Mohammadi, H. Evidence for the Involvement of the Dopaminergic System in Seizure and Oxidative Damage Induced by Tramadol. Int. J. Toxicol. 2018, 37, 164–170. [Google Scholar] [CrossRef]
  107. Sylantyev, S.; Jensen, T.P.; Ross, R.A.; Rusakov, D.A. Cannabinoid- and lysophosphatidylinositol-sensitive receptor GPR55 boosts neurotransmitter release at central synapses. Proc. Natl. Acad. Sci. USA 2013, 110, 5193–5198. [Google Scholar] [CrossRef] [Green Version]
  108. Jones, N.A.; Hill, A.J.; Smith, I.; Bevan, S.A.; Williams, C.M.; Whalley, B.J.; Stephens, G.J. Cannabidiol displays antiepileptiform and antiseizure properties in vitro and in vivo. J. Phamacol. Exp. Ther. 2010, 332, 569–577. [Google Scholar] [CrossRef] [Green Version]
  109. Kaplan, J.S.; Stella, N.; Catterall, W.A.; Westenbroek, R.E. Cannabidiol attenuates seizures and social deficits in a mouse model of Dravet syndrome. Proc. Natl. Acad. Sci. USA 2017, 114, 11229–11234. [Google Scholar] [CrossRef] [Green Version]
  110. Sperk, G.; Furtinger, S.; Schwarzer, C.; Pirker, S. GABA and its receptors in epilepsy. Adv. Exp. Med. Biol. 2004, 548, 92–103. [Google Scholar] [CrossRef]
  111. Sarlo, G.L.; Holton, K.F. Brain concentrations of glutamate and GABA in human epilepsy: A review. Seizure 2021, 91, 213–227. [Google Scholar] [CrossRef]
  112. Cifelli, P.; Ruffolo, G.; De Felice, E.; Alfano, V.; van Vliet, E.A.; Aronica, E.; Palma, E. Phytocannabinoids in Neurological Diseases: Could They Restore a Physiological GABAergic Transmission? Int. J. Mol. Sci. 2020, 21, 723. [Google Scholar] [CrossRef] [Green Version]
  113. Mori, F.; Ribolsi, M.; Kusayanagi, H.; Monteleone, F.; Mantovani, V.; Buttari, F.; Marasco, E.; Bernardi, G.; Maccarrone, M.; Centonze, D. TRPV1 channels regulate cortical excitability in humans. J. Neurosci. 2012, 32, 873–879. [Google Scholar] [CrossRef] [Green Version]
  114. Bhaskaran, M.D.; Smith, B.N. Effects of TRPV1 activation on synaptic excitation in the dentate gyrus of a mouse model of temporal lobe epilepsy. Exp. Neurol. 2010, 223, 529–536. [Google Scholar] [CrossRef] [Green Version]
  115. Gonzalez-Reyes, L.E.; Ladas, T.P.; Chiang, C.C.; Durand, D.M. TRPV1 antagonist capsazepine suppresses 4-AP-induced epileptiform activity in vitro and electrographic seizures in vivo. Exp. Neurol. 2013, 250, 321–332. [Google Scholar] [CrossRef] [Green Version]
  116. Shirazi, M.; Izadi, M.; Amin, M.; Rezvani, M.E.; Roohbakhsh, A.; Shamsizadeh, A. Involvement of central TRPV1 receptors in pentylenetetrazole and amygdala-induced kindling in male rats. Neurol. Sci. 2014, 35, 1235–1241. [Google Scholar] [CrossRef]
  117. Socała, K.; Nieoczym, D.; Pieróg, M.; Wlaź, P. α-Spinasterol, a TRPV1 receptor antagonist, elevates the seizure threshold in three acute seizure tests in mice. J. Neural. Transm. 2015, 122, 1239–1247. [Google Scholar] [CrossRef] [Green Version]
  118. Saffarzadeh, F.; Eslamizade, M.J.; Mousavi, S.M.; Abraki, S.B.; Hadjighassem, M.R.; Gorji, A. TRPV1 receptors augment basal synaptic transmission in CA1 and CA3 pyramidal neurons in epilepsy. Neuroscience 2016, 314, 170–178. [Google Scholar] [CrossRef]
  119. Cho, S.J.; Vaca, M.A.; Miranda, C.J.; N’Gouemo, P. Inhibition of transient potential receptor vanilloid type 1 suppresses seizure susceptibility in the genetically epilepsy-prone rat. CNS Neurosci. Ther. 2018, 24, 18–28. [Google Scholar] [CrossRef] [Green Version]
  120. Iannotti, F.A.; Hill, C.L.; Leo, A.; Alhusaini, A.; Soubrane, C.; Mazzarella, E.; Russo, E.; Whalley, B.J.; Di Marzo, V.; Stephens, G.J. Nonpsychotropic plant cannabinoids, cannabidivarin (CBDV) and cannabidiol (CBD), activate and desensitize transient receptor potential vanilloid 1 (TRPV1) channels in vitro: Potential for the treatment of neuronal hyperexcitability. ACS Chem. Neurosci. 2014, 5, 1131–1141. [Google Scholar] [CrossRef]
  121. Vilela, L.R.; Lima, I.V.; Kunsch, É.B.; Pinto, H.P.P.; de Miranda, A.S.; Vieira, É.L.M.; de Oliveira, A.C.P.; Moraes, M.F.D.; Teixeira, A.L.; Moreira, F.A. Anticonvulsant effect of cannabidiol in the pentylenetetrazole model: Pharmacological mechanisms, electroencephalographic profile, and brain cytokine levels. Epilepsy Behav. 2017, 75, 29–35. [Google Scholar] [CrossRef]
  122. Armijo, J.A.; Shushtarian, M.; Valdizan, E.M.; Cuadrado, A.; de las Cuevas, I.; Adín, J. Ion channels and epilepsy. Curr. Pharm. Des. 2005, 11, 1975–2003. [Google Scholar] [CrossRef] [PubMed]
  123. Xu, J.H.; Tang, F.R. Voltage-Dependent Calcium Channels, Calcium Binding Proteins, and Their Interaction in the Pathological Process of Epilepsy. Int. J. Mol. Sci. 2018, 19, 2735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ghovanloo, M.R.; Shuart, N.G.; Mezeyova, J.; Dean, R.A.; Ruben, P.C.; Goodchild, S.J. Inhibitory effects of cannabidiol on voltage-dependent sodium currents. J. Biol. Chem. 2018, 293, 16546–16558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Mason, E.R.; Cummins, T.R. Differential Inhibition of Human Nav1.2 Resurgent and Persistent Sodium Currents by Cannabidiol and GS967. Int. J. Mol. Sci. 2020, 21, 2454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Lazarini-Lopes, W.; Do Val-da Silva, R.A.; da Silva-Júnior, R.M.P.; Leite, J.P.; Garcia-Cairasco, N. The anticonvulsant effects of cannabidiol in experimental models of epileptic seizures: From behavior and mechanisms to clinical insights. Neurosci. Biobehav. Rev. 2020, 111, 166–182. [Google Scholar] [CrossRef]
  127. Wallace, M.J.; Wiley, J.L.; Martin, B.R.; DeLorenzo, R.J. Assessment of the role of CB1 receptors in cannabinoid anticonvulsant effects. Eur. J. Pharmacol. 2001, 428, 51–57. [Google Scholar] [CrossRef]
  128. Klein, B.D.; Jacobson, C.A.; Metcalf, C.S.; Smith, M.D.; Wilcox, K.S.; Hampson, A.J.; Kehne, J.H. Evaluation of Cannabidiol in Animal Seizure Models by the Epilepsy Therapy Screening Program (ETSP). Neurochem. Res. 2017, 42, 1939–1948. [Google Scholar] [CrossRef]
  129. Patra, P.H.; Barker-Haliski, M.; White, H.S.; Whalley, B.J.; Glyn, S.; Sandhu, H.; Jones, N.; Bazelot, M.; Williams, C.M.; McNeish, A.J. Cannabidiol reduces seizures and associated behavioral comorbidities in a range of animal seizure and epilepsy models. Epilepsia 2019, 60, 303–314. [Google Scholar] [CrossRef] [Green Version]
  130. Socała, K.; Wyska, E.; Szafarz, M.; Nieoczym, D.; Wlaź, P. Acute effect of cannabidiol on the activity of various novel antiepileptic drugs in the maximal electroshock- and 6 Hz-induced seizures in mice: Pharmacodynamic and pharmacokinetic studies. Neuropharmacology 2019, 158, 107733. [Google Scholar] [CrossRef]
  131. Gray, R.A.; Stott, C.G.; Jones, N.A.; Di Marzo, V.; Whalley, B.J. Anticonvulsive Properties of Cannabidiol in a Model of Generalized Seizure Are Transient Receptor Potential Vanilloid 1 Dependent. Cannabis Cannabinoid Res. 2020, 5, 145–149. [Google Scholar] [CrossRef] [Green Version]
  132. Patel, D.C.; Wallis, G.; Fujinami, R.S.; Wilcox, K.S.; Smith, M.D. Cannabidiol reduces seizures following CNS infection with Theiler’s murine encephalomyelitis virus. Epilepsia Open 2019, 4, 431–442. [Google Scholar] [CrossRef] [Green Version]
  133. Shapiro, L.; Escayg, A.; Wong, J.C. Cannabidiol Increases Seizure Resistance and Improves Behavior in an Scn8a Mouse Model. Front. Pharmacol. 2022, 13, 815950. [Google Scholar] [CrossRef]
  134. Anderson, L.L.; Low, I.K.; McGregor, I.S.; Arnold, J.C. Interactions between cannabidiol and Δ9 -tetrahydrocannabinol in modulating seizure susceptibility and survival in a mouse model of Dravet syndrome. Br. J. Pharmacol. 2020, 177, 4261–4274. [Google Scholar] [CrossRef]
  135. Chuang, S.H.; Westenbroek, R.E.; Stella, N.; Catterall, W.A. Combined Antiseizure Efficacy of Cannabidiol and Clonazepam in a Conditional Mouse Model of Dravet Syndrome. J. Exp. Neurol. 2021, 2, 81–85. [Google Scholar] [CrossRef]
  136. Yu, Y.; Yang, Z.; Jin, B.; Qin, X.; Zhu, X.; Sun, J.; Huo, L.; Wang, R.; Shi, Y.; Jia, Z.; et al. Cannabidiol inhibits febrile seizure by modulating AMPA receptor kinetics through its interaction with the N-terminal domain of GluA1/GluA2. Pharmacol. Res. 2020, 161, 105128. [Google Scholar] [CrossRef]
  137. Cabral-Pereira, G.; Sánchez-Benito, D.; Díaz-Rodríguez, S.M.; Gonçalves, J.; Sancho, C.; Castellano, O.; Muñoz, L.J.; López, D.E.; Gómez-Nieto, R. Behavioral and Molecular Effects Induced by Cannabidiol and Valproate Administration in the GASH/Sal Model of Acute Audiogenic Seizures. Front. Behav. Neurosci. 2021, 14, 612624. [Google Scholar] [CrossRef]
  138. Uttl, L.; Hložek, T.; Mareš, P.; Páleníček, T.; Kubová, H. Anticonvulsive Effects and Pharmacokinetic Profile of Cannabidiol (CBD) in the Pentylenetetrazol (PTZ) or N-Methyl-D-Aspartate (NMDA) Models of Seizures in Infantile Rats. Int. J. Mol. Sci. 2021, 23, 94. [Google Scholar] [CrossRef]
  139. Friedman, L.K.; Wongvravit, J.P. Anticonvulsant and Neuroprotective Effects of Cannabidiol During the Juvenile Period. J. Neuropathol. Exp. Neurol. 2018, 77, 904–919. [Google Scholar] [CrossRef]
  140. Jones, N.A.; Glyn, S.E.; Akiyama, S.; Hill, T.D.; Hill, A.J.; Weston, S.E.; Burnett, M.D.; Yamasaki, Y.; Stephens, G.J.; Whalley, B.J.; et al. Cannabidiol exerts anti-convulsant effects in animal models of temporal lobe and partial seizures. Seizure 2012, 21, 344–352. [Google Scholar] [CrossRef]
  141. Lima, I.V.A.; Bellozi, P.M.Q.; Batista, E.M.; Vilela, L.R.; Brandão, I.L.; Ribeiro, F.M.; Moraes, M.F.D.; Moreira, F.A.; de Oliveira, A.C.P. Cannabidiol anticonvulsant effect is mediated by the PI3Kγ pathway. Neuropharmacology 2020, 176, 108156. [Google Scholar] [CrossRef]
  142. Frías-Soria, C.L.; Pérez-Pérez, D.; Orozco-Suárez, S.; Rocha, L. Cannabidiol modifies the seizure expression and effects of antiseizure drugs in a rat model of recurrent severe seizures. Seizure 2021, 90, 67–73. [Google Scholar] [CrossRef] [PubMed]
  143. Kundrick, E.R.; Marrero-Rosado, B.M.; de Araujo Furtado, M.; Stone, M.; Schultz, C.R.; Lumley, L.A. Cannabidiol reduces soman-induced lethality and seizure severity in female plasma carboxylesterase knockout mice treated with midazolam. Neurotoxicology 2021, 82, 130–136. [Google Scholar] [CrossRef] [PubMed]
  144. Do Val-da Silva, R.A.; Peixoto-Santos, J.E.; Kandratavicius, L.; De Ross, J.B.; Esteves, I.; De Martinis, B.S.; Alves, M.N.; Scandiuzzi, R.C.; Hallak, J.E.; Zuardi, A.W.; et al. Protective Effects of Cannabidiol against Seizures and Neuronal Death in a Rat Model of Mesial Temporal Lobe Epilepsy. Front. Pharmacol. 2017, 8, 131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fallah, M.S.; Dlugosz, L.; Scott, B.W.; Thompson, M.D.; Burnham, W.M. Antiseizure effects of the cannabinoids in the amygdala-kindling model. Epilepsia 2021, 62, 2274–2282. [Google Scholar] [CrossRef] [PubMed]
  146. Mao, K.; You, C.; Lei, D.; Zhang, H. High dosage of cannabidiol (CBD) alleviates pentylenetetrazole-induced epilepsy in rats by exerting an anticonvulsive effect. Int. J. Clin. Exp. Med. 2015, 8, 8820–8827. [Google Scholar]
  147. Gu, B.; Zhu, M.; Glass, M.R.; Rougié, M.; Nikolova, V.D.; Moy, S.S.; Carney, P.R.; Philpot, B.D. Cannabidiol attenuates seizures and EEG abnormalities in Angelman syndrome model mice. J. Clin. Investig. 2019, 129, 5462–5467. [Google Scholar] [CrossRef] [Green Version]
  148. Barton, M.E.; Klein, B.D.; Wolf, H.H.; White, H.S. Pharmacological characterization of the 6 Hz psychomotor seizure model of partial epilepsy. Epilepsy Res. 2001, 47, 217–227. [Google Scholar] [CrossRef]
  149. Anderson, L.L.; Ametovski, A.; Lin Luo, J.; Everett-Morgan, D.; McGregor, I.S.; Banister, S.D.; Arnold, J.C. Cannabichromene, Related Phytocannabinoids, and 5-Fluoro-cannabichromene Have Anticonvulsant Properties in a Mouse Model of Dravet Syndrome. ACS Chem. Neurosci. 2021, 12, 330–339. [Google Scholar] [CrossRef]
  150. Anderson, L.L.; Heblinski, M.; Absalom, N.L.; Hawkins, N.A.; Bowen, M.T.; Benson, M.J.; Zhang, F.; Bahceci, D.; Doohan, P.T.; Chebib, M.; et al. Cannabigerolic acid, a major biosynthetic precursor molecule in cannabis, exhibits divergent effects on seizures in mouse models of epilepsy. Br. J. Pharmacol. 2021, 178, 4826–4841. [Google Scholar] [CrossRef]
  151. Hill, A.J.; Mercier, M.S.; Hill, T.D.; Glyn, S.E.; Jones, N.A.; Yamasaki, Y.; Futamura, T.; Duncan, M.; Stott, C.G.; Stephens, G.J.; et al. Cannabidivarin is anticonvulsant in mouse and rat. Br. J. Pharmacol. 2012, 167, 1629–1642. [Google Scholar] [CrossRef] [Green Version]
  152. Anderson, L.L.; Low, I.K.; Banister, S.D.; McGregor, I.S.; Arnold, J.C. Pharmacokinetics of Phytocannabinoid Acids and Anticonvulsant Effect of Cannabidiolic Acid in a Mouse Model of Dravet Syndrome. J. Nat. Prod. 2019, 82, 3047–3055. [Google Scholar] [CrossRef] [Green Version]
  153. Citti, C.; Pacchetti, B.; Vandelli, M.A.; Forni, F.; Cannazza, G. Analysis of cannabinoids in commercial hemp seed oil and decarboxylation kinetics studies of cannabidiolic acid (CBDA). J. Pharm. Biomed. Anal. 2018, 149, 532–540. [Google Scholar] [CrossRef]
  154. Goerl, B.; Watkins, S.; Metcalf, C.; Smith, M.; Beenhakker, M. Cannabidiolic acid exhibits entourage-like improvements of anticonvulsant activity in an acute rat model of seizures. Epilepsy Res. 2021, 169, 106525. [Google Scholar] [CrossRef]
  155. Garcia, G.A.; Kube, S.; Carrera-Justiz, S.; Tittle, D.; Wakshlag, J.J. Safety and efficacy of cannabidiol-cannabidiolic acid rich hemp extract in the treatment of refractory epileptic seizures in dogs. Front. Vet. Sci. 2022, 9, 939966. [Google Scholar] [CrossRef]
  156. Amada, N.; Yamasaki, Y.; Williams, C.M.; Whalley, B.J. Cannabidivarin (CBDV) suppresses pentylenetetrazole (PTZ)-induced increases in epilepsy-related gene expression. PeerJ 2013, 1, e214. [Google Scholar] [CrossRef]
  157. Hill, T.D.M.; Cascio, M.G.; Romano, B.; Duncan, M.; Pertwee, R.G.; Williams, C.M.; Whalley, B.J.; Hill, A.J. Cannabidivarin-rich cannabis extracts are anticonvulsant in mouse and rat via a CB1 receptor-independent mechanism. Br. J. Pharmacol. 2013, 170, 679–692. [Google Scholar] [CrossRef] [Green Version]
  158. Huizenga, M.N.; Sepulveda-Rodriguez, A.; Forcelli, P.A. Preclinical safety and efficacy of cannabidivarin for early life seizures. Neuropharmacology 2019, 148, 189–198. [Google Scholar] [CrossRef]
  159. Whalley, B.J.; Stott, C.; Gray, R.A.; Jones, N.A. The human metabolite of cannabidiol, 7-hydroxy cannabidiol, but not 7-carboxy cannabidiol, is anticonvulsant in the maximal electroshock seizure threshold test (MEST) in mouse. [abstract No. 1.435]. In Proceedings of the 71st Annual Meeting of the American Epilepsy Society, Washington, DC, USA, 1–5 December 2017; Available online: https://cms.aesnet.org/abstractslisting/the-human-metabolite-of-cannabidiol--7-hydroxy-cannabidiol--but-not-7-carboxy-cannabidiol--is-anticonvulsant-in-the-maximal-electroshock-seizure-threshold-test-(mest)-in-mouse (accessed on 21 September 2022).
  160. Thornton, C.; Dickson, K.E.; Carty, D.R.; Ashpole, N.M.; Willett, K.L. Cannabis constituents reduce seizure behavior in chemically-induced and scn1a-mutant zebrafish. Epilepsy Behav. 2020, 110, 107152. [Google Scholar] [CrossRef]
  161. Costa, A.M.; Senn, L.; Anceschi, L.; Brighenti, V.; Pellati, F.; Biagini, G. Antiseizure Effects of Fully Characterized Non-Psychoactive Cannabis sativa L. Extracts in the Repeated 6-Hz Corneal Stimulation Test. Pharmaceuticals 2021, 14, 1259. [Google Scholar] [CrossRef]
  162. Morales, P.; Reggio, P.H.; Jagerovic, N. An Overview on Medicinal Chemistry of Synthetic and Natural Derivatives of Cannabidiol. Front. Pharmacol. 2017, 8, 422. [Google Scholar] [CrossRef]
  163. ElSohly, M.A.; Slade, D. Chemical constituents of marijuana: The complex mixture of natural cannabinoids. Life Sci. 2005, 78, 539–548. [Google Scholar] [CrossRef] [PubMed]
  164. Ujváry, I.; Hanuš, L. Human metabolites of cannabidiol: A review on their formation, biological activity, and relevance in therapy. Cannabis Cannabinoid Res. 2016, 1, 90–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Brodie, M.J.; Czapinski, P.; Pazdera, L.; Sander, J.W.; Toledo, M.; Napoles, M.; Sahebkar, F.; Schreiber, A. GWEP1330 Study Group. A Phase 2 Randomized Controlled Trial of the Efficacy and Safety of Cannabidivarin as Add-on Therapy in Participants with Inadequately Controlled Focal Seizures. Cannabis Cannabinoid Res. 2021, 6, 528–536. [Google Scholar] [CrossRef] [PubMed]
  166. Hurley, E.N.; Ellaway, C.J.; Johnson, A.M.; Truong, L.; Gordon, R.; Galettis, P.; Martin, J.H.; Lawson, J.A. Efficacy and safety of cannabidivarin treatment of epilepsy in girls with Rett syndrome: A phase 1 clinical trial. Epilepsia 2022, 63, 1736–1747. [Google Scholar] [CrossRef] [PubMed]
  167. Huizenga, M.N.; Wicker, E.; Beck, V.C.; Forcelli, P.A. Anticonvulsant effect of cannabinoid receptor agonists in models of seizures in developing rats. Epilepsia 2017, 58, 1593–1602. [Google Scholar] [CrossRef] [Green Version]
  168. Mascal, M.; Hafezi, N.; Wang, D.; Hu, Y.; Serra, G.; Dallas, M.L.; Spencer, J.P.E. Synthetic, non-intoxicating 8,9-dihydrocannabidiol for the mitigation of seizures. Sci. Rep. 2019, 9, 7778. [Google Scholar] [CrossRef] [Green Version]
  169. Klotz, K.A.; Grob, D.; Hirsch, M.; Metternich, B.; Schulze-Bonhage, A.; Jacobs, J. Efficacy and Tolerance of Synthetic Cannabidiol for Treatment of Drug Resistant Epilepsy. Front. Neurol. 2019, 10, 1313. [Google Scholar] [CrossRef]
  170. Mattson, M.P.; Gleichmann, M.; Cheng, A. Mitochondria in neuroplasticity and neurological disorders. Neuron 2008, 60, 748–766. [Google Scholar] [CrossRef] [Green Version]
  171. Wu, Y.; Chen, M.; Jiang, J. Mitochondrial dysfunction in neurodegenerative diseases and drug targets via apoptotic signaling. Mitochondrion 2019, 49, 35–45. [Google Scholar] [CrossRef]
  172. Kovac, S.; Dinkova Kostova, A.T.; Herrmann, A.M.; Melzer, N.; Meuth, S.G.; Gorji, A. Metabolic and Homeostatic Changes in Seizures and Acquired Epilepsy-Mitochondria, Calcium Dynamics and Reactive Oxygen Species. Int. J. Mol. Sci. 2017, 18, 1935. [Google Scholar] [CrossRef]
  173. Kunz, W.S.; Kudin, A.P.; Vielhaber, S.; Blümcke, I.; Zuschratter, W.; Schramm, J.; Beck, H.; Elger, C.E. Mitochondrial complex I deficiency in the epileptic focus of patients with temporal lobe epilepsy. Ann. Neurol. 2000, 48, 766–773. [Google Scholar] [CrossRef]
  174. Folbergrová, J.; Druga, R.; Otáhal, J.; Haugvicová, R.; Mares, P.; Kubová, H. Seizures induced in immature rats by homocysteic acid and the associated brain damage are prevented by group II metabotropic glutamate receptor agonist (2R,4R)-4-aminopyrrolidine-2,4-dicarboxylate. Exp. Neurol. 2005, 192, 420–436. [Google Scholar] [CrossRef]
  175. Henshall, D.C. Apoptosis signalling pathways in seizure-induced neuronal death and epilepsy. Biochem. Soc. Trans. 2007, 35 Pt 2, 421–423. [Google Scholar] [CrossRef] [Green Version]
  176. Jarrett, S.G.; Liang, L.P.; Hellier, J.L.; Staley, K.J.; Patel, M. Mitochondrial DNA damage and impaired base excision repair during epileptogenesis. Neurobiol. Dis. 2008, 30, 130–138. [Google Scholar] [CrossRef] [Green Version]
  177. Vielhaber, S.; Niessen, H.G.; Debska-Vielhaber, G.; Kudin, A.P.; Wellmer, J.; Kaufmann, J.; Schönfeld, M.A.; Fendrich, R.; Willker, W.; Leibfritz, D.; et al. Subfield-specific loss of hippocampal N-acetyl aspartate in temporal lobe epilepsy. Epilepsia 2008, 49, 40–50. [Google Scholar] [CrossRef] [PubMed]
  178. Waldbaum, S.; Liang, L.P.; Patel, M. Persistent impairment of mitochondrial and tissue redox status during lithium-pilocarpine-induced epileptogenesis. J. Neurochem. 2010, 115, 1172–1182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Folbergrová, J.; Kunz, W.S. Mitochondrial dysfunction in epilepsy. Mitochondrion 2012, 12, 35–40. [Google Scholar] [CrossRef] [PubMed]
  180. Liang, L.P.; Waldbaum, S.; Rowley, S.; Huang, T.T.; Day, B.J.; Patel, M. Mitochondrial oxidative stress and epilepsy in SOD2 deficient mice: Attenuation by a lipophilic metalloporphyrin. Neurobiol. Dis. 2012, 45, 1068–1076. [Google Scholar] [CrossRef] [Green Version]
  181. Ristić, A.J.; Savić, D.; Sokić, D.; Bogdanović Pristov, J.; Nestorov, J.; Baščarević, V.; Raičević, S.; Savić, S.; Spasojević, I. Hippocampal antioxidative system in mesial temporal lobe epilepsy. Epilepsia 2015, 56, 789–799. [Google Scholar] [CrossRef]
  182. Wesół-Kucharska, D.; Rokicki, D.; Jezela-Stanek, A. Epilepsy in Mitochondrial Diseases-Current State of Knowledge on Aetiology and Treatment. Children 2021, 8, 532. [Google Scholar] [CrossRef]
  183. Ryan, D.; Drysdale, A.J.; Lafourcade, C.; Pertwee, R.G.; Platt, B. Cannabidiol targets mitochondria to regulate intracellular Ca2+ levels. J. Neurosci. 2009, 29, 2053–2063. [Google Scholar] [CrossRef] [Green Version]
  184. Rimmerman, N.; Ben-Hail, D.; Porat, Z.; Juknat, A.; Kozela, E.; Daniels, M.P.; Connelly, P.S.; Leishman, E.; Bradshaw, H.B.; Shoshan-Barmatz, V.; et al. Direct modulation of the outer mitochondrial membrane channel, voltage-dependent anion channel 1 (VDAC1) by cannabidiol: A novel mechanism for cannabinoid-induced cell death. Cell Death Dis. 2013, 4, e949. [Google Scholar] [CrossRef] [Green Version]
  185. Chan, J.Z.; Duncan, R.E. Regulatory Effects of Cannabidiol on Mitochondrial Functions: A Review. Cells 2021, 10, 1251. [Google Scholar] [CrossRef]
  186. Bénard, G.; Massa, F.; Puente, N.; Lourenço, J.; Bellocchio, L.; Soria-Gómez, E.; Matias, I.; Delamarre, A.; Metna-Laurent, M.; Cannich, A.; et al. Mitochondrial CB1 receptors regulate neuronal energy metabolism. Nat. Neurosci. 2012, 15, 558–564. [Google Scholar] [CrossRef] [PubMed]
  187. Hebert-Chatelain, E.; Reguero, L.; Puente, N.; Lutz, B.; Chaouloff, F.; Rossignol, R.; Piazza, P.V.; Benard, G.; Grandes, P.; Marsicano, G. Cannabinoid control of brain bioenergetics: Exploring the subcellular localization of the CB1 receptor. Mol. Metab. 2014, 3, 495–504. [Google Scholar] [CrossRef]
  188. Fišar, Z.; Singh, N.; Hroudová, J. Cannabinoid-induced changes in respiration of brain mitochondria. Toxicol. Lett. 2014, 231, 62–71. [Google Scholar] [CrossRef]
  189. Singh, N.; Hroudová, J.; Fišar, Z. Cannabinoid-Induced Changes in the Activity of Electron Transport Chain Complexes of Brain Mitochondria. J. Mol. Neurosci. 2015, 56, 926–931. [Google Scholar] [CrossRef]
  190. Valvassori, S.S.; Bavaresco, D.V.; Scaini, G.; Varela, R.B.; Streck, E.L.; Chagas, M.H.; Hallak, J.E.; Zuardi, A.W.; Crippa, J.A.; Quevedo, J. Acute and chronic administration of cannabidiol increases mitochondrial complex and creatine kinase activity in the rat brain. Braz. J. Psychiatry 2013, 35, 380–386. [Google Scholar] [CrossRef] [Green Version]
  191. Drummond-Main, C.D.; Ahn, Y.; Kesler, M.; Gavrilovici, C.; Kim, D.Y.; Kiroski, I.; Baglot, S.L.; Chen, A.; Sharkey, K.A.; Hill, M.N.; et al. Cannabidiol Impairs Brain Mitochondrial Metabolism and Neuronal Integrity. Cannabis Cannabinoid Res. 2022. [Google Scholar] [CrossRef]
  192. Ashrafi, G.; Schwarz, T.L. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 2013, 20, 31–42. [Google Scholar] [CrossRef] [PubMed]
  193. Li, L.L.; Jin, M.F.; Ni, H. Zinc/CaMK II Associated-Mitophagy Signaling Contributed to Hippocampal Mossy Fiber Sprouting and Cognitive Deficits Following Neonatal Seizures and Its Regulation by Chronic Leptin Treatment. Front. Neurol. 2018, 9, 802. [Google Scholar] [CrossRef]
  194. Wu, M.; Liu, X.; Chi, X.; Zhang, L.; Xiong, W.; Chiang, S.M.V.; Zhou, D.; Li, J. Mitophagy in Refractory Temporal Lobe Epilepsy Patients with Hippocampal Sclerosis. Cell. Mol. Neurobiol. 2018, 38, 479–486. [Google Scholar] [CrossRef]
  195. Zhang, Y.; Lian, Y.; Lian, X.; Zhang, H.; Chen, Y.; Sheng, H.; Feng, R. FUNDC1 Mediated Mitophagy in Epileptic Hippocampal Neuronal Injury Induced by Magnesium-Free Fluid. Neurochem. Res. 2022. [Google Scholar] [CrossRef]
  196. Fang, Q.; Zheng, S.; Chen, Q.; Chen, L.; Yang, Y.; Wang, Y.; Zhang, H.; Chen, J. The protective effect of inhibiting mitochondrial fission on the juvenile rat brain following PTZ kindling through inhibiting the BCL2L13/LC3 mitophagy pathway. Metab. Brain Dis. 2022. [Google Scholar] [CrossRef]
  197. Ramirez, A.; Old, W.; Selwood, D.L.; Liu, X. Cannabidiol activates PINK1-Parkin-dependent mitophagy and mitochondrial-derived vesicles. Eur. J. Cell Biol. 2022, 101, 151185. [Google Scholar] [CrossRef]
  198. Sugiura, A.; McLelland, G.L.; Fon, E.A.; McBride, H.M. A new pathway for mitochondrial quality control: Mitochondrial-derived vesicles. EMBO J. 2014, 33, 2142–2156. [Google Scholar] [CrossRef] [Green Version]
  199. Pickrell, A.M.; Youle, R.J. The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson’s disease. Neuron 2015, 85, 257–273. [Google Scholar] [CrossRef] [Green Version]
  200. Zavala-Tecuapetla, C.; Cuellar-Herrera, M.; Luna-Munguia, H. Insights into Potential Targets for Therapeutic Intervention in Epilepsy. Int. J. Mol. Sci. 2020, 21, 8573. [Google Scholar] [CrossRef]
  201. Baybis, M.; Yu, J.; Lee, A.; Golden, J.A.; Weiner, H.; McKhann, G., 2nd; Aronica, E.; Crino, P.B. mTOR cascade activation distinguishes tubers from focal cortical dysplasia. Ann. Neurol. 2004, 56, 478–487. [Google Scholar] [CrossRef] [PubMed]
  202. Zeng, L.H.; Xu, L.; Gutmann, D.H.; Wong, M. Rapamycin prevents epilepsy in a mouse model of tuberous sclerosis complex. Ann. Neurol. 2008, 63, 444–453. [Google Scholar] [CrossRef]
  203. Zhang, B.; Wong, M. Pentylenetetrazole-induced seizures cause acute, but not chronic, mTOR pathway activation in rat. Epilepsia 2012, 53, 506–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Van Vliet, E.A.; Forte, G.; Holtman, L.; den Burger, J.C.; Sinjewel, A.; de Vries, H.E.; Aronica, E.; Gorter, J.A. Inhibition of mammalian target of rapamycin reduces epileptogenesis and blood-brain barrier leakage but not microglia activation. Epilepsia 2012, 53, 1254–1263. [Google Scholar] [CrossRef] [PubMed]
  205. Talos, D.M.; Jacobs, L.M.; Gourmaud, S.; Coto, C.A.; Sun, H.; Lim, K.C.; Lucas, T.H.; Davis, K.A.; Martinez-Lage, M.; Jensen, F.E. Mechanistic target of rapamycin complex 1 and 2 in human temporal lobe epilepsy. Ann. Neurol. 2018, 83, 311–327. [Google Scholar] [CrossRef] [PubMed]
  206. Theilmann, W.; Gericke, B.; Schidlitzki, A.; Muneeb Anjum, S.M.; Borsdorf, S.; Harries, T.; Roberds, S.L.; Aguiar, D.J.; Brunner, D.; Leiser, S.C.; et al. Novel brain permeant mTORC1/2 inhibitors are as efficacious as rapamycin or everolimus in mouse models of acquired partial epilepsy and tuberous sclerosis complex. Neuropharmacology 2020, 180, 108297. [Google Scholar] [CrossRef] [PubMed]
  207. Zhao, X.F.; Liao, Y.; Alam, M.M.; Mathur, R.; Feustel, P.; Mazurkiewicz, J.E.; Adamo, M.A.; Zhu, X.C.; Huang, Y. Microglial mTOR is Neuronal Protective and Antiepileptogenic in the Pilocarpine Model of Temporal Lobe Epilepsy. J. Neurosci. 2020, 40, 7593–7608. [Google Scholar] [CrossRef]
  208. Huang, X.Y.; Hu, Q.P.; Shi, H.Y.; Zheng, Y.Y.; Hu, R.R.; Guo, Q. Everolimus inhibits PI3K/Akt/mTOR and NF-kB/IL-6 signaling and protects seizure-induced brain injury in rats. J. Chem. Neuroanat. 2021, 114, 101960. [Google Scholar] [CrossRef]
  209. Chmielewska, N.; Wawer, A.; Osuch, B.; Maciejak, P.; Szyndler, J. mTOR and HDAC2 are simultaneously activated during electrically induced kindling of seizures. Epilepsy Res. 2022, 185, 106991. [Google Scholar] [CrossRef]
  210. Devinsky, O.; Patel, A.D.; Thiele, E.A.; Wong, M.H.; Appleton, R.; Harden, C.L.; Greenwood, S.; Morrison, G.; Sommerville, K. GWPCARE1 Part A Study Group. Randomized, dose-ranging safety trial of cannabidiol in Dravet syndrome. Neurology 2018, 90, e1204–e1211. [Google Scholar] [CrossRef] [Green Version]
  211. Thiele, E.A.; Bebin, E.M.; Bhathal, H.; Jansen, F.E.; Kotulska, K.; Lawson, J.A.; O’Callaghan, F.J.; Wong, M.; Sahebkar, F.; Checketts, D.; et al. Add-on Cannabidiol Treatment for Drug-Resistant Seizures in Tuberous Sclerosis Complex: A Placebo-Controlled Randomized Clinical Trial. JAMA Neurol. 2021, 78, 285–292. [Google Scholar] [CrossRef]
  212. GW Pharmaceuticals. Press Release: GW Pharmaceuticals Receives Approval for EPIDYOLEX® (Cannabidiol) from the MHRA for the Treatment of Seizures Associated with Tuberous Sclerosis Complex, London, Aug. 10, 2021. Available online: https://investor.jazzpharma.com/news-releases/news-release-details/gw-pharmaceuticals-receives-approval-epidyolexr-cannabidiol-mhra (accessed on 6 December 2022).
  213. Devinsky, O.; Marsh, E.; Friedman, D.; Thiele, E.; Laux, L.; Sullivan, J.; Miller, I.; Flamini, R.; Wilfong, A.; Filloux, F.; et al. Cannabidiol in patients with treatment-resistant epilepsy: An open-label interventional trial. Lancet Neurol. 2016, 15, 270–278. [Google Scholar] [CrossRef]
  214. Devinsky, O.; Cross, J.H.; Laux, L.; Marsh, E.; Miller, I.; Nabbout, R.; Scheffer, I.E.; Thiele, E.A.; Wright, S.; Cannabidiol in Dravet Syndrome Study Group. Trial of Cannabidiol for Drug-Resistant Seizures in the Dravet Syndrome. N. Engl. J. Med. 2017, 25, 376, 2011–2020. [Google Scholar] [CrossRef] [Green Version]
  215. Devinsky, O.; Nabbout, R.; Miller, I.; Laux, L.; Zolnowska, M.; Wright, S.; Roberts, C. Long-term cannabidiol treatment in patients with Dravet syndrome: An open-label extension trial. Epilepsia 2019, 60, 294–302. [Google Scholar] [CrossRef]
  216. Hess, E.J.; Moody, K.A.; Geffrey, A.L.; Pollack, S.F.; Skirvin, L.A.; Bruno, P.L.; Paolini, J.L.; Thiele, E.A. Cannabidiol as a new treatment for drug-resistant epilepsy in tuberous sclerosis complex. Epilepsia 2016, 57, 1617–1624. [Google Scholar] [CrossRef]
  217. Szaflarski, J.P.; Hernando, K.; Bebin, E.M.; Gaston, T.E.; Grayson, L.E.; Ampah, S.B.; Moreadith, R. Higher cannabidiol plasma levels are associated with better seizure response following treatment with a pharmaceutical grade cannabidiol. Epilepsy Behav. 2019, 95, 131–136. [Google Scholar] [CrossRef]
  218. Thiele, E.A.; Marsh, E.D.; French, J.A.; Mazurkiewicz-Beldzinska, M.; Benbadis, S.R.; Joshi, C.; Lyons, P.D.; Taylor, A.; Roberts, C.; Sommerville, K.; et al. Cannabidiol in patients with seizures associated with Lennox-Gastaut syndrome (GWPCARE4): A randomised, double-blind, placebo-controlled phase 3 trial. Lancet 2018, 391, 1085–1096. [Google Scholar] [CrossRef]
  219. Thiele, E.; Marsh, E.; Mazurkiewicz-Beldzinska, M.; Halford, J.J.; Gunning, B.; Devinsky, O.; Checketts, D.; Roberts, C. Cannabidiol in patients with Lennox-Gastaut syndrome: Interim analysis of an open-label extension study. Epilepsia 2019, 60, 419–428. [Google Scholar] [CrossRef] [Green Version]
  220. 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]
  221. Gaston, T.E.; Ampah, S.B.; Martina Bebin, E.; Grayson, L.P.; Cutter, G.R.; Hernando, K.; Szaflarski, J.P.; UAB CBD Program. Long-term safety and efficacy of highly purified cannabidiol for treatment refractory epilepsy. Epilepsy Behav. 2021, 117, 107862. [Google Scholar] [CrossRef]
  222. Chesney, E.; Oliver, D.; Green, A.; Sovi, S.; Wilson, J.; Englund, A.; Freeman, T.P.; McGuire, P. Adverse effects of cannabidiol: A systematic review and meta-analysis of randomized clinical trials. Neuropsychopharmacology 2020, 45, 1799–1806. [Google Scholar] [CrossRef]
  223. Privitera, M.; Bhathal, H.; Wong, M.; Cross, J.H.; Wirrell, E.; Marsh, E.D.; Mazurkiewicz-Beldzinska, M.; Villanueva, V.; Checketts, D.; Knappertz, V.; et al. Time to onset of cannabidiol (CBD) treatment effect in Lennox-Gastaut syndrome: Analysis from two randomized controlled trials. Epilepsia 2021, 62, 1130–1140. [Google Scholar] [CrossRef]
  224. Silvestro, S.; Mammana, S.; Cavalli, E.; Bramanti, P.; Mazzon, E. Use of Cannabidiol in the Treatment of Epilepsy: Efficacy and Security in Clinical Trials. Molecules 2019, 24, 1459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Silva, G.D.; Del Guerra, F.B.; de Oliveira Lelis, M.; Pinto, L.F. Cannabidiol in the Treatment of Epilepsy: A Focused Review of Evidence and Gaps. Front. Neurol. 2020, 11, 531939. [Google Scholar] [CrossRef] [PubMed]
  226. De Carvalho Reis, R.; Almeida, K.J.; da Silva Lopes, L.; de Melo Mendes, C.M.; Bor-Seng-Shu, E. Efficacy and adverse event profile of cannabidiol and medicinal cannabis for treatment-resistant epilepsy: Systematic review and meta-analysis. Epilepsy Behav. 2020, 102, 106635. [Google Scholar] [CrossRef] [PubMed]
  227. Von Wrede, R.; Helmstaedter, C.; Surges, R. Cannabidiol in the Treatment of Epilepsy. Clin. Drug Investig. 2021, 41, 211–220. [Google Scholar] [CrossRef]
  228. Lattanzi, S.; Trinka, E.; Striano, P.; Rocchi, C.; Salvemini, S.; Silvestrini, M.; Brigo, F. Highly Purified Cannabidiol for Epilepsy Treatment: A Systematic Review of Epileptic Conditions Beyond Dravet Syndrome and Lennox-Gastaut Syndrome. CNS Drugs 2021, 35, 265–281. [Google Scholar] [CrossRef]
Figure 1. Some molecular targets of CBD’s actions in the brain. CBD acts as a negative allosteric modulator of CB1, MOR, and DOR receptors; as an inverse agonist of CB2 receptors; as an agonist of A2A, 5-HT1A, D2, PPARү, TRPV1, TRPV2, and GABAA receptors; as an antagonist of GPR55 receptor; and as an inhibitor of ENT1. CBD, cannabidiol; CB1, cannabinoid receptor type 1; CB2, cannabinoid receptor type 2; A2AR, adenosine receptor 2A; ENT1, equilibrative nucleoside transporter 1; 5-HT1AR, serotonin receptor 1A; D2, D2-like dopamine receptor; GPR55, G-protein-coupled receptor 55; PPARγ, peroxisome proliferator-activated receptor gamma; GABAAR, gamma-aminobutyric acid receptor A; MOR, mu opioid receptor; DOR, delta opioid receptor; TRPV1, transient receptor potential vanilloid type 1; TRPV2, transient receptor potential vanilloid type 2.
Figure 1. Some molecular targets of CBD’s actions in the brain. CBD acts as a negative allosteric modulator of CB1, MOR, and DOR receptors; as an inverse agonist of CB2 receptors; as an agonist of A2A, 5-HT1A, D2, PPARү, TRPV1, TRPV2, and GABAA receptors; as an antagonist of GPR55 receptor; and as an inhibitor of ENT1. CBD, cannabidiol; CB1, cannabinoid receptor type 1; CB2, cannabinoid receptor type 2; A2AR, adenosine receptor 2A; ENT1, equilibrative nucleoside transporter 1; 5-HT1AR, serotonin receptor 1A; D2, D2-like dopamine receptor; GPR55, G-protein-coupled receptor 55; PPARγ, peroxisome proliferator-activated receptor gamma; GABAAR, gamma-aminobutyric acid receptor A; MOR, mu opioid receptor; DOR, delta opioid receptor; TRPV1, transient receptor potential vanilloid type 1; TRPV2, transient receptor potential vanilloid type 2.
Ijms 23 16181 g001
Figure 2. Examples of the chemical structures of some natural cannabinoids (green box) and of synthetic CBD derivatives (brown box) that recently have attracted attention in the search for anti-seizure medications. CBDVA, cannabidivarinic acid; CBDA, cannabidiolic acid; CBDV, cannabidivarin; HU-308, pinene dimethoxy-DMH-CBD derivative; O-1602, a synthetic cannabinoid; H2CBD, 8,9-dihydrocannabidiol.
Figure 2. Examples of the chemical structures of some natural cannabinoids (green box) and of synthetic CBD derivatives (brown box) that recently have attracted attention in the search for anti-seizure medications. CBDVA, cannabidivarinic acid; CBDA, cannabidiolic acid; CBDV, cannabidivarin; HU-308, pinene dimethoxy-DMH-CBD derivative; O-1602, a synthetic cannabinoid; H2CBD, 8,9-dihydrocannabidiol.
Ijms 23 16181 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zavala-Tecuapetla, C.; Luna-Munguia, H.; López-Meraz, M.-L.; Cuellar-Herrera, M. Advances and Challenges of Cannabidiol as an Anti-Seizure Strategy: Preclinical Evidence. Int. J. Mol. Sci. 2022, 23, 16181. https://doi.org/10.3390/ijms232416181

AMA Style

Zavala-Tecuapetla C, Luna-Munguia H, López-Meraz M-L, Cuellar-Herrera M. Advances and Challenges of Cannabidiol as an Anti-Seizure Strategy: Preclinical Evidence. International Journal of Molecular Sciences. 2022; 23(24):16181. https://doi.org/10.3390/ijms232416181

Chicago/Turabian Style

Zavala-Tecuapetla, Cecilia, Hiram Luna-Munguia, María-Leonor López-Meraz, and Manola Cuellar-Herrera. 2022. "Advances and Challenges of Cannabidiol as an Anti-Seizure Strategy: Preclinical Evidence" International Journal of Molecular Sciences 23, no. 24: 16181. https://doi.org/10.3390/ijms232416181

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