Next Article in Journal
Molecular Design of Interfaces of Model Food Nanoemulsions: A Combined Experimental and Theoretical Approach
Next Article in Special Issue
Air-Frying Is a Better Thermal Processing Choice for Improving Antioxidant Properties of Brassica Vegetables
Previous Article in Journal
How the Competition for Cysteine May Promote Infection of SARS-CoV-2 by Triggering Oxidative Stress
Previous Article in Special Issue
Effects of UV Stress in Promoting Antioxidant Activities in Fungal Species Тrametes versicolor (L.) Lloyd and Flammulina velutipes (Curtis) Singer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cannabidiol: Bridge between Antioxidant Effect, Cellular Protection, and Cognitive and Physical Performance

1
Department of Pharmacology and Clinical Pharmacy, Faculty of Pharmacy, George Emil Palade University of Medicine, Pharmacy, Science and Technology of Târgu Mureș, 540139 Târgu Mureș, Romania
2
Doctoral School of Medicine and Pharmacy, I.O.S.U.D, George Emil Palade University of Medicine, Pharmacy, Science and Technology of Târgu Mureș, 540139 Târgu Mureș, Romania
3
Department of Biochemistry, Faculty of Pharmacy, George Emil Palade University of Medicine, Pharmacy, Science and Technology of Târgu Mureș, 540139 Târgu Mureș, Romania
*
Author to whom correspondence should be addressed.
Antioxidants 2023, 12(2), 485; https://doi.org/10.3390/antiox12020485
Submission received: 13 January 2023 / Revised: 8 February 2023 / Accepted: 13 February 2023 / Published: 14 February 2023

Abstract

:
The literature provides scientific evidence for the beneficial effects of cannabidiol (CBD), and these effects extend beyond epilepsy treatment (e.g., Lennox–Gastaut and Dravet syndromes), notably the influence on oxidative status, neurodegeneration, cellular protection, cognitive function, and physical performance. However, products containing CBD are not allowed to be marketed everywhere in the world, which may ultimately have a negative effect on health as a result of the uncontrolled CBD market. After the isolation of CBD follows the discovery of CB1 and CB2 receptors and the main enzymatic components (diacylglycerol lipase (DAG lipase), monoacyl glycerol lipase (MAGL), fatty acid amino hydrolase (FAAH)). At the same time, the antioxidant potential of CBD is due not only to the molecular structure but also to the fact that this compound increases the expression of the main endogenous antioxidant systems, superoxide dismutase (SOD), and glutathione peroxidase (GPx), through the nuclear complex erythroid 2-related factor (Nrf2)/Keep1. Regarding the role in the control of inflammation, this function is exercised by inhibiting (nuclear factor kappa B) NF-κB, and also the genes that encode the expression of molecules with a pro-inflammatory role (cytokines and metalloproteinases). The other effects of CBD on cognitive function and physical performance should not be excluded. In conclusion, the CBD market needs to be regulated more thoroughly, given the previously listed properties, with the mention that the safety profile is a very good one.

Graphical Abstract

1. Introduction

In light of the increased interest in natural products and the emergence of a self-medication attitude among the population, more awareness should be raised around a relatively new concept related to the natural compounds of various origins (fruits, vegetables, algae) included in food or in various pharmaceutical forms, with the aim of providing health benefits (treatment or prevention). Namely, these are substances considered nutraceuticals [1]. Consequently, the nutraceuticals market follows a similar upward trend [2,3]. There is a degree of controversy over nutraceuticals, as these are neither nutritious nor pharmacologically active, yet are assumed to be safe. In addition, the regulation of these compounds differs from country to country, and when it comes to other substances that may seem illegal at first glance (e.g., cannabidiol) their marketing may even be banned because these are considered substances of abuse. Since the regulation of the use of nutraceuticals and/or cannabidiol is not the subject of this article, for a review see [4].
The use of the cannabis plant for various purposes represents an increasingly frequent topic, and at the same time increasingly controversial, given its composition and rising rate of consumption. In various oriental cultures (e.g., India, China) the use of this plant is mentioned since ancient times for medicinal purposes, to treat various conditions such as pain, constipation, epilepsy, or bacterial conditions such as malaria or tuberculosis [5]. Its use is also mentioned in other conditions, this time in the psychiatric area, such as depression, and anxiety [6,7], and also as a tranquilizer and hypnotic [8,9]. Despite these considerations, the use of cannabis was considered an illegal practice, given the fact that compounds specific to the conditions in question were developed.
The cannabis plant has a complex composition that also includes terpenes, amides, carbohydrates, fatty acids, phytosterols, and specific compounds considered cannabinoids (cannabigerol (CBG), cannabichromene (CBC), cannabinol (CBN)) [10,11,12].
There are three species of cannabis and cultivars differ widely in their Δ9-THC and CBD levels as described below:
  • Type I, with high content of Δ9-THC;
  • Type II, containing different ratios of both CBD and Δ9-THC (predominantly CBD);
  • Type III, with high content of CBD, and low content in Δ9-THC;
In the 1980s, following some in vitro experiments, the cannabinoid receptor type 1 (CB1) was discovered, and then, in 1990, it was also identified in the human brain. Additionally, in this period, the existence of another receptor subtype was demonstrated, namely the CB2 receptor. Following these discoveries, endogenous compounds that can stimulate these receptors were also isolated, such as anandamide (N-arachidonylethanolamine AEA, Sanskrit name, which means internal bliss, isolated from pig brain) and 2-arachidonylglycerol (2-AG, isolated from rat brain and canine gut) [13]. Therefore, the discovery of the two components (receptors and endocannabinoids) as well as other enzyme components (diacylglycerol lipase (DAG lipase), alpha/beta-hydrolase domain containing 6/2-arachidonoylglycerol hydrolase (ABHD6), N-arachidonoyl phosphatidyl ethanolamine phospholipase D (NAPE-PLD), monoacyl glycerol lipase (MAGL), fatty acid amino hydrolase (FAAH)) involved in synthesis or metabolism, led to the discovery of the endocannabinoid system [14,15,16,17,18].
Thus, the endocannabinoid system has an important role in the development of the central nervous system (CNS) and the neural circuits and also appears to be interconnected with other signaling pathways involved in the regulation of multiple neural functions, cognition, control of motor function, modulation of pain, and/or eating behavior [19,20,21]. Because this aspect is beyond the scope of this review, the presentation of cannabinoid receptors, CB1 and CB2 will be briefly presented in this paragraph. These two receptors are coupled with Gi/o proteins (GPCRs) with an inhibitory role and have the ability to reduce the activity of adenylate cyclase and some voltage-dependent calcium channels, but also to increase the enzymatic activity of mitogen-activated protein kinases (MAPKs) and inwardly rectifying potassium channels (GIRKs) [22,23].
As a localization, CB1 receptors are found in the liver, adipose tissue, skin, and predominantly in the CNS (cortex, hippocampus, caudate-putamen, substantia nigra pars reticulata, globus pallidus, cerebellum, spinal cord), especially in GABAergic interneurons, and glutamatergic, cholinergic, glycinergic, adrenergic, opioid, cholecystokinin, and finally serotoninergic, in the synaptic endings [20,24]. These receptors are also found on astrocytes, oligodendrocytes, and microglia, but their exact role is not defined [21]. CB2 receptors are located in the cells of the immune system [25], in microglia [26], in the vasculature [27], in pancreatic beta cells [28], at the central level, in case of the existence of a pathological condition (neuroinflammation) [29], and are also involved in the processes of atherosclerosis and bone remodeling [30]. It is important to mention that the endocannabinoid system plays a key role as an inhibitory/stimulating neuromodulator, yet its mechanism is not completely defined. Targeted areas are located both peripherally and centrally, especially in brain regions involved in affective behavior. However, in psychiatric conditions, dysregulation in endocannabinoid-mediated cell signaling could be incriminated. A hyperfunction of the endocannabinoid system associated with a massive expression of CB1 receptors was observed in the case of suicidal patients diagnosed with depression. In contrast, patients diagnosed with major depression and anxiety had lower receptor density. Concomitantly, increased AEA, peripheral brain-derived neurotrophic factor (BDNF) in physically active female patients is observed, which suggests the important role of physical exercise in neuroplasticity (and depression), directly linked to the endocannabinoid system, for a review see [31,32,33,34,35].
Regarding the mechanism of action of the endocannabinoid system, it is based on retrograde signaling. Thus, postsynaptic activity causes the production of endocannabinoids that travel to the presynaptic level and stimulate CB1 receptors with the consecutive inhibition of neurotransmitter release. There are also data in the literature characterizing non-retrograde signaling with modulation of neuronal function via transient receptor potential vanilloid receptor type 1 (TRPV1) and CB1 [36].
Based on the previously mentioned considerations, the purpose of this review is to create an image of the mechanisms by which nonpsychoactive compounds, such as CBD, regulate oxidative status, cognitive function, and physical performance are considered nutraceutical molecules.

2. Cannabidiol: Pharmacological Targets

In 1970 in Israel, Raphel Mechoulam discovered other compounds, called cannabinoids, that interfere with the activity of Δ9-THC, especially CBD, which is found in considerable amounts in Cannabis sativa. This molecule consists of a core of phenolic terpenes, consisting of 21 carbon atoms. CBD is one of the cannabinoids holding an important therapeutic potential among the multitude of compounds with null psychoactive properties. This derives from preclinical studies, which attest to the presence of antioxidant, neuro- and cardioprotective, anxiolytic, and anti-inflammatory characteristics [37,38,39,40]. Clinical trials involving CBD and monitoring its effect are scarce. At the same time, many of the preparations containing this compound are not approved by the authorities; the only preparation accepted by the Food and Drug Administration (FDA), which is found on the market, is Epidiolex® (GW Pharmaceuticals, Cambridge, UK). Despite the fact that cannabinoid receptors are widely distributed in the body, CBD acts as a negative allosteric modulator at the CB1 level [41,42,43,44] and as an inverse agonist on CB2 receptors, through which the anti-inflammatory, reduction in self-administration of cocaine, and antiepileptic actions can be explained [41,45,46,47]. Another proposed mechanism of action is represented by the inhibition of FAAH, with the consequent increase in AEA [48].
Even if the action on cannabinoid receptors is limited, other receptors have been identified through which the pharmacodynamic properties of CBD can be explained. Thus, the affinity and agonistic activity of the TRPV1 receptor was described [48]. The literature also shows that CBD activates other vanilloid receptor subtypes (TRPV1, TRPV2, TRPV3, and TRPV4), transient receptor potential ankyrin 1 (TRPA1) [49,50,51], but the neuroprotective activity is limited to the TRPV1 receptor [52]. The CBD-induced effects which are not directly related to CB1 and/or CB2 receptors, but with transient receptor potential channels are an important topic to be discussed since it has been demonstrated that the CBD activation of receptors such as TRPV1 and TRPV2 produces several effects. Therefore, animal studies show anxiolytic, analgesic, anti-inflammatory, and anticonvulsant effects correlated with TRPV1 receptor activation. Studies carried out on different cell cultures demonstrated a reduction in inflammatory processes, consequent to diminishing levels of IL-6, IL-8, and TNF-α. CBD is the phytocannabinoid that seems to have the highest affinity to the TRPV2 receptor, and consequently the highest potency, having a pro-apoptotic effect. In contrast, the link between CBD and the TRPV3 receptor is not sufficiently studied, with little data available in the literature. Besides TRP channels, CBD acts as an allosteric modulator of GABAA receptors and also presents an inhibitory action over Na+ channels, for a review see [53].
In addition, studies on rodents state that the stimulation of 5-hydroxytryptamine subtype 1A (5-HT1A) receptors is responsible for alleviating anxiolytic behavior, fear-associated freezing behavior, reducing the autonomic stress response [54], and also for inhibiting 5-HT3 receptors. Other therapeutic targets include protein structures involved in Ca2+ homeostasis, G Protein-Coupled Receptors 55 (GPR55), GPR18, and GPR119 receptors (antagonistic action), and GPR3, GPR6, and GPR12 (inverse agonistic action), glycinergic receptors α1 and α1β, peroxisome proliferator-activated receptor gamma (PPARγ), adenosine receptors A1 and A2, lipoxygenase and cyclooxygenase type 2 (COX2) [55,56].
Regarding the therapeutic purpose of CBD, its use extends to multiple psychiatric and neurological pathologies. Perhaps the most frequently mentioned therapeutic purpose of CBD is anxiety, its anxiolytic effect being demonstrated in various clinical studies [57,58,59,60,61]. As presented throughout this review, the endocannabinoid system acts as an inhibitor through retrograde signaling at synaptic levels of GABAergic and glutamatergic systems and thus can have both anxiolytic and anxiogenic effects [62,63]. Nevertheless, the influence on other neurons should not be excluded. Such neurons are the ones that express the D1 dopaminergic receptors, though the exact connection is not completely elucidated. The D1 receptors are located in CNS regions (hippocampus, amygdala, nucleus accumbens) that are responsible for memorization processes and aversive behavior, and the alteration of the endocannabinoid system can affect the abovementioned processes and behaviors [64,65].
In this context of psychiatric diseases, the use of CBD for schizophrenia is also described in the literature [66,67,68]. Concrete examples of CBD involvement in schizophrenia are found in studies that indicate that CBD regulates dopaminergic activity in the mesolimbic system, alleviating the behavioral effects, though the mechanism through which D2 receptors expression is regulated is not clearly defined [69,70,71]. The possible partial agonist activity of CBD towards D2 receptors must be mentioned [72]. Unlike classical compounds used for the treatment of schizophrenia, CBD presents the advantages of improving cognitive impairments and a better safety profile. Moreover, a study conducted by Stark et al. on an animal model of schizophrenia (methylazoxymethanol acetate), revealed that CBD, administered in doses that prevent the disease onset can normalize the upregulation of D3 receptors in different regions of the brain by regulating the expression of this receptor [73,74].
Regarding neurological disorders, most pieces of evidence support the use of CBD in epilepsy, which is why the only existing product in CBD-based therapy, Epidiolex®, was approved [75,76,77]. Beyond this condition, CBD’s effects on Parkinson’s and Huntington’s disease have been intensively researched. The obtained data support the improvement of Parkinsonian symptoms, but more research in this field is needed [78,79]. In addition, preclinical studies in vitro and on animal models (rodents) support and demonstrate a beneficial effect of CBD in Alzheimer’s disease (AD), with the reduction in neuroinflammation and at the same time the promotion of neurogenesis. The proposed mechanisms are related to the inhibition of the production of pro-inflammatory cytokines (IL-1β, IL-6, TNF-α) by blocking Nuclear Factor-Κappa B (NF-κB), reducing the level of reactive oxygen species (ROS), and reducing β-amyloid peptide synthesis and apoptosis, as a result of PPARγ activation [80,81,82].

3. Link between Cannabidiol and Oxidative Stress

This beneficial effect of CBD on pathological conditions (mainly neurodegenerative diseases) has been and is increasingly being exploited, which is why it is interesting to see the mechanisms by which this compound modulates the oxidative status, considering that these conditions are characterized among others by the presence of oxidative stress. The antioxidant potential is a result of the molecular structure of CBD, which converts reactive species into compounds with weaker or inert reactivity. The aromatic nucleus (which confers the molecule electrophilic character) and the hydroxyl group on the phenolic nucleus give the antioxidant properties, as demonstrated by Hampson et al., by the fact that CBD behaves similarly to known antioxidants (vitamin E, butylated hydroxytoluene, and BHT) [83,84,85].
As we can infer from the above paragraph, the idea is suggested that CBD exerts its antioxidant activity both through a direct mechanism (which is directly related to the molecular structure) and indirectly through influencing some molecular mechanisms involved in the regulation of redox homeostasis. Thus, CBD decreases the production of ROS primarily through the property of chelating transition metal ions that enter the Fenton reaction, from which free radicals result [86,87]. Associated with this mechanism, CBD increases the gene expression of the main endogenous antioxidant systems, superoxide dismutase (SOD) and glutathione peroxidase (GPx), via the nuclear erythroid 2-related factor (Nrf2)/Keap1 complex [88,89,90,91,92]. Additionally, CBD prevents the depletion of Zn and Se, which are known to have an important involvement in the enzymatic activity of SOD and GPx, respectively [93], and this antioxidant activity is assumed to be greater compared to vitamin E and vitamin C [94]. One of the most common reactions involving oxidative stress is the peroxidation of lipids, resulting in malondialdehyde (MDA) and 4-hydroxynonenal (4-HNE), compounds that provide electrophilic character and can bind to DNA, lipids, proteins and may decrease the ratio of reduced glutathione/oxidized glutathione (GSH/GSSG) [95,96]. In addition, the presence of these molecules, along with the changes they induce in other structures (formation of sulfoxides and disulfides from cysteine, kynurenine from tryptophane, tyrosine nitration), alters cellular signaling [97,98,99]. The 3-hydroxykynurenine/kynurenic acid ratio is a marker of neurotoxicity, also characterizing the oxidative status [100]. Many studies that attest to the antioxidant effect of CBD manifested in the brain are reported in the literature. This effect is demonstrated by the low level of MDA resulting from hypoxia and reperfusion conditions, respectively, and the decreased level of degraded proteins in the case of the administration of compounds with oxidant potential [101,102,103]. On the same note, in an AD model, CBD reduced the level of polyunsaturated fatty acids (PUFA) cyclization products [104].
The cannabinoid receptor-mediated antioxidant mechanism of CBD has received limited attention due to the fact that the stimulation of CB1 receptors increases the production of ROS and TNF-α [105]. On CB2 receptors, CBD exerts a weak agonist action, but data suggest an inverse agonist action [106,107]. The thing that drew attention to the latter receptor is the result of its activation, exerting opposite effects of the stimulation of CB1; namely, it produces a decrease in ROS and TNF-α [105]. The involvement in maintaining redox homeostasis is also carried out through other receptor pathways. For example, a link between CBD binding to TRPV1 and oxidative stress is suggested, as ROS can influence the activity of this receptor by oxidizing thiol groups [108,109]. Through TRP, CBD regulates Ca2+ homeostasis, important in regulating the inflammatory response, via the nuclear factor of the activated T cells (NFAT) pathway [110,111]. Of major importance is that CBD stimulates PPARγ, the receptor through which the transcription of pro-inflammatory proteins (COX2) is inhibited. Additionally, it inhibits other factors involved in inflammatory response signaling (NF-κB). In addition, PPARγ cooperates with Nrf2 and demonstrates its cytoprotective properties by binding to the specific region of the genes encoding the antioxidant proteins, catalase (CAT), Mn-SOD, and heme-oxygenase-1 (HO-1). Last but not least, it is important to state that PPARγ expression is controlled by Nrf2, by binding the latter to the antioxidant response element (ARE) sequence [112,113,114,115]. CBD also demonstrates its antioxidant properties by inhibiting the degradation of AEA and 2-AG, which are otherwise known to stimulate PPARγ [116]. An indirect way for CBD to reduce the generation of ROS is represented by the GPR55 receptor, towards which it has an antagonistic behavior and can thus modulate the level of Ca2+, depending on the excitability of the neuronal cell, which is important in the case of pathologies such as epilepsy or AD [117]. Another receptor to which CBD shows affinity is the 5-HT1A membrane receptor, through which it limits the oxidative changes resulting from the lipid peroxidation reaction [118,119]. Moreover, the fact that CBD has the ability to activate adenosine A2A receptors should not be excluded. Following the activation of these receptors, the degree of oxidative damage can be improved as a result of reperfusion and it reduces the level of vascular cell adhesion molecules (VCAM-1), which gives anti-inflammatory properties in the case of multiple sclerosis [120]. Figure 1 shows the general mechanisms underlying the antioxidant properties.
Apart from the mechanisms proposed and described above, other data attesting to the antioxidant effect of CBD are presented in the literature. Thus, the induction of HO-1 by CBD [121,122,123], and the regulation of the GSH/GSSG ratio, by increasing the level of GSH, as well as SOD and GPx, is demonstrated in various experimental models, affecting both the CNS and other organs [93,101]. On the same note, it is stated that CBD decreases the expression of several isoforms of reduced nicotinamide adenine dinucleotide phosphate oxidase form (NOX) in various experimental models testing antioxidant properties [124,125,126]. Additionally, CBD reduces the general stress of reactive nitrogen species (RNS) and decreases the expression of Fas ligand (after binding to the specific receptor initiates apoptosis) and caspase-3 [93]. Last but not least, the reduction in ROS through CBD also protects the other non-enzymatic antioxidant mechanisms, represented by vitamins, with the ultimate goal of improving the oxidative status of the entire organism. One thing that should not be excluded is the oxidative status that the cell possesses, knowing that this is not always a negative factor [99], as suggested in a study led by Massi et al., who noted that CBD exerts its anti-proliferative properties depending on the level of ROS; thus, in human glioma cells, they show pro-oxidant properties, whereas, in non-transformed glial cells, they do not behave in this way [127]. These pro-oxidant properties are also dose-dependent and, paradoxically, may enhance the activity of antioxidant systems, which is important, especially in neurons [128].

4. Cannabidiol Involvement in Neurodegeneration and Cellular Protection

The link between neurodegeneration and oxidative stress (regardless of its origin, mitochondria, arachidonic acid metabolism, nitric oxide synthase, xanthine oxidase (XO), NOX, etc.) is a long-debated topic and the involvement of reactive species in these pathologies is generally accepted. A question that derives from the previous topic is whether ROS are at the origin of neurodegeneration, the generated reactive species are responsible for disease progression, or both. Moreover, glutamate-induced neurotoxicity is not only limited to the presence of Ca2+ and the activation of caspases, but also through the inhibition of the cystine/glutamate antiport (xc), with the consequent decrease in GSH and the accumulation of ROS [129]. CBD reduces the degree of ROS-mediated neurotoxic damage in the cortical region, regardless of its origin, whether it is generated by N-methyl-D-aspartate (NMDA) receptors, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) or kainate, which suggests either an inhibitory action on these receptors or an action on downstream proteins [112]. At the same time, previous studies have suggested and even stated that CBD has an antioxidant capacity similar to BHT, and in addition, it does not promote protumor effects, thus making it one of the most promising agents in this regard [83]. For example, in Parkinson’s disease (PD), the region affected by ROS is considered to be the substantia nigra pars compacta (SNpc), where the highest density of dopaminergic neurons is found, where the basal level of ROS is higher than in other areas of the brain, as a result of the intense metabolism of dopamine. Thus, these neurons are susceptible to the presence of oxidative stress, the decrease in their number causing the appearance of motor, but also cognitive, memory, and learning symptoms [99]. Thus, the addition of CBD in therapy increases the number of unaffected dopaminergic neurons, also suggesting the importance of the HO-1 system in neuroprotection [123]. At the same time, the continuous administration of CBD in preclinical studies led to the observation of decreased levels of MDA at the cortical and striatal level, as well as increased activity of antioxidant mechanisms in other areas (cortex and striatum) [130]. Similar effects could also be observed in various preclinical models of PD [131,132].
The presence of neuroinflammation within controlled limits is imperative for the restoration of neuronal tissue, but such a particular state continues to be associated with the development of neurodegenerative diseases [133]. Without a well-known mechanism, the density of CB2 receptors in microglia increases under oxidative stress conditions, as seen in multiple sclerosis, AD, and PD. The activation of these receptors causes the inhibition of microglia activity, hence the reduction in neuronal toxicity mediated by ROS and RNS [134,135], by decreasing glial fibrillary acidic protein expression (GFAP) [136]. At the same time, CB2 activation improves blood–brain barrier (BBB) function and reduces inducible nitric oxide synthase (iNOS) expression as a result of extracellular signal-regulated kinase 1/2 inhibition (ERK 1/2) [137]. It is suggested that CBD exerts its neuroprotective and antioxidant properties in a unique manner that is independent of CB2, TRPV1, or PPARγ receptors. In experimental studies aimed at testing the effect of CBD in AD, it reduced iNOS activity, decreased tau protein hyperphosphorylation, MDA levels, and caspase 3 activity [138]. In the case of Huntington’s disease, 3-nitropropionic acid (3-NP) is used to generate a model as close as possible to the pathology. The administration of CBD led to the rescue of neurons, independent of CB2 receptors, the most likely reason being the antioxidant effect [139].
Other mechanisms involved in the control of inflammation are the inhibition of NF-κB and also of the genes that encode the expression of molecules with a pro-inflammatory role (iNOS, COX2, cytokines, and metalloproteinases) [140]. Additionally, the inhibitory control of NF-κB can be achieved by reducing the phosphorylation of kinases involved in the transcription of this factor such as p38 mitogen-activated protein kinase (p38 MAPK) [137]. Additionally, in the same category of mechanisms with neuroprotective potential are included 5-HT1A receptors, adenosine reuptake inhibition, and the WNT/β-catenin signaling pathway, which has an important role in β-amyloid (Aβ)-induced glycogen synthase kinase-3 beta activation (GSK-3β) and tau hyperphosphorylation [141,142].
Several studies have focused on investigating how CBD works in neurodegenerative diseases. Thus, in animal models (rats) of AD, CBD treatment prevented cognitive impairments and reduced the risk of progression via PPARγ and SOD. It is also not excluded that AEA, whose level increases in the presence of CBD, and reduces the formation of β-amyloid, as suggested in cell culture studies, also has an important role. As previously stated, other studies also support the inhibition of the GSK-3β protein complex through TRPV1-mediated phosphoinositide-3-kinase/protein kinase B (PI3K/Akt) activation and protection of neuronal plasticity by reducing ROS and inflammation [143,144,145,146].
Literature reports suggest that the accumulation of Aβ destabilizes the redox balance, generates massive amounts of ROS (promotes lipid peroxidation reactions, protein, and DNA oxidation) [147], and stimulates pro-inflammatory reactions, which ultimately result in disease progression [148]. On the same note, NF-κB, a factor sensitive to changes in oxidative homeostasis, which is activated by the family of stress-activated protein kinases (SAPKs), which includes p38 MAPK, regulates the transcription of pro-inflammatory factors and the immune response, but also induces iNOS in neurons affected by the presence of Aβ, both in cell cultures and in post-mortem studies on the brains of AD patients [149,150,151]. Thus, it is demonstrated that CBD exerts an inhibitory effect on both NF-κB and p38 MAPK, ultimately determining the reduction in iNOS expression with the limitation of the disruptive effects of oxidative stress, and at the same time being responsible for the protection of PC12 cells from the negative effect of Aβ [137]. The idea is strengthened by the fact that CBD protects PC12 neurons in an antioxidant manner and also reduces the hyperphosphorylation of tau proteins through the WNT/β-catenin pathway [112], diminishing the inflammatory markers generated by Aβ, such as GFAP and IL-1β, these effects being the sum of complex and dynamic processes [152,153,154,155]. Another cause of the vicious circle created by ROS and the progression of Aβ plaque aggregation (which generates in vitro, H2O2, and O 2 ), is the presence of increased concentrations of transition metal ions (Cu2+, Fe3+) in the brain of patients (postmortem) diagnosed with AD [156,157,158]. In this sense, the chelating properties of CBD, along with the antioxidant ones, can constitute an alternative treatment [159].
In the case of cell cultures (PC12 cell line), treated with MPP+, the inclusion of CBD under these conditions resulted in increased viability of these cells and also the expression of the axonal protein, Growth Associated Protein 43 (GAP-43). CBD also stimulates neurite formation but, in a manner, is closely related to the tropomyosin kinase A receptor (TrkA) and not to the nerve growth factor (NGF) [157]. Depending on the time of administration, i.e., the time elapsed since the injury occurred, CBD has the ability to restore the dopamine level, as demonstrated in the case of the injection of 6-hydroxydopamine (6-OHDA), and is proposed as a mechanism, the activation of the antioxidant systems and CB2 receptors, respectively, TRPV1 expressed at nigrostriatal level, as well as COX2 inhibition [132].
Preclinical models based on 3-NP induce striatal toxicity, but apparently, CBD can increase gamma-aminobutyric acid (GABA) levels and protect GABAergic neuron projections in the SN, increase BDNF, reduce ROS, and also reduce cellular signaling via the PI3K/Akt pathway, independent of CB2 or TRPV1 receptors, whereas the reduction in iNOS is mediated by CB2 [102,139,160,161,162]. It is very likely that in this case too, the beneficial effects of CBD are mediated through the Nrf2/ARE pathway, as it is known that in the case of 3-NP intoxication, the expression of Nrf2 is increased [163,164,165].
An important aspect to specify is that oxidative stress is not only found in neurodegenerative conditions but also in particular psychiatric conditions. Thus, regarding ROS, it doesn’t need to constitute the origin of these conditions, but rather the aggravation of symptoms and brain functions. In this sense, mental states characterized by stress induce ROS by altering the balance of excitatory neurotransmitters (adrenaline, glutamate), serotonin, and GABA, which also modulate the immune response and maintain the body’s homeostasis and the functionality of the hypothalamic-pituitary-adrenal axis (HPA) [166]. For this purpose, in post-traumatic stress syndrome (PTSD), CBD finds applicability by blocking FAAH and increasing available AEA [167,168,169]; in depression, CBD alleviates symptoms by stimulating serotoninergic neurotransmission, mediated by 5-HT1A, and by activating the BDNF-TrkB complex [170,171,172], whereas in anxiety and fear, these beneficial effects are mediated by the same 5-HT1A receptor, but also by GABAA receptors, respectively, by inhibiting the enzymatic activity of iNOS and FAAH [173].
These positive effects exerted by CBD (antioxidant, neuroprotective) have also been investigated outside the context of neurodegenerative diseases. Thus, in the case of toxicity mediated by alcohol consumption (in the rat binge alcohol model), CBD limited the pathological changes in the hippocampal region, the proposed mechanism is based on antagonizing the effect of glutamate, beyond the stimulation of NMDA receptors (by glutamate), being demonstrated neuronal decline by alcohol, through the influence on mitochondria, resulting in ROS [174]. Additionally, in a cardiotoxicity model using doxorubicin, CBD limited this effect by enhancing mitochondrial complex I activity and also GPx activity [175]. The same antioxidant behavior of CBD could also be observed in a cisplatin-induced nephrotoxicity model [124].
Another area of interest in the use of CBD is the effect of this compound on the skin. Thus, it is known that ultraviolet radiations (UV), UVA, and UVB produce a redox imbalance, which results in the degradation of the normal structure of the skin with possible consequences such as photoaging and photocarcinogenesis [176,177]. Apparently, at this level as well (keratinocytes), CBD can induce Nrf2 and decrease ROS. In addition, an increase in the level of thioredoxin and the activity of thioredoxin reductase is observed, a system through which the activity of apoptosis-regulating kinase (ASK-1) is inhibited, thus protecting cells from apoptosis induced by ROS and irradiation [178,179,180,181]. Moreover, CBD reduces the signaling of apoptotic pathways by restoring Ca2+ homeostasis at the mitochondrial level [166]. Under physiological conditions, CBD activates A2 receptors with a reduction in NF-κB activity and a consequent decrease in TNF-α [182,183]. Another proposed mechanism is that through which CBD stimulates the formation of an anti-inflammatory prostaglandin, 15d-PGJ2, which regulates COX activity and also the redox balance by facilitating the dissociation of Nrf2 from Keap1, an effect that promotes HO-1 transcription and finally the rescue of irradiated keratinocytes [184,185,186,187]. Related to Nrf2, it seems that continuous and prolonged activation of this factor increases the risk of malignancies and creates an environment favorable to the development, proliferation, and creation of resistance to chemotherapy and radiotherapy through antioxidant mechanisms [188,189] At the same time, it is suggested that CBD reduces the transcription of Nrf2, manifesting in this way the protection of cells [190].
In addition to influencing Nrf2 and HO-1, various studies state that CBD has an inhibitory effect on BTB Domain and CNC Homolog 1 (BACH1), a transcription factor involved in ROS generation [90,121,188,191,192,193]. This BACH1 can be considered as a functional antagonist of Nrf2, in the absence of oxidative stress, by binding to Maf recognition elements (MAREs), regions indispensable for the coding of genes with an antioxidant role, following that the same Nrf2 will restore the levels of BACH1, decreased due to oxidative stress [194] (see Figure 2). Other studies claim that, in fact, in keratinocytes, CBD is a weak activator of Nrf2 and rather a strong inhibitor of BACH1 [193].

5. Cannabidiol and Cognitive Performance Related to Oxidative Stress

In the absence of a condition affecting cognitive function (attention, memory, ability to use memories), CBD does not show an improvement in this function. However, in the case of pathological situations that induce cognitive impairments (AD, neuroinflammatory conditions, ischemic states, epilepsy, schizophrenia, hepatic encephalopathy), studies show an improvement in these events [195,196,197,198,199]. In addition, in studies on human subjects, cannabis users with increased concentration in CBD, no memory impairments were reported [200]. Most often, glutamate toxicity is described in the inability to form and use memories, which triggers irreversible processes resulting in lipid peroxidation resulting in ROS production, vascular microlesions, and last but not least, cell death [201]. According to this hypothesis, it is suggested that CBD administration alleviates glutamate-induced toxicity by reducing available Ca2+. In addition, alongside this effect, others such as the antioxidant and anti-inflammatory effects can be counted, and it is also considered that the combination of THC and CBD is more effective than CBD alone [202,203,204], although there are studies that state that CBD protects memory function of the negative effects of THC [55,205] being observed that in the case of people who consume cannabis with a high content of THC and low content of CBD, they have a smaller hippocampal region, associated with low cognitive performance [206]. In a study conducted by Friedman et al., it is considered that in the case of most brain traumas, they are moderate, and cognitive impairment can be improved by applying CBD. Regarding the size of the lesions, CBD has the ability to limit this and also improve vestibulomotor function and learning as well as memory. On the other hand, what cannot be stated with certainty is whether the beneficial effects are the result of the anti-inflammatory or antioxidant effect [207].
Another proposed mechanism for improving cognitive capacities is based on antagonizing the pro-oxidant effect of transition metal ions through their chelation, a topic discussed in the previous section. In the case of the presence of pathologies affecting the brain, changes in memory function are most often observed. A study led by Fagherazzi et al. suggests that CBD acts mainly in the memory consolidation phase and not in the acquisition and retrieval phase [208]. Another study reported that CBD administration was beneficial in reversing cognitive deficits in sepsis, and this is due to its antioxidant properties [193]. Regarding this particular condition (sepsis), it is shown that ROS are generated leading to oxidative damage not only in the periphery but also in the CNS [209,210]. In this sense, the administration of CBD, due to its antioxidant properties, limits these lesions in the hippocampus, in the initial stages of sepsis, and prevents memory loss through an antioxidant mechanism, in animal models [211]. In another study, this time focusing on patients with schizophrenia, it is observed that it does not improve cognitive function, based on the consideration that the effect of CBD is crucial during critical periods of the disease, and not in chronic schizophrenia [212]. The data on this subject are quite contradictory, so Blaes et al. presented a beneficial effect of using cannabis in inhaled form (smoke), and the proposed mechanism is based on prefrontal cortex pyramidal neurons. Thus, in these neurons, GABAergic neurotransmission plays an important role [213]. Additionally, there are CB1 receptors in this area, whose primary role is to inhibit the release of GABA and promote working memory, thus preventing the suppression of pyramidal neurons by GABA [214]. This beneficial effect observed in the case of THC may be due to differences in pharmacokinetics, given the fact that by inhaling the smoke, a smaller amount of the substance is absorbed [215]. Another paper in the literature on hepatic encephalopathy shows that motor and cognitive functions are improved through the 5-HT1A and A2A receptor pathways [160].
An immunological mechanism with implications in AD and at the same time in cognitive performance is related to IL-33, a centrally expressed cytokine, especially in the conditions of altered BBB function. Along with this marker, the triggering receptor expressed on myeloid cells 2 (TREM2), is another important factor in the development of AD, suggesting that the latter has a neuroprotective role. In preclinical studies on mice, a considerable increase in IL-33 and TREM2 could be observed after CBD administration, with a parallel decrease in Aβ production, while also improving contextual memory [215]. Beneficial effects on memory could also be noted in an animal model of PD induced by reserpine treatment [216]. In the case of the study conducted by Razavi et al., CBD treatment during methamphetamine-induced abstinence resulted in improved spatial memory [217]. These data are evidence that CBD intervenes in cognitive processes and through antioxidant mechanisms, knowing the potential of amphetamines to generate oxidative stress [103]. Concerning drug addiction, in part to substances in the class of amphetamines, CBD seems to have the ability to attenuate the hyperactivity of dopaminergic neurotransmission in the mesolimbic region through an agonistic behavior towards dopaminergic D2 receptors [218].
The study conducted by García-Baos et al. demonstrates the protective effect of CBD on cognitive function (effect exerted in the hippocampal region and the prefrontal cortex) in the case of mice exposed to alcohol during the prenatal period, respectively, during the lactation period, most likely through the mediation of NF-κB, TNF-α, COX2, and caspase-3 expression, events also noted in other studies [219,220,221,222], but also in ischemia situations [223,224]. The same study states that CBD does not improve recognition memory, but rather spatial memory, assumptions reinforced by other studies in this field [205,219,225]. Another possible mechanism is the inhibition of AEA metabolism. Thus, the two molecules (CBD and AEA) can activate the TRPV1 receptors and cause glutamate release [226]. Table 1 states the results of studies on cognitive function after CBD administration.

6. Cannabidiol and Physical Performance

Regarding CBD use in athletes, the World Anti-Doping Agency (WADA) has included this compound in the list of substances prohibited during competitions [231]. Inclusion on this list is based on the hypothesis that CBD possesses the ability to enhance physical performance through various mechanisms [232,233,234]. A detail that must be considered is that there is a risk that CBD-containing products are also contaminated with THC (a compound found on the same prohibited list), which would result in a positive result in the anti-doping test [235,236].
Although multiple studies in the literature suggest improved physical performance as a result of the calming/relaxing effect of CBD [60,237,238], the present study focuses on antioxidant and muscle recovery mechanisms. Thus, it is known that high-intensity physical exercise correlates with inflammation and damage of skeletal muscle, and at the same time, with a decrease in physical performance [239]. For these reasons, athletes seek and resort to methods that shorten their recovery period [240,241,242]. These beneficial effects on muscle appear to be due to the inhibition of NF-κB and activation of the JAK/STAT (Janus Kinase/Signal Transducer and Activator of Transcription) pathway. However, a study conducted by Isenmann et al. noted that a single dose of CBD had minimal effect on creatine kinase (CK) and myoglobin levels after 72 h. It is therefore suggested that for a more pronounced effect, multiple doses are likely to be more effective [243]. In addition, anti-inflammatory properties are correlated with decreased levels of cytokines, prostaglandin E2 (PGE2), and NO, and are assumed to be mediated by 5-HT1A and TRPV1 receptors [244]. In another study, this time conducted by Ianotti et al., in a model of muscular dystrophy (Duchenne muscular dystrophy), it was shown that the use of CBD limited motor dysfunction, improved muscle strength, and also reduced pro-inflammatory markers (IL-6, TNF-α, TGF-β1, iNOS) and autophagy (autophagy-regulating protease 4 and 12 (Atg4, Atg12), Unc-51 like autophagy activating kinase (ULK1)). In the same study, the authors observed that on C2C12 myoblasts, CBD exerted its pro-differentiating effect via TRPV1 [245,246]. At the same time, in a study on human subjects, to evaluate soreness and performance, it was observed that the administration of CBD (150 mg), regardless of the time of administration (24 and 48 h) did not lead to an improvement in muscle condition [247], whereas, in another study, the administration of CBD (16.67 mg given with 1 mL of medium chain triglyceride) at 24, 48, and 72 h visibly reduced the muscle pain felt [248].
These considerations are based on the idea that inflammation, exercise-related damage proliferation, and cell differentiation are closely related to ROS, which gives rise to another hypothesis, that the reduction in oxidative stress can have an important role in sports [249]. On the same note, CBD, by regulating cortisol release via CB1, CB2, and A2 receptors, decreases the level of immune cells and cytokines (IL-1, TNF-α), and in addition, favors the release of arachidonic acid with the stimulation of healing capacity, promoted by growth and anti-inflammatory signals (lipoxin A4, 15d-PGJ2) [250,251].
Despite the fact that the results of the studies carried out to date do not provide conclusive data on whether the use of CBD helps or improves physical performance, thus attracting controversy regarding sports ethics, this compound seems to be increasingly used, and for this reason, more extensive studies are needed, both clinical and preclinical, to establish as clearly as possible the mechanisms, respectively, the pharmacodynamic effects in sports (when used before, during, and/or after training). Thus, these studies should consider several aspects, including the type of muscle pain, the fatigue of the individuals, the type of physical effort/sport, as well as gender differences.

7. Cannabidiol and Autophagy

The autophagy process represents a cellular mechanism by which damaged organelles and non-functional protein aggregates are degraded to obtain energy or recycle them (after degradation, amino acids are used for new synthesis). This is a very precise process aimed at maintaining the body’s homeostasis, normal growth, and development, as well as regulating inflammatory processes and immunity, and protecting against viral and/or bacterial infections. There are three types of autophagy (macroautophagy, microautophagy, and chaperone-mediated autophagy), but macroautophagy is currently thought to be actually autophagy [252,253,254].
Thus, in the case of neurodegenerative diseases, it has been observed that most of the time a mitochondrial dysfunction occurs that determines the progression of the disease [255]. In an experimental model of MPP+-induced PD in SH-SY5Y cell cultures, CBD administration protected these cells by increasing the expression of silent mating type information regulation 2 homolog 1 (SIRT1), to inhibit NF-κB. By inducing autophagy, attenuation of Tyrosine Hydroxylase (TH) loss and α-synuclein accumulation could be observed [256]. At the same time, it is known that in cases where mitochondrial dysfunction occurs, the expression of Nrf2 and antioxidant mechanisms are decreased, which implies a progression of oxidative damage [257]. Furthermore, pretreatment of SH-SY5Y cells with CBD mediates oxidative stress damage through the activation of PINK-1/parkin and DJ-1 proteins [256]. For this reason, the presence of the autophagic mechanism is necessary, the neurons use the energy resulting from the affected organelles, and the synaptic remodeling is also stimulated, but these beneficial effects also extend to other pathologies [252]. At the same time, the opposite extreme is not excluded, a dysregulation of autophagy can also lead to neuronal impairments and neurodegenerative diseases [258]. In a recent study, CBD-induced autophagy was concentration-dependent and required communication between extracellular signal-regulated protein kinases 1 and 2 (ERK1/2) and Akt [259]. Another subject in which autophagy is brought to the fore is the effect on the aging process of CBD, being known that autophagy is closely related to caloric restriction and the anti-aging effect, as suggested in multiple studies [260,261,262]. In a recent study on Caenorhabditis elegans, it is shown that CBD increases autophagic activity during aging and improves the health span and morphology of neurons in the context of aging [263], knowing that the cytoarchitecture of neurons changes during this process, and at the same time there is a decline of cognitive abilities [264]. Therefore, it is suggested that CBD does not necessarily improve these functions, but at least keeps them within normal limits. Another aspect that must be considered for the proper functioning of cells is mitochondrial homeostasis and the influence that CBD has on this organelle. In this case, two situations can occur in cases of high metabolic stress, namely mitochondrial fusion and fission [265]. Thus, mitochondrial fusion protects the organelle from age-related mtDNA mutations and is mediated by a protein, Optic atrophy 1 (OPA1), whereas mitochondrial fission is mediated by Dynamin-related peptide 1 (Drp-1) [99]. The study conducted by da Silva et al. evaluating iron-induced toxicity through oxidative stress in the brain demonstrates the beneficial effect of CBD by bringing the protein expression of Drp-1, caspase 3, and synaptophysin to values similar to those observed in the control group, without altering OPA1 [266]. In this regard, other studies have also shown an improvement in memory and a reduction in synaptophysin levels in cases of iron intoxication after administration of CBD [207], whereas an animal model of hypoxic-ischemic injury demonstrated a reduction in caspase 9 levels [267], and a study on PC12 cells treated with Aβ reduced caspase 3 levels [150]. Another detail that should be highlighted is the possibility that CBD manifests its effects on mitochondrial dynamics, noting that in the hippocampus Drp-1 levels are reduced, whereas OPA1 is increased [266].
These data support the idea that CBD exerts its neuroprotective effect not only through its antioxidant potential but also through its antiapoptotic properties.

8. Conclusions

In this article, we have highlighted the positive effects of CBD on several functions, which are closely related to its antioxidant properties and effects.
In conclusion, the endocannabinoid system and cannabinoid derivatives, such as CBD, will be increasingly researched and will certainly show more and more therapeutic importance, given the scientific evidence obtained both in preclinical studies and clinical ones.

Author Contributions

Writing—original draft preparation, G.J.; writing—review and editing, G.J., B.E.Ő., C.E.V., A.T.-V., C.-M.R. and A.P.; visualization, G.J. and C.-M.R.; funding acquisition, G.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by George Emil Palade University of Medicine, Pharmacy, Science, and Technology of Târgu Mureș Research Grant number 510/19/17.01.2022.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moldes, A.B.; Vecino, X.; Cruz, J.M. Nutraceuticals and Food Additives. In Current Developments in Biotechnology and Bioengineering; Pandey, A., Sanromán, M.Á., Du, G., Soccol, C.R., Dussap, C.-G., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 143–164. ISBN 9780444636669. [Google Scholar] [CrossRef]
  2. Williamson, E.M.; Liu, X.; Izzo, A.A. Trends in use, pharmacology, and clinical applications of emerging herbal nutraceuticals. Br. J. Pharmacol. 2020, 177, 1227–1240. [Google Scholar] [CrossRef]
  3. Sachdeva, V.; Roy, A.; Bharadvaja, N. Current Prospects of Nutraceuticals: A Review. Curr. Pharm. Biotechnol. 2020, 21, 884–896. [Google Scholar] [CrossRef]
  4. Current FDA Regulation of Hemp/CBD in Foods and Dietary Supplements. Available online: https://www.fdli.org/wp-content/uploads/2022/03/140-240-Preparing-for-a-world.pdf (accessed on 5 November 2022).
  5. Zuardi, A.W. History of cannabis as a medicine: A review. Braz. J. Psychiatry 2006, 28, 153–157. [Google Scholar] [CrossRef] [PubMed]
  6. Hillard, C.J.; Weinlander, K.M.; Stuhr, K.L. Contributions of endocannabinoid signaling to psychiatric disorders in humans: Genetic and biochemical evidence. Neuroscience 2012, 204, 207–229. [Google Scholar] [CrossRef] [PubMed]
  7. Mechoulam, R.; Parker, L.A. The endocannabinoid system and the brain. Annu. Rev. Psychol. 2013, 64, 21–47. [Google Scholar] [CrossRef]
  8. Mechoulam, R.; Shani, A.; Edery, H.; Grunfeld, Y. Chemical basis of hashish activity. Science 1970, 169, 611–612. [Google Scholar] [CrossRef] [PubMed]
  9. Russo, E.; Guy, G.W. A tale of two cannabinoids: The therapeutic rationale for combining tetrahydrocannabinol and cannabidiol. Med. Hypotheses 2006, 66, 234–246. [Google Scholar] [CrossRef]
  10. Andre, C.M.; Hausman, J.F.; Guerriero, G. Cannabis sativa: The Plant of the Thousand and One Molecules. Front. Plant Sci. 2016, 7, 19. [Google Scholar] [CrossRef]
  11. Brighenti, V.; Pellati, F.; Steinbach, M.; Maran, D.; Benvenuti, S. Development of a new extraction technique and HPLC method for the analysis of non-psychoactive cannabinoids in fibre-type Cannabis sativa L. (hemp). J. Pharm. Biomed. Anal. 2017, 143, 228–236. [Google Scholar] [CrossRef]
  12. Pellati, F.; Brighenti, V.; Sperlea, J.; Marchetti, L.; Bertelli, D.; Benvenuti, S. New Methods for the Comprehensive Analysis of Bioactive Compounds in Cannabis sativa L. (hemp). Molecules 2018, 23, 2639. [Google Scholar] [CrossRef]
  13. Scherma, M.; Masia, P.; Satta, V.; Fratta, W.; Fadda, P.; Tanda, G. Brain activity of anandamide: A rewarding bliss? Acta Pharmacol. Sin. 2019, 40, 309–323. [Google Scholar] [CrossRef] [PubMed]
  14. Di Marzo, V.; Fontana, A.; Cadas, H.; Schinelli, S.; Cimino, G.; Schwartz, J.C.; Piomelli, D. Formation and inactivation of endogenous cannabinoid anandamide in central neurons. Nature 1994, 372, 686–691. [Google Scholar] [CrossRef]
  15. Marrs, W.R.; Blankman, J.L.; Horne, E.A.; Thomazeau, A.; Lin, Y.H.; Coy, J.; Bodor, A.L.; Muccioli, G.G.; Hu, S.S.; Woodruff, G.; et al. The serine hydrolase ABHD6 controls the accumulation and efficacy of 2-AG at cannabinoid receptors. Nat. Neurosci. 2010, 13, 951–957. [Google Scholar] [CrossRef]
  16. Cravatt, B.F.; Demarest, K.; Patricelli, M.P.; Bracey, M.H.; Giang, D.K.; Martin, B.R.; Lichtman, A.H. Supersensitivity to anandamide and enhanced endogenous cannabinoid signaling in mice lacking fatty acid amide hydrolase. Proc. Natl. Acad. Sci. USA 2001, 98, 9371–9376. [Google Scholar] [CrossRef] [PubMed]
  17. Blankman, J.L.; Simon, G.M.; Cravatt, B.F. A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem. Biol. 2007, 14, 1347–1356. [Google Scholar] [CrossRef]
  18. Kohnz, R.A.; Nomura, D.K. Chemical approaches to therapeutically target the metabolism and signaling of the endocannabinoid 2-AG and eicosanoids. Chem. Soc. Rev. 2014, 43, 6859–6869. [Google Scholar] [CrossRef] [PubMed]
  19. Bodor, A.L.; Katona, I.; Nyíri, G.; Mackie, K.; Ledent, C.; Hájos, N.; Freund, T.F. Endocannabinoid signaling in rat somatosensory cortex: Laminar differences and involvement of specific interneuron types. J. Neurosci. 2005, 25, 6845–6856. [Google Scholar] [CrossRef]
  20. Hu, S.S.; Mackie, K. Distribution of the Endocannabinoid System in the Central Nervous System. In Endocannabinoids; Handbook of Experimental Pharmacology; Springer: Cham, Switzerland, 2015; Volume 231, pp. 59–93. [Google Scholar] [CrossRef]
  21. Lu, H.C.; Mackie, K. Review of the Endocannabinoid System. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2021, 6, 607–615. [Google Scholar] [CrossRef]
  22. Walsh, K.B.; Holmes, A.E. Pharmacology of Minor Cannabinoids at the Cannabinoid CB1 Receptor: Isomer- and Ligand-Dependent Antagonism by Tetrahydrocannabivarin. Receptors 2022, 1, 3–12. [Google Scholar] [CrossRef]
  23. Franco-Vadillo, A.; Toledo-Blass, M.; Rivera-Herrera, Z.; Guevara-Balcazar, G.; Orihuela-Rodriguez, O.; Morales-Carmona, J.A.; Kormanovski-Kovzova, A.; Lopez-Sanchez, P.; Rubio-Gayosso, I.; Castillo-Hernandez, M.D.C. Cannabidiol-mediated RISK PI3K/AKT and MAPK/ERK pathways decreasing reperfusion myocardial damage. Pharmacol. Res. Perspect. 2021, 9, e00784. [Google Scholar] [CrossRef]
  24. Kano, M.; Ohno-Shosaku, T.; Hashimotodani, Y.; Uchigashima, M.; Watanabe, M. Endocannabinoid-mediated control of synaptic transmission. Physiol. Rev. 2009, 9, 309–380. [Google Scholar] [CrossRef] [PubMed]
  25. Galiègue, S.; Mary, S.; Marchand, J.; Dussossoy, D.; Carrière, D.; Carayon, P.; Bouaboula, M.; Shire, D.; Le Fur, G.; Casellas, P. Expression of central and peripheral cannabinoid receptors in human immune tissues and leukocyte subpopulations. Eur. J. Biochem. 1995, 232, 54–61. [Google Scholar] [CrossRef] [PubMed]
  26. Cabral, G.A.; Ferreira, G.A.; Jamerson, M.J. Endocannabinoids and the Immune System in Health and Disease. In Endocannabinoids; Handbook of Experimental Pharmacology; Springer: Cham, Switzerland, 2015; Volume 231, pp. 185–211. [Google Scholar] [CrossRef]
  27. Ofek, O.; Karsak, M.; Leclerc, N.; Fogel, M.; Frenkel, B.; Wright, K.; Tam, J.; Attar-Namdar, M.; Kram, V.; Shohami, E.; et al. Peripheral cannabinoid receptor, CB2, regulates bone mass. Proc. Natl. Acad. Sci. USA 2006, 103, 696–701. [Google Scholar] [CrossRef]
  28. Juan-Picó, P.; Fuentes, E.; Bermúdez-Silva, F.J.; Javier Díaz-Molina, F.; Ripoll, C.; Rodríguez de Fonseca, F.; Nadal, A. Cannabinoid receptors regulate Ca(2+) signals and insulin secretion in pancreatic beta-cell. Cell Calcium 2006, 39, 155–162. [Google Scholar] [CrossRef]
  29. Atwood, B.K.; Mackie, K. CB2: A cannabinoid receptor with an identity crisis. Br. J. Pharmacol. 2010, 160, 467–479. [Google Scholar] [CrossRef]
  30. Appendino, G.; Chianese, G.; Taglialatela-Scafati, O. Cannabinoids: Occurrence and medicinal chemistry. Curr. Med. Chem. 2011, 18, 1085–1099. [Google Scholar] [CrossRef]
  31. Micale, V.; Di Marzo, V.; Sulcova, A.; Wotjak, C.T.; Drago, F. Endocannabinoid system and mood disorders: Priming a target for new therapies. Pharmacol. Ther. 2013, 138, 18–37. [Google Scholar] [CrossRef] [PubMed]
  32. Kucerova, J.; Tabiova, K.; Drago, F.; Micale, V. Therapeutic potential of cannabinoids in schizophrenia. Recent Pat. CNS Drug Discov. 2014, 9, 13–25. [Google Scholar] [CrossRef]
  33. Micale, V.; Tabiova, K.; Kucerova, J.; Drago, F. Role of the endocannabinoid system in depression: From preclinical to clinical evidence. In Cannabinoid Modulation of Emotion, Memory, and Motivation; Campolongo, P., Fattore, L., Eds.; Springer: New York, NY, USA, 2015; pp. 97–129. [Google Scholar] [CrossRef]
  34. Micale, V.; Drago, F. Endocannabinoid system, stress and HPA axis. Eur. J. Pharmacol. 2018, 834, 230–239. [Google Scholar] [CrossRef]
  35. Stark, T.; Di Martino, S.; Drago, F.; Wotjak, C.T.; Micale, V. Phytocannabinoids and schizophrenia: Focus on adolescence as a critical window of enhanced vulnerability and opportunity for treatment. Pharmacol. Res. 2021, 174, 105938. [Google Scholar] [CrossRef]
  36. Castillo, P.E.; Younts, T.J.; Chávez, A.E.; Hashimotodani, Y. Endocannabinoid signaling and synaptic function. Neuron 2012, 76, 70–81. [Google Scholar] [CrossRef] [PubMed]
  37. Fernández-Ruiz, J.; Moreno-Martet, M.; Rodríguez-Cueto, C.; Palomo-Garo, C.; Gómez-Cañas, M.; Valdeolivas, S.; Guaza, C.; Romero, J.; Guzmán, M.; Mechoulam, R.; et al. Prospects for cannabinoid therapies in basal ganglia disorders. Br. J. Pharmacol. 2011, 163, 1365–1378. [Google Scholar] [CrossRef] [PubMed]
  38. Alexander, S.P. Therapeutic potential of cannabis-related drugs. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 64, 157–166. [Google Scholar] [CrossRef] [PubMed]
  39. Campos, A.C.; Fogaça, M.V.; Sonego, A.B.; Guimarães, F.S. Cannabidiol, neuroprotection and neuropsychiatric disorders. Pharmacol. Res. 2016, 112, 119–127. [Google Scholar] [CrossRef]
  40. Pereira, S.R.; Hackett, B.; O’Driscoll, D.N.; Sun, M.C.; Downer, E.J. Cannabidiol modulation of oxidative stress and signalling. Neuronal Signal. 2021, 5, NS20200080. [Google Scholar] [CrossRef] [PubMed]
  41. Pisanti, S.; Malfitano, A.M.; Ciaglia, E.; Lamberti, A.; Ranieri, R.; Cuomo, G.; Abate, M.; Faggiana, G.; Proto, M.C.; Fiore, D.; et al. Cannabidiol: State of the art and new challenges for therapeutic applications. Pharmacol. Ther. 2017, 175, 133–150. [Google Scholar] [CrossRef]
  42. Chung, H.; Fierro, A.; Pessoa-Mahana, C.D. Cannabidiol binding and negative allosteric modulation at the cannabinoid type 1 receptor in the presence of delta-9-tetrahydrocannabinol: An In Silico study. PLoS ONE 2019, 14, e0220025. [Google Scholar] [CrossRef]
  43. Laprairie, R.B.; Bagher, A.M.; Kelly, M.E.; 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]
  44. 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]
  45. Galaj, E.; Bi, G.H.; Yang, H.J.; Xi, Z.X. Cannabidiol attenuates the rewarding effects of cocaine in rats by CB2, 5-HT1A and TRPV1 receptor mechanisms. Neuropharmacology 2020, 167, 107740. [Google Scholar] [CrossRef]
  46. 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] [PubMed]
  47. Milligan, C.J.; Anderson, L.L.; Bowen, M.T.; Banister, S.D.; McGregor, I.S.; Arnold, J.C.; Petrou, S.A. nutraceutical product, extracted from Cannabis sativa, modulates voltage-gated sodium channel function. J. Cannabis Res. 2022, 4, 30. [Google Scholar] [CrossRef] [PubMed]
  48. De Petrocellis, L.; Ligresti, A.; Moriello, A.S.; Allarà, 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]
  49. Nabissi, M.; Morelli, M.B.; Santoni, M.; Santoni, G. Triggering of the TRPV2 channel by cannabidiol sensitizes glioblastoma cells to cytotoxic chemotherapeutic agents. Carcinogenesis 2013, 34, 48–57. [Google Scholar] [CrossRef] [PubMed]
  50. Nabissi, M.; Morelli, M.B.; Amantini, C.; Liberati, S.; Santoni, M.; Ricci-Vitiani, L.; Pallini, R.; Santoni, G. Cannabidiol stimulates Aml-1a-dependent glial differentiation and inhibits glioma stem-like cells proliferation by inducing autophagy in a TRPV2-dependent manner. Int. J. Cancer 2015, 137, 1855–1869. [Google Scholar] [CrossRef]
  51. Zhong, H.; Gutkin, D.W.; Han, B.; Ma, Y.; Keskinov, A.A.; Shurin, M.R.; Shurin, G.V. Origin and pharmacological modulation of tumor-associated regulatory dendritic cells. Int. J. Cancer 2014, 134, 2633–2645. [Google Scholar] [CrossRef] [Green Version]
  52. 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] [PubMed]
  53. Etemad, L.; Karimi, G.; Alavi, M.S.; Roohbakhsh, A. Pharmacological effects of cannabidiol by transient receptor potential channels. Life Sci. 2022, 300, 120582. [Google Scholar] [CrossRef] [PubMed]
  54. Soares, V.P.; Campos, A.C. Evidences for the Anti-panic Actions of Cannabidiol. Curr. Neuropharmacol. 2017, 15, 291–299. [Google Scholar] [CrossRef]
  55. Atalay, S.; Jarocka-Karpowicz, I.; Skrzydlewska, E. Antioxidative and Anti-Inflammatory Properties of Cannabidiol. Antioxidants 2019, 9, 21. [Google Scholar] [CrossRef]
  56. Oláh, A.; Szekanecz, Z.; Bíró, T. Targeting Cannabinoid Signaling in the Immune System: "High"-ly Exciting Questions, Possibilities, and Challenges. Front. Immunol. 2017, 8, 1487. [Google Scholar] [CrossRef] [PubMed]
  57. Zuardi, A.W.; Cosme, R.A.; Graeff, F.G.; Guimarães, F.S. Effects of ipsapirone and cannabidiol on human experimental anxiety. J. Psychopharmacol. 1993, 7, 82–88. [Google Scholar] [CrossRef] [PubMed]
  58. Crippa, J.A.; Derenusson, G.N.; Ferrari, T.B.; Wichert-Ana, L.; Duran, F.L.; Martin-Santos, R.; Simões, M.V.; Bhattacharyya, S.; Fusar-Poli, P.; Atakan, Z.; et al. Neural basis of anxiolytic effects of cannabidiol (CBD) in generalized social anxiety disorder: A preliminary report. J. Psychopharmacol. 2011, 25, 121–130. [Google Scholar] [CrossRef]
  59. Zuardi, A.W.; Rodrigues, N.P.; Silva, A.L.; Bernardo, S.A.; Hallak, J.E.C.; Guimarães, F.S.; Crippa, J.A.S. Inverted U-Shaped Dose-Response Curve of the Anxiolytic Effect of Cannabidiol during Public Speaking in Real Life. Front. Pharmacol. 2017, 8, 259. [Google Scholar] [CrossRef] [PubMed]
  60. Linares, I.M.; Zuardi, A.W.; Pereira, L.C.; Queiroz, R.H.; Mechoulam, R.; Guimarães, F.S.; Crippa, J.A. Cannabidiol presents an inverted U-shaped dose-response curve in a simulated public speaking test. Braz. J. Psychiatry 2019, 41, 9–14. [Google Scholar] [CrossRef]
  61. Appiah-Kusi, E.; Petros, N.; Wilson, R.; Colizzi, M.; Bossong, M.G.; Valmaggia, L.; Mondelli, V.; McGuire, P.; Bhattacharyya, S. Effects of short-term cannabidiol treatment on response to social stress in subjects at clinical high risk of developing psychosis. Psychopharmacology 2020, 237, 1121–1130. [Google Scholar] [CrossRef] [Green Version]
  62. Llorente-Berzal, A.; Terzian, A.L.B.; di Marzo, V.; Micale, V.; Viveros, M.P.; Wotjak, C.T. 2-AG promotes the expression of conditioned fear via cannabinoid receptor type 1 on GABAergic neurons. Psychopharmacology 2015, 232, 2811–2825. [Google Scholar] [CrossRef]
  63. Terzian, A.L.; Micale, V.; Wotjak, C.T. Cannabinoid receptor type 1 receptors on GABAergic vs. glutamatergic neurons differentially gate sex-dependent social interest in mice. Eur. J. Neurosci. 2014, 40, 2293–2298. [Google Scholar] [CrossRef]
  64. 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]
  65. Micale, V.; Stepan, J.; Jurik, A.; Pamplona, F.A.; Marsch, R.; Drago, F.; Eder, M.; Wotjak, C.T. Extinction of avoidance behavior by safety learning depends on endocannabinoid signaling in the hippocampus. J. Psychiatr. Res. 2017, 90, 46–59. [Google Scholar] [CrossRef]
  66. Hallak, J.E.; Machado-de-Sousa, J.P.; Crippa, J.A.; Sanches, R.F.; Trzesniak, C.; Chaves, C.; Bernardo, S.A.; Regalo, S.C.; Zuardi, A.W. Performance of schizophrenic patients in the Stroop Color Word Test and electrodermal responsiveness after acute administration of cannabidiol (CBD). Braz. J. Psychiatry 2010, 32, 56–61. [Google Scholar] [CrossRef]
  67. Leweke, F.M.; Piomelli, D.; Pahlisch, F.; Muhl, D.; Gerth, C.W.; Hoyer, C.; Klosterkötter, J.; Hellmich, M.; Koethe, D. Cannabidiol enhances anandamide signaling and alleviates psychotic symptoms of schizophrenia. Transl. Psychiatry 2012, 2, e94. [Google Scholar] [CrossRef] [PubMed]
  68. Arkell, T.R.; Lintzeris, N.; Kevin, R.C.; Ramaekers, J.G.; Vandrey, R.; Irwin, C.; Haber, P.S.; McGregor, I.S. Cannabidiol (CBD) content in vaporized cannabis does not prevent tetrahydrocannabinol (THC)-induced impairment of driving and cognition. Psychopharmacology 2019, 236, 2713–2724. [Google Scholar] [CrossRef] [PubMed]
  69. Stark, T.; Ruda-Kucerova, J.; Iannotti, F.A.; D’Addario, C.; Di Marco, R.; Pekarik, V.; Drazanova, E.; Piscitelli, F.; Bari, M.; Babinska, Z.; et al. Peripubertal cannabidiol treatment rescues behavioral and neurochemical abnormalities in the MAM model of schizophrenia. Neuropharmacology 2019, 146, 212–221. [Google Scholar] [CrossRef] [PubMed]
  70. Di Bartolomeo, M.; Stark, T.; Maurel, O.M.; Iannotti, F.A.; Kuchar, M.; Ruda-Kucerova, J.; Piscitelli, F.; Laudani, S.; Pekarik, V.; Salomone, S.; et al. Crosstalk between the transcriptional regulation of dopamine D2 and cannabinoid CB1 receptors in schizophrenia: Analyses in patients and in perinatal Δ9-tetrahydrocannabinol-exposed rats. Pharmacol. Res. 2021, 164, 105357. [Google Scholar] [CrossRef] [PubMed]
  71. Brancato, A.; Castelli, V.; Lavanco, G.; Tringali, G.; Micale, V.; Kuchar, M.; D’Amico, C.; Pizzolanti, G.; Feo, S.; Cannizzaro, C. Binge-like Alcohol Exposure in Adolescence: Behavioural, Neuroendocrine and Molecular Evidence of Abnormal Neuroplasticity… and Return. Biomedicines 2021, 9, 1161. [Google Scholar] [CrossRef]
  72. Seeman, P. Cannabidiol is a partial agonist at dopamine D2High receptors, predicting its antipsychotic clinical dose. Transl. Psychiatry 2016, 6, e920. [Google Scholar] [CrossRef]
  73. Stark, T.; Di Bartolomeo, M.; Di Marco, R.; Drazanova, E.; Platania, C.B.M.; Iannotti, F.A.; Ruda-Kucerova, J.; D’Addario, C.; Kratka, L.; Pekarik, V.; et al. Altered dopamine D3 receptor gene expression in MAM model of schizophrenia is reversed by peripubertal cannabidiol treatment. Biochem. Pharmacol. 2020, 177, 114004. [Google Scholar] [CrossRef]
  74. Micale, V.; Di Bartolomeo, M.; Di Martino, S.; Stark, T.; Dell’Osso, B.; Drago, F.; D’Addario, C. Are the epigenetic changes predictive of therapeutic efficacy for psychiatric disorders? A translational approach towards novel drug targets. Pharmacol. Ther. 2023, 241, 108279. [Google Scholar] [CrossRef]
  75. 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, 376, 2011–2020. [Google Scholar] [CrossRef]
  76. 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]
  77. 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]
  78. Zuardi, A.W.; Crippa, J.A.; Hallak, J.E.; Pinto, J.P.; Chagas, M.H.; Rodrigues, G.G.; Dursun, S.M.; Tumas, V. Cannabidiol for the treatment of psychosis in Parkinson’s disease. J. Psychopharmacol. 2009, 23, 979–983. [Google Scholar] [CrossRef]
  79. Chagas, M.H.; Zuardi, A.W.; Tumas, V.; Pena-Pereira, M.A.; Sobreira, E.T.; Bergamaschi, M.M.; dos Santos, A.C.; Teixeira, A.L.; Hallak, J.E.; Crippa, J.A. Effects of cannabidiol in the treatment of patients with Parkinson’s disease: An exploratory double-blind trial. J. Psychopharmacol. 2014, 28, 1088–1098. [Google Scholar] [CrossRef] [PubMed]
  80. Guo, K.; Mou, X.; Huang, J.; Xiong, N.; Li, H. Trans-caryophyllene suppresses hypoxia-induced neuroinflammatory responses by inhibiting NF-κB activation in microglia. J. Mol. Neurosci. 2014, 54, 41–48. [Google Scholar] [CrossRef] [PubMed]
  81. 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] [Green Version]
  82. Scuderi, C.; Steardo, L.; Esposito, G. Cannabidiol promotes amyloid precursor protein ubiquitination and reduction of beta amyloid expression in SHSY5YAPP+ cells through PPARγ involvement. Phytother. Res. 2014, 28, 1007–1013. [Google Scholar] [CrossRef]
  83. Hampson, A.J.; Grimaldi, M.; Axelrod, J.; Wink, D. Cannabidiol and (-)Delta9-tetrahydrocannabinol are neuroprotective antioxidants. Proc. Natl. Acad. Sci. USA 1998, 95, 8268–8273. [Google Scholar] [CrossRef]
  84. Borges, R.S.; Batista, J., Jr.; Viana, R.B.; Baetas, A.C.; Orestes, E.; Andrade, M.A.; Honório, K.M.; da Silva, A.B. Understanding the molecular aspects of tetrahydrocannabinol and cannabidiol as antioxidants. Molecules 2013, 18, 12663–12674. [Google Scholar] [CrossRef]
  85. Borges, R.S.; da Silva, A.B.F. Chapter e12—Cannabidiol as an Antioxidant. In Handbook of Cannabis and Related Pathologies; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. e122–e130. ISBN 9780128007563. [Google Scholar] [CrossRef]
  86. Hamelink, C.; Hampson, A.; Wink, D.A.; Eiden, L.E.; Eskay, R.L. Comparison of cannabidiol, antioxidants, and diuretics in reversing binge ethanol-induced neurotoxicity. J. Pharmacol. Exp. Ther. 2005, 314, 780–788. [Google Scholar] [CrossRef] [PubMed]
  87. Wheal, A.J.; Jadoon, K.; Randall, M.D.; O’Sullivan, S.E. In Vivo Cannabidiol Treatment Improves Endothelium-Dependent Vasorelaxation in Mesenteric Arteries of Zucker Diabetic Fatty Rats. Front. Pharmacol. 2017, 8, 248. [Google Scholar] [CrossRef] [PubMed]
  88. Costa, B.; Trovato, A.E.; Comelli, F.; Giagnoni, G.; Colleoni, M. The non-psychoactive cannabis constituent cannabidiol is an orally effective therapeutic agent in rat chronic inflammatory and neuropathic pain. Eur. J. Pharmacol. 2007, 556, 75–83. [Google Scholar] [CrossRef]
  89. Fernández-Ruiz, J.; García, C.; Sagredo, O.; Gómez-Ruiz, M.; de Lago, E. The endocannabinoid system as a target for the treatment of neuronal damage. Expert Opin. Ther. Targets 2010, 14, 387–404. [Google Scholar] [CrossRef]
  90. Juknat, A.; Pietr, M.; Kozela, E.; Rimmerman, N.; Levy, R.; Gao, F.; Coppola, G.; Geschwind, D.; Vogel, Z. Microarray and pathway analysis reveal distinct mechanisms underlying cannabinoid-mediated modulation of LPS-induced activation of BV-2 microglial cells. PLoS ONE 2013, 8, e61462. [Google Scholar] [CrossRef]
  91. Vomund, S.; Schäfer, A.; Parnham, M.J.; Brüne, B.; von Knethen, A. Nrf2, the Master Regulator of Anti-Oxidative Responses. Int. J. Mol. Sci. 2017, 18, 2772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Li, L.; Xuan, Y.; Zhu, B.; Wang, X.; Tian, X.; Zhao, L.; Wang, Y.; Jiang, X.; Wen, N. Protective Effects of Cannabidiol on Chemotherapy-Induced Oral Mucositis via the Nrf2/Keap1/ARE Signaling Pathways. Oxid. Med. Cell. Longev. 2022, 2022, 4619760. [Google Scholar] [CrossRef] [PubMed]
  93. Fouad, A.A.; Albuali, W.H.; Al-Mulhim, A.S.; Jresat, I. Cardioprotective effect of cannabidiol in rats exposed to doxorubicin toxicity. Environ. Toxicol. Pharmacol. 2013, 36, 347–357. [Google Scholar] [CrossRef]
  94. Iffland, K.; Grotenhermen, F. An Update on Safety and Side Effects of Cannabidiol: A Review of Clinical Data and Relevant Animal Studies. Cannabis Cannabinoid Res. 2017, 2, 139–154. [Google Scholar] [CrossRef]
  95. Jîtcă, G.; Fogarasi, E.; Ősz, B.E.; Vari, C.E.; Tero-Vescan, A.; Miklos, A.; Bătrînu, M.G.; Rusz, C.M.; Croitoru, M.D.; Dogaru, M.T. A Simple HPLC/DAD Method Validation for the Quantification of Malondialdehyde in Rodent’s Brain. Molecules 2021, 26, 5066. [Google Scholar] [CrossRef] [PubMed]
  96. Jîtcă, G.; Fogarasi, E.; Ősz, B.E.; Vari, C.E.; Fülöp, I.; Croitoru, M.D.; Rusz, C.M.; Dogaru, M.T. Profiling the Concentration of Reduced and Oxidized Glutathione in Rat Brain Using HPLC/DAD Chromatographic System. Molecules 2021, 26, 6590. [Google Scholar] [CrossRef]
  97. 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]
  98. Gianazza, E.; Crawford, J.; Miller, I. Detecting oxidative post-translational modifications in proteins. Amino Acids 2007, 33, 51–56. [Google Scholar] [CrossRef] [PubMed]
  99. Jîtcă, G.; Ősz, B.E.; Tero-Vescan, A.; Miklos, A.P.; Rusz, C.M.; Bătrînu, M.G.; Vari, C.E. Positive Aspects of Oxidative Stress at Different Levels of the Human Body: A Review. Antioxidants 2022, 11, 572. [Google Scholar] [CrossRef] [PubMed]
  100. di Giacomo, V.; Chiavaroli, A.; Orlando, G.; Cataldi, A.; Rapino, M.; Di Valerio, V.; Leone, S.; Brunetti, L.; Menghini, L.; Recinella, L.; et al. Neuroprotective and Neuromodulatory Effects Induced by Cannabidiol and Cannabigerol in Rat Hypo-E22 cells and Isolated Hypothalamus. Antioxidants 2020, 9, 71. [Google Scholar] [CrossRef] [PubMed]
  101. Sun, S.; Hu, F.; Wu, J.; Zhang, S. Cannabidiol attenuates OGD/R-induced damage by enhancing mitochondrial bioenergetics and modulating glucose metabolism via pentose-phosphate pathway in hippocampal neurons. Redox Biol. 2017, 11, 577–585. [Google Scholar] [CrossRef]
  102. Valvassori, S.S.; Elias, G.; de Souza, B.; Petronilho, F.; Dal-Pizzol, F.; Kapczinski, F.; Trzesniak, C.; Tumas, V.; Dursun, S.; Chagas, M.H.; et al. Effects of cannabidiol on amphetamine-induced oxidative stress generation in an animal model of mania. J. Psychopharmacol. 2011, 25, 274–280. [Google Scholar] [CrossRef]
  103. Jîtcă, G.; Ősz, B.E.; Tero-Vescan, A.; Vari, C.E. Psychoactive Drugs-From Chemical Structure to Oxidative Stress Related to Dopaminergic Neurotransmission. A Review. Antioxidants 2021, 10, 381. [Google Scholar] [CrossRef]
  104. Cheng, D.; Low, J.K.; Logge, W.; Garner, B.; Karl, T. Chronic cannabidiol treatment improves social and object recognition in double transgenic APPswe/PS1∆E9 mice. Psychopharmacology 2014, 231, 3009–3017. [Google Scholar] [CrossRef]
  105. Han, K.H.; Lim, S.; Ryu, J.; Lee, C.W.; Kim, Y.; Kang, J.H.; Kang, S.S.; Ahn, Y.K.; Park, C.S.; Kim, J.J. CB1 and CB2 cannabinoid receptors differentially regulate the production of reactive oxygen species by macrophages. Cardiovasc. Res. 2009, 84, 378–386. [Google Scholar] [CrossRef]
  106. Muller, C.; Morales, P.; Reggio, P.H. Cannabinoid Ligands Targeting TRP Channels. Front. Mol. Neurosci. 2019, 11, 487. [Google Scholar] [CrossRef]
  107. Pertwee, R.G. The diverse CB1 and CB2 receptor pharmacology of three plant cannabinoids: Delta9-tetrahydrocannabinol, cannabidiol and delta9-tetrahydrocannabivarin. Br. J. Pharmacol. 2008, 153, 199–215. [Google Scholar] [CrossRef]
  108. Ogawa, N.; Kurokawa, T.; Fujiwara, K.; Polat, O.K.; Badr, H.; Takahashi, N.; Mori, Y. Functional and Structural Divergence in Human TRPV1 Channel Subunits by Oxidative Cysteine Modification. J. Biol. Chem. 2016, 291, 4197–4210. [Google Scholar] [CrossRef]
  109. Hellenthal, K.E.M.; Brabenec, L.; Gross, E.R.; Wagner, N.M. TRP Channels as Sensors of Aldehyde and Oxidative Stress. Biomolecules 2021, 11, 1401. [Google Scholar] [CrossRef] [PubMed]
  110. Minke, B. TRP channels and Ca2+ signaling. Cell Calcium 2006, 40, 261–275. [Google Scholar] [CrossRef] [PubMed]
  111. Bujak, J.K.; Kosmala, D.; Szopa, I.M.; Majchrzak, K.; Bednarczyk, P. Inflammation, Cancer and Immunity-Implication of TRPV1 Channel. Front. Oncol. 2019, 9, 1087. [Google Scholar] [CrossRef]
  112. Vallée, A.; Lecarpentier, Y.; Guillevin, R.; Vallée, J.N. Effects of cannabidiol interactions with Wnt/β-catenin pathway and PPARγ on oxidative stress and neuroinflammation in Alzheimer’s disease. Acta Biochim. Biophys. Sin. 2017, 49, 853–866. [Google Scholar] [CrossRef]
  113. Paunkov, A.; Chartoumpekis, D.V.; Ziros, P.G.; Sykiotis, G.P. A Bibliometric Review of the Keap1/Nrf2 Pathway and its Related Antioxidant Compounds. Antioxidants 2019, 8, 353. [Google Scholar] [CrossRef]
  114. Cho, H.Y.; Gladwell, W.; Wang, X.; Chorley, B.; Bell, D.; Reddy, S.P.; Kleeberger, S.R. Nrf2-regulated PPAR{gamma} expression is critical to protection against acute lung injury in mice. Am. J. Respir. Crit. Care Med. 2010, 182, 170–182. [Google Scholar] [CrossRef]
  115. Lee, C. Collaborative Power of Nrf2 and PPARγActivators against Metabolic and Drug-Induced Oxidative Injury. Oxid. Med. Cell. Longev. 2017, 2017, 1378175. [Google Scholar] [CrossRef]
  116. O’Sullivan, S.E. An update on PPAR activation by cannabinoids. Br. J. Pharmacol. 2016, 173, 1899–1910. [Google Scholar] [CrossRef] [PubMed]
  117. 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]
  118. Booz, G.W. Cannabidiol as an emergent therapeutic strategy for lessening the impact of inflammation on oxidative stress. Free Radic. Biol. Med. 2011, 51, 1054–1061. [Google Scholar] [CrossRef]
  119. Azouzi, S.; Santuz, H.; Morandat, S.; Pereira, C.; Côté, F.; Hermine, O.; El Kirat, K.; Colin, Y.; Le Van Kim, C.; Etchebest, C.; et al. Antioxidant and Membrane Binding Properties of Serotonin Protect Lipids from Oxidation. Biophys. J. 2017, 112, 1863–1873. [Google Scholar] [CrossRef]
  120. Mecha, M.; Feliú, 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] [PubMed]
  121. Casares, L.; García, V.; Garrido-Rodríguez, M.; Millán, E.; Collado, J.A.; García-Martín, A.; Peñarando, J.; Calzado, M.A.; de la Vega, L.; Muñoz, E. Cannabidiol induces antioxidant pathways in keratinocytes by targeting BACH1. Redox Biol. 2020, 28, 101321. [Google Scholar] [CrossRef] [PubMed]
  122. Bublitz, K.; Böckmann, S.; Peters, K.; Hinz, B. Cannabinoid-Induced Autophagy and Heme Oxygenase-1 Determine the Fate of Adipose Tissue-Derived Mesenchymal Stem Cells under Stressful Conditions. Cells 2020, 9, 2298. [Google Scholar] [CrossRef] [PubMed]
  123. Duvigneau, J.C.; Trovato, A.; Müllebner, A.; Miller, I.; Krewenka, C.; Krenn, K.; Zich, W.; Moldzio, R. Cannabidiol Protects Dopaminergic Neurons in Mesencephalic Cultures against the Complex I Inhibitor Rotenone Via Modulation of Heme Oxygenase Activity and Bilirubin. Antioxidants 2020, 9, 135. [Google Scholar] [CrossRef]
  124. Pan, H.; Mukhopadhyay, P.; Rajesh, M.; Patel, V.; Mukhopadhyay, B.; Gao, B.; Haskó, G.; Pacher, P. Cannabidiol attenuates cisplatin-induced nephrotoxicity by decreasing oxidative/nitrosative stress, inflammation, and cell death. J. Pharmacol. Exp. Ther. 2009, 328, 708–714. [Google Scholar] [CrossRef]
  125. Wang, Y.; Mukhopadhyay, P.; Cao, Z.; Wang, H.; Feng, D.; Haskó, G.; Mechoulam, R.; Gao, B.; Pacher, P. Cannabidiol attenuates alcohol-induced liver steatosis, metabolic dysregulation, inflammation and neutrophil-mediated injury. Sci. Rep. 2017, 7, 12064. [Google Scholar] [CrossRef]
  126. Dos-Santos-Pereira, M.; Guimarães, F.S.; Del-Bel, E.; Raisman-Vozari, R.; Michel, P.P. Cannabidiol prevents LPS-induced microglial inflammation by inhibiting ROS/NF-κB-dependent signaling and glucose consumption. Glia 2020, 68, 561–573. [Google Scholar] [CrossRef]
  127. Massi, P.; Vaccani, A.; Bianchessi, S.; Costa, B.; Macchi, P.; Parolaro, D. The non-psychoactive cannabidiol triggers caspase activation and oxidative stress in human glioma cells. Cell. Mol. Life Sci. 2006, 63, 2057–2066. [Google Scholar] [CrossRef] [PubMed]
  128. di Giacomo, V.; Chiavaroli, A.; Recinella, L.; Orlando, G.; Cataldi, A.; Rapino, M.; Di Valerio, V.; Ronci, M.; Leone, S.; Brunetti, L.; et al. Antioxidant and Neuroprotective Effects Induced by Cannabidiol and Cannabigerol in Rat CTX-TNA2 Astrocytes and Isolated Cortexes. Int. J. Mol. Sci. 2020, 21, 3575. [Google Scholar] [CrossRef] [PubMed]
  129. Xia, Y.; Xing, J.Z.; Krukoff, T.L. Neuroprotective effects of R,R-tetrahydrochrysene against glutamate-induced cell death through anti-excitotoxic and antioxidant actions involving estrogen receptor-dependent and -independent pathways. Neuroscience 2009, 162, 292–306. [Google Scholar] [CrossRef] [PubMed]
  130. Khaksar, S.; Bigdeli, M.; Samiee, A.; Shirazi-Zand, Z. Antioxidant and anti-apoptotic effects of cannabidiol in model of ischemic stroke in rats. Brain Res. Bull. 2022, 180, 118–130. [Google Scholar] [CrossRef]
  131. Lastres-Becker, I.; Molina-Holgado, F.; Ramos, J.A.; Mechoulam, R.; Fernández-Ruiz, J. Cannabinoids provide neuroprotection against 6-hydroxydopamine toxicity in vivo and in vitro: Relevance to Parkinson’s disease. Neurobiol. Dis. 2005, 19, 96–107. [Google Scholar] [CrossRef]
  132. García-Arencibia, M.; González, S.; de Lago, E.; Ramos, J.A.; Mechoulam, R.; Fernández-Ruiz, J. Evaluation of the neuroprotective effect of cannabinoids in a rat model of Parkinson’s disease: Importance of antioxidant and cannabinoid receptor-independent properties. Brain Res. 2007, 1134, 162–170. [Google Scholar] [CrossRef]
  133. Vivekanantham, S.; Shah, S.; Dewji, R.; Dewji, A.; Khatri, C.; Ologunde, R. Neuroinflammation in Parkinson’s disease: Role in neurodegeneration and tissue repair. Int. J. Neurosci. 2015, 125, 717–725. [Google Scholar] [CrossRef] [PubMed]
  134. Aso, E.; Ferrer, I. CB2 Cannabinoid Receptor as Potential Target against Alzheimer’s Disease. Front. Neurosci. 2016, 10, 243. [Google Scholar] [CrossRef]
  135. Concannon, R.M.; Okine, B.N.; Finn, D.P.; Dowd, E. Upregulation of the cannabinoid CB2 receptor in environmental and viral inflammation-driven rat models of Parkinson’s disease. Exp. Neurol. 2016, 283, 204–212. [Google Scholar] [CrossRef]
  136. Scuderi, C.; Filippis, D.D.; Iuvone, T.; Blasio, A.; Steardo, A.; Esposito, G. Cannabidiol in medicine: A review of its therapeutic potential in CNS disorders. Phytother. Res. 2009, 23, 597–602. [Google Scholar] [CrossRef]
  137. Esposito, G.; De Filippis, D.; Maiuri, M.C.; De Stefano, D.; Carnuccio, R.; Iuvone, T. Cannabidiol inhibits inducible nitric oxide synthase protein expression and nitric oxide production in beta-amyloid stimulated PC12 neurons through p38 MAP kinase and NF-kappaB involvement. Neurosci. Lett. 2006, 399, 91–95. [Google Scholar] [CrossRef]
  138. Esposito, G.; De Filippis, D.; Carnuccio, R.; Izzo, A.A.; Iuvone, T. The marijuana component cannabidiol inhibits beta-amyloid-induced tau protein hyperphosphorylation through Wnt/beta-catenin pathway rescue in PC12 cells. J. Mol. Med. 2006, 84, 253–258. [Google Scholar] [CrossRef]
  139. Sagredo, O.; Ramos, J.A.; Decio, A.; Mechoulam, R.; Fernández-Ruiz, J. Cannabidiol reduced the striatal atrophy caused 3-nitropropionic acid in vivo by mechanisms independent of the activation of cannabinoid, vanilloid TRPV1 and adenosine A2A receptors. Eur. J. Neurosci. 2007, 26, 843–851. [Google Scholar] [CrossRef] [PubMed]
  140. Hill, A.J.; Williams, C.M.; Whalley, B.J.; Stephens, G.J. Phytocannabinoids as novel therapeutic agents in CNS disorders. Pharmacol. Ther. 2012, 133, 79–97. [Google Scholar] [CrossRef]
  141. Fernández-Ruiz, J.; Sagredo, O.; Pazos, M.R.; García, C.; Pertwee, R.; Mechoulam, R.; Martínez-Orgado, J. Cannabidiol for neurodegenerative disorders: Important new clinical applications for this phytocannabinoid? Br. J. Clin. Pharmacol. 2013, 75, 323–333. [Google Scholar] [CrossRef]
  142. Vallée, A.; Vallée, J.N.; Lecarpentier, Y. Potential role of cannabidiol in Parkinson’s disease by targeting the WNT/β-catenin pathway, oxidative stress and inflammation. Aging 2021, 13, 10796–10813. [Google Scholar] [CrossRef]
  143. Hughes, B.; Herron, C.E. Cannabidiol Reverses Deficits in Hippocampal LTP in a Model of Alzheimer’s Disease. Neurochem. Res. 2019, 44, 703–713. [Google Scholar] [CrossRef]
  144. Cheng, D.; Spiro, A.S.; Jenner, A.M.; Garner, B.; Karl, T. Long-term cannabidiol treatment prevents the development of social recognition memory deficits in Alzheimer’s disease transgenic mice. J. Alzheimer’s Dis. 2014, 42, 1383–1396. [Google Scholar] [CrossRef] [PubMed]
  145. Libro, R.; Diomede, F.; Scionti, D.; Piattelli, A.; Grassi, G.; Pollastro, F.; Bramanti, P.; Mazzon, E.; Trubiani, O. Cannabidiol Modulates the Expression of Alzheimer’s Disease-Related Genes in Mesenchymal Stem Cells. Int. J. Mol. Sci. 2016, 18, 26. [Google Scholar] [CrossRef] [PubMed]
  146. Esposito, G.; Scuderi, C.; Savani, C.; Steardo, L., Jr.; De Filippis, D.; Cottone, P.; Iuvone, T.; Cuomo, V.; Steardo, L. Cannabidiol in vivo blunts beta-amyloid induced neuroinflammation by suppressing IL-1beta and iNOS expression. Br. J. Pharmacol. 2007, 151, 1272–1279. [Google Scholar] [CrossRef]
  147. Butterfield, D.A.; Drake, J.; Pocernich, C.; Castegna, A. Evidence of oxidative damage in Alzheimer’s disease brain: Central role for amyloid β-peptide. Trends Mol. Med. 2001, 7, 548–554. [Google Scholar] [CrossRef]
  148. Turner, B.J.; Li, Q.X.; Laughton, K.M.; Masters, C.L.; Lopes, E.C.; Atkin, J.D.; Cheema, S.S. Brain beta-amyloid accumulation in transgenic mice expressing mutant superoxide dismutase 1. Neurochem. Res. 2004, 29, 2281–2286. [Google Scholar] [CrossRef] [PubMed]
  149. Haas, J.; Storch-Hagenlocher, B.; Biessmann, A.; Wildemann, B. Inducible nitric oxide synthase and argininosuccinate synthetase: Co-induction in brain tissue of patients with Alzheimer’s dementia and following stimulation with β-amyloid 1–42 in vitro. Neurosci. Lett. 2002, 322, 121–125. [Google Scholar] [CrossRef] [PubMed]
  150. Iuvone, T.; Esposito, G.; Esposito, R.; Santamaria, R.; Di Rosa, M.; Izzo, A.A. Neuroprotective effect of cannabidiol, a non-psychoactive component from Cannabis sativa, on beta-amyloid-induced toxicity in PC12 cells. J. Neurochem. 2004, 89, 134–141. [Google Scholar] [CrossRef] [PubMed]
  151. Moncada, S.; Rees, D.D.; Schulz, R.; Palmer, R.M. Development and mechanism of a specific supersensitivity to nitrovasodilators after inhibition of vascular nitric oxide synthesis in vivo. Proc. Natl. Acad. Sci. USA 1991, 88, 2166–2170. [Google Scholar] [CrossRef] [Green Version]
  152. Esposito, G.; Scuderi, C.; Valenza, M.; Togna, G.I.; Latina, V.; De Filippis, D.; Cipriano, M.; Carratù, M.R.; Iuvone, T.; Steardo, L. Cannabidiol reduces Aβ-induced neuroinflammation and promotes hippocampal neurogenesis through PPARγ involvement. PLoS ONE 2011, 6, e28668. [Google Scholar] [CrossRef] [PubMed]
  153. Bedse, G.; Romano, A.; Cianci, S.; Lavecchia, A.M.; Lorenzo, P.; Elphick, M.R.; Laferla, F.M.; Vendemiale, G.; Grillo, C.; Altieri, F.; et al. Altered expression of the CB1 cannabinoid receptor in the triple transgenic mouse model of Alzheimer’s disease. J. Alzheimer’s Dis. 2014, 40, 701–712. [Google Scholar] [CrossRef]
  154. Gallelli, C.A.; Calcagnini, S.; Romano, A.; Koczwara, J.B.; de Ceglia, M.; Dante, D.; Villani, R.; Giudetti, A.M.; Cassano, T.; Gaetani, S. Modulation of the Oxidative Stress and Lipid Peroxidation by Endocannabinoids and Their Lipid Analogues. Antioxidants 2018, 7, 93. [Google Scholar] [CrossRef]
  155. Cassano, T.; Villani, R.; Pace, L.; Carbone, A.; Bukke, V.N.; Orkisz, S.; Avolio, C.; Serviddio, G. From Cannabis sativa to Cannabidiol: Promising Therapeutic Candidate for the Treatment of Neurodegenerative Diseases. Front. Pharmacol. 2020, 11, 124. [Google Scholar] [CrossRef]
  156. Ahmadi, S.; Zhu, S.; Sharma, R.; Wu, B.; Soong, R.; Dutta Majumdar, R.; Wilson, D.J.; Simpson, A.J.; Kraatz, H.B. Aggregation of Microtubule Binding Repeats of Tau Protein is Promoted by Cu2. ACS Omega 2019, 4, 5356–5366. [Google Scholar] [CrossRef]
  157. Santos, N.A.; Martins, N.M.; Sisti, F.M.; Fernandes, L.S.; Ferreira, R.S.; Queiroz, R.H.; Santos, A.C. The neuroprotection of cannabidiol against MPP⁺-induced toxicity in PC12 cells involves trkA receptors, upregulation of axonal and synaptic proteins, neuritogenesis, and might be relevant to Parkinson’s disease. Toxicol. In Vitro 2015, 30, 231–240. [Google Scholar] [CrossRef]
  158. Ahmadi, S.; Zhu, S.; Sharma, R.; Wilson, D.J.; Kraatz, H.B. Interaction of metal ions with tau protein. The case for a metal-mediated tau aggregation. J. Inorg. Biochem. 2019, 194, 44–51. [Google Scholar] [CrossRef]
  159. Wei, B.; McGuffey, J.E.; Blount, B.C.; Wang, L. Sensitive Quantification of Cannabinoids in Milk by Alkaline Saponification-Solid Phase Extraction Combined with Isotope Dilution UPLC-MS/MS. ACS Omega 2016, 1, 1307–1313. [Google Scholar] [CrossRef] [PubMed]
  160. Magen, I.; Avraham, Y.; Ackerman, Z.; Vorobiev, L.; Mechoulam, R.; Berry, E.M. Cannabidiol ameliorates cognitive and motor impairments in bile-duct ligated mice via 5-HT1A receptor activation. Br. J. Pharmacol. 2010, 159, 950–957. [Google Scholar] [CrossRef] [PubMed]
  161. Blázquez, C.; Chiarlone, A.; Bellocchio, L.; Resel, E.; Pruunsild, P.; García-Rincón, D.; Sendtner, M.; Timmusk, T.; Lutz, B.; Galve-Roperh, I.; et al. The CB₁ cannabinoid receptor signals striatal neuroprotection via a PI3K/Akt/mTORC1/BDNF pathway. Cell Death Differ. 2015, 22, 1618–1629. [Google Scholar] [CrossRef] [PubMed]
  162. Valdeolivas, S.; Satta, V.; Pertwee, R.G.; Fernández-Ruiz, J.; Sagredo, O. Sativex-like combination of phytocannabinoids is neuroprotective in malonate-lesioned rats, an inflammatory model of Huntington’s disease: Role of CB1 and CB2 receptors. ACS Chem. Neurosci. 2012, 3, 400–406. [Google Scholar] [CrossRef]
  163. Calkins, M.J.; Jakel, R.J.; Johnson, D.A.; Chan, K.; Kan, Y.W.; Johnson, J.A. Protection from mitochondrial complex II inhibition in vitro and in vivo by Nrf2-mediated transcription. Proc. Natl. Acad. Sci. USA 2005, 102, 244–249. [Google Scholar] [CrossRef]
  164. Calkins, M.J.; Townsend, J.A.; Johnson, D.A.; Johnson, J.A. Cystamine protects from 3-nitropropionic acid lesioning via induction of nf-e2 related factor 2 mediated transcription. Exp. Neurol. 2010, 224, 307–317. [Google Scholar] [CrossRef]
  165. Sagredo, O.; Pazos, M.R.; Satta, V.; Ramos, J.A.; Pertwee, R.G.; Fernández-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]
  166. Borovac Štefanović, L.; Kalinić, D.; Mimica, N.; Beer Ljubić, B.; Aladrović, J.; Mandelsamen Perica, M.; Curić, M.; Grošić, P.F.; Delaš, I. Oxidative status and the severity of clinical symptoms in patients with post-traumatic stress disorder. Ann. Clin. Biochem. 2015, 52, 95–104. [Google Scholar] [CrossRef]
  167. Fidelman, S.; Mizrachi Zer-Aviv, T.; Lange, R.; Hillard, C.J.; Akirav, I. Chronic treatment with URB597 ameliorates post-stress symptoms in a rat model of PTSD. Eur. Neuropsychopharmacol. 2018, 28, 630–642. [Google Scholar] [CrossRef] [PubMed]
  168. Orsolini, L.; Chiappini, S.; Volpe, U.; Berardis, D.; Latini, R.; Papanti, G.D.; Corkery, A.J.M. Use of Medicinal Cannabis and Synthetic Cannabinoids in Post-Traumatic Stress Disorder (PTSD): A Systematic Review. Medicina 2019, 55, 525. [Google Scholar] [CrossRef]
  169. de Munter, J.; Pavlov, D.; Gorlova, A.; Sicker, M.; Proshin, A.; Kalueff, A.V.; Svistunov, A.; Kiselev, D.; Nedorubov, A.; Morozov, S.; et al. Increased Oxidative Stress in the Prefrontal Cortex as a Shared Feature of Depressive- and PTSD-Like Syndromes: Effects of a Standardized Herbal Antioxidant. Front. Nutr. 2021, 8, 661455. [Google Scholar] [CrossRef] [PubMed]
  170. Sales, A.J.; Fogaça, M.V.; Sartim, A.G.; Pereira, V.S.; Wegener, G.; Guimarães, F.S.; Joca, S.R.L. Cannabidiol Induces Rapid and Sustained Antidepressant-Like Effects Through Increased BDNF Signaling and Synaptogenesis in the Prefrontal Cortex. Mol. Neurobiol. 2019, 56, 1070–1081. [Google Scholar] [CrossRef]
  171. Linge, R.; Jiménez-Sánchez, L.; Campa, L.; Pilar-Cuéllar, F.; Vidal, R.; Pazos, A.; Adell, A.; Díaz, Á. Cannabidiol induces rapid-acting antidepressant-like effects and enhances cortical 5-HT/glutamate neurotransmission: Role of 5-HT1A receptors. Neuropharmacology 2016, 103, 16–26. [Google Scholar] [CrossRef] [Green Version]
  172. 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]
  173. Surkin, P.N.; Gallino, S.L.; Luce, V.; Correa, F.; Fernandez-Solari, J.; De Laurentiis, A. Pharmacological augmentation of endocannabinoid signaling reduces the neuroendocrine response to stress. Psychoneuroendocrinology 2018, 87, 131–140. [Google Scholar] [CrossRef]
  174. Mansouri, A.; Demeilliers, C.; Amsellem, S.; Pessayre, D.; Fromenty, B. Acute ethanol administration oxidatively damages and depletes mitochondrial dna in mouse liver, brain, heart, and skeletal muscles: Protective effects of antioxidants. J. Pharmacol. Exp. Ther. 2001, 298, 737–743. [Google Scholar]
  175. Hao, E.; Mukhopadhyay, P.; Cao, Z.; Erdélyi, K.; Holovac, E.; Liaudet, L.; Lee, W.S.; Haskó, G.; Mechoulam, R.; Pacher, P. Cannabidiol Protects against Doxorubicin-Induced Cardiomyopathy by Modulating Mitochondrial Function and Biogenesis. Mol. Med. 2015, 21, 38–45. [Google Scholar] [CrossRef] [PubMed]
  176. Chen, H.; Weng, Q.Y.; Fisher, D.E. UV signaling pathways within the skin. J. Investig. Dermatol. 2014, 134, 2080–2085. [Google Scholar] [CrossRef]
  177. Atalay, S.; Gęgotek, A.; Wroński, A.; Domigues, P.; Skrzydlewska, E. Therapeutic application of cannabidiol on UVA and UVB irradiated rat skin. A proteomic study. J. Pharm. Biomed. Anal. 2021, 192, 113656. [Google Scholar] [CrossRef]
  178. Holmgren, A.; Lu, J. Thioredoxin and thioredoxin reductase: Current research with special reference to human disease. Biochem. Biophys. Res. Commun. 2010, 396, 120–124. [Google Scholar] [CrossRef]
  179. Łuczaj, W.; Domingues, M.D.R.; Domingues, P.; Skrzydlewska, E. Changes in Lipid Profile of Keratinocytes from Rat Skin Exposed to Chronic UVA or UVB Radiation and Topical Application of Cannabidiol. Antioxidants 2020, 9, 1178. [Google Scholar] [CrossRef] [PubMed]
  180. Petrovici, A.R.; Simionescu, N.; Sandu, A.I.; Paraschiv, V.; Silion, M.; Pinteala, M. New Insights on Hemp Oil Enriched in Cannabidiol: Decarboxylation, Antioxidant Properties and In Vitro Anticancer Effect. Antioxidants 2021, 10, 738. [Google Scholar] [CrossRef]
  181. Biernacki, M.; Jastrząb, A.; Skrzydlewska, E. Changes in Hepatic Phospholipid Metabolism in Rats under UV Irradiation and Topically Treated with Cannabidiol. Antioxidants 2021, 10, 1157. [Google Scholar] [CrossRef]
  182. Olivas-Aguirre, M.; Torres-López, L.; Valle-Reyes, J.S.; Hernández-Cruz, A.; Pottosin, I.; Dobrovinskaya, O. Cannabidiol directly targets mitochondria and disturbs calcium homeostasis in acute lymphoblastic leukemia. Cell Death Dis. 2019, 10, 779. [Google Scholar] [CrossRef] [Green Version]
  183. 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]
  184. 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] [PubMed]
  185. Costa, B.; Colleoni, M.; Conti, S.; Parolaro, D.; Franke, C.; Trovato, A.E.; Giagnoni, G. Oral anti-inflammatory activity of cannabidiol, a non-psychoactive constituent of cannabis, in acute carrageenan-induced inflammation in the rat paw. Naunyn Schmiedebergs Arch. Pharmacol. 2004, 369, 294–299. [Google Scholar] [CrossRef]
  186. Oh, J.Y.; Giles, N.; Landar, A.; Darley-Usmar, V. Accumulation of 15-deoxy-delta(12,14)-prostaglandin J2 adduct formation with Keap1 over time: Effects on potency for intracellular antioxidant defence induction. Biochem. J. 2008, 411, 297–306. [Google Scholar] [CrossRef] [PubMed]
  187. Gilroy, D.W. Eicosanoids and the endogenous control of acute inflammatory resolution. Int. J. Biochem. Cell Biol. 2010, 42, 524–528. [Google Scholar] [CrossRef]
  188. Jastrząb, A.; Gęgotek, A.; Skrzydlewska, E. Cannabidiol Regulates the Expression of Keratinocyte Proteins Involved in the Inflammation Process through Transcriptional Regulation. Cells 2019, 8, 827. [Google Scholar] [CrossRef] [PubMed]
  189. Jaramillo, M.C.; Zhang, D.D. The emerging role of the Nrf2-Keap1 signaling pathway in cancer. Genes Dev. 2013, 27, 2179–2191. [Google Scholar] [CrossRef] [PubMed]
  190. Pi, J.; Diwan, B.A.; Sun, Y.; Liu, J.; Qu, W.; He, Y.; Styblo, M.; Waalkes, M.P. Arsenic-induced malignant transformation of human keratinocytes: Involvement of Nrf2. Free Radic. Biol. Med. 2008, 45, 651–658. [Google Scholar] [CrossRef] [PubMed]
  191. Gęgotek, A.; Rybałtowska-Kawałko, P.; Skrzydlewska, E. Rutin as a Mediator of Lipid Metabolism and Cellular Signaling Pathways Interactions in Fibroblasts Altered by UVA and UVB Radiation. Oxid. Med. Cell Longev. 2017, 2017, 4721352. [Google Scholar] [CrossRef] [PubMed]
  192. Juknat, A.; Pietr, M.; Kozela, E.; Rimmerman, N.; Levy, R.; Coppola, G.; Geschwind, D.; Vogel, Z. Differential transcriptional profiles mediated by exposure to the cannabinoids cannabidiol and Δ9-tetrahydrocannabinol in BV-2 microglial cells. Br. J. Pharmacol. 2012, 165, 2512–2528. [Google Scholar] [CrossRef] [Green Version]
  193. Chianese, G.; Sirignano, C.; Benetti, E.; Marzaroli, V.; Collado, J.A.; de la Vega, L.; Appendino, G.; Muñoz, E.; Taglialatela-Scafati, O. A Nrf-2 Stimulatory Hydroxylated Cannabidiol Derivative from Hemp (Cannabis sativa). J. Nat. Prod. 2022, 85, 1089–1097. [Google Scholar] [CrossRef]
  194. Zhang, X.; Guo, J.; Wei, X.; Niu, C.; Jia, M.; Li, Q.; Meng, D. Bach1: Function, Regulation, and Involvement in Disease. Oxid. Med. Cell Longev. 2018, 2018, 1347969. [Google Scholar] [CrossRef]
  195. Zhornitsky, S.; Potvin, S. Cannabidiol in humans-the quest for therapeutic targets. Pharmaceuticals 2012, 5, 529–552. [Google Scholar] [CrossRef]
  196. Pazos, M.R.; Cinquina, V.; Gómez, A.; Layunta, R.; Santos, M.; Fernández-Ruiz, J.; Martínez-Orgado, J. Cannabidiol administration after hypoxia-ischemia to newborn rats reduces long-term brain injury and restores neurobehavioral function. Neuropharmacology 2012, 63, 776–783. [Google Scholar] [CrossRef]
  197. Osborne, A.L.; Solowij, N.; Weston-Green, K. A systematic review of the effect of cannabidiol on cognitive function: Relevance to schizophrenia. Neurosci. Biobehav. Rev. 2017, 72, 310–324. [Google Scholar] [CrossRef]
  198. Martin, R.C.; Gaston, T.E.; Thompson, M.; Ampah, S.B.; Cutter, G.; Bebin, E.M.; Szaflarski, J.P. Cognitive functioning following long-term cannabidiol use in adults with treatment-resistant epilepsy. Epilepsy Behav. 2019, 97, 105–110. [Google Scholar] [CrossRef] [PubMed]
  199. Leweke, F.M.; Rohleder, C.; Gerth, C.W.; Hellmich, M.; Pukrop, R.; Koethe, D. Cannabidiol and Amisulpride Improve Cognition in Acute Schizophrenia in an Explorative, Double-Blind, Active-Controlled, Randomized Clinical Trial. Front. Pharmacol. 2021, 12, 614811. [Google Scholar] [CrossRef]
  200. Morgan, C.J.; Schafer, G.; Freeman, T.P.; Curran, H.V. Impact of cannabidiol on the acute memory and psychotomimetic effects of smoked cannabis: Naturalistic study: Naturalistic study [corrected]. Br. J. Psychiatry 2010, 197, 285–290. [Google Scholar] [CrossRef]
  201. McGuire, J.L.; DePasquale, E.A.K.; Watanabe, M.; Anwar, F.; Ngwenya, L.B.; Atluri, G.; Romick-Rosendale, L.E.; McCullumsmith, R.E.; Evanson, N.K. Chronic Dysregulation of Cortical and Subcortical Metabolism After Experimental Traumatic Brain Injury. Mol. Neurobiol. 2019, 56, 2908–2921. [Google Scholar] [CrossRef]
  202. Tasker, J.G.; Chen, C.; Fisher, M.O.; Fu, X.; Rainville, J.R.; Weiss, G.L. Endocannabinoid Regulation of Neuroendocrine Systems. Int. Rev. Neurobiol. 2015, 125, 163–201. [Google Scholar] [CrossRef] [PubMed]
  203. Ueberall, M.A.; Essner, U.; Mueller-Schwefe, G.H. Effectiveness and tolerability of THC:CBD oromucosal spray as add-on measure in patients with severe chronic pain: Analysis of 12-week open-label real-world data provided by the German Pain e-Registry. J. Pain Res. 2019, 12, 1577–1604. [Google Scholar] [CrossRef]
  204. Sarris, J.; Sinclair, J.; Karamacoska, D.; Davidson, M.; Firth, J. Medicinal cannabis for psychiatric disorders: A clinically-focused systematic review. BMC Psychiatry 2020, 20, 24. [Google Scholar] [CrossRef]
  205. Osborne, A.L.; Solowij, N.; Babic, I.; Huang, X.F.; Weston-Green, K. Improved Social Interaction, Recognition and Working Memory with Cannabidiol Treatment in a Prenatal Infection (poly I:C) Rat Model. Neuropsychopharmacology 2017, 42, 1447–1457. [Google Scholar] [CrossRef] [PubMed]
  206. Demirakca, T.; Sartorius, A.; Ende, G.; Meyer, N.; Welzel, H.; Skopp, G.; Mann, K.; Hermann, D. Diminished gray matter in the hippocampus of cannabis users: Possible protective effects of cannabidiol. Drug Alcohol Depend. 2011, 114, 242–245. [Google Scholar] [CrossRef]
  207. Friedman, L.K.; Peng, H.; Zeman, R.J. Cannabidiol reduces lesion volume and restores vestibulomotor and cognitive function following moderately severe traumatic brain injury. Exp. Neurol. 2021, 346, 113844. [Google Scholar] [CrossRef] [PubMed]
  208. Fagherazzi, E.V.; Garcia, V.A.; Maurmann, N.; Bervanger, T.; Halmenschlager, L.H.; Busato, S.B.; Hallak, J.E.; Zuardi, A.W.; Crippa, J.A.; Schröder, N. Memory-rescuing effects of cannabidiol in an animal model of cognitive impairment relevant to neurodegenerative disorders. Psychopharmacology 2012, 219, 1133–1140. [Google Scholar] [CrossRef] [PubMed]
  209. Cassol, O.J., Jr.; Comim, C.M.; Silva, B.R.; Hermani, F.V.; Constantino, L.S.; Felisberto, F.; Petronilho, F.; Hallak, J.E.; De Martinis, B.S.; Zuardi, A.W.; et al. Treatment with cannabidiol reverses oxidative stress parameters, cognitive impairment and mortality in rats submitted to sepsis by cecal ligation and puncture. Brain Res. 2010, 1348, 128–138. [Google Scholar] [CrossRef]
  210. Halliwell, B. Oxidative stress and neurodegeneration: Where are we now? J. Neurochem. 2006, 97, 1634–1658. [Google Scholar] [CrossRef] [PubMed]
  211. Tuon, L.; Comim, C.M.; Petronilho, F.; Barichello, T.; Izquierdo, I.; Quevedo, J.; Dal-Pizzol, F. Time-dependent behavioral recovery after sepsis in rats. Intensive Care Med. 2008, 34, 1724–1731. [Google Scholar] [CrossRef]
  212. Boggs, D.L.; Surti, T.; Gupta, A.; Gupta, S.; Niciu, M.; Pittman, B.; Schnakenberg Martin, A.M.; Thurnauer, H.; Davies, A.; D’Souza, D.C.; et al. The effects of cannabidiol (CBD) on cognition and symptoms in outpatients with chronic schizophrenia a randomized placebo controlled trial. Psychopharmacology 2018, 235, 1923–1932. [Google Scholar] [CrossRef]
  213. Blaes, S.L.; Orsini, C.A.; Holik, H.M.; Stubbs, T.D.; Ferguson, S.N.; Heshmati, S.C.; Bruner, M.M.; Wall, S.C.; Febo, M.; Bruijnzeel, A.W.; et al. Enhancing effects of acute exposure to cannabis smoke on working memory performance. Neurobiol. Learn. Mem. 2019, 157, 151–162. [Google Scholar] [CrossRef]
  214. McLaughlin, R.J.; Hill, M.N.; Gorzalka, B.B. A critical role for prefrontocortical endocannabinoid signaling in the regulation of stress and emotional behavior. Neurosci. Biobehav. Rev. 2014, 42, 116–131. [Google Scholar] [CrossRef]
  215. Khodadadi, H.; Salles, É.L.; Jarrahi, A.; Costigliola, V.; Khan, M.B.; Yu, J.C.; Morgan, J.C.; Hess, D.C.; Vaibhav, K.; Dhandapani, K.M.; et al. Cannabidiol Ameliorates Cognitive Function via Regulation of IL-33 and TREM2 Upregulation in a Murine Model of Alzheimer’s Disease. J. Alzheimer’s Dis. 2021, 80, 973–977. [Google Scholar] [CrossRef]
  216. Peres, F.F.; Levin, R.; Suiama, M.A.; Diana, M.C.; Gouvêa, D.A.; Almeida, V.; Santos, C.M.; Lungato, L.; Zuardi, A.W.; Hallak, J.E.; et al. Cannabidiol Prevents Motor and Cognitive Impairments Induced by Reserpine in Rats. Front. Pharmacol. 2016, 7, 343. [Google Scholar] [CrossRef]
  217. Razavi, Y.; Shabani, R.; Mehdizadeh, M.; Haghparast, A. Neuroprotective effect of chronic administration of cannabidiol during the abstinence period on methamphetamine-induced impairment of recognition memory in the rats. Behav. Pharmacol. 2020, 31, 385–396. [Google Scholar] [CrossRef] [PubMed]
  218. Renard, J.; Loureiro, M.; Rosen, L.G.; Zunder, J.; de Oliveira, C.; Schmid, S.; Rushlow, W.J.; Laviolette, S.R. Cannabidiol Counteracts Amphetamine-Induced Neuronal and Behavioral Sensitization of the Mesolimbic Dopamine Pathway through a Novel mTOR/p70S6 Kinase Signaling Pathway. J. Neurosci. 2016, 36, 5160–5169. [Google Scholar] [CrossRef]
  219. 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]
  220. Barichello, T.; Ceretta, R.A.; Generoso, J.S.; Moreira, A.P.; Simões, L.R.; Comim, C.M.; Quevedo, J.; Vilela, M.C.; Zuardi, A.W.; Crippa, J.A.; et al. Cannabidiol reduces host immune response and prevents cognitive impairments in Wistar rats submitted to pneumococcal meningitis. Eur. J. Pharmacol. 2012, 697, 158–164. [Google Scholar] [CrossRef] [PubMed]
  221. Martín-Moreno, A.M.; Reigada, D.; Ramírez, 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] [PubMed]
  222. García-Baos, A.; Puig-Reyne, X.; García-Algar, Ó.; Valverde, O. Cannabidiol attenuates cognitive deficits and neuroinflammation induced by early alcohol exposure in a mice model. Biomed. Pharmacother. 2021, 141, 111813. [Google Scholar] [CrossRef]
  223. Mori, M.A.; Meyer, E.; Soares, L.M.; Milani, H.; Guimarães, F.S.; de Oliveira, R.M.W. Cannabidiol reduces neuroinflammation and promotes neuroplasticity and functional recovery after brain ischemia. Prog. Neuropsychopharmacol. Biol. Psychiatry 2017, 75, 94–105. [Google Scholar] [CrossRef] [PubMed]
  224. Mori, M.A.; Meyer, E.; da Silva, F.F.; Milani, H.; Guimarães, F.S.; Oliveira, R.M.W. Differential contribution of CB1, CB2, 5-HT1A, and PPAR-γ receptors to cannabidiol effects on ischemia-induced emotional and cognitive impairments. Eur. J. Neurosci. 2021, 53, 1738–1751. [Google Scholar] [CrossRef]
  225. Magen, I.; Avraham, Y.; Ackerman, Z.; Vorobiev, L.; Mechoulam, R.; Berry, E.M. Cannabidiol ameliorates cognitive and motor impairments in mice with bile duct ligation. J. Hepatol. 2009, 51, 528–534. [Google Scholar] [CrossRef]
  226. Moreira, F.A.; Aguiar, D.C.; Terzian, A.L.; Guimarães, F.S.; Wotjak, C.T. Cannabinoid type 1 receptors and transient receptor potential vanilloid type 1 channels in fear and anxiety-two sides of one coin? Neuroscience 2012, 204, 186–192. [Google Scholar] [CrossRef] [PubMed]
  227. Kozela, E.; Krawczyk, M.; Kos, T.; Juknat, A.; Vogel, Z.; Popik, P. Cannabidiol Improves Cognitive Impairment and Reverses Cortical Transcriptional Changes Induced by Ketamine, in Schizophrenia-Like Model in Rats. Mol. Neurobiol. 2020, 57, 1733–1747. [Google Scholar] [CrossRef] [PubMed]
  228. Corrê, M.D.S.; de Freitas, B.S.; Machado, G.D.B.; Pires, V.N.; Bromberg, E.; Hallak, J.E.C.; Zuardi, A.W.; Crippa, J.A.S.; Schröder, N. Cannabidiol reverses memory impairments and activates components of the Akt/GSK3β pathway in an experimental model of estrogen depletion. Behav. Brain Res. 2022, 417, 113555. [Google Scholar] [CrossRef] [PubMed]
  229. de Paula Faria, D.; Estessi de Souza, L.; Duran, F.L.S.; Buchpiguel, C.A.; Britto, L.R.; Crippa, J.A.S.; Filho, G.B.; Real, C.C. Cannabidiol Treatment Improves Glucose Metabolism and Memory in Streptozotocin-Induced Alzheimer’s Disease Rat Model: A Proof-of-Concept Study. Int. J. Mol. Sci. 2022, 23, 1076. [Google Scholar] [CrossRef] [PubMed]
  230. Kruk-Slomka, M.; Biala, G. Cannabidiol Attenuates MK-801-Induced Cognitive Symptoms of Schizophrenia in the Passive Avoidance Test in Mice. Molecules 2021, 26, 5977. [Google Scholar] [CrossRef]
  231. The Prohibited List. Available online: https://www.wada-ama.org/en/prohibited-list#search-anchor (accessed on 5 November 2022).
  232. Bergamaschi, M.M.; Crippa, J.A. Why should Cannabis be Considered Doping in Sports? Front. Psychiatry 2013, 4, 32. [Google Scholar] [CrossRef]
  233. Kennedy, M.C. Cannabis: Exercise performance and sport. A systematic review. J. Sci. Med. Sport 2017, 20, 825–829. [Google Scholar] [CrossRef]
  234. Ware, M.A.; Jensen, D.; Barrette, A.; Vernec, A.; Derman, W. Cannabis and the Health and Performance of the Elite Athlete. Clin J. Sport Med. 2018, 28, 480–484. [Google Scholar] [CrossRef]
  235. WADA. Summary of Major Modifications and Explanatory Notes. 2020 Prohibited List; S8 Cannabinoids; World Anti-Doping Agency: Montreal, QC, Canada, 2019; Available online: https://www.wada-ama.org/sites/default/files/wada_2020_english_summary_of_modifications_.pdf (accessed on 5 November 2022).
  236. Lachenmeier, D.W.; Diel, P. A Warning against the Negligent Use of Cannabidiol in Professional and Amateur Athletes. Sports 2019, 7, 251. [Google Scholar] [CrossRef]
  237. Hindocha, C.; Freeman, T.P.; Schafer, G.; Gardener, C.; Das, R.K.; Morgan, C.J.; Curran, H.V. Acute effects of delta-9-tetrahydrocannabinol, cannabidiol and their combination on facial emotion recognition: A randomised, double-blind, placebo-controlled study in cannabis users. Eur. Neuropsychopharmacol. 2015, 25, 325–334. [Google Scholar] [CrossRef]
  238. Masataka, N. Anxiolytic Effects of Repeated Cannabidiol Treatment in Teenagers with Social Anxiety Disorders. Front. Psychol. 2019, 10, 2466. [Google Scholar] [CrossRef]
  239. Isenmann, E.; Blume, F.; Bizjak, D.A.; Hundsdörfer, V.; Pagano, S.; Schibrowski, S.; Simon, W.; Schmandra, L.; Diel, P. Comparison of Pro-Regenerative Effects of Carbohydrates and Protein Administrated by Shake and Non-Macro-Nutrient Matched Food Items on the Skeletal Muscle after Acute Endurance Exercise. Nutrients 2019, 11, 744. [Google Scholar] [CrossRef] [PubMed]
  240. Lieberman, H.R.; Marriott, B.P.; Williams, C.; Judelson, D.A.; Glickman, E.L.; Geiselman, P.J.; Dotson, L.; Mahoney, C.R. Patterns of dietary supplement use among college students. Clin. Nutr. 2015, 34, 976–985. [Google Scholar] [CrossRef]
  241. Tavares, F.; Smith, T.B.; Driller, M. Fatigue and Recovery in Rugby: A Review. Sports Med. 2017, 47, 1515–1530. [Google Scholar] [CrossRef] [PubMed]
  242. Wilson, L.J.; Cockburn, E.; Paice, K.; Sinclair, S.; Faki, T.; Hills, F.A.; Gondek, M.B.; Wood, A.; Dimitriou, L. Recovery following a marathon: A comparison of cold water immersion, whole body cryotherapy and a placebo control. Eur. J. Appl. Physiol. 2018, 118, 153–163. [Google Scholar] [CrossRef] [PubMed]
  243. Isenmann, E.; Veit, S.; Starke, L.; Flenker, U.; Diel, P. Effects of Cannabidiol Supplementation on Skeletal Muscle Regeneration after Intensive Resistance Training. Nutrients 2021, 13, 3028. [Google Scholar] [CrossRef] [PubMed]
  244. Philpott, H.T.; O’Brien, M.; McDougall, J.J. Attenuation of early phase inflammation by cannabidiol prevents pain and nerve damage in rat osteoarthritis. Pain 2017, 158, 2442–2451. [Google Scholar] [CrossRef]
  245. Iannotti, F.A.; Pagano, E.; Moriello, A.S.; Alvino, F.G.; Sorrentino, N.C.; D’Orsi, L.; Gazzerro, E.; Capasso, R.; De Leonibus, E.; De Petrocellis, L.; et al. Effects of non-euphoric plant cannabinoids on muscle quality and performance of dystrophic mdx mice. Br. J. Pharmacol. 2019, 176, 1568–1584. [Google Scholar] [CrossRef]
  246. Langer, H.T.; Mossakowski, A.A.; Pathak, S.; Mascal, M.; Baar, K. Cannabidiol Does Not Impair Anabolic Signaling Following Eccentric Contractions in Rats. Int. J. Sport Nutr. Exerc. Metab. 2021, 31, 93–100. [Google Scholar] [CrossRef]
  247. Cochrane-Snyman, K.C.; Cruz, C.; Morales, J.; Coles, M. The Effects of Cannabidiol Oil on Noninvasive Measures of Muscle Damage in Men. Med. Sci. Sports Exerc. 2021, 53, 1460–1472. [Google Scholar] [CrossRef]
  248. Hatchett, A.; Armstrong, K.; Hughes, B.; Parr, B.B. The influence cannabidiol on delayed onset of muscle soreness. Int. J. Phys. Educ. Sports Health 2020, 7, 89–94. [Google Scholar]
  249. Gamelin, F.X.; Cuvelier, G.; Mendes, A.; Aucouturier, J.; Berthoin, S.; Di Marzo, V.; Heyman, E. Cannabidiol in sport: Ergogenic or else? Pharmacol. Res. 2020, 156, 104764. [Google Scholar] [CrossRef]
  250. Burstein, S. Cannabidiol (CBD) and its analogs: A review of their effects on inflammation. Bioorg. Med. Chem. 2015, 23, 1377–1385. [Google Scholar] [CrossRef]
  251. McCartney, D.; Benson, M.J.; Desbrow, B.; Irwin, C.; Suraev, A.; McGregor, I.S. Cannabidiol and Sports Performance: A Narrative Review of Relevant Evidence and Recommendations for Future Research. Sports Med. Open 2020, 6, 27. [Google Scholar] [CrossRef]
  252. Shrivastava, A.; Kuzontkoski, P.M.; Groopman, J.E.; Prasad, A. Cannabidiol induces programmed cell death in breast cancer cells by coordinating the cross-talk between apoptosis and autophagy. Mol. Cancer Ther. 2011, 10, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  253. Ohsumi, Y. Historical landmarks of autophagy research. Cell Res. 2014, 24, 9–23. [Google Scholar] [CrossRef] [PubMed]
  254. Lee, X.C.; Werner, E.; Falasca, M. Molecular Mechanism of Autophagy and Its Regulation by Cannabinoids in Cancer. Cancers 2021, 13, 1211. [Google Scholar] [CrossRef]
  255. Wang, W.; Wang, X.; Fujioka, H.; Hoppel, C.; Whone, A.L.; Caldwell, M.A.; Cullen, P.J.; Liu, J.; Zhu, X. Parkinson’s disease-associated mutant VPS35 causes mitochondrial dysfunction by recycling DLP1 complexes. Nat. Med. 2016, 2, 54–63. [Google Scholar] [CrossRef] [Green Version]
  256. Kang, S.; Li, J.; Yao, Z.; Liu, J. Cannabidiol Induces Autophagy to Protects Neural Cells from Mitochondrial Dysfunction by Upregulating SIRT1 to Inhibits NF-κB and NOTCH Pathways. Front. Cell. Neurosci. 2021, 15, 654340. [Google Scholar] [CrossRef] [PubMed]
  257. Choi, H.I.; Kim, H.J.; Park, J.S.; Kim, I.J.; Bae, E.H.; Ma, S.K.; Kim, S.W. PGC-1α attenuates hydrogen peroxide-induced apoptotic cell death by upregulating Nrf-2 via GSK3β inactivation mediated by activated p38 in HK-2 Cells. Sci. Rep. 2017, 7, 4319. [Google Scholar] [CrossRef]
  258. Moors, T.E.; Hoozemans, J.J.; Ingrassia, A.; Beccari, T.; Parnetti, L.; Chartier-Harlin, M.C.; van de Berg, W.D. Therapeutic potential of autophagy-enhancing agents in Parkinson’s disease. Mol. Neurodegener. 2017, 12, 11. [Google Scholar] [CrossRef]
  259. Vrechi, T.A.M.; Leão, A.H.F.F.; Morais, I.B.M.; Abílio, V.C.; Zuardi, A.W.; Hallak, J.E.C.; Crippa, J.A.; Bincoletto, C.; Ureshino, R.P.; Smaili, S.S.; et al. Cannabidiol induces autophagy via ERK1/2 activation in neural cells. Sci. Rep. 2021, 11, 5434. [Google Scholar] [CrossRef]
  260. Pandelides, Z.; Thornton, C.; Faruque, A.S.; Whitehead, A.P.; Willett, K.L.; Ashpole, N.M. Developmental exposure to cannabidiol (CBD) alters longevity and health span of zebrafish (Danio rerio). Geroscience 2020, 42, 785–800. [Google Scholar] [CrossRef] [PubMed]
  261. Land, M.H.; Toth, M.L.; MacNair, L.; Vanapalli, S.A.; Lefever, T.W.; Peters, E.N.; Bonn-Miller, M.O. Effect of Cannabidiol on the Long-Term Toxicity and Lifespan in the Preclinical Model Caenorhabditis elegans. Cannabis Cannabinoid Res. 2021, 6, 522–527. [Google Scholar] [CrossRef]
  262. Wang, Z.; Zheng, P.; Xie, Y.; Chen, X.; Solowij, N.; Green, K.; Chew, Y.L.; Huang, X.F. Cannabidiol regulates CB1-pSTAT3 signaling for neurite outgrowth, prolongs lifespan, and improves health span in Caenorhabditis elegans of Aβ pathology models. FASEB J. 2021, 35, e21537. [Google Scholar] [CrossRef] [PubMed]
  263. Wang, Z.; Zheng, P.; Chen, X.; Xie, Y.; Weston-Green, K.; Solowij, N.; Chew, Y.L.; Huang, X.F. Cannabidiol induces autophagy and improves neuronal health associated with SIRT1 mediated longevity. Geroscience 2022, 44, 1505–1524. [Google Scholar] [CrossRef] [PubMed]
  264. Chen, C.H.; Chen, Y.C.; Jiang, H.C.; Chen, C.K.; Pan, C.L. Neuronal aging: Learning from C. elegans. J. Mol. Signal. 2013, 8, 14. [Google Scholar] [CrossRef]
  265. Chan, J.Z.; Duncan, R.E. Regulatory Effects of Cannabidiol on Mitochondrial Functions: A Review. Cells 2021, 10, 1251. [Google Scholar] [CrossRef]
  266. da Silva, V.K.; de Freitas, B.S.; da Silva Dornelles, A.; Nery, L.R.; Falavigna, L.; Ferreira, R.D.; Bogo, M.R.; Hallak, J.E.; Zuardi, A.W.; Crippa, J.A.; et al. Cannabidiol normalizes caspase 3, synaptophysin, and mitochondrial fission protein DNM1L expression levels in rats with brain iron overload: Implications for neuroprotection. Mol. Neurobiol. 2014, 49, 222–233. [Google Scholar] [CrossRef]
  267. Castillo, A.; Tolón, M.R.; Fernández-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]
Figure 1. Cannabidiol has the property of inhibiting the fatty acid amide hydrolase (FAAH) enzyme with a consequent increase in the availability of anandamide (AEA) and 2-arachidonoylglycerol (2-AG), which are substances of the endocannabinoid system considered to be reverse mediators of depolarization-induced inhibition. Thus, it prevents the release of glutamate, which no longer acts on AMPA and NMDA receptors. At the same time, CBD, through TRPV1, GPR55, and CB2 receptors, prevents the generation of cytokines with a pro-inflammatory role, and in addition, through the PPARγ receptor, it increases the expression of endogenous antioxidant systems via the Nrf2 pathway. Created with BioRender.com.
Figure 1. Cannabidiol has the property of inhibiting the fatty acid amide hydrolase (FAAH) enzyme with a consequent increase in the availability of anandamide (AEA) and 2-arachidonoylglycerol (2-AG), which are substances of the endocannabinoid system considered to be reverse mediators of depolarization-induced inhibition. Thus, it prevents the release of glutamate, which no longer acts on AMPA and NMDA receptors. At the same time, CBD, through TRPV1, GPR55, and CB2 receptors, prevents the generation of cytokines with a pro-inflammatory role, and in addition, through the PPARγ receptor, it increases the expression of endogenous antioxidant systems via the Nrf2 pathway. Created with BioRender.com.
Antioxidants 12 00485 g001
Figure 2. Under the action of reactive oxygen species (ROS), the Keap1-Nrf2 complex cleaves. Once cleaved, Nrf2 enters the nucleus at the DNA level, where it displaces Bach1 from Maf, in order to synthesize HO-1. Created with BioRender.com.
Figure 2. Under the action of reactive oxygen species (ROS), the Keap1-Nrf2 complex cleaves. Once cleaved, Nrf2 enters the nucleus at the DNA level, where it displaces Bach1 from Maf, in order to synthesize HO-1. Created with BioRender.com.
Antioxidants 12 00485 g002
Table 1. Experimental results of cognitive behavioral tests performed on rodents.
Table 1. Experimental results of cognitive behavioral tests performed on rodents.
Animal ModelCognitive
Impairment Model
TaskTreatment (Dose)DurationObservationsRef.
Sabra female mice (8 weeks)Hepatic encephalopathy (bile duct ligation)Eight-arm maze testCBD i.p.
(5 mg/kg)
21 daysRestored cognitive function (p = 0.001). CBD had no effect on sham animals.[160]
Male and female Wistar rats (7–10 days old)Hypoxia-ischemia (left carotid artery electrocoagulation)Novel object recognition test (NOR)CBD s.c.
(1 mg/kg)
3 daysCBD-treated rats test performance was similar to
that of sham animals (p < 0.05).
[196]
Adult male and female Sprague–Dawley and Wistar ratsControlled cortical contusion injuryT-maze test
Novel object recognition test (NOR)
CBD i.p. (40 mg/kg, 20 mg/kg)15 daysCBD treatment increased the discrimination index compared to the traumatic brain injury group, being able to discriminate the novel object.[207]
Male Wistar rats (3 weeks)Iron neonatal treatmentInhibitory avoidance task
Novel object recognition test (NOR)
CBD i.p. (2.5, 5, 10 mg/kg)14 days The 10 mg/kg dose of CBD was
able to reverse iron-induced memory deficits, recognition index of this group being significantly
higher when compared to the iron treated group (p < 0.0001).
[208]
Male Wistar rats (3–4 months)Sepsis induction-cecal ligation and perforationInhibitory avoidance testCBD i.p. (2.5, 5, 10 mg/kg)9 daysCBD in all doses
reverted the memory
alteration (p = 0.001).
[209]
Male and female Long Evans rats-Working memory testCannabis cigarettes (5.6% THC, 0% CBD, 0.4% CBN)-In male rats, cannabis smoke exposure did not affect choice accuracy. In female rats, however, exposure to cannabis smoke significantly enhanced choice accuracy. The effect on working memory accuracy, cannabis smoke increased the number of trials completed in females (p = 0.01) but not in males (p = 0.48) and decreased locomotor activity in both sexes (males: p = 0.04; females; p = 0.03). It should be noted that the performance-enhancing effect after exposure to cannabis smoke was evident only in female rats, whose initial performance was significantly lower than in males.[213]
5xFAD male mice (9–12 months)Transgenic miceNovel object recognition test (NOR)CBD i.p. (10 mg/kg)2 weeks (one dose every other day)CBD treatment improved cognitive function as
measured by NOR (Discrimination Index increased to 0.5 ± 0.9 from
−0.2 ± 0.8, p ≤ 0.04).
[215]
Male Wistar rats (3 months)Motor and cognitive impairments induced by reserpinePlus maze discriminative avoidance taskCBD i.p (0.5 and 5 mg/kg)4 daysThe time spent in aversive enclosed arm is lower than the time spent in the non-aversive enclosed arm for CBD 0.5 (p < 0.05), but not for CBD 5.[216]
Male Wistar ratsMethamphetamine chronic exposure (10 days, twice/day)The Y-maze (YM) test
Novel object recognition test (NOR)
CBD ICV microinjection
(32 and 162 nmol)
10 daysBoth doses
of CBD significantly improved spatial memory.
The higher dose of CBD
was more effective, p < 0.001). A high dose of CBD(160 nmol) could improve long-term memory p < 0.025).
[217]
Male and female offspring C57BL/6JAlcohol exposureThe Y-maze (YM) test
Novel object recognition test (NOR)
CBD i.p. (20 mg/kg)10 daysCBD-treated group showed a significantly higher preference
for the novel arm as compared to alcohol-exposed group (p < 0.05), revealing that CBD treatment hampers the detrimental effect on reference memory
caused by alcohol. CBD does not affect the recognition memory (NOR), but CBD treatment impedes the deleterious effects of alcohol on object location memory.
[222]
Male C57BL/6J mice (2–3 months)Bilateral common carotid artery occlusion (BCCAO)The Y-maze (YM) testCBD i.p. (10 mg/kg)0.5, 3, 24, 48 h after the surgeryCBD increased
in the % of the time in the novel arm of the YM when compared
to BCCAO animals treated only with vehicle p < 0.01.
[224]
Male Sprague–Dawley ratsSchizophrenia-like cognitive deficits induced by repeated ketamine administrationNovel object recognition test (NOR)CBD i.p. (7.5 mg/kg)6 days Ketamine-induced cognitive deficits were restored by CBD (p < 0.001). [227]
Adult female Wistar ratsEstrogen depletion (surgery)Inhibitory avoidanceCBD i.p. (10 mg/kg)14 daysOvariectomized rats treated with CBD had a higher latency to step down (p = 0.001).[228]
Male Wistar rats (3 weeks)ICV injection of streptozotocinNovel object recognition test (NOR)CBD i.p. (20 mg/kg)7 daysCBD-treated group showed a better performance both on short- and long-term memory (p < 0.0001).[229]
Swiss male mice (4 weeks)MK-801 injectionPassive avoidance testCBD i.p. (1, 5, 30 mg/kg)Acute experimentCBD treatment in dose of 30 mg/kg appeared to have a better performance (p < 0.001).[230]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jîtcă, G.; Ősz, B.E.; Vari, C.E.; Rusz, C.-M.; Tero-Vescan, A.; Pușcaș, A. Cannabidiol: Bridge between Antioxidant Effect, Cellular Protection, and Cognitive and Physical Performance. Antioxidants 2023, 12, 485. https://doi.org/10.3390/antiox12020485

AMA Style

Jîtcă G, Ősz BE, Vari CE, Rusz C-M, Tero-Vescan A, Pușcaș A. Cannabidiol: Bridge between Antioxidant Effect, Cellular Protection, and Cognitive and Physical Performance. Antioxidants. 2023; 12(2):485. https://doi.org/10.3390/antiox12020485

Chicago/Turabian Style

Jîtcă, George, Bianca E. Ősz, Camil E. Vari, Carmen-Maria Rusz, Amelia Tero-Vescan, and Amalia Pușcaș. 2023. "Cannabidiol: Bridge between Antioxidant Effect, Cellular Protection, and Cognitive and Physical Performance" Antioxidants 12, no. 2: 485. https://doi.org/10.3390/antiox12020485

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