Next Article in Journal
UGT72, a Major Glycosyltransferase Family for Flavonoid and Monolignol Homeostasis in Plants
Previous Article in Journal
Ensemble Deep-Learning-Enabled Clinical Decision Support System for Breast Cancer Diagnosis and Classification on Ultrasound Images
Previous Article in Special Issue
Potential and Limits of Cannabinoids in Alzheimer’s Disease Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endocannabinoid Modulation in Neurodegenerative Diseases: In Pursuit of Certainty

by
Alexandru Vasincu
1,†,
Răzvan-Nicolae Rusu
1,*,
Daniela-Carmen Ababei
1,†,
Mădălina Larion
2,3,
Walther Bild
4,5,
Gabriela Dumitrița Stanciu
6,
Carmen Solcan
7 and
Veronica Bild
1,5,6
1
Department of Pharmacodynamics and Clinical Pharmacy, “Grigore T Popa” University of Medicine and Pharmacy, 16 Universitatii Street, 700115 Iasi, Romania
2
Department of Anaesthesiology Intensive Therapy, Regional Institute of Gastroenterology and Hepatology “Prof. Dr. Octavian Fodor”, 19 Croitorilor Street, 400162 Cluj-Napoca, Romania
3
Department of Anaesthetics, Midland Regional Hospital, Longford Road, Mullingar, N91 NA43 Co. Westmeath, Ireland
4
Department of Physiology, “Grigore T Popa” University of Medicine and Pharmacy, 16 Universitatii Street, 700115 Iasi, Romania
5
Center of Biomedical Research of the Romanian Academy, 700506 Iasi, Romania
6
Center for Advanced Research and Development in Experimental Medicine (CEMEX), “Grigore T. Popa” University of Medicine and Pharmacy, 16 Universitatii Street, 700115 Iasi, Romania
7
Preclinics Department, “Ion Ionescu de la Brad” University of Life Sciences, 8 M. Sadoveanu Alley, 700489 Iasi, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biology 2022, 11(3), 440; https://doi.org/10.3390/biology11030440
Submission received: 30 January 2022 / Revised: 4 March 2022 / Accepted: 10 March 2022 / Published: 14 March 2022
(This article belongs to the Special Issue The Therapeutic Use of Cannabinoids in Neurodegenerative Disorders)

Abstract

:

Simple Summary

Neurodegenerative diseases represent an important cause of morbidity and mortality worldwide. Existing therapeutic options are limited and focus mostly on improving symptoms and reducing exacerbations. The endocannabinoid system is involved in the pathophysiology of such disorders, an idea which has been highlighted by recent scientific work. The current work focusses its attention on the importance and implications of this system and its synthetic and natural ligands in disorders such as Alzheimer’s, Parkinson’s, Huntington’s and multiple sclerosis.

Abstract

Neurodegenerative diseases are an increasing cause of global morbidity and mortality. They occur in the central nervous system (CNS) and lead to functional and mental impairment due to loss of neurons. Recent evidence highlights the link between neurodegenerative and inflammatory diseases of the CNS. These are typically associated with several neurological disorders. These diseases have fundamental differences regarding their underlying physiology and clinical manifestations, although there are aspects that overlap. The endocannabinoid system (ECS) is comprised of receptors (type-1 (CB1R) and type-2 (CB2R) cannabinoid-receptors, as well as transient receptor potential vanilloid 1 (TRPV1)), endogenous ligands and enzymes that synthesize and degrade endocannabinoids (ECBs). Recent studies revealed the involvement of the ECS in different pathological aspects of these neurodegenerative disorders. The present review will explore the roles of cannabinoid receptors (CBRs) and pharmacological agents that modulate CBRs or ECS activity with reference to Alzheimer’s Disease (AD), Parkinson’s Disease (PD), Huntington’s Disease (HD) and multiple sclerosis (MS).

1. Introduction

Neurodegenerative diseases are an increasing cause of global morbidity and mortality [1]. Pressure exerted on healthcare systems by diseases such as PD or AD make them a priority for healthcare workers [2].
It was found that in the USA, between 2016 and 2017 there were approximately 4.7 to 6.0 million individuals living with neurodegenerative disorders, which were responsible for 272,644 deaths [3]. In 2014, 1.6% of the US population lived with AD and related dementias. In 2020, the cost for AD, amyotrophic lateral sclerosis (ALS), PD and spinal muscular atrophy was $655 billion [4,5]. It is estimated that by 2060, the prevalence will double to 3.3%, meaning that approximately 13.9 million adults will be living with AD and related dementias [6]. Regarding the EU, studies show that in almost twenty years (1994–2013), deaths attributed to AD increased from 41,255 to 86,822 [7]. These numbers highlight the significant impact of these disorders on society and patients.
The neurodegenerative processes behind these disorders occur in the CNS and are associated with the alteration of neuronal morphology and functionality, thus leading to functional and mental impairment due to reduced neuronal survival and increased neuronal death [8,9].
Recent evidence also highlights the link between neurodegenerative diseases and inflammatory diseases of the CNS. Neuroinflammation is a protective mechanism of the CNS due to the activation and proliferation of B- and T-cells, neutrophils, macrophages, microglia and mast cells within the spinal cord and brain, with removal or inactivation of noxious agents such as viruses, bacterial lipopolysaccharides and inhibition of different pathogens. The persistence of the neuroinflammatory state activates microglia and astrocytes [10]. Axonal degeneration, synaptic impairment, mitochondria dysfunction, degradation of proteins and apoptosis are due to the release of various neurotoxic cytokines and chemokines [11].
These conditions are recognized in different neurological disorders, including AD, PD, HD and MS [12]. While MS is a classical example of a neuroinflammatory disease with aspects of neurodegeneration, the other three disorders have as a common feature the aggregation of misfolded proteins [13]. The degenerative disorders are demonstrably affected by changes in their conformation, thereby gaining toxicity or losing physiological functions [14], and by inflammation [15]. These diseases are the most commonly occurring forms and have fundamental differences regarding their underlying physiology and clinical manifestations, although there are aspects that overlap, such as changes in the morphology of microglia, progressive neuronal loss, oxidative stress and elevated cytokine levels [16].
This complex cascade of underlying mechanisms shows that neurodegeneration as well as neuroinflammation play important roles in these types of disorders. By tackling these two components, novel therapeutic strategies could be developed.
Cannabis sativa (family Cannabaceae), commonly called marijuana, has been widely used throughout history for medical and therapeutic purposes. There is evidence that its use dates back to 5000 years ago in what is today known as Romania [17].
The waxy resin obtained from leaves and flowers of female plants contains approximately 100 lipid soluble compounds with pharmacologically relevant activity (“cannabinoids” or “phytocannabinoids”) [18,19]. Two of these compounds are of major significance: cannabidiol (CBD) and Δ9-tetrahydrocannabinol (Δ9-THC). The structure of Δ9-THC, which is the main psychoactive phytocannabinoid, was determined in 1964 by Mechoulam and Gaoni [20,21]. Results of their work determined further research and, in the early 1990s, specific membrane receptors of Δ9-THC were discovered and opened the way to the revelation of a novel receptor system, a system that is known to us today as the ECS [22].
Δ9-THC and CBD are two extensively studied phytocannabinoids with positive effects in different neurodegenerative diseases [23]. The main difference between Δ9-THC and CBD is the fact Δ9-THC induces psychotropic effects while CBD is non-psychoactive and could reduce the side-effects of Δ9-THC [24].
Recent studies have highlighted the importance of the ECS in regard to mechanisms that underlie neurodegenerative diseases [1]. The complexity of ECS offers new perspectives into the mechanisms of neurodegenerative and neuroinflammatory components of diseases such as AD, PD, HD and MS, contributing to a more profound view on the pathophysiological aspects and thus allowing for the discovery of potential targets for their treatment [15].
The present review will explore the roles of CB1R and CB2R, and pharmacological agents that modulate CBRs or ECS activity, with reference to AD, PD, HD and MS diseases.

2. The Endocanabinoid System

ECS is a ubiquitous and complex component that influences numerous parts of the physiology and pathology of mammals [1].
This system influences a number of cognitive and neurophysiological processes such as motor function, memory, learning, energy metabolism, inflammation, nociception and neuroprotection, making it an attractive pharmacological target [25,26].
ECS consists of receptors, enzymes and transporters that regulate ECB-actions at the central level, as well as in the periphery. Important receptors are those that bind extracellularly, which are type 1 and type 2 receptors (e.g., CB1R and CB2R), ones that bind intracellularly (e.g., TRPV1), as well as receptors that are coupled to G-protein [27]. Regarding endogenous ligands, anadamide (AEA) and 2-arachidonoylglycerol (2-AG) are of importance, while the main enzymes that contribute to the metabolism of ECBs are N-acylphosphatidylethanolamine phospholipase D (NAPE-PLD), monoacylglycerol lipase (MAGL), and fatty acid amide hydrolase (FAAH) [28,29,30,31].

2.1. Endocannabinoids

Endocannabinoids (ECBs) are messengers of endogenous lipids, their name being attributed to their property to bind to CBRs.
As mentioned above, the most widely studied ECBs are AEA and 2-AG, which are derived from arachidonic acid [25]. These are involved in retrograde synaptic transmission by inhibiting the release of neurotransmitters at presynaptic level [32]. Other ECBs are virodhamine, noladin ether and N-arachidonoyldopamine, besides homo-linolenylethanolamide and docosatetraenylethanolamide [33].
The first endogenous compound identified was AEA but studies have shown that 2-AG would be a more appropriate endogenous ligand of CBRs, being the most abundant ECB in the brain [25,34]. They act as neurotransmitters being synthesized and released by neurons but differ from classical neurotransmitters in that they are synthesized and released when required, without being stored in synaptic vesicles, in a retrograde manner from postsynaptic neurons, and subsequently act on presynaptic CBRs [35,36,37].
The mechanism of action of ECBs is unconventional because it consists in synaptic retrograde transmission. It involves second messenger pathways such as cyclic AMP-increased intracellular Ca2+ concentrations that activates different protein kinases using as activation molecules the cannabinoid ligands, with particular importance placed on the control of multiple metabolic pathways carrying endogenous cannabinoids [38,39].
By gating postsynaptic-located glutamate receptors, glutamate is released under the flow of calcium ions into the synaptic cleft. The release of glutamate from the synaptic cleft into the presynaptic space is due to excessive presynaptic activity [25,37]. The process involves the activation of presynaptic metabotropic receptors triggering enzymatic activity with production of 2-AG. The ECB, released postsynaptically, moves retrogradely through the synaptic cleft, has intrinsic activity and activates CB1R and CB2R presynaptically, whereas AEA has low affinity for CB1R and very low affinity for CB2R [38]. This process involves second messenger pathways such as cyclic AMP-increased intracellular Ca2+ concentrations that activate different protein kinases [40,41].
Generally, ECBs act as retrograde synaptic signals by stimulating type 1 receptors causing suppression of neurotransmitter release. Regardless of the presynaptically released neurotransmitter (gamma-aminobutyric acid (GABA), glutamate) or the endogenous or exogenous cannabinoid ligand activating CB1R, the stimulation will always lead to decreased presynaptic transmission via retrograde control [42].
ECBs exert their neuroprotective roles while released in response to pathological factors. AEA and 2-AG have been found in the ventral tegmental area (VTA), where they act as retrograde messengers by activating CB1 receptors located on excitatory as well as inhibitory synapses. The main neurons in the VTA contain dopamine and play a major role in reward, cognition and motivational behaviors [26].

2.2. Endocanabinoid Receptors

Among numerous receptors of ECBs ligands described in the literature, the most studied are mainly CB1R, encoded by CNR1 gene and CB2R, encoded by CNR2 gene [43,44]. Through adenylyl cyclases inhibition, voltage-gated calcium channels, inward rectifying potassium channels and the activation of mitogen-activated protein kinases (MAPK), CB1R and CB2R signal transduction is mediated [45,46].
It has been suggested that cannabinoid ligands can bind and activate other targets such as GPR 18, GPR 55, GPR 119, TRPV1 and TRPV2 receptors [47].
The most abundant receptor is CB1R, which is mainly expressed centrally in neurons and glial cells. Neurological functions (memory, cognition) and emotional and pain control are modulated by the activation of this type of receptor [48,49,50].
Through proteomic system analysis and information from different sources on the mRNA, it has been shown that CB1R can be found at the mRNA level in the brain, in the pituitary gland, in adipocytes, and at the protein level in the central nervous system. CB2R can be found in blood and at the mRNA level in lymphoid tissue while also being present in different protein levels in the majority of tissues [39].
Likewise, it was shown that CB1R was mainly present in the plasma membrane and in the actin filaments, while CB2R was also found in the plasma membrane as well as in B- and T- type lymphocytes and in alveolar type 2 cells. These findings were suggested by studies that used single-cell RNA sequencing information from human normal tissues as well as protein localization information obtained from antibody-based profiling in the Cell Atlas and the HPA Cell Type Atlas [51].
In terms of localization at the central level, CB1Rs are found almost in all synapses, mainly in the basal ganglia, hippocampus, cerebellum, neocortex and hypothalamus [25,52]. CB1R can be found in presynaptic or postsynaptic levels in glutamatergic synapses, being able to control excessive-glutamate release and an overactivation of glutamatergic receptors [53]. At the same time, it can also limit the influx of calcium as well as the activation of destructive-pathways, which are calcium-dependent [54]. These receptors are also present in GABA-ergic neurons, where they are silenced under excitotoxic conditions, unlike in glutamatergic neurons, where they are activated [55,56].
CB1R can also be found in the peripheral tissue such as in the liver, kidneys and lungs, where it modulates metabolism and energy balance [57,58]. Its signaling is involved in different pathologic processes of movement disorders reproduced in animal models [59].
CB2Rs were considered “peripheral cannabinoid receptors” due to the fact that they are mainly located on endocrine and immune cells and peripheral organs (e.g., muscle, intestine, liver and testicles) [53]. High CB2R mRNA was also expressed in the spleen [48,60]. Its expression can increase up to 100-fold in inflammatory processes, after injuring the tissues, in brain trauma, or in various pathological states [25,52]. CB2R can also be found in CNS within astrocytes, brainstem and microglia in stressful conditions such as neuroinflammation. It is not present in resting microglia [61,62,63]. Upon the activation of CB2R from microglial cells, the release of anti-inflammatory cytokines has been induced, while proinflammatory cytokines and their release are inhibited [64].
Cannabinoids also have affinity to another group of receptors called vanilloid or capsaicin receptors (TRPV1) that are located on peripheral nervous system. Their presence has also been observed in sensory neurons in the skin, heart, blood vessels or lungs, and in CNS. TRPV1 receptors play an important role in transmission and modulation of pain. Their activation by AEA releases substance P and calcitonin gene-related with promotion of local vasodilation and inflammation [49,65].

2.3. Neuroprotection Roles

Components of ECS play a major role as key regulators of the immune system within neurodegenerative diseases, with important immunomodulatory effects in the brain, promoting neuroprotection by decrease of neuroinflammatory process and of its important triggers, such as oxidative stress, cellular apoptosis or toxic neural excitability.
The reduction in the immune response is mostly due to CB2R activation in microglia and brain-infiltrating immune cells. However, recent studies also demonstrate that CB1R helps regulate the immune system in the case of MS or AD models or traumatic brain injury [66,67,68,69].
Increased levels of ECBs (AEA and 2-AG) in the brain through inhibition of FAAH and MAGL enzymes are a successful therapeutic option for controlling the immune response in AD, HD and MS disorders [12,68].
Several cellular mechanisms such as inhibition of both glutamate release at presynaptic level and NMDA-induced intracellular Ca2+ release, antioxidant activity, protein kinase A (PKA) signaling and nitric oxide generation contribute to the neuroprotective effect of type 1 CBR agonists on in vitro excitotoxicity, which is induced by N-methyl-D-aspartic acid (NMDA). Thus, neuroprotective properties of cannabinoids suggest their potential therapeutic use in managing neurodegenerative disorders (e.g., multiple sclerosis) [53].
Nowadays, the anti-inflammatory and anti-nociceptive roles of ECBs are widely documented [70].
ECB molecules modulate nociception through central excitability and control of transmission of pain information to the CNS. The relationship between elevated concentrations of ECBs and pain is not fully established due to its involvement in both analgesic and pro-algesic phenomena. The analgesic effects of ECBs are due to CBRs activation, while the pro-algesic effects are exerted by production of arachidonic acid and of its active metabolites such as prostaglandin E2 (PGE2). Cyclooxygenase (COX)-2 pathway of ECBs metabolism could promote these antagonistic effects, depending on the metabolite produced [71].
Additionally, their role in neuroprotection has been shown by their involvement in modulation of the Blood–Brain Barrier (BBB) and further control of pathogens and toxic compounds, which circulate the blood at the brain level. In vitro/in vivo models of multiple sclerosis, chronic head injury and ischaemia/reperfusion have revealed that AEA and 2-AG have a regulating effect on its permeability [72].
This is of importance because the main pathological features of neurological diseases such as chronic inflammatory demyelinating polyneuropathy, AD, MS and neuromyelitis optica are different dysregulations of the BBB. In addition to the neuroprotective effects mediated by CB1R and CB2R (e.g., decrease of inflammatory state and reduction of cellular death as well as oxidative stress), it was also shown that they are capable of restoring and also improving the integrity of the BBB, contributing to the protective effects in psychiatric and neurological diseases [73].
Another way through which these receptors exert their protective effect on the BBB is through clearing it from Aβ depositions by contributing to their efflux. Because of a dysregulation in the BBB, Aβ cannot be transported across the barrier (Aβ-depositions are an important feature in AD). The involvement of the ECS in the transport of Aβ across the BBB was highlighted by Bachmeier et al. in a study that used animal models of AD in which depositions of Aβ were reduced in the AD brain, while cognitive behavior was improved [30].
As mentioned beforehand, the most common neurodegenerative disorders are AD and PD, age-dependent conditions characterized by memory and cognitive impairments (in AD) or affected ability to move, speak and breathe (in PD). Both diseases have common pathophysiological factors, including neuroinflammation and oxidative stress [74,75].
Due to their ability to modulate a wide range of cognitive functions (e.g. memory, learning, affective parameters, etc.) via their interactions with neurotransmitters, cannabinoids could have a positive effect on these type of diseases [52,76].

3. Cannabinoids and Alzheimer’s Disease

AD is one of the most serious CNS diseases, characterized by neuronal degeneration with irreversible evolution, and clinically it is manifested by impairment of cognitive functions often accompanied by other disorders such as behavioral, verbal, motor, mood and functional disorders, altering patients’ quality of life [75,77,78]. Major brain lesions are caused by the formation of extracellular senile plaques. Their structure includes insoluble β-amyloid (Aβ) peptides that initiate a neurodegenerative cascade resulting in neuronal cell death. Another specific feature of AD is intracellular tau proteins associated with microtubules, folded and hyperphosphorylated, which clump and form neurofibrillary tangles (NTFs), disrupting nutrient transport and causing neuronal senescence [35,75,79,80]. Based on these aspects, cannabinoids have been found to have neuroprotective effects.
Some preclinical studies suggest that CB1R activation causes hyperpolarization of the neuronal membrane, modulating neurotransmitter and cytokine release. Increased CB1R density in AD may be considered as a neuroprotective and anti-inflammatory response that compensates for excitotoxicity, with these findings suggesting that CB1R agonists could promote neuronal survival and could also be considered as therapeutic targets for treating agitation in AD [79].
Learning and memory impairment can be considered the most relevant side effects of cannabinoids, effects mainly mediated by CB1R through their activation. These issues have been observed in numerous studies conducted on rats and mice, which have evaluated the effects of agonists such as arachidonyl-cyclopropyl amide (ACPA) and antagonists (AM251) of CB1Rs, on spatial memory and learning processes, where agonists caused cognitive impairments and antagonists often enhanced memory and learning processes [52,81].
Low serotonin synthesis is involved in determining the occurrence of mood and behavioral disorders, as CBD and THC may contribute to increased synthesis of tryptophan, a precursor for serotonin synthesis. Compensating for the degradation of tryptophan could be an important mechanism by which CBD and THC can improve status in diseases associated with inflammation because it passes through one of the metabolism pathways into serotonin, a neurotransmitter involved in the pathogenesis of mood disorders [82].
Cannabinoids are involved in the reduction of oxidative stress and of the tumor necrosis factor-α (TNF-α); they may have synergistic effect on acetylcholinesterase (AchE) inhibitors, preventing Aβ formation, and they may improve circadian rhythm disturbances commonly seen in AD [79,83].
The proteolytic cleavage of APP generates Aβ1-42 monomer, which, under pathological conditions, leads to the formation of oligomers that activate astrocytes and microglia, release kinases, release nitric oxide, produce inflammatory cytokines and promote the phosphorylation of tau proteins.
Microglia and astrocytes in the presence of inflammatory processes synthesize ECBs (AEA and 2-AG) that suppress cytokine synthesis via CB1/CB2 receptors. In AD progression, nuclear factor-κB (NF-κB) activates the β-secretase (BACE1) promoter favoring increased synthesis of Aβ1-42 [84]. The neuronal protective role of ECBs against beta-amyloid-induced neurodegeneration has been demonstrated on primary hippocampal neuron cell cultures from rat embryos. After sacrifice, embryos were removed, from which hippocampi were collected and prepared for determination of TUNEL (neurodegenerative marker)-positive neurons. The number of TUNEL-positive neurons was significantly increased in the Aβ 25-35-treated culture, but Aβ 25-35-induced neurodegeneration and apoptosis were prevented by application of 2-AG as this protective effect appears to be mediated by CB1R. This ECB is metabolized in the brain by hydroxylation reaction in the presence of monoacylglycerol lipase (MAGL). It was observed that administration of MAGL inhibitors (URB 602 and JZL 184) increased endogenous levels of 2-AG with significant reduction of TUNEL-positive neurons in cultures treated with Aβ 25-35 or Aβ1-42. The results of this study suggest that 2-AG has anti-apoptotic effects that are concentration-dependent [85]. In another study, it was shown that this ECB is able to suppress the expression of COX-2, an enzyme involved in neuroinflammation, action also mediated by CB1R [86].
Another mechanism through which ECS confers neuroprotective effects is related to the direct clearance of Aβ via CB2-activated, macrophage-induced phagocytosis and through the inhibition of glycogen synthase kinase-3β (GSK-3β), which in turn has an attenuating effect on the hyperphosphorylation of tau protein [75].
CB2R is vital for memory as cognitive process, while a depletion or disruption of these receptors in rodents has been shown to induce long-lasting memory deficits. Administration of CB2R antagonists, such as AM630, in rats showed negative effects such as impaired memory. However, administration of CB2R agonists (JWH-133) in rats significantly reduced spatial memory impairment, memory impaired by administration of okadaic acid (OKA), which mimics AD due to tau pathology. The beneficial effects of this agonist were also manifested by reduced neuroinflammation and neurodegeneration [87].
The neuroprotective effects of CBD have been demonstrated in β-amyloid-induced AD, with the formation of Aβ peptide aggregates in the brain. In this condition, massive loss of hippocampal neurons occurs, leading to biochemical communication breakdown between neurons, subsequently followed by neuronal apoptosis [88]. CBD interacts with the PPAR-γ receptor, attenuating beta-amyloid (Aβ)-induced neuroinflammation and also promoting neurogenesis in the hippocampus [89].
In recent decades, research or studies have shown that CBD may be a potential drug for improvement of cognition within AD. A recent study attempted to discover the molecular mechanism by which CBD improves the pathology of AD.
Hippocampal transcriptome sequencing from 6-month-old APP/PS1 mice after CBD treatment (intraperitoneal (i.p.) administration, daily for one month) was used as research methods to investigate AD progression. Autophagosomes were analyzed by electron microscopy showing significant enhancement, and Aβ plaques were investigated by immunohistochemistry, these being significantly lower in the CBD treated group [90].
Another study showed the cerebroprotective action of some cannabinoids, which may be due to the mechanism of inhibition of myeloperoxidase independent of the cannabinoid receptor. The study was performed on a murine model (male mice) induced with focal cerebral ischaemia. In this study, Δ9-THC and CBD significantly reduced the size of cerebral infarction given by middle cerebral artery (MCA) occlusion. Δ9-THC interacts with central CB1Rs and peripheral CB2Rs, leading to affective, cognitive, sensory and somatic effects. It also has neuroprotective properties in cerebral ischaemia and hypoxia, improving ocular circulation [52,75]. Treatment with Δ9-THC before ischemia challenge reduced infarct size, while pre- and post-ischemic treatment with CBD showed strong and longer-lasting neuroprotection. CBD also inhibited myeloperoxidase activity in neutrophils, acting through an anti-inflammatory mechanism independent of cannabinoid receptors. The results of this study indicate that CBD alone has a pre- and post-ischaemic neuroprotective effect independent of cannabinoid receptors, offering the possibility of a new therapeutic perspective for neurodegenerative disorders [91].
The CBD proved its anti-inflammatory actions through its ability to decrease reactive oxygen species (ROS), thereby inhibiting inflammation not only in AD but also in murine models of type 1 diabetic cardiomyopathy, colitis, inflammatory bowel syndrome and pneumococcal meningitis [89].
The use of exogenous cannabinoids such as CBD has led to positive results in neuronal cell cultures exposed to β-amyloid. Neuroprotective properties have been demonstrated in numerous studies of mice and rats whose Aβ-induced memory deficits were prevented by exogenous cannabinoids, while CBD administration determined/promoted neurogenesis, highlighting the potential beneficial effects in AD. Other studies have shown that cannabinoids improve brain perfusion, behavioral disturbances in AD [35]. An overview of recent literature data regarding the involvement of ECS in various pathological aspects of AD is presented in Table 1.

4. Cannabinoids and Parkinson’s Disease

PD is a chronic neurodegenerative disease, its main feature being the loss of dopaminergic neurons in the substantia nigra of the basal ganglia. This loss is generated by various factors such as inflammation, oxidative stress, genetic mutations and misfolded protein aggregates.
In terms of clinical manifestations, the ones specific to this condition are muscle rigidity, bradykinesia, postural instability, tremor, sleep disturbances, fatigue, uncoordinated body movements, amnesia, voice disorders and cognitive dysfunction [75,101].
In patients with PD, an increase in CB2R was observed compared to controls with unchanged CB1R, with these differences suggesting the main role of CB2R in the pathogenesis of PD [101].
Endogenous and exogenous cannabinoids play a pivotal and complex role in basal ganglia circuitry through interactions with dopaminergic, glutamatergic and GABAergic mechanisms. Dopaminergic transmission at this level is modulated by the ECS and by activation of of presynaptic CB1R, and the release of GABA from striatal terminals that project to dopaminergic neurons is decreased. By activating CB1R and CB2R, an inhibition of the release of glutamate from the cortical axon head that projects to the striatum is observed. By influencing the release of neurotransmitter, dopamine synthesis in the dorsal striatum is indirectly affected, even though CB1R are not found on dopaminergic neurons [37]. In PD, motor activity is intensely affected. In order to better understand this disorder, we will review the role of the basal ganglia, which are primarily responsible for motor activity but also for executive behavioral functions [102].
Spinal projection neurons (MSN) receive signals from the glutamatergic cerebral cortex, an excitatory neurotransmitter; these neurons are modulated by nigrostriatal dopaminergic transmissions by stimulating D1 and D2 receptors followed by the generation of excitatory or inhibitory effects [37].
Presynaptic CB1R activation in the striatal area has an inhibitory effect on motor activity. CB1R are presynaptic regulators of synaptic transmission, which means that controlling excess presynaptic activity should be one of the main applications. It is well known that excitotoxicity causes lesions [25,37].
Thus, ECS disruption is involved in the neurodegenerative component of PD due to its’ impairment in the basal ganglia [101].
The neuroprotective effect of cannabinoids has been highlighted by numerous studies on animal models of PD. Experimental research has shown that three weeks after injecting mice with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), CB2R have been down-regulated in the hippocampus and in the substantia nigra [103]. Thus, on the MPTP-induced mouse model, administration of JW015, a CB2R agonist, protects neuronal loss as well as MPTP-induced motor deficits. Administration of AM1241, a selective CB2R agonist, regenerated dopaminergic neurons in MPTP-induced PD mice [104].
It is also well known that chronic administration of levodopa (L-DOPA) for motor symptoms can induce dyskinesia, with the ECS being a target for attenuation of dyskinesia, which has been demonstrated in animal models [37].
Studies in experimental rodent and primate models of PD have shown profound changes in CB1R signalling in basal ganglia circuits. Other studies in animal models and in PD patients have shown that the disease is linked to an overactivation of the ECB signalling system, with increased levels of AEA and 2-AG, as well as an increase in the CB1R density, an aspect which has been shown post-mortem in the striatum of PD patients.
This increase in CB1R density can be considered as the compensatory response that occurs due to dopamine depletion in PD.
CB2Rs also play a particularly important role and are involved in neuroinflammation in PD. In post-mortem studies in PD patients, it has been shown that CB receptors in microglial cells manifested increased density [25,105].
Numerous studies in animal models of PD have shown that pharmacological activation of CB2Rs suppressed the release of pro-inflammatory cytokines, selective agonists of CB2R with the ability to reduce inflammation in the brain of mice treated with MPTP, while pharmacological inactivation of these receptors resulted in reversal of this effect. Furthermore, CB2R-deficient mice showed an exacerbation of PD pathology [25].
Currently, the treatment for PD aims for the management of symptoms, with its efficacy being limited by its side effects. It is well known that L-DOPA therapy can cause dyskinesia, which could be antagonized or prevented by the administration of cannabinoids exerting their anti-dyskinetic effect. GABA-induced signal transmission was enhanced by cannabinoid agonists through inhibition of its uptake at glutamatergic synapses and suppression of excitation produced by NMDA receptors and AMPA receptors on dopaminergic neurons, contributing to the anti-dyskinetic effect [75]. Modulation of dopaminergic activity by the ECS in the basal ganglia has been supported by various behavioral, electrophysiological and neurochemical studies [101].
The neuroprotective properties of CBD have been increasingly scientifically evaluated in neurodegenerative diseases, including PD and AD [106,107]. Although studies in animal models of PD are promising, clinical studies evaluating the neuroprotective effects of CBD are few. A study from 2014 reports that 21 PD patients selected from a sample of 119 patients received a daily dose of 300 mg CBD. Patients were assessed before the start of treatment as well as in the last week of treatment, with motor score, symptoms present in PD and change in patients’ lives for possible neuroprotective effects assessed. The results of this study suggest that CBD could improve patients’ quality of life but only in those cases free of comorbidities [106]. An overview of recent literature data regarding the implications of ECS in various pathological aspects of PD is presented in Table 2.

5. Cannabinoids and Huntington’s Disease

HD, also called as Huntington’s chorea [110], is a rare neurodegenerative autosomal dominant disorder, being the most common monogenic neurological disorder in the developed world [111]. The mean age at onset is 40 years, whereas recent prevalence studies of disease in population was between 5 and 10 persons per 100,000 [112]. The survival of patients with HD is about 17 years [113].
HD has autosomal dominant inheritance [114,115], being characterized by a repeated expansion of CAG trinucleotide in the first exon of the Huntingtin (HTT) gene, located on chromosome 4/band 16.3 [116]. The gene encodes HTT affected proteins, in which the expansion of CAG repeats leads to expansion of glutamine repeats [115,117]. The hallmark of the disease consists in accumulation of mutant Huntingtin (mHTT) aggregates and inclusions throughout the brain. Despite being caused by unrelated proteins with different expression patterns, other disorders such as AD, PD and ALS have a set of common characteristics with HD [112].
The symptoms and signs of HD consists of motor disturbances, behavioral and psychiatric symptoms and cognitive disturbances [118].
Numerous agents and interventions have been assessed for their efficacy in suppressing HD-chorea. These agents include antagonists and agonists of dopamine, drugs that deplete dopamine, inhibitors of acetylcholinesterase, benzodiazepines, glutamate antagonists, antiepileptic drugs, fetal cell transplantation and deep brain stimulation [119,120,121,122]. The current available options are symptomatic, without an influence on the course of the disease. Pharmacological interventions usually address the movement disorders associated with the disease [123]. The chorea-component is treated using tetrabenazine, which was approved by the FDA for this use in 2008. In 2017, FDA approved deutetrabenazine, a second therapeutic option for the treatment of chorea and for tardive dyskinesia, with improved pharmacokinetic profile. Excitotoxicity and neuronal death can be caused by an excessive increase in glutamate release; thus, treatments may involve the blockage of glutamatergic receptors or the inhibition of glutamate release. Drugs that demonstrated efficacy in clinical trials and act upon glutamatergic transmission are memantine, amantadine, lamotrigine and remacemide [124]. In the case of patients with Westphal variant, an akinetic form of the disease, drugs such as dopamine agonists, amantadine, levodopa and antiparkinsonians could be of use [123,125,126,127].
Studies using mouse-models of HD demonstrated an impairment in the function of ECS. In patients with HD, studies that have been realized on postmortem brains have shown a decrease of CB1-immunoreactivity throughout the course of the disease in the globus pallidus as well as in the putamen.
In vivo PET imaging has shown that even in early manifest phases in patients with the manifest form of the disease, a loss of CB1R binding is present [128]. This suggests that a reduction of type-1 receptors could be seen even in the pre-HD phase, hypothesis also supported by observing transgenic mouse models of the disease in the pre-HD phase, in which a reduction of levels of CB1R protein and messenger RNA was present. During the manifest course of the disease, a further decline is noticed [129]. Ceccarini et al. has also managed to demonstrate that a reduction of CB1R is present in pre-HD. This could be caused by the mHTT and its repressive effect on the transcription of the CB1R gene [130]. The reduction was considered a potential compensatory response because it could reduce GABA release in the striatum [131,132].
An early neurochemical alteration, which can be seen in HD patients, refers to a reduction of striatum-CB1R binding. It seems that CB1R mRNA decrease appears before the appearance of motor symptoms in transgenic rodent models of HD [133]. Glass et al. has shown that environmental enrichment determines an upregulation of CB1R binding in the R6/1 transgenic model of HD, thus providing behavioral improvement [134]. It is thus suggested that a protective effect could be obtained by activating CB1R, while a reduction of these receptors can negatively influence the disease. By analyzing post-mortem brains of patients suffering from HD, a decreased of CB1R expression was observed. This reduction was also seen in genetic and phenotypic models of the disease. It was shown that an up-regulation of CB1R led to an improvement of motor dysfunction, leading to the idea that an early reduction of these receptors is of high importance in HD development [135].
Work conducted by Stephane et al. on phenotypic 3NP and N171-82Q transgenic model of the disease has highlighted the involvement of CB1R through the inactivation of its specific gene: in the N171-82Q model, loss of CB1R is associated with an early appearance of motor symptoms and an exacerbation of these alterations, in a similar way to that seen in R6/2 model [133]. It also determines an increase in striatal aggregation frequency in N171-82Q model similar to the R6/2 model [136]. Regarding the excitotoxic-dependent 3NP model, it has been shown that CB1R is necessary to counteract neuronal degeneration [133].
In a study realized by Laprairie et al., it has been shown that CB1-positive allosteric modulators (PAM) (GAT211, GAT228 and GAT229) determined improvements in R6/2 mouse model of the disease through their beneficial influence on the viability of mHTT-expressing cultured medium spiny projection neurons. The CB1-selective aimed to evaluate the positive outcomes of allosteric modulation of type 1 receptor in HD. GAT211 and its enantiomer GAT229 determine CB1-dependent effects in vivo, leading to a reduction in psychoactivity without tolerance or dependence [137,138,139]. Thus, positive allosteric modulation of CB1 could reduce the progression of the disease while also delaying it [138].
Unlike neuronal CB1R, microglial CB2R are induced in HD animal models and patients. Palazuelos et al. have shown that in a transgenic model of a neurodegenerative disease, CB2R ablation determined an exacerbated activation of the microglia, contributing to the onset of symptoms [140]. In an experimental malonate-rat model of HD, an increase in CB2R could be observed in reactive microglial cells and in activated astrocytes within the striatum [141]. Unlike CB1R orthosteric agonists, which induce psychotropic effects, mood alterations, cognitive and motor impairments and acute psychosis, CB2R ligands appear to be promising drugs for treating neuro-inflammatory diseases [142] because they do not induce the undesirable psychotropic effects. Thus, CB2R activation is associated with a neuroprotective effect in HD models through controlling the damaging activity of the microglia [140].
CB2R selective agonists may represent an important therapeutic alternative to CB1R due to the early reduction of CB1R during HD and to the lack of psychotropic effects of CB2R agonists. They are viable options for the potential treatment of the disease because of their ability to reduce non-cell autonomous degeneration of neurons by active microglia [140]. The major challenge regarding CB2R is related to the selective targeting of the brain CB2Rs without influencing peripheral CB2R. It is known that CB2R levels are much higher in peripheral tissues, with this type of receptor being considered a “peripheral cannabinoid receptor”.
Drugs that are currently used in clinics and activate CB1R and CB2R are nabilone, dronabinol, Δ9-THC and combination of Δ9-THC with CBD. Considering that the side effects of these agonists are due to the activation of CB1R, it is anticipated that new drugs that selectively modulate CB2Rs will emerge [143].
A 44-patient trial done by Curtis et al. has demonstrated beneficial effects of the cannabinoid-treatment on aspects such as chorea, behavior, cognition, and motor and neuropsychiatric symptoms [144].
By administering AM404 and UCM707 (ECB re-uptake inhibitors), a reduction of hyperkinetic activity and restoration of neurochemical alterations in an HD rat model with bilateral striatal injection of 3-nitropropionic acid (3-NP) was observed. In the case of AM404, the activation of TRPV1 receptor was responsible for the effect, without the involvement of CB1R [1,145]. 3-NP has a damaging effect on the striatum due to its inhibition of a component of the respiratory chain [146].
CP55,940, which is a CB1R agonist, has demonstrated beneficial anti-hyperkinetic effects with no influence on GABA and dopamine levels [1]. Chronic administration of WIN 55,212-2 (another CB1R agonist) in a R6/1 transgenic mouse model of the disease demonstrated a protective effect against motor impairment. Concerning nabilone, studies have shown that its administration did not have a beneficial effect on the hyperkinetic component associated with the disease. This could be due to the fact that it does not determine the activation of TRPV1 receptors, which are responsible for the anti-hyperkinetic activity in the rat model of the disease. It is important to note the involvement and beneficial aspects of using compounds that can activate both CB1 and TRPV1 receptors [147].
In regard to the neurodegenerative aspects of HD, studies realized on rodent models of the disease have demonstrated the beneficial effect of compounds that act as CB1R agonists [54]. In a rat model of HD in which excitotoxicity was increased through striatal injection of quinolinic acid, by administering WIN 55,212-2, a reduction of both glutamate levels, as well as quinolinic acid effect on corticostriatal local field potential recordings, was observed in vitro. The observed-effects were CB1-dependent, which was demonstrated by the administration of AM 251 (CB1R antagonist), with this compound blocking the effects of WIN 55,212-2 [148].
Δ9-THC determined an improvement of symptoms, neuropathology and molecular pathology related to R6/2 model of HD [136]. These findings support the role of CB1R and CB1 agonists as useful options for reducing or delaying the progression of HD.
The administration of HU-308 (CB2R agonist) has shown promising neuroprotective effects in quinolinic-acid lesioned mice as well as in the malonate-lesion rat model of HD through the reduction of neuronal damage in the striatum, by attenuating glial activation [140]. R6/2 mice, which are deficient in CB2R, were associated with a faster progression of the disease phenotype, with the ablation leading to an increase in the activation of the glia as well as to a higher sensitivity to excitotoxic-induced neurodegeneration of the striatum [140].
Another way through which some cannabinoids can protect against neurodegeneration is independent of CB1R and CB2R and involves ROS blockage (e.g. Δ9-THC) [149].
In a study realized by Heim et al., the use of nabilone was assessed in five patients with HD for alleviating therapy-resistant symptoms. All patients reported improvement of symptoms, suggesting that nabilone is well tolerated and could be an efficient adjuvant in the treatment of HD [150].
In a literature review done by Akinyemi et al., it has been found that for the selected human studies, medical marijuana demonstrated therapeutic effects for movement disorder symptomatology. These studies used nabilone (synthetic cannabinoid capsule that is a selective agonist for CB1R and CB2R), and Sativex®, which is an oral spray that contains a combination of Δ9-THC and CBD. The majority of studies demonstrated statistically significant results favoring the use of medical marijuana for improving quality of sleep and motor symptoms. Positive effects have also been shown for symptoms such as spasticity and tremor, although patients that presented these symptoms had other neurodegenerative diseases and not HD (e.g., spasticity improvement was demonstrated in patients with autoimmune demyelinative disease and MS, while tremor was improved in patients with PD following administration of medical marijuana). The positive effects shown in patients with HD were improvements in chorea; improvements in the neuropsychiatric index; and trend for improvements in the Unified Huntington’s Disease Rating Scale motor score, the dystonia subscore and behavior score [151].
In another report done by Saft et al., 7 patients with early-onset HD, in which dystonia was the main manifestation, were treated with cannabinoids. Five of these patients were in advanced disease stages. It was shown that the treatment demonstrated beneficial effects on dystonia, thus improving motor symptoms. Improvements that were reported by participants include nail and dental care, social-life activities, fine motor skills and movements such as head-lifting, thus leading to better life-quality. Three of these patients have also reported behavioral changes, with less irritability and apathy [152].
Another compound of interest is cannabigerol (CBG), which shares characteristics with CBD and Δ9-THC regarding affinity and activity but also presents uniqueness in that it interacts with α-2 adrenoreceptors and with 5-hydroxytryptamine (5-HT1A). Research, which used this compound as well as VCE-003.2 (a synthetic quinone derivative), demonstrated in vitro neuroprotective potential in disease-models such as HD, ALS, PD and MS, reducing the severity of neurologic illnesses. In a 3-NP animal model of HD, it has shown that the use of CBG prevented neuronal death within the striatum, also improving motor deficits and reducing inflammatory markers [153].
Another study that used VCE-003.2 discovered the beneficial effects of the compound on the symptoms of the disease in the quinolinic acid murine model. The animals showed better Rotarod performance. In the 3-NP model of HD, VCE-003.2 demonstrated good results for the improvement of motor deficits [154].
The specific cannabinoid tetrad (hypothermia, hypomotility, antinociception and catalepsy) is not induced by CBD; therefore, the compound could be of use for the management of movement disorders, aspect that is also highlighted by a series of pre-clinical and clinical studies. In a cell culture model of HD in which cells expressed mHTT, CBD along with other tested compounds (Δ8-THC, Δ9-THC and CBD) have shown a protective effect against cell-death induced by HTT [155], although these effects might be independent of CB1 and due to antioxidant mechanisms. CBD has also exerted protective effects in 3-NP animal models of the disease by reversing or attenuating the alterations that were induced by 3-nitropropionic acid [146]. Recent data from literature regarding the involvement of ECS in various pathological aspects of HD is presented in Table 3.

6. Cannabinoids and Multiple Sclerosis

MS is a neuroinflammatory condition, characterized by chronic disability of the CNS, being spread in young adults aged between 20 and 40 years [156]. Recent prevalence studies reveal that an estimated population of 2.8 million people worldwide is suffering from MS [157,158,159]. The prevalence of the disease is reported to be of 35.9 cases per 100,000 people and has been constantly growing in recent decades; it was observed that women are more affected than men by the disease, by a ratio of up to 3:1 [160].
It causes significant multifocal demyelination in both gray and white matter, and axonal damage via an autoimmune mechanism directly linked to neuroinflammation [161]. The stimulation of the immune system is not completely understood [162], but it is believed that both genetic and environmental factors are the main triggers [12].
Both adaptive and innate immunity are involved in the inflammatory process [163,164]. Myelin-specific CD4 and CD8 positive T-lymphocytes of adaptive immunity have been identified within active MS lesions. B-cells play a key role in sustained inflammation of progressive MS by stimulation of T cells response [165]. The negative influence on myelin sheaths through their infiltration into CNS by crossing BBB is due to release of pro-inflammatory cytokines during macrophage activation, including TNF-α, lymphotoxin, interleukin-6 and metalloproteinases [49,166]. Recent studies indicate a risen presence of ectopic B-cell follicles in the cerebral meninges that is associated with irreversible disability and cortical demyelination [165,167].
Oligodendrocyte degeneration in the CNS is another important mechanism involved in disease pathology [168]. Recent research shows that oligodendrocytes, also known as myelin-forming cells located in CNS, play a key role in remyelination process that occurs in early phases of the disease that temporarily decreases the symptoms in relapsing-remitting MS [169,170]. Microglia, the innate immune cells in the CNS and activation of astrocytes, a sub-type of glial cells in CNS, represent early events in lesion development, loss of blood–brain barrier function and CNS inflammation [171]. A release of proteolytic enzymes, cytokines, reactive oxygen and nitrogen species is observed with negative influence on oligodendrocytes and myelin [172].
The condition is characterized by episodic relapses of different intensity and remissions and leads to progressive reduction of neurological function due to disruption of neural transmission [12]. The consequences are represented by prolonged and progressive physical, psychological and cognitive impairments [173,174]. The signs and clinical symptoms of MS vary depending on the CNS damage and include paresthesia, dysesthesia, weakness and visual disturbances [175]. Spasticity is the mainly observed symptom in MS. It is associated with spasms, pain and sleep disturbance [176,177]. Another symptom frequently reported by up to 80% of patients suffering from MS is urinary incontinence due to bladder dysfunction [178].
According to International Advisory Committee on Clinical Trials of MS in 2013, based on the classification made by Lublin et al., four main clinical courses of the disease have been identified, depending on its evolution: relapsing-remitting type (RRMS), progressive relapsing type (PRMS), primary-progressive type (PPMS) and secondary-progressive type (SPMS) [179,180]. RRMS is reported to have the highest prevalence, being encountered in approximately 85% of cases [179]. The PPMS and SPMS pathophysiology is different from that of RRMS [18].
Even though is known that MS has no cure and drugs used for progressive forms are currently limited, the main therapeutic goal remains the reduction of the exacerbations of inflammatory phenomena and neurological disability of the patients [181]. This is achieved by modulation of innate and adaptive immune responses with disease modifying therapies (DMTs) in long-term treatment [182]. Being a chronic disease, it is very important to optimize MS pharmacotherapy in order to obtain the best clinical response from the patient [183]. Using immunomodulating or immunosuppressive drugs is unfortunately accompanied by high risk of cancer development due to involvement of immune system in recognizing and eliminating cancer cells [182]. On the other hand, some available MS DMTs have been used for years in cancer therapy (e.g., rituximab [184], cladribine [185] and methotrexate [186]), while other current DMTs are being evaluated in the present for their anti-tumor potential (e.g., dimethylfumarate [187], fingolimod [188] and teriflunomide [189]).
Cannabinoids have positive effects in reducing MS symptoms and can limit the inflammatory processes ongoing in the CNS and progressive neurodegeneration in MS, due to anti-inflammatory properties and neuroprotective effects [12,190]. While activation of CB2R demonstrates the modulation of inflammatory process; through inhibition of neuroinflammatory signaling pathways, CB1 agonists are involved in neuroprotection [191]. Both exogenous or endogenous cannabinoids also influence humoral and cellular immune response [50].
A combination of equal amounts of Δ9-THC and CBD was recently approved in some countries under brand name Nabiximol (Sativex®) [192], which is used for the management of bladder dysfunction, neuropathic pain and spasticity associated with MS [193]. Wade et al. reported significant improvement in spasticity after administration of a daily dose below 120 mg of Sativex [194]. Surprisingly, some researchers reported that MS-induced spasticity was controlled only by Δ9-THC action on CB1Rs and not by CBD [195]. Despite the fact that CBD has low affinity to both CB1R and CB2R, recent studies showed the ability to potentiate the effect of Δ9-THC due to its allosteric modulation property and CBR indirect antagonism [196]. Another randomized double-blind placebo-controlled study reported that the treatment with a daily dose consisting of 48 sprays of Sativex® during 5 weeks reduced MS-related pain [197]. The patients enrolled in this study were tracked for the next 2 years, and it was observed that the effect was maintained, with no signs of tolerance [198].
Preclinical studies on different experimental animal models of MS show a wide alteration of CB1R and CB2R expression [199]. Phytocannabinoids promote reduction of inflammation and neuroprotection [19]. Their anti-inflammatory and antioxidant properties mediate the neuroprotection towards modulation of signaling systems, receptors or channels involved in neurodegenerative diseases [200]. It was also reported that agonists of both ECB receptors can limit relapses, axonal degeneration and neuroinflammation in rodent models of MS [192].
It is of interest to elucidate the mechanism through which they influence neuroinflammation and neurodegeneration since this could contribute to the development of new therapeutic approaches.
Experimental autoimmune encephalomyelitis (EAE) is frequently used as rodent model of autoimmune disease of the CNS that is characterized by brain inflammation associated with neurodegenerative pathology, including the development of cutaneous mechanical and cold hypersensitivity. It initiates demyelination and neurological dysfunction within acute monophasic RRMS and SPMS [201]. Within this model, animals are immunized with myelin antigens such as proteolipid and myelin basic proteins; encephalitogenic peptides; or myelin oligodendrocyte glycoprotein (MOG). Another possibility of inducing EAE is by T lymphocyte sensitization to myelin proteins due to adoptive transfer of encephalitogenic T cells [202]. The immune response consists of demyelination process and progression of CNS inflammation, involving paralysis of animal tail, limb and forelimb.
The influence of cannabinoids on EAE-induced neurodegeneration phenomenon is due to their affinity on both CB1R and CB2R.
In a report done by Baker et al., mice suffering from chronic relapsing EAE were treated with Δ9-THC, methanandamide (analogue of AEA), WIN 55,212-2 (agonist of both CB1R and CB2R) and JWH-133 (agonist of CBR type 2). The treatment reduced motor symptoms, including limb spasticity, tremor and paralysis [203]. It was hypothesized that it could control clinical EAE symptoms through inhibition of synaptic transmission [204].
Similar studies were conducted by different research teams identifying the importance of these receptors in controlling neuroinflammation [205,206,207,208,209,210,211]. Ni et al. showed that administration of WIN 55,212-2 in C57BL/6 mice immunized with MOG35–55 attenuated EAE progression [209]. In addition, Maresz et al. revealed the effects of co-administration of CB1R/CB2R antagonists and WIN 55,212-2 on EAE progression. Thus, administration of CB1R antagonist SR141716A had no influence on the protective effect, whereas the CB2R antagonist SR144528 reduced it, suggesting that CB2Rs play a key role in the protective effect of WIN 55,212-2 [212]. Their activation produces inhibitory effects on leukocyte/endothelial interactions that have beneficial effects in disease management [209]. On the other hand, CB1R activation may provide neuroprotective effects in later stages of the disease, although CB1R blockers did not reduce the beneficial effects of WIN 55,212-2 [212]. These aspects suggest that only neuronal but not T lymphocyte CB1 expression plays an important role in mediating EAE suppression [213].
It was also observed that during chronic relapsing, EAE mice deficient in CB1Rsare are more susceptible to neurodegeneration following immune attack. In this process, both markers of axonal damage such as caspase 3 activation and presence of nonphosphorylated form of neurofilament H epitope are involved [49]. In another experimental model using Lewis rats, it was observed that EAE induction reduced CB1R activation in striatal and cortical neurons [214,215].
Neuroprotective CB1R-mediated effects are explained due to modulation of glutamate activity, control of NMDA receptor-induced calcium influx and decrease in oxidative injury by antioxidant properties [67]. Increased glutamate level is associated with EAE neurodegeneration due to its excitotoxicity [210,216]. Several studies reported that Δ9-THC and its synthetic analogues demonstrated antiglutamatergic effects by inhibiting the release of glutamate in rat hippocampus cultures [217].
The inflammatory response in EAE model is mediated by CB2Rs expressed by infiltrating T-cells and monocytes [218]. CB2Rs are also found in microglial cells; thus, their activation inhibits the production of proinflammatory cytokines and oxygen and nitrogen reactive species. It was observed elevated local levels of interleukin-6 (IL-6) in the CNS of animals with induced EAE, although it was not present in their spleens [219].
Kozela et al. administrated 5 mg/kg b.w. CBD (i.p.) in C57BL/6 mice for 3 consecutive days. The dose that was used in the experiment is similar to the one used in studies on rheumatic arthritis. The histological analysis showed a reduction of the severity of EAE through different mechanisms: diminishment of axonal damage and inflammation, decrease of spinal glial activation and inhibition of T-cell recruitment in the spinal cord [220]. Gonzáles-García et al. performed a similar study and observed that different negative aspects of the disease such as microglial activity, cell infiltration and demyelination, axonal damage and levels of pro-inflammatory cytokine IL-6 were significantly reduced after administration of a dose of 50 mg/kg b.w. CBD [221]. Similarly, Giacoppo et al. revealed the neuroprotective effect of CBD on apoptotic markers associated with neurodegeneration in MS [222].
Theiler’s murine encephalomyelitis virus (TMEV) belongs to Picornaviridae family, being a nonenveloped RNA virus. This model has been used as viral model for MS [223]. TMEV infection induces inflammatory demyelination process attributed to infiltration of both CD4 and CD8 positive T cells into CNS. Whilst CD4 positive T-cells are important in development of demyelinating disease at later stages, CD8 positive T-cells make a significant contribution in early phase of immune response against the virus [224]. It is supposed that autoimmune response in propagating the disease is due to myelin antigens/epitope spreading [225]. On the other hand, it was also observed that depletion of CD8+ T-cells reduced chronic disease at the same time, with chronic axonal injury or neurodegeneration [223].
Different data indicate that cannabinoids can reduce MS progression induced by this murine model. In this respect, Molina-Holgado et al. showed that ECB AEA was found to potentiate IL-6 (immunosuppressive and anti-inflammatory cytokine) production by TMEV-infected astrocytes in a dose-dependent manner [219].
Other researchers demonstrated the anti-inflammatory role of AEA in MS due to inhibition of IL-1β, IL-6, IL-12 and IL-23 release in myeloid dendritic cells and inhibition of microglial activation [161]. IL-12 leads to induction of Th1 regulatory cells, while IL-23 is a maintenance factor of Th17 phenotype [226].
The neuroprotective effect of AEA against microglial neurotoxicity was mediated by a mechanism between neurons (with reference to membrane glycoproteins) and microglia/macrophages that involves CD200 ligand-receptor interaction [227].
Another study done by Loria et al. described that treatment with palmithoylethanolamide (PEA), an analogue of AEA with anti-inflammatory activity, reduced the expression of IL-1, TNF-α and microglial activation in the spinal cord of mice with TMEV-induced demyelinating disease [228].
Significant improvement in motor function and neurological deficits was obtained following the administration of UCM707, a selective AEA uptake inhibitor in the same animal model or different synthetic cannabinoids such as JWH-015, WIN 55,212-2 and ACEA. The proposed common mechanisms through which the compounds influence disease progression were the reduction of microglial activation, inhibition of major histocompatibility complex (MHC) class II antigen expression, depletion of spinal-cord-infiltrating CD4 positive T cells or decrease of the production of pro-inflammatory (TNF-α, IL-1β and IL-6) cytokines [229,230].
Cannabinoids produce immunosuppresion in astrocyte reactivity. During MS progression, astrocytes are activated and contribute to inflammatory response by cytokines, chemokines and nitric oxide release.
It was observed that AEA inhibited production of IL-6 and astrocytes activation through CB1-mediated pathway, in a study on primary astrocytes infected with TMEV [231].
By using different mouse models of MS, the above-mentioned findings suggest that both endogenous and synthetic CBR ligands can influence the development of progressive forms of MS.
In the following, a summary of recent literature data regarding the implications of ECS in various pathological aspects of MS is presented in Table 4.

7. Conclusions

Neurodegenerative diseases affect millions of people worldwide. Whilst there are currently no available cures, the management of symptoms and improvements of physiological and cognitive deficits remain a clinical challenge. The ECS is an attractive research field involved in different types of interactions with neurotransmitters involved in the pathological aspects of neurological diseases. The complexity of ECB functions offers new insights into the mechanisms that underlie the neuroinflammatory and neurodegenerative components of diseases such as AD, PD, HD and MS. The current review provides information regarding the CBR and their endogenous or exogenous ligands, contributing to a better understanding of pathophysiological features as well as allowing the identification of promising and effective therapeutic candidates for the treatment of these disorders. Despite the advantages of the reviewed compounds, further preclinical and clinical studies that evaluate safety, efficacy and pharmacokinetics are necessary.

Author Contributions

Conceptualization, A.V., V.B., D.-C.A. and G.D.S.; methodology, A.V. and V.B.; validation, A.V. and R.-N.R.; formal analysis, A.V. and R.-N.R.; investigation, A.V., V.B., D.-C.A., R.-N.R. and M.L.; resources, A.V., V.B., W.B., D.-C.A., R.-N.R., M.L., G.D.S. and C.S.; writing—original draft preparation, A.V., R.-N.R., D.-C.A. and M.L.; writing—review and editing, A.V., R.-N.R., W.B. and V.B.; visualization, V.B., W.B., G.D.S. and C.S.; supervision, V.B. and C.S.; project administration, A.V, R.-N.R., D.-C.A. and V.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

2-AG2-Arachidonoylglycerol
3-NP3-Nitropropionic acid
5-HT1A5-Hydroxytryptamine
AchEAcetylcholinesterase
ACPAArachidonyl-cyclopropyl amide
ADAlzheimer’s Disease
AEAAnandamide
ALSAmyotrophic lateral sclerosis
APPAmyloid precursor protein
BBBBlood–Brain Barrier
BLABasolateral amygdala
CA1Dorsal hippocampus
CB1RCannabinoid-receptors type 1
CB2RCannabinoid-receptors type 2
CBDCannabidiol
CBD-DMHCannabidiol dimethylheptyl
CBGCannabigerol
CBRCannabinoid receptor
CIDPChronic inflammatory demyelinating polyneuropathy
CNSCentral nervous system
COXCyclooxygenase
DMTDisease modifying therapy
EAEExperimental autoimmune encephalomyelitis
ECBEndocannabinoid
ECSEndocannabinoid system
FAAHFatty acid amide hydrolase
GABAGamma-aminobutyric acid
GBSGuillain–Barré syndrome
GPRCG-Coupled protein receptor
GSK-3βGlycogen synthase kinase-3β
HDHuntington’s Disease
HTTHuntingtin
ILInterleukin
i.p.Intraperitoneal
L-DOPALevodopa
MAGLMonoacylglycerol lipase
MAPKMitogen-activated protein kinases
MCAMiddle cerebral artery
MHCMajor histocompatibility complex
mHTTMutant huntingtin
MOGMyelin oligodendrocyte glycoprotein
MPTP1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine
MSMultiple Sclerosis
MSNMedium spiny neurons
NAcNucleus accumbens
NAPE-PLDN-arachidonoyl phosphatidylethanolamine-specific phospholipase D
NF-kBNuclear factor-kB
NMDAN-Methyl-D-aspartic acid
NTFNeurofibrillary tangles
OKAOkadaic acid
PDParkinson’s Disease
PEAPalmithoylethanolamide
PFCPrefrontal cortex
PGProstaglandin
PKAProtein kinase A
PPMSPrimary-progressive MS
PRMSProgressive-relapsing MS
ROSReactive oxygen species
RRMSRelapsing-remitting MS
SPMSSecondary-progressive MS
TMEVTheiler’s murine encephalomyelitis virus
TNF-αTumor necrosis factor-α
TRPVTransient receptor potential vanilloid
VTAVentral tegmental area
Δ9-THCΔ9-Tetrahydrocannabinol

References

  1. Morera-Herreras, T.; Miguelez, C.; Aristieta, A.; Torrecilla, M.; Ruiz-Ortega, J.Á.; Ugedo, L. Cannabinoids and Motor Control of the Basal Ganglia: Therapeutic Potential in Movement Disorders. In Cannabinoids in Health and Disease; Meccariello, R., Ed.; InTech Open: London, UK, 2016; pp. 60–92. [Google Scholar]
  2. Erkkinen, M.G.; Kim, M.O.; Geschwind, M.D. Clinical Neurology and Epidemiology of the Major Neurodegenerative Diseases. Cold Spring Harb. Perspect. Biol. 2018, 10, a033118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nichols, E.; Szoeke, C.E.I.; Vollset, S.E.; Abbasi, N.; Abd-Allah, F.; Abdela, J.; Aichour, M.T.E.; Akinyemi, R.O.; Alahdab, F.; Asgedom, S.W.; et al. Global, regional, and national burden of Alzheimer’s disease and other dementias, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2019, 18, 88–106. [Google Scholar] [CrossRef] [Green Version]
  4. Yang, W.; Hamilton, J.L.; Kopil, C.; Beck, J.C.; Tanner, C.M.; Albin, R.L.; Ray Dorsey, E.; Dahodwala, N.; Cintina, I.; Hogan, P.; et al. Current and projected future economic burden of Parkinson’s disease in the U.S. NPJ Parkinsons Dis. 2020, 6, 15. [Google Scholar] [CrossRef] [PubMed]
  5. Alzheimer’s Association. 2021 Alzheimer’s disease facts and figures. Alzheimer’s Dement. 2021, 17, 327–406. [Google Scholar] [CrossRef] [PubMed]
  6. Matthews, K.A.; Xu, W.; Gaglioti, A.H.; Holt, J.B.; Croft, J.B.; Mack, D.; McGuire, L.C. Racial and ethnic estimates of Alzheimer’s disease and related dementias in the United States (2015–2060) in adults aged >/=65 years. Alzheimer’s Dement. 2019, 15, 17–24. [Google Scholar] [CrossRef] [PubMed]
  7. Niu, H.; Alvarez-Alvarez, I.; Guillen-Grima, F.; Al-Rahamneh, M.J.; Aguinaga-Ontoso, I. Trends of mortality from Alzheimer’s disease in the European Union, 1994–2013. Eur. J. Neurol. 2017, 24, 858–866. [Google Scholar] [CrossRef] [PubMed]
  8. Kempuraj, D.; Thangavel, R.; Natteru, P.; Selvakumar, G.; Saeed, D.; Zahoor, H.; Zaheer, S.; Iyer, S.; Zaheer, A. Neuroinflammation Induces Neurodegeneration. J. Neurol. Neurosurg. Spine 2016, 1, 1003. [Google Scholar] [PubMed]
  9. Ransohoff, R.M. How neuroinflammation contributes to neurodegeneration. Science 2016, 353, 777–783. [Google Scholar] [CrossRef] [PubMed]
  10. Ward, R.J.; Dexter, D.T.; Crichton, R.R. Ageing, neuroinflammation and neurodegeneration. Front. Biosci. 2015, 7, 189–204. [Google Scholar] [CrossRef] [PubMed]
  11. Chevalier-Larsen, E.; Holzbaur, E. Axonal transport and neurodegenerative disease. Biochim. Biophys. Acta 2006, 1762, 1094–1108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chiurchiù, V.; van der Stelt, M.; Centonze, D.; Maccarrone, M. The endocannabinoid system and its therapeutic exploitation in multiple sclerosis: Clues for other neuroinflammatory diseases. Prog. Neurobiol. 2018, 160, 82–100. [Google Scholar] [CrossRef] [PubMed]
  13. Lanuti, P.; Ciccocioppo, F.; Bologna, G.; Ercolino, E.; Pierdomenico, L.; Simeone, P.; Pieragostino, D.; Del Boccio, P.; Marchisio, M.; Miscia, S. Neurodegenerative diseases as proteinopathies-driven immune disorders. Neural Regen. Res. 2020, 15, 850–856. [Google Scholar] [CrossRef] [PubMed]
  14. Hussain, R.; Zubair, H.; Pursell, S.; Shahab, M. Neurodegenerative Diseases: Regenerative Mechanisms and Novel Therapeutic Approaches. Brain Sci. 2018, 8, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Rossi, S.; Bernardi, G.; Centonze, D. The endocannabinoid system in the inflammatory and neurodegenerative processes of multiple sclerosis and of amyotrophic lateral sclerosis. Exp. Neurol. 2010, 224, 92–102. [Google Scholar] [CrossRef] [PubMed]
  16. Lull, M.E.; Block, M.L. Microglial activation and chronic neurodegeneration. Neurotherapeutics 2010, 7, 354–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Bridgeman, M.B.; Abazia, D.T. Medicinal Cannabis: History, Pharmacology, and Implications for the Acute Care Setting. Pharm. Ther. 2017, 42, 180–188. [Google Scholar]
  18. Maurya, N.; Velmurugan, B.K. Therapeutic applications of cannabinoids. Chem. Biol. Interact. 2018, 293, 77–88. [Google Scholar] [CrossRef] [PubMed]
  19. 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] [Green Version]
  20. Jones, É.; Vlachou, S. A Critical Review of the Role of the Cannabinoid Compounds Δ9-Tetrahydrocannabinol (Δ9-THC) and Cannabidiol (CBD) and their Combination in Multiple Sclerosis Treatment. Molecules 2020, 25, 4930. [Google Scholar] [CrossRef]
  21. Mechoulam, R.; Gaoni, Y. A Total Synthesis of dl-Δ1-Tetrahydrocannabinol, the Active Constituent of Hashish. J. Am. Chem. Soc. 1965, 87, 3273–3275. [Google Scholar] [CrossRef]
  22. Crocq, M.-A. History of cannabis and the endocannabinoid system. Dialogues Clin. Neurosci. 2020, 22, 223–228. [Google Scholar] [CrossRef] [PubMed]
  23. Silvestro, S.; Schepici, G.; Bramanti, P.; Mazzon, E. Molecular Targets of Cannabidiol in Experimental Models of Neurological Disease. Molecules 2020, 25, 5186. [Google Scholar] [CrossRef] [PubMed]
  24. Koppel, B.S.; Brust, J.C.; Fife, T.; Bronstein, J.; Youssof, S.; Gronseth, G.; Gloss, D. Systematic review: Efficacy and safety of medical marijuana in selected neurologic disorders: Report of the Guideline Development Subcommittee of the American Academy of Neurology. Neurology 2014, 82, 1556–1563. [Google Scholar] [CrossRef] [PubMed]
  25. Cilia, R. Molecular Imaging of the Cannabinoid System in Idiopathic Parkinson’s Disease. In International Review of Neurobiology; Elsevier: Amsterdam, The Netherlands, 2018; Volume 141, pp. 305–345. [Google Scholar] [CrossRef]
  26. Melis, M.; Pillolla, G.; Bisogno, T.; Minassi, A.; Petrosino, S.; Perra, S.; Muntoni, A.L.; Lutz, B.; Gessa, G.L.; Marsicano, G. Protective activation of the endocannabinoid system during ischemia in dopamine neurons. Neurobiol. Dis. 2006, 24, 15–27. [Google Scholar] [CrossRef] [PubMed]
  27. Maccarrone, M. Missing Pieces to the Endocannabinoid Puzzle. Trends Mol. Med. 2020, 26, 263–272. [Google Scholar] [CrossRef] [PubMed]
  28. Daris, B.; Tancer Verboten, M.; Knez, Z.; Ferk, P. Cannabinoids in cancer treatment: Therapeutic potential and legislation. Bosn. J. Basic Med. Sci. 2019, 19, 14–23. [Google Scholar] [CrossRef] [PubMed]
  29. Di Marzo, V.; Bifulco, M.; De Petrocellis, L. The endocannabinoid system and its therapeutic exploitation. Nat. Rev. Drug Discov. 2004, 3, 771–784. [Google Scholar] [CrossRef]
  30. Lowe, H.; Toyang, N.; Steele, B.; Bryant, J.; Ngwa, W. The Endocannabinoid System: A Potential Target for the Treatment of Various Diseases. Int. J. Mol. Sci. 2021, 22, 9472. [Google Scholar] [CrossRef]
  31. Munro, S.; Thomas, K.L.; Abu-Shaar, M. Molecular characterization of a peripheral receptor for cannabinoids. Nature 1993, 365, 61–65. [Google Scholar] [CrossRef]
  32. Trautmann, S.M.; Sharkey, K.A. The Endocannabinoid System and Its Role in Regulating the Intrinsic Neural Circuitry of the Gastrointestinal Tract. In International Review of Neurobiology; Elsevier: Amsterdam, The Netherlands, 2015; Volume 125, pp. 85–126. [Google Scholar] [CrossRef]
  33. Wilhelmsen, K.; Khakpour, S.; Tran, A.; Sheehan, K.; Schumacher, M.; Xu, F.; Hellman, J. The Endocannabinoid/Endovanilloid N-Arachidonoyl Dopamine (NADA) and Synthetic Cannabinoid WIN55,212-2 Abate the Inflammatory Activation of Human Endothelial Cells. J. Biol. Chem. 2014, 289, 13079–13100. [Google Scholar] [CrossRef] [Green Version]
  34. Sugiura, T.; Kondo, S.; Sukagawa, A.; Nakane, S.; Shinoda, A.; Itoh, K.; Yamashita, A.; Waku, K. 2-Arachidonoylglycerol: A possible endogenous cannabinoid receptor ligand in brain. Biochem. Biophys. Res. Commun. 1995, 215, 89–97. [Google Scholar] [CrossRef] [PubMed]
  35. Aso, E.; Ferrer, I. Cannabinoids for treatment of Alzheimer’s disease: Moving toward the clinic. Front. Pharmacol. 2014, 5, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Everett, T.J.; Gomez, D.M.; Hamilton, L.R.; Oleson, E.B. Endocannabinoid modulation of dopamine release during reward seeking, interval timing, and avoidance. Prog. Neuropsychopharmacol. Biol. Psychiatry 2021, 104, 110031. [Google Scholar] [CrossRef] [PubMed]
  37. Giuffrida, A.; Martinez, A. Chapter 3—The Endocannabinoid System and Parkinson Disease. In The Endocannabinoid System. Genetics, Biochemistry, Brain Disorders, and Therapy; Murillo-Rodríguez, E., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 63–81. [Google Scholar]
  38. Contino, M.; McCormick, P.J. Editorial: The Canonical and Non-Canonical Endocannabinoid System as a Target in Cancer and Acute and Chronic Pain. Front. Pharmacol. 2020, 11, 312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Moreno, E.; Cavic, M.; Canela, E.I. Functional Fine-Tuning of Metabolic Pathways by the Endocannabinoid System-Implications for Health and Disease. Int. J. Mol. Sci. 2021, 22, 3661. [Google Scholar] [CrossRef] [PubMed]
  40. Heinbockel, T.; Wang, Z.-J.; Brown, E.A.; Austin, P.T. Endocannabinoid Signaling in Neural Circuits of the Olfactory and Limbic System. In Cannabinoids in Health and Disease; InTechOpen: London, UK, 2016; pp. 11–37. [Google Scholar]
  41. Medeiros, D.d.C.; Cota, V.R.; Oliveira, A.C.P.; Moreira, F.A.; Moraes, M.F.D. The Endocannabinoid System Activation as a Neural Network Desynchronizing Mediator for Seizure Suppression. Front. Behav. Neurosci. 2020, 14, 603245. [Google Scholar] [CrossRef]
  42. Katona, I.; Freund, T.F. Endocannabinoid signaling as a synaptic circuit breaker in neurological disease. Nat. Med. 2008, 14, 923–930. [Google Scholar] [CrossRef] [PubMed]
  43. HUGO(a), G.N.C. CNR1. Available online: https://www.genenames.org/data/gene-symbol-report/#!/hgnc_id/2159 (accessed on 5 January 2022).
  44. HUGO(b), G.N.C. CNR2. Available online: https://www.genenames.org/data/gene-symbol-report/#!/hgnc_id/HGNC:2160 (accessed on 5 January 2022).
  45. Howlett, A.C.; Barth, F.; Bonner, T.I.; Cabral, G.; Casellas, P.; Devane, W.A.; Felder, C.C.; Herkenham, M.; Mackie, K.; Martin, B.R.; et al. International Union of Pharmacology. XXVII. Classification of cannabinoid receptors. Pharmacol. Rev. 2002, 54, 161–202. [Google Scholar] [CrossRef] [PubMed]
  46. Lu, H.-C.; Mackie, K. An Introduction to the Endogenous Cannabinoid System. Biol. Psychiatry 2016, 79, 516–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Veilleux, A.; Di Marzo, V.; Silvestri, C. The Expanded Endocannabinoid System/Endocannabinoidome as a Potential Target for Treating Diabetes Mellitus. Curr. Diab. Rep. 2019, 19, 117. [Google Scholar] [CrossRef]
  48. Cassano, T.; Calcagnini, S.; Pace, L.; De Marco, F.; Romano, A.; Gaetani, S. Cannabinoid Receptor 2 Signaling in Neurodegenerative Disorders: From Pathogenesis to a Promising Therapeutic Target. Front. Neurosci. 2017, 11, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Katchan, V.; David, P.; Shoenfeld, Y. Cannabinoids and autoimmune diseases: A systematic review. Autoimmun. Rev. 2016, 15, 513–528. [Google Scholar] [CrossRef] [PubMed]
  50. Kubajewska, I.; Constantinescu, C.S. Cannabinoids and experimental models of multiple sclerosis. Immunobiology 2010, 215, 647–657. [Google Scholar] [CrossRef]
  51. Moreno, E.; Cavic, M.; Krivokuca, A.; Canela, E.I. The Interplay between Cancer Biology and the Endocannabinoid System-Significance for Cancer Risk, Prognosis and Response to Treatment. Cancers 2020, 12, 3275. [Google Scholar] [CrossRef] [PubMed]
  52. Vaseghi, S.; Nasehi, M.; Zarrindast, M.-R. How do stupendous cannabinoids modulate memory processing via affecting neurotransmitter systems? Neurosci. Biobehav. Rev. 2021, 120, 173–221. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, Q.; Bhat, M.; Bowen, W.D.; Cheng, J. Signaling pathways from cannabinoid receptor-1 activation to inhibition of N-methyl-D-aspartic acid mediated calcium influx and neurotoxicity in dorsal root ganglion neurons. J. Pharmacol. Exp. Ther. 2009, 331, 1062–1070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Sagredo, O.; Pazos, M.R.; Valdeolivas, S.; Fernandez-Ruiz, J. Cannabinoids: Novel medicines for the treatment of Huntington’s disease. Recent Pat. CNS Drug Discov. 2012, 7, 41–48. [Google Scholar] [CrossRef] [PubMed]
  55. Földy, C.; Neu, A.; Jones, M.V.; Soltesz, I. Presynaptic, Activity-Dependent Modulation of Cannabinoid Type 1 Receptor-Mediated Inhibition of GABA Release. J. Neurosci. 2006, 26, 1465–1469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Monory, K.; Massa, F.; Egertová, M.; Eder, M.; Blaudzun, H.; Westenbroek, R.; Kelsch, W.; Jacob, W.; Marsch, R.; Ekker, M.; et al. The Endocannabinoid System Controls Key Epileptogenic Circuits in the Hippocampus. Neuron 2006, 51, 455–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Howlett, A.C. Cannabinoid receptor signaling. In Cannabinoids: Handbook of Experimental Pharmacology; Pertwee, R.G., Ed.; Springer: Berlin/Heidelberg, Germany, 2005; Volume 168, pp. 53–79. [Google Scholar] [CrossRef]
  58. Silvestri, C.; Di Marzo, V. The Endocannabinoid System in Energy Homeostasis and the Etiopathology of Metabolic Disorders. Cell Metab. 2013, 17, 475–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kreitzer, A.C.; Malenka, R.C. Endocannabinoid-mediated rescue of striatal LTD and motor deficits in Parkinson’s disease models. Nature 2007, 445, 643–647. [Google Scholar] [CrossRef] [PubMed]
  60. Liu, Q.R.; Pan, C.H.; Hishimoto, A.; Li, C.Y.; Xi, Z.X.; Llorente-Berzal, A.; Viveros, M.P.; Ishiguro, H.; Arinami, T.; Onaivi, E.S.; et al. Species differences in cannabinoid receptor 2 (CNR2gene): Identification of novel human and rodent CB2 isoforms, differential tissue expression and regulation by cannabinoid receptor ligands. Genes Brain Behav. 2009, 8, 519–530. [Google Scholar] [CrossRef] [PubMed]
  61. Cabral, G.A.; Griffin-Thomas, L. Emerging role of the cannabinoid receptor CB2in immune regulation: Therapeutic prospects for neuroinflammation. Expert Rev. Mol. Med. 2009, 11, e3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Núñez, E.; Benito, C.; Tolón, R.M.; Hillard, C.J.; Griffin, W.S.T.; Romero, J. Glial expression of cannabinoid CB2 receptors and fatty acid amide hydrolase are beta amyloid–linked events in Down’s syndrome. Neuroscience 2008, 151, 104–110. [Google Scholar] [CrossRef] [PubMed]
  63. Van Sickle, M.D.; Duncan, M.; Kingsley, P.J.; Mouihate, A.; Urbani, P.; Mackie, K.; Stella, N.; Makriyannis, A.; Piomelli, D.; Davison, J.S.; et al. Identification and Functional Characterization of Brainstem Cannabinoid CB 2 Receptors. Science 2005, 310, 329–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Aso, E.; Ferrer, I. CB2 Cannabinoid Receptor as Potential Target against Alzheimer’s Disease. Front. Neurosci. 2016, 10, 243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Price, T.J.; Patwardhan, A.; Akopian, A.N.; Hargreaves, K.M.; Flores, C.M. Modulation of trigeminal sensory neuron activity by the dual cannabinoid-vanilloid agonists anandamide, N-arachidonoyl-dopamine and arachidonyl-2-chloroethylamide. Br. J. Pharmacol. 2004, 141, 1118–1130. [Google Scholar] [CrossRef] [PubMed]
  66. Aso, E.; Palomer, E.; Juves, S.; Maldonado, R.; Munoz, F.J.; Ferrer, I. CB1 agonist ACEA protects neurons and reduces the cognitive impairment of AbetaPP/PS1 mice. J. Alzheimer’s Dis. 2012, 30, 439–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Croxford, J.L.; Pryce, G.; Jackson, S.J.; Ledent, C.; Giovannoni, G.; Pertwee, R.G.; Yamamura, T.; Baker, D. Cannabinoid-mediated neuroprotection, not immunosuppression, may be more relevant to multiple sclerosis. J. Neuroimmunol. 2008, 193, 120–129. [Google Scholar] [CrossRef] [PubMed]
  68. Leuti, A.; Fazio, D.; Fava, M.; Piccoli, A.; Oddi, S.; Maccarrone, M. Bioactive lipids, inflammation and chronic diseases. Adv. Drug Deliv. Rev. 2020, 159, 133–169. [Google Scholar] [CrossRef]
  69. Panikashvili, D.; Shein, N.A.; Mechoulam, R.; Trembovler, V.; Kohen, R.; Alexandrovich, A.; Shohami, E. The endocannabinoid 2-AG protects the blood-brain barrier after closed head injury and inhibits mRNA expression of proinflammatory cytokines. Neurobiol. Dis. 2006, 22, 257–264. [Google Scholar] [CrossRef] [PubMed]
  70. Klein, T.W. Cannabinoid-based drugs as anti-inflammatory therapeutics. Nat. Rev. Immunol. 2005, 5, 400–411. [Google Scholar] [CrossRef] [PubMed]
  71. Buisseret, B.; Alhouayek, M.; Guillemot-Legris, O.; Muccioli, G.G. Endocannabinoid and Prostanoid Crosstalk in Pain. Trends Mol. Med. 2019, 25, 882–896. [Google Scholar] [CrossRef] [PubMed]
  72. Hind, W.H.; Tufarelli, C.; Neophytou, M.; Anderson, S.I.; England, T.J.; O’Sullivan, S.E. Endocannabinoids modulate human blood-brain barrier permeability in vitro. Br. J. Pharmacol. 2015, 172, 3015–3027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Gris, J.C.; Nobile, B.; Bouvier, S. Neuropsychiatric presentations of antiphospholipid antibodies. Thromb. Res. 2015, 135, S56–S59. [Google Scholar] [CrossRef]
  74. Gitler, A.D.; Dhillon, P.; Shorter, J. Neurodegenerative disease: Models, mechanisms, and a new hope. Dis. Model. Mech. 2017, 10, 499–502. [Google Scholar] [CrossRef] [Green Version]
  75. Pugazhendhi, A.; Suganthy, N.; Chau, T.P.; Sharma, A.; Unpaprom, Y.; Ramaraj, R.; Karuppusamy, I.; Brindhadevi, K. Cannabinoids as anticancer and neuroprotective drugs: Structural insights and pharmacological interactions—A review. Process. Biochem. 2021, 111, 9–31. [Google Scholar] [CrossRef]
  76. García, C.; Palomo-Garo, C.; Gómez-Gálvez, Y.; Fernández-Ruiz, J. Cannabinoid–dopamine interactions in the physiology and physiopathology of the basal ganglia. Br. J. Pharmacol. 2016, 173, 2069–2079. [Google Scholar] [CrossRef]
  77. Loera-Valencia, R.; Cedazo-Minguez, A.; Kenigsberg, P.A.; Page, G.; Duarte, A.I.; Giusti, P.; Zusso, M.; Robert, P.; Frisoni, G.B.; Cattaneo, A.; et al. Current and emerging avenues for Alzheimer’s disease drug targets. J. Intern. Med. 2019, 286, 398–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Stanciu, G.D.; Rusu, R.N.; Bild, V.; Filipiuc, L.E.; Tamba, B.-I.; Ababei, D.C. Systemic Actions of SGLT2 Inhibition on Chronic mTOR Activation as a Shared Pathogenic Mechanism between Alzheimer’s Disease and Diabetes. Biomedicines 2021, 9, 576. [Google Scholar] [CrossRef]
  79. Outen, J.D.; Burhanullah, M.H.; Vandrey, R.; Amjad, H.; Harper, D.G.; Patrick, R.E.; May, R.L.; Agronin, M.E.; Forester, B.P.; Rosenberg, P.B. Cannabinoids for Agitation in Alzheimer’s Disease. Am. J. Geriatr. Psychiatry 2021, 29, 1253–1263. [Google Scholar] [CrossRef] [PubMed]
  80. Stanciu, G.D.; Luca, A.; Rusu, R.N.; Bild, V.; Beschea Chiriac, S.I.; Solcan, C.; Bild, W.; Ababei, D.C. Alzheimer’s Disease Pharmacotherapy in Relation to Cholinergic System Involvement. Biomolecules 2020, 10, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Vaseghi, S.; Babapour, V.; Nasehi, M.; Zarrindast, M.-R. The role of CA1 CB1 receptors on lithium-induced spatial memory impairment in rats. EXCLI J. 2018, 17, 916–934. [Google Scholar] [CrossRef] [PubMed]
  82. Jenny, M.; Schröcksnadel, S.; Überall, F.; Fuchs, D. The Potential Role of Cannabinoids in Modulating Serotonergic Signaling by Their Influence on Tryptophan Metabolism. Pharmaceutical 2010, 3, 2647–2660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Eubanks, L.M.; Rogers, C.J.; Beuscher, A.E.; Koob, G.F.; Olson, A.J.; Dickerson, T.J.; Janda, K.D. A Molecular Link Between the Active Component of Marijuana and Alzheimer’s Disease Pathology. Mol. Pharm. 2006, 3, 773–777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bajaj, S.; Jain, S.; Vyas, P.; Bawa, S.; Vohora, D. The role of endocannabinoid pathway in the neuropathology of Alzheimer’s disease: Can the inhibitors of MAGL and FAAH prove to be potential therapeutic targets against the cognitive impairment associated with Alzheimer’s disease? Brain Res. Bull. 2021, 174, 305–322. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, X.; Zhang, J.; Chen, C. Endocannabinoid 2-arachidonoylglycerol protects neurons against β-amyloid insults. Neurosci. Biobehav. Rev. 2011, 178, 159–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zhang, J.; Chen, C. Endocannabinoid 2-Arachidonoylglycerol Protects Neurons by Limiting COX-2 Elevation. J. Biol. Chem. 2008, 283, 22601–22611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Çakır, M.; Tekin, S.; Doğanyiğit, Z.; Erden, Y.; Soytürk, M.; Çiğremiş, Y.; Sandal, S. Cannabinoid type 2 receptor agonist JWH-133, attenuates Okadaic acid induced spatial memory impairment and neurodegeneration in rats. Life Sci. 2019, 217, 25–33. [Google Scholar] [CrossRef] [PubMed]
  88. Han, Q.-W.; Yuan, Y.-H.; Chen, N.-H. The therapeutic role of cannabinoid receptors and its agonists or antagonists in Parkinson’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 2020, 96, 109745. [Google Scholar] [CrossRef]
  89. Burstein, S. Cannabidiol (CBD) and its analogs: A review of their effects on inflammation. Bioorg. Med. Chem. 2015, 23, 1377–1385. [Google Scholar] [CrossRef] [PubMed]
  90. Hao, F.; Feng, Y. Cannabidiol (CBD) enhanced the hippocampal immune response and autophagy of APP/PS1 Alzheimer’s mice uncovered by RNA-seq. Life Sci. 2021, 264, 118624. [Google Scholar] [CrossRef] [PubMed]
  91. Hayakawa, K.; Mishima, K.; Nozako, M.; Hazekawa, M.; Irie, K.; Fujioka, M.; Orito, K.; Abe, K.; Hasebe, N.; Egashira, N.; et al. Delayed treatment with cannabidiol has a cerebroprotective action via a cannabinoid receptor-independent myeloperoxidase-inhibiting mechanism. J. Neurochem. 2007, 102, 1488–1496. [Google Scholar] [CrossRef] [PubMed]
  92. Li, C.; Shi, J.; Wang, B.; Li, J.; Jia, H. CB2 cannabinoid receptor agonist ameliorates novel object recognition but not spatial memory in transgenic APP/PS1 mice. Neurosci. Lett. 2019, 707, 134286. [Google Scholar] [CrossRef] [PubMed]
  93. Kuljis, R.O. Lesions in the Pulvinar in Patients with Alzheimer’s Disease. J. Neuropathol. Exp. Neurol. 1994, 53, 202–211. [Google Scholar] [CrossRef] [PubMed]
  94. Chou, X.-l.; Fang, Q.; Yan, L.; Zhong, W.; Peng, B.; Li, H.; Wei, J.; Tao, H.W.; Zhang, L.I. Contextual and cross-modality modulation of auditory cortical processing through pulvinar mediated suppression. eLife 2020, 9, e54157. [Google Scholar] [CrossRef]
  95. Fang, Q.; Chou, X.-l.; Peng, B.; Zhong, W.; Zhang, L.I.; Tao, H.W. A Differential Circuit via Retino-Colliculo-Pulvinar Pathway Enhances Feature Selectivity in Visual Cortex through Surround Suppression. Neuron 2020, 105, 355–369.e356. [Google Scholar] [CrossRef] [PubMed]
  96. Ibrahim, L.A.; Mesik, L.; Ji, X.-Y.; Fang, Q.; Li, H.-F.; Li, Y.-T.; Zingg, B.; Zhang, L.I.; Tao, H.W. Cross-Modality Sharpening of Visual Cortical Processing through Layer-1-Mediated Inhibition and Disinhibition. Neuron 2016, 89, 1031–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Amin, M.; Ontaneda, D. Thalamic Injury and Cognition in Multiple Sclerosis. Front. Neurol. 2021, 11, 623914. [Google Scholar] [CrossRef] [PubMed]
  98. Erskine, D.; Thomas, A.J.; Attems, J.; Taylor, J.P.; McKeith, I.G.; Morris, C.M.; Khundakar, A.A. Specific patterns of neuronal loss in the pulvinar nucleus in dementia with lewy bodies. Mov. Disord. 2017, 32, 414–422. [Google Scholar] [CrossRef] [Green Version]
  99. Micaelo-Fernandes, C.; Bouskila, J.; Bouchard, J.-F.; Ptito, M. Presence of the Endocannabinoid System in the Inferior Pulvinar of the Vervet Monkey. Brain Sci. 2021, 11, 770. [Google Scholar] [CrossRef] [PubMed]
  100. Platt, B.; Kamboj, S.; Morgan, C.J.A.; Curran, H.V. Processing dynamic facial affect in frequent cannabis-users: Evidence of deficits in the speed of identifying emotional expressions. Drug Alcohol Depend. 2010, 112, 27–32. [Google Scholar] [CrossRef] [PubMed]
  101. Baul, H.S.; Manikandan, C.; Sen, D. Cannabinoid receptor as a potential therapeutic target for Parkinson’s Disease. Brain Res. Bull. 2019, 146, 244–252. [Google Scholar] [CrossRef] [PubMed]
  102. Lanciego, J.L.; López, I.P.; Rico, A.J.; Aymerich, M.S.; Pérez-Manso, M.; Conte, L.; Combarro, C.; Roda, E.; Molina, C.; Gonzalo, N.; et al. The search for a role of the caudal intralaminar nuclei in the pathophysiology of Parkinson’s disease. Brain Res. Bull. 2009, 78, 55–59. [Google Scholar] [CrossRef] [PubMed]
  103. Kibret, B.G.; Ishiguro, H.; Horiuchi, Y.; Onaivi, E.S. New Insights and Potential Therapeutic Targeting of CB2 Cannabinoid Receptors in CNS Disorders. Int. J. Mol. Sci. 2022, 23, 975. [Google Scholar] [CrossRef] [PubMed]
  104. Shi, J.; Cai, Q.; Zhang, J.; He, X.; Liu, Y.; Zhu, R.; Jin, L. AM1241 alleviates MPTP-induced Parkinson’s disease and promotes the regeneration of DA neurons in PD mice. Oncotarget 2017, 8, 67837–67850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Gómez-Gálvez, Y.; Palomo-Garo, C.; Fernández-Ruiz, J.; García, C. Potential of the cannabinoid CB2 receptor as a pharmacological target against inflammation in Parkinson’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 64, 200–208. [Google Scholar] [CrossRef]
  106. Chagas, M.H.N.; 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.C.; Crippa, J.A.S. 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]
  107. Hawes, E.M.; Lee, C.R.; Brackney, D.E.; Ensley, T.G.; Kidd, J.; Page, C. Cannabidiol Products: Review of the Regulatory and Clinical Considerations. J. Nurse Pract. 2020, 16, 747–755. [Google Scholar] [CrossRef]
  108. Brotchie, J.M. CB1 cannabinoid receptor signalling in Parkinson’s disease. Curr. Opin. Pharmacol. 2003, 3, 54–61. [Google Scholar] [CrossRef]
  109. Chung, Y.C.; Bok, E.; Huh, S.H.; Park, J.Y.; Yoon, S.H.; Kim, S.R.; Kim, Y.S.; Maeng, S.; Park, S.H.; Jin, B.K. Cannabinoid receptor type 1 protects nigrostriatal dopaminergic neurons against MPTP neurotoxicity by inhibiting microglial activation. J. Immunol. 2011, 187, 6508–6517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Zuccato, C.; Valenza, M.; Cattaneo, E. Molecular Mechanisms and Potential Therapeutical Targets in Huntington’s Disease. Physiol. Rev. 2010, 90, 905–981. [Google Scholar] [CrossRef] [PubMed]
  111. Bates, G.P.; Dorsey, R.; Gusella, J.F.; Hayden, M.R.; Kay, C.; Leavitt, B.R.; Nance, M.; Ross, C.A.; Scahill, R.I.; Wetzel, R.; et al. Huntington disease. Nat. Rev. Dis. Primers 2015, 1, 15005. [Google Scholar] [CrossRef] [PubMed]
  112. Labbadia, J.; Morimoto, R.I. Huntington’s disease: Underlying molecular mechanisms and emerging concepts. Trends Biochem. Sci. 2013, 38, 378–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Demetriou, C.A.; Heraclides, A.; Salafori, C.; Tanteles, G.A.; Christodoulou, K.; Christou, Y.; Zamba-Papanicolaou, E. Epidemiology of Huntington disease in Cyprus: A 20-year retrospective study. Clin. Genet. 2018, 93, 656–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Cloud, L.J.; Rosenblatt, A.; Margolis, R.L.; Ross, C.A.; Pillai, J.A.; Corey-Bloom, J.; Tully, H.M.; Bird, T.; Panegyres, P.K.; Nichter, C.A.; et al. Seizures in juvenile Huntington’s disease: Frequency and characterization in a multicenter cohort. Mov. Disord. 2012, 27, 1797–1800. [Google Scholar] [CrossRef] [PubMed]
  115. Douglas, I.; Evans, S.; Rawlins, M.D.; Smeeth, L.; Tabrizi, S.J.; Wexler, N.S. Juvenile Huntington’s disease: A population-based study using the General Practice Research Database. BMJ Open 2013, 3, e002085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Quarrell, O.W.; Nance, M.A.; Nopoulos, P.; Paulsen, J.S.; Smith, J.A.; Squitieri, F. Managing juvenile Huntington’s disease. Neurodegener. Dis. Manag. 2013, 3, 267–276. [Google Scholar] [CrossRef] [Green Version]
  117. Letort, D.; Gonzalez-Alegre, P. Huntington’s disease in children. Handb. Clin. Neurol. 2013, 113, 1913–1917. [Google Scholar] [CrossRef] [PubMed]
  118. Ajitkumar, A.; De Jesus, O. Huntington Disease; StatPearls Publishing LLC.: Treasure Island, FL, USA, 2022. [Google Scholar]
  119. Armstrong, M.J.; Miyasaki, J.M. Evidence-based guideline: Pharmacologic treatment of chorea in Huntington disease: Report of the guideline development subcommittee of the American Academy of Neurology. Neurology 2012, 79, 597–603. [Google Scholar] [CrossRef] [Green Version]
  120. Bagchi, S.P. Differential interactions of phencyclidine with tetrabenazine and reserpine affecting intraneuronal dopamine. Biochem. Pharmacol. 1983, 32, 2851–2856. [Google Scholar] [CrossRef]
  121. Paleacu, D. Tetrabenazine in the treatment of Huntington’s disease. Neuropsychiatr. Dis. Treat. 2007, 3, 545–551. [Google Scholar] [PubMed]
  122. Pidgeon, C.; Rickards, H. The pathophysiology and pharmacological treatment of Huntington disease. Behav. Neurol. 2013, 26, 245–253. [Google Scholar] [CrossRef] [PubMed]
  123. Frank, S. Treatment of Huntington’s Disease. Neurotherapeutics 2013, 11, 153–160. [Google Scholar] [CrossRef] [Green Version]
  124. Potkin, K.T.; Potkin, S.G. New directions in therapeutics for Huntington disease. Future Neurol. 2018, 13, 101–121. [Google Scholar] [CrossRef] [PubMed]
  125. Bonelli, R.M.; Niederwieser, G.; Diez, J.; Gruber, A.; Koltringer, P. Pramipexole ameliorates neurologic and psychiatric symptoms in a Westphal variant of Huntington’s disease. Clin. Neuropharmacol. 2002, 25, 58–60. [Google Scholar] [CrossRef] [PubMed]
  126. Low, P.A.; Allsop, J.L.; Halmacyi, G.M. Huntington’s chorea: The rigid form (Westphal variant) treated with levodopa. Med. J. Aust. 1974, 1, 393–394. [Google Scholar] [CrossRef] [PubMed]
  127. Magnet, M.K.; Bonelli, R.M.; Kapfhammer, H.-P. Amantadine in the Akinetic-Rigid Variant of Huntington’s Disease. Ann. Pharmacother. 2004, 38, 1194–1196. [Google Scholar] [CrossRef]
  128. Van Laere, K.; Casteels, C.; Dhollander, I.; Goffin, K.; Grachev, I.; Bormans, G.; Vandenberghe, W. Widespread Decrease of Type 1 Cannabinoid Receptor Availability in Huntington Disease In Vivo. J. Nucl. Med. 2010, 51, 1413–1417. [Google Scholar] [CrossRef] [Green Version]
  129. Ooms, M.; Rietjens, R.; Rangarajan, J.R.; Vunckx, K.; Valdeolivas, S.; Maes, F.; Himmelreich, U.; Fernandez-Ruiz, J.; Bormans, G.; Van Laere, K.; et al. Early decrease of type 1 cannabinoid receptor binding and phosphodiesterase 10A activity in vivo in R6/2 Huntington mice. Neurobiol. Aging 2014, 35, 2858–2869. [Google Scholar] [CrossRef] [Green Version]
  130. Ceccarini, J.; Ahmad, R.; Van de Vliet, L.; Casteels, C.; Vandenbulcke, M.; Vandenberghe, W.; Van Laere, K. Behavioral Symptoms in Premanifest Huntington Disease Correlate with Reduced Frontal CB1R Levels. J. Nucl. Med. 2019, 60, 115–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Allen, K.L.; Waldvogel, H.J.; Glass, M.; Faull, R.L.M. Cannabinoid (CB1), GABAA and GABAB receptor subunit changes in the globus pallidus in Huntington’s disease. J. Chem. Neuroanat. 2009, 37, 266–281. [Google Scholar] [CrossRef]
  132. Bisogno, T.; Martire, A.; Petrosino, S.; Popoli, P.; Di Marzo, V. Symptom-related changes of endocannabinoid and palmitoylethanolamide levels in brain areas of R6/2 mice, a transgenic model of Huntington’s disease. Neurochem. Int. 2008, 52, 307–313. [Google Scholar] [CrossRef] [PubMed]
  133. Mievis, S.; Blum, D.; Ledent, C. Worsening of Huntington disease phenotype in CB1 receptor knockout mice. Neurobiol. Dis. 2011, 42, 524–529. [Google Scholar] [CrossRef] [PubMed]
  134. Glass, M.; Dragunow, M.; Faull, R.L. The pattern of neurodegeneration in Huntington’s disease: A comparative study of cannabinoid, dopamine, adenosine and GABA(A) receptor alterations in the human basal ganglia in Huntington’s disease. Neuroscience 2000, 97, 505–519. [Google Scholar] [CrossRef]
  135. Glass, M.; van Dellen, A.; Blakemore, C.; Hannan, A.J.; Faull, R.L.M. Delayed onset of huntington′s disease in mice in an enriched environment correlates with delayed loss of cannabinoid CB1 receptors. Neuroscience 2004, 123, 207–212. [Google Scholar] [CrossRef]
  136. Blázquez, C.; Chiarlone, A.; Sagredo, O.; Aguado, T.; Pazos, M.R.; Resel, E.; Palazuelos, J.; Julien, B.; Salazar, M.; Börner, C.; et al. Loss of striatal type 1 cannabinoid receptors is a key pathogenic factor in Huntington’s disease. Brain 2011, 134, 119–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ignatowska-Jankowska, B.M.; Baillie, G.L.; Kinsey, S.; Crowe, M.; Ghosh, S.; Owens, R.A.; Damaj, I.M.; Poklis, J.; Wiley, J.L.; Zanda, M.; et al. A Cannabinoid CB1 Receptor-Positive Allosteric Modulator Reduces Neuropathic Pain in the Mouse with No Psychoactive Effects. Neuropsychopharmacology 2015, 40, 2948–2959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Laprairie, R.B.; Bagher, A.M.; Rourke, J.L.; Zrein, A.; Cairns, E.A.; Kelly, M.E.M.; Sinal, C.J.; Kulkarni, P.M.; Thakur, G.A.; Denovan-Wright, E.M. Positive allosteric modulation of the type 1 cannabinoid receptor reduces the signs and symptoms of Huntington’s disease in the R6/2 mouse model. Neuropharmacology 2019, 151, 1–12. [Google Scholar] [CrossRef]
  139. Slivicki, R.A.; Xu, Z.; Kulkarni, P.M.; Pertwee, R.G.; Mackie, K.; Thakur, G.A.; Hohmann, A.G. Positive Allosteric Modulation of Cannabinoid Receptor Type 1 Suppresses Pathological Pain Without Producing Tolerance or Dependence. Biol. Psychiatry 2018, 84, 722–733. [Google Scholar] [CrossRef] [Green Version]
  140. Palazuelos, J.; Aguado, T.; Pazos, M.R.; Julien, B.; Carrasco, C.; Resel, E.; Sagredo, O.; Benito, C.; Romero, J.; Azcoitia, I.; et al. Microglial CB2 cannabinoid receptors are neuroprotective in Huntington’s disease excitotoxicity. Brain 2009, 132, 3152–3164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Navarro, G.; Morales, P.; Rodríguez-Cueto, C.; Fernández-Ruiz, J.; Jagerovic, N.; Franco, R. Targeting Cannabinoid CB2 Receptors in the Central Nervous System. Medicinal Chemistry Approaches with Focus on Neurodegenerative Disorders. Front. Neurosci. 2016, 10, 406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Polini, B.; Cervetto, C.; Carpi, S.; Pelassa, S.; Gado, F.; Ferrisi, R.; Bertini, S.; Nieri, P.; Marcoli, M.; Manera, C. Positive Allosteric Modulation of CB1 and CB2 Cannabinoid Receptors Enhances the Neuroprotective Activity of a Dual CB1R/CB2R Orthosteric Agonist. Life 2020, 10, 333. [Google Scholar] [CrossRef] [PubMed]
  143. Chen, D.-J.; Gao, M.; Gao, F.-F.; Su, Q.-X.; Wu, J. Brain cannabinoid receptor 2: Expression, function and modulation. Acta Pharmacol. Sin. 2017, 38, 312–316. [Google Scholar] [CrossRef] [PubMed]
  144. Curtis, A.; Mitchell, I.; Patel, S.; Ives, N.; Rickards, H. A pilot study using nabilone for symptomatic treatment in Huntington’s disease. Mov. Disord. 2009, 24, 2254–2259. [Google Scholar] [CrossRef] [PubMed]
  145. Lastres-Becker, I.; Hansen, H.H.; Berrendero, F.; De Miguel, R.; Pérez-Rosado, A.; Manzanares, J.; Ramos, J.A.; Fernández-Ruiz, J. Alleviation of motor hyperactivity and neurochemical deficits by endocannabinoid uptake inhibition in a rat model of Huntington’s disease. Synapse 2002, 44, 23–35. [Google Scholar] [CrossRef] [PubMed]
  146. Peres, F.F.; Lima, A.C.; Hallak, J.E.C.; Crippa, J.A.; Silva, R.H.; Abílio, V.C. Cannabidiol as a Promising Strategy to Treat and Prevent Movement Disorders? Front. Pharmacol. 2018, 9, 482. [Google Scholar] [CrossRef] [Green Version]
  147. Fernández-Ruiz, J. The endocannabinoid system as a target for the treatment of motor dysfunction. Br. J. Pharmacol. 2009, 156, 1029–1040. [Google Scholar] [CrossRef] [Green Version]
  148. Pintor, A.; Tebano, M.T.; Martire, A.; Grieco, R.; Galluzzo, M.; Scattoni, M.L.; Pèzzola, A.; Coccurello, R.; Felici, F.; Cuomo, V.; et al. The cannabinoid receptor agonist WIN 55,212-2 attenuates the effects induced by quinolinic acid in the rat striatum. Neuropharmacology 2006, 51, 1004–1012. [Google Scholar] [CrossRef] [Green Version]
  149. Sagredo, O.; García-Arencibia, M.; de Lago, E.; Finetti, S.; Decio, A.; Fernández-Ruiz, J. Cannabinoids and Neuroprotection in Basal Ganglia Disorders. Mol. Neurobiol. 2007, 36, 82–91. [Google Scholar] [CrossRef]
  150. Heim, B.; Bajaj, S.; Marzi, R.D.; Mangesius, S.; Djamshidian, A.; Poewe, W.; Seppi, K. M6 Nabilone in huntington’s disease: A case series of five patients. J. Neurol. Neurosurg. Psychiatry 2016, 87, A103. [Google Scholar] [CrossRef]
  151. Akinyemi, E.; Randhawa, G.; Longoria, V.; Zeine, R. Medical Marijuana Effects in Movement Disorders, Focus on Huntington Disease; A Literature Review. J. Pharm. Pharm. Sci. 2020, 23, 389–395. [Google Scholar] [CrossRef] [PubMed]
  152. Saft, C.; von Hein, S.M.; Lücke, T.; Thiels, C.; Peball, M.; Djamshidian, A.; Heim, B.; Seppi, K. Cannabinoids for Treatment of Dystonia in Huntington’s Disease. J. Huntington’s Dis. 2018, 7, 167–173. [Google Scholar] [CrossRef] [PubMed]
  153. Nachnani, R.; Raup-Konsavage, W.M.; Vrana, K.E. The Pharmacological Case for Cannabigerol. J. Pharmacol. Exp. Ther. 2021, 376, 204–212. [Google Scholar] [CrossRef] [PubMed]
  154. Díaz-Alonso, J.; Paraíso-Luna, J.; Navarrete, C.; del Río, C.; Cantarero, I.; Palomares, B.; Aguareles, J.; Fernández-Ruiz, J.; Bellido, M.L.; Pollastro, F.; et al. VCE-003.2, a novel cannabigerol derivative, enhances neuronal progenitor cell survival and alleviates symptomatology in murine models of Huntington’s disease. Sci. Rep. 2016, 6, 29789. [Google Scholar] [CrossRef] [Green Version]
  155. Aiken, C.T.; Tobin, A.J.; Schweitzer, E.S. A cell-based screen for drugs to treat Huntington’s disease. Neurobiol. Dis. 2004, 16, 546–555. [Google Scholar] [CrossRef]
  156. Maker-Clark, G.; Patel, S. Integrative therapies for multiple sclerosis. Dis. Mon. 2013, 59, 290–301. [Google Scholar] [CrossRef]
  157. Al-Ghezi, Z.Z.; Miranda, K.; Nagarkatti, M.; Nagarkatti, P.S. Combination of Cannabinoids, Δ9-Tetrahydrocannabinol and Cannabidiol, Ameliorates Experimental Multiple Sclerosis by Suppressing Neuroinflammation Through Regulation of miRNA-Mediated Signaling Pathways. Front. Immunol. 2019, 10, 1921. [Google Scholar] [CrossRef]
  158. Ford, H. Clinical presentation and diagnosis of multiple sclerosis. Clin. Med. 2020, 20, 380–383. [Google Scholar] [CrossRef]
  159. Miclea, A.; Salmen, A.; Zoehner, G.; Diem, L.; Kamm, C.P.; Chaloulos-Iakovidis, P.; Miclea, M.; Briner, M.; Kilidireas, K.; Stefanis, L.; et al. Age-dependent variation of female preponderance across different phenotypes of multiple sclerosis: A retrospective cross-sectional study. CNS Neurosci. Ther. 2018, 25, 527–531. [Google Scholar] [CrossRef] [Green Version]
  160. Walton, C.; King, R.; Rechtman, L.; Kaye, W.; Leray, E.; Marrie, R.A.; Robertson, N.; La Rocca, N.; Uitdehaag, B.; van der Mei, I.; et al. Rising prevalence of multiple sclerosis worldwide: Insights from the Atlas of MS, third edition. Mult. Scler. J. 2020, 26, 1816–1821. [Google Scholar] [CrossRef] [PubMed]
  161. Correa, F.; Hernangómez-Herrero, M.; Mestre, L.; Loría, F.; Docagne, F.; Guaza, C. The endocannabinoid anandamide downregulates IL-23 and IL-12 subunits in a viral model of multiple sclerosis: Evidence for a cross-talk between IL-12p70/IL-23 axis and IL-10 in microglial cells. Brain Behav. Immun. 2011, 25, 736–749. [Google Scholar] [CrossRef] [PubMed]
  162. Rabiei, Z. Phytotherapy as a Complementary Medicine for Multiple Sclerosis. Turk. J. Pharm. Sci. 2019, 16, 246–251. [Google Scholar] [CrossRef] [PubMed]
  163. Huseby, E.S.; Huseby, P.G.; Shah, S.; Smith, R.; Stadinski, B.D. Pathogenic CD8 T Cells in Multiple Sclerosis and Its Experimental Models. Front. Immunol. 2012, 3, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Huseby, E.S.; Liggitt, D.; Brabb, T.; Schnabel, B.; Öhlén, C.; Goverman, J. A Pathogenic Role for Myelin-Specific Cd8+ T Cells in a Model for Multiple Sclerosis. J. Exp. Med. 2001, 194, 669–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Fraussen, J.; de Bock, L.; Somers, V. B cells and antibodies in progressive multiple sclerosis: Contribution to neurodegeneration and progression. Autoimmun. Rev. 2016, 15, 896–899. [Google Scholar] [CrossRef]
  166. Zhou, T.; Ahmad, T.K.; Alrushaid, S.; Pozdirca, M.; Ethans, K.; Intrater, H.; Le, T.; Burczynski, F.; Kong, J.; Namaka, M. Therapeutic impact of orally administered cannabinoid oil extracts in an experimental autoimmune encephalomyelitis animal model of multiple sclerosis. Biochem. Biophys. Res. Commun. 2019, 516, 373–380. [Google Scholar] [CrossRef] [PubMed]
  167. Magliozzi, R.; Howell, O.; Vora, A.; Serafini, B.; Nicholas, R.; Puopolo, M.; Reynolds, R.; Aloisi, F. Meningeal B-cell follicles in secondary progressive multiple sclerosis associate with early onset of disease and severe cortical pathology. Brain 2006, 130, 1089–1104. [Google Scholar] [CrossRef] [PubMed]
  168. Jackson, S.J.; Diemel, L.T.; Pryce, G.; Baker, D. Cannabinoids and neuroprotection in CNS inflammatory disease. J. Neurol. Sci. 2005, 233, 21–25. [Google Scholar] [CrossRef] [PubMed]
  169. Peferoen, L.; Kipp, M.; van der Valk, P.; van Noort, J.M.; Amor, S. Oligodendrocyte-microglia cross-talk in the central nervous system. Immunology 2014, 141, 302–313. [Google Scholar] [CrossRef]
  170. Kipp, M. Oligodendrocyte Physiology and Pathology Function. Cells 2020, 9, 2078. [Google Scholar] [CrossRef] [PubMed]
  171. Correale, J.; Farez, M.F. The Role of Astrocytes in Multiple Sclerosis Progression. Front. Neurol. 2015, 6, 180. [Google Scholar] [CrossRef] [PubMed]
  172. Fitzpatrick, J.-M.K.; Downer, E.J. Toll-like receptor signalling as a cannabinoid target in Multiple Sclerosis. Neuropharmacology 2017, 113, 618–626. [Google Scholar] [CrossRef] [PubMed]
  173. Gustavsen, S.; Søndergaard, H.; Linnet, K.; Thomsen, R.; Rasmussen, B.; Sorensen, P.; Sellebjerg, F.; Oturai, A. Safety and efficacy of low-dose medical cannabis oils in multiple sclerosis. Mult. Scler. Relat. Disord. 2021, 48, 102708. [Google Scholar] [CrossRef]
  174. Kappos, L.; Fox, R.J.; Burcklen, M.; Freedman, M.S.; Havrdová, E.K.; Hennessy, B.; Hohlfeld, R.; Lublin, F.; Montalban, X.; Pozzilli, C.; et al. Ponesimod Compared with Teriflunomide in Patients with Relapsing Multiple Sclerosis in the Active-Comparator Phase 3 OPTIMUM Study. JAMA Neurol. 2021, 78, 558. [Google Scholar] [CrossRef] [PubMed]
  175. Lee, J.Y.; Taghian, K.; Petratos, S. Axonal degeneration in multiple sclerosis: Can we predict and prevent permanent disability? Acta Neuropathol. Commun. 2014, 2, 97. [Google Scholar] [CrossRef] [PubMed]
  176. da Rovare, V.P.; Magalhães, G.P.A.; Jardini, G.D.A.; Beraldo, M.L.; Gameiro, M.O.; Agarwal, A.; Luvizutto, G.J.; Paula-Ramos, L.; Camargo, S.E.A.; de Oliveira, L.D.; et al. Cannabinoids for spasticity due to multiple sclerosis or paraplegia: A systematic review and meta-analysis of randomized clinical trials. Complement. Ther. Med. 2017, 34, 170–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Novotna, A.; Mares, J.; Ratcliffe, S.; Novakova, I.; Vachova, M.; Zapletalova, O.; Gasperini, C.; Pozzilli, C.; Cefaro, L.; Comi, G.; et al. A randomized, double-blind, placebo-controlled, parallel-group, enriched-design study of nabiximols * (Sativex®), as add-on therapy, in subjects with refractory spasticity caused by multiple sclerosis. Eur. J. Neurol. 2011, 18, 1122–1131. [Google Scholar] [CrossRef]
  178. Kim-Fine, S.; Greenfield, J.; Chaput, K.H.; Robert, M.; Metz, L.M. Cannabinoids and bladder symptoms in multiple sclerosis. Mult. Scler. Relat. Disord. 2021, 54, 103105. [Google Scholar] [CrossRef]
  179. Lublin, F.D.; Reingold, S.C.; Cohen, J.A.; Cutter, G.R.; Sørensen, P.S.; Thompson, A.J.; Wolinsky, J.S.; Balcer, L.J.; Banwell, B.; Barkhof, F.; et al. Defining the clinical course of multiple sclerosis. The 2013 revisions. Neurology 2014, 83, 278–286. [Google Scholar] [CrossRef] [Green Version]
  180. Rasoul, A.A.; Khudhur, Z.O.; Hamad, M.S.; Ismaeal, Y.S.; Smail, S.W.; Rasul, M.F.; Mohammad, K.A.; Bapir, A.A.; Omar, S.A.; Qadir, M.K.; et al. The role of oxidative stress and haematological parameters in relapsing-remitting multiple sclerosis in Kurdish population. Mult. Scler. Relat. Disord. 2021, 56, 103228. [Google Scholar] [CrossRef] [PubMed]
  181. Grand′Maison, F.; Yeung, M.; Morrow, S.; Lee, L.; Emond, F.; Ward, B.; Laneuville, P.; Schecter, R. Sequencing of high-efficacy disease-modifying therapies in multiple sclerosis: Perspectives and approaches. Neural Regen. Res. 2018, 13, 1871. [Google Scholar] [CrossRef] [PubMed]
  182. Melamed, E.; Lee, M.W. Multiple Sclerosis and Cancer: The Ying-Yang Effect of Disease Modifying Therapies. Front. Immunol. 2020, 10, 2954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Saposnik, G.; Montalban, X. Therapeutic Inertia in the New Landscape of Multiple Sclerosis Care. Front. Neurol. 2018, 9, 174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Pierpont, T.M.; Limper, C.B.; Richards, K.L. Past, Present, and Future of Rituximab—The World’s First Oncology Monoclonal Antibody Therapy. Front. Oncol. 2018, 8, 163. [Google Scholar] [CrossRef]
  185. Wang, B.; Lyu, H.; Pei, S.; Song, D.; Ni, J.; Liu, B. Cladribine in combination with entinostat synergistically elicits anti-proliferative/anti-survival effects on multiple myeloma cells. Cell Cycle 2018, 17, 985–996. [Google Scholar] [CrossRef]
  186. Wei, C.W.; Yu, Y.L.; Chen, Y.H.; Hung, Y.T.; Yiang, G.T. Anticancer effects of methotrexate in combination with α-tocopherol and α-tocopherol succinate on triple-negative breast cancer. Oncol. Rep. 2019, 41, 2060–2066. [Google Scholar] [CrossRef] [PubMed]
  187. Saidu, N.E.B.; Bretagne, M.; Mansuet, A.L.; Just, P.-A.; Leroy, K.; Cerles, O.; Chouzenoux, S.; Nicco, C.; Damotte, D.; Alifano, M.; et al. Dimethyl Fumarate Is Highly Cytotoxic in Kras Mutated Cancer Cells but Spares Non-Tumorigenic Cells. Oncotarget 2018, 9, 9088–9099. [Google Scholar] [CrossRef] [Green Version]
  188. Rupp, T.; Pelouin, O.; Genest, L.; Legrand, C.; Froget, G.; Castagné, V. Therapeutic potential of Fingolimod in triple negative breast cancer preclinical models. Transl. Oncol. 2021, 14, 100926. [Google Scholar] [CrossRef] [PubMed]
  189. Huang, O.; Zhang, W.; Zhi, Q.; Xue, X.; Liu, H.; Shen, D.; Geng, M.; Xie, Z.; Jiang, M. Featured Article: Teriflunomide, an immunomodulatory drug, exerts anticancer activity in triple negative breast cancer cells. Exp. Biol. Med. 2014, 240, 426–437. [Google Scholar] [CrossRef]
  190. Chiurchiù, V.; Leuti, A.; Maccarrone, M. Cannabinoid Signaling and Neuroinflammatory Diseases: A Melting pot for the Regulation of Brain Immune Responses. J. Neuroimm. Pharmacol. 2015, 10, 268–280. [Google Scholar] [CrossRef] [PubMed]
  191. Bie, B.; Wu, J.; Foss, J.F.; Naguib, M. An overview of the cannabinoid type 2 receptor system and its therapeutic potential. Curr. Opin. Anaesthesiol. 2018, 31, 407–414. [Google Scholar] [CrossRef] [PubMed]
  192. Zou, S.; Kumar, U. Cannabinoid Receptors and the Endocannabinoid System: Signaling and Function in the Central Nervous System. Int. J. Mol. Sci. 2018, 19, 833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Mecha, M.; Carrillo-Salinas, F.J.; Feliú, A.; Mestre, L.; Guaza, C. Perspectives on Cannabis-Based Therapy of Multiple Sclerosis: A Mini-Review. Front. Cell. Neurosci. 2020, 14, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wade, D.T.; Makela, P.; Robson, P.; House, H.; Bateman, C. Do cannabis-based medicinal extracts have general or specific effects on symptoms in multiple sclerosis? A double-blind, randomized, placebo-controlled study on 160 patients. Mult. Scler. 2004, 10, 434–441. [Google Scholar] [CrossRef]
  195. Pryce, G.; Visintin, C.; Ramagopalan, S.V.; Al-Izki, S.; De Faveri, L.E.; Nuamah, R.A.; Mein, C.A.; Montpetit, A.; Hardcastle, A.J.; Kooij, G.; et al. Control of spasticity in a multiple sclerosis model using central nervous system-excluded CB1 cannabinoid receptor agonists. FASEB J. 2013, 28, 117–130. [Google Scholar] [CrossRef] [PubMed]
  196. Laprairie, R.B.; Bagher, A.M.; Kelly, M.E.M.; Denovan-Wright, E.M. Cannabidiol is a negative allosteric modulator of the cannabinoid CB1 receptor. Br. J. Pharmacol. 2015, 172, 4790–4805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Rog, D.J.; Nurmikko, T.J.; Friede, T.; Young, C.A. Randomized, controlled trial of cannabis-based medicine in central pain in multiple sclerosis. Neurology 2005, 65, 812–819. [Google Scholar] [CrossRef] [PubMed]
  198. Rog, D.J.; Nurmikko, T.J.; Young, C.A. Oromucosal delta9-tetrahydrocannabinol/cannabidiol for neuropathic pain associated with multiple sclerosis: An uncontrolled, open-label, 2-year extension trial. Clin. Ther. 2007, 29, 2068–2079. [Google Scholar] [CrossRef] [PubMed]
  199. Maccarrone, M.; Maldonado, R.; Casas, M.; Henze, T.; Centonze, D. Cannabinoids therapeutic use: What is our current understanding following the introduction of THC, THC: CBD oromucosal spray and others? Expert. Rev. Clin. Pharmacol. 2017, 10, 443–455. [Google Scholar] [CrossRef]
  200. Mannucci, C.; Navarra, M.; Calapai, F.; Spagnolo, E.V.; Busardò, F.P.; Cas, R.D.; Ippolito, F.M.; Calapai, G. Neurological Aspects of Medical Use of Cannabidiol. CNS Neurol. Disord. Drug Targets 2017, 16, 541–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Constantinescu, C.S.; Farooqi, N.; O’Brien, K.; Gran, B. Experimental autoimmune encephalomyelitis (EAE) as a model for multiple sclerosis (MS). Br. J. Pharmacol. 2011, 164, 1079–1106. [Google Scholar] [CrossRef] [PubMed]
  202. Burrows, D.J.; McGown, A.; Jain, S.A.; De Felice, M.; Ramesh, T.M.; Sharrack, B.; Majid, A. Animal models of multiple sclerosis: From rodents to zebrafish. Mult. Scler. J. 2018, 25, 306–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Baker, D.; Pryce, G.; Croxford, J.L.; Brown, P.; Pertwee, R.G.; Huffman, J.W.; Layward, L. Cannabinoids control spasticity and tremor in a multiple sclerosis model. Nature 2000, 404, 84–87. [Google Scholar] [CrossRef] [PubMed]
  204. Zhang, M.; Martin, B.R.; Adler, M.W.; Razdan, R.J.; Kong, W.; Ganea, D.; Tuma, R.F. Modulation of Cannabinoid Receptor Activation as a Neuroprotective Strategy for EAE and Stroke. J. Neuroimm. Pharmacol. 2009, 4, 249–259. [Google Scholar] [CrossRef] [PubMed]
  205. Bernal-Chico, A.; Canedo, M.; Manterola, A.; Victoria Sánchez-Gómez, M.; Pérez-Samartín, A.; Rodríguez-Puertas, R.; Matute, C.; Mato, S. Blockade of monoacylglycerol lipase inhibits oligodendrocyte excitotoxicity and prevents demyelination in vivo. Glia 2015, 63, 163–176. [Google Scholar] [CrossRef] [PubMed]
  206. Croxford, A.L.; Kurschus, F.C.; Waisman, A. Mouse models for multiple sclerosis: Historical facts and future implications. Biochim. Biophys. Acta Mol. Basis Dis. 2011, 1812, 177–183. [Google Scholar] [CrossRef] [Green Version]
  207. Hasseldam, H.; Johansen, F.F. Neuroprotection without immunomodulation is not sufficient to reduce first relapse severity in experimental autoimmune encephalomyelitis. Neuroimmunomodulation 2010, 17, 252–264. [Google Scholar] [CrossRef]
  208. Hernández-Torres, G.; Cipriano, M.; Hedén, E.; Björklund, E.; Canales, A.; Zian, D.; Feliú, A.; Mecha, M.; Guaza, C.; Fowler, C.; et al. A reversible and selective inhibitor of monoacylglycerol lipase ameliorates multiple sclerosis. Angew. Chem. Int. Ed. Engl. 2014, 53, 13765–13770. [Google Scholar] [CrossRef]
  209. Ni, X.; Geller, E.B.; Eppihimer, M.J.; Eisenstein, T.K.; Adler, M.W.; Tuma, R.F. Win 55212-2, a cannabinoid receptor agonist, attenuates leukocyte/endothelial interactions in an experimental autoimmune encephalomyelitis model. Mult. Scler. J. 2004, 10, 158–164. [Google Scholar] [CrossRef]
  210. Pryce, G.; Ahmed, Z.; Hankey, D.J.R.; Jackson, S.J.; Croxford, J.L.; Pocock, J.M.; Ledent, C.; Petzold, A.; Thompson, A.J.; Giovannoni, G.; et al. Cannabinoids inhibit neurodegeneration in models of multiple sclerosis. Brain 2003, 126, 2191–2202. [Google Scholar] [CrossRef] [PubMed]
  211. Webb, M.; Luo, L.; Ma, J.Y.; Tham, C. Genetic deletion of Fatty Acid Amide Hydrolase results in improved long-term outcome in chronic autoimmune encephalitis. Neurosci. Lett. 2008, 439, 106–110. [Google Scholar] [CrossRef] [PubMed]
  212. Maresz, K.; Pryce, G.; Ponomarev, E.D.; Marsicano, G.; Croxford, J.L.; Shriver, L.P.; Ledent, C.; Cheng, X.; Carrier, E.J.; Mann, M.K.; et al. Direct suppression of CNS autoimmune inflammation via the cannabinoid receptor CB1 on neurons and CB2 on autoreactive T cells. Nat. Med. 2007, 13, 492–497. [Google Scholar] [CrossRef] [PubMed]
  213. Cabral, G.A.; Griffin-Thomas, L. Cannabinoids as therapeutic agents for ablating neuroinflammatory disease. Endocr. Metab. Immune Disord. Drug Targets 2008, 8, 159–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Walter, L.; Stella, N. Cannabinoids and neuroinflammation. Br. J. Pharmacol. 2004, 141, 775–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Berard, J.L.; Wolak, K.; Fournier, S.; David, S. Characterization of relapsing–remitting and chronic forms of experimental autoimmune encephalomyelitis in C57BL/6 mice. Glia 2010, 58, 434–445. [Google Scholar] [CrossRef] [PubMed]
  216. Centonze, D.; Bari, M.; Rossi, S.; Prosperetti, C.; Furlan, R.; Fezza, F.; DeChiara, V.; Battistini, L.; Bernardi, G.; Bernardini, S.; et al. The endocannabinoid system is dysregulated in multiple sclerosis and in experimental autoimmune encephalomyelitis. Brain 2007, 130, 2543–2553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Shen, M.; Piser, T.M.; Seybold, V.S.; Thayer, S.A. Cannabinoid receptor agonists inhibit glutamatergic synaptic transmission in rat hippocampal cultures. J. Neurosci. 1996, 72, 169–177. [Google Scholar] [CrossRef] [Green Version]
  218. Joseph, D.; Schulze, J. Cannabinoid Activity—Is There a Causal Connection to Spasmolysis in Clinical Studies? Biomolecules 2021, 11, 826. [Google Scholar] [CrossRef]
  219. Molina-Holgado, F.; Molina-Holgado, E.; Guaza, C. The endogenous cannabinoid anandamide potentiates interleukin-6 production by astrocytes infected with Theiler’s murine encephalomyelitis virus by a receptor-mediated pathway. FEBS Lett. 1998, 433, 139–142. [Google Scholar] [CrossRef]
  220. Kozela, E.; Lev, N.; Kaushansky, N.; Eilam, R.; Rimmerman, N.; Levy, R.; Ben-Nun, A.; Juknat, A.; Vogel, Z. Cannabidiol inhibits pathogenic T cells, decreases spinal microglial activation and ameliorates multiple sclerosis-like disease in C57BL/6 mice. Br. J. Pharmacol. 2011, 163, 1507–1519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. González-García, C.; Torres, I.M.; García-Hernández, R.; Campos-Ruíz, L.; Esparragoza, L.R.; Coronado, M.J.; Grande, A.G.; García-Merino, A.; Sánchez López, A.J. Mechanisms of action of cannabidiol in adoptively transferred experimental autoimmune encephalomyelitis. Exp. Neurol. 2017, 298, 57–67. [Google Scholar] [CrossRef] [PubMed]
  222. Giacoppo, S.; Galuppo, M.; Pollastro, F.; Grassi, G.; Bramanti, P.; Mazzon, E. A new formulation of cannabidiol in cream shows therapeutic effects in a mouse model of experimental autoimmune encephalomyelitis. DARU J. Pharm. Sci. 2015, 23, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Lassmann, H.; Bradl, M. Multiple sclerosis: Experimental models and reality. Acta Neuropathol. 2016, 133, 223–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Olson, J.K.; Eagar, T.N.; Miller, S.D. Functional activation of myelin-specific T cells by virus-induced molecular mimicry. J. Immunol. 2002, 169, 2719–2726. [Google Scholar] [CrossRef] [Green Version]
  225. Oleszak, E.L.; Chang, J.R.; Friedman, H.; Katsetos, C.D.; Platsoucas, C.D. Theiler’s Virus Infection: A Model for Multiple Sclerosis. Clin. Microbiol. Rev. 2004, 17, 174–207. [Google Scholar] [CrossRef] [Green Version]
  226. Nagarkatti, P.; Pandey, R.; Rieder, S.A.; Hegde, V.L.; Nagarkatti, M. Cannabinoids as novel anti-inflammatory drugs. Future Med. Chem. 2009, 1, 1333–1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Hernangómez, M.; Mestre, L.; Correa, F.G.; Loría, F.; Mecha, M.; Iñigo, P.M.; Docagne, F.; Williams, R.O.; Borrell, J.; Guaza, C. CD200-CD200R1 interaction contributes to neuroprotective effects of anandamide on experimentally induced inflammation. Glia 2012, 60, 1437–1450. [Google Scholar] [CrossRef] [Green Version]
  228. Loría, F.; Petrosino, S.; Hernangómez, M.; Mestre, L.; Spagnolo, A.; Correa, F.; Di Marzo, V.; Docagne, F.; Guaza, C. An endocannabinoid tone limits excitotoxicity in vitro and in a model of multiple sclerosis. Neurobiol. Dis. 2010, 37, 166–176. [Google Scholar] [CrossRef] [PubMed]
  229. Arevalo-Martin, A.; Vela, J.M.; Molina-Holgado, E.; Borrell, J.; Guaza, C. Therapeutic action of cannabinoids in a murine model of multiple sclerosis. J. Neurosci. 2003, 23, 2511–2516. [Google Scholar] [CrossRef]
  230. Eljaschewitsch, E.; Witting, A.; Mawrin, C.; Lee, T.; Schmidt, P.M.; Wolf, S.; Hoertnagl, H.; Raine, C.S.; Schneider-Stock, R.; Nitsch, R.; et al. The endocannabinoid anandamide protects neurons during CNS inflammation by induction of MKP-1 in microglial cells. Neuron 2006, 49, 67–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Sheng, W.S.; Hu, S.; Min, X.; Lokensgard, G.A.C.J.R.; Peterson, P.K. Synthetic cannabinoid WIN55,212-2 inhibits generation of inflammatory mediators by IL-1beta-stimulated human astrocytes. Glia 2005, 49, 211–219. [Google Scholar] [CrossRef] [PubMed]
  232. Ceña, V.; Sulkowski, G.; Dąbrowska-Bouta, B.; Salińska, E.; Strużyńska, L. Modulation of Glutamate Transport and Receptor Binding by Glutamate Receptor Antagonists in EAE Rat Brain. PLoS ONE 2014, 9, e113954. [Google Scholar] [CrossRef] [Green Version]
Table 1. Involvement of ECS in various pathological aspects of AD.
Table 1. Involvement of ECS in various pathological aspects of AD.
Mechanisms of
AD Pathogenesis
Implications of the ECS in ADCannabinoid Receptor Ligands with Potential Benefits in
Therapeutic Management of AD
Target
Components
Physiological
Function
Disease Model and SpeciesCompoundBIOLOGICAL EFFECT
β-amyloid (Aβ) peptides →neurodegenerative cascade →
neuronal cell death
[35,75,79,80]
CB1R activation
Biology 11 00440 i001
hyperpolarization of the neuronal membrane
modulating of neurotransmitter and cytokine release [79]
Rats
intra-CA1 microinjection
i.p.
intra-NAc
intra-mPFC
ACPA
(agonist of CB1R)
cognitive impairments [52,81]
Biology 11 00440 i002
[35,75,79,80]
↑ CB1R densityneuroprotective and anti-inflammatory response [79]Rats
intra-CA1 microinjection
intra-NAc
intra-BLA
AM251
(antagonist of CB1R)
enhanced memory and learning processes [52,81]
Rat
hippocampus
CBD and Δ9-THC↑ synthesis of tryptophan → improvement of the disease [82]
In vitro
Methods
Molecular docking
Cannabinoids
(CBD, CBD-DMH)
↓ of oxidative stress
↓ of TNF-α
synergistic effect with AchE inhibitors →
  • preventing Aβ formation
  • improvement of circadian rhythm disturbances commonly seen in AD [79,83]
proteolytic cleavage of β-APP →
formation of Aβ1-42 monomers →
activation of astrocytes and microglia →
release of inflammatory cytokines, kinases and nitric oxide →
phosphorylation of tau proteins [84]
Endocannabinoids
(2-AG)
CB1R/CB2R
activation
synthesized by microglia and astrocytes →
suppress cytokine synthesis via CB1/CB2 receptors
Primary hippocampal neuron cell cultures from rat embryosMAGL inhibitors
(URB 602 andJZL 184)
↑ endogenous levels of 2-AG →
↓ of TUNEL-positive neurons
  • anti-apoptotic effects
  • mediated through CB1R [85]
(added space)
GSK-3β promotes tau proteinhyperphosphorylation →
formation of NFT →
impairing the axonal transport →
neuronal atrophy [75]
neuroprotective effects → targeting this pathway with key roles in AD pathogenesisPC12 cells treated with AβCBD
AEA
WIN 55,212–2
attenuation of tau hyperphosphorylation by inhibiting Gsk-3β [75]
increased COX-2 levels → involvement in neuroinflammation [86]2-AG is substrate for COX-2
2-AG suppress elevation of hippocampal COX-2 expression
MAGL inhibitor (URB602) and
nonselective MAGL
inhibitor (ATFMK)
↑ 2-AG levels → suppresses the expression of COX-2
action mediated by CB1R [86]
tau pathology [87]CB2R activationrole in memory processing
its activation → vital for cognitive processes
a depletion or disruption of these receptors in rodents →induces long-lasting memory deficits
RatsAM630 (CB2R antagonist)negative effects such as impaired memory [87]
Rats → impaired memory by administration of okadaic acidJWH-133 (CB2R agonist)reduced spatial memory impairment
reduced neuroinflammation and neurodegeneration [87]
formation of Aβ peptide aggregates in the brain [88]
PPAR- γ involvment in disease management [89]
CB2R activationinvolved in controlling InflammationBeta-amyloid challenged astrocytesCBD ± PPAR- γ
antagonist (MK886 or GW9662)
interacts with the PPAR-γ receptor →
  • attenuate beta-amyloid (Aβ)-induced neuroinflammation
  • promote neurogenesis in the hippocampus [89]
Aβ peptide-induced neurotoxicity, oxidative stress and inflammatory statusCB2R activationenhances immune system response andautophagy pathwayAnalyses of transcriptome of APP/PS1 mice hypocampusCBD,
chronic i.p. injection
(30 days)
improvement of the neuroinflammation and oxidative stress level [90]
neuroinflammatory
mechanism
CB2R activationameliorate the neuroinflammation and cognitive impairments of ADAPP/PS1 miceJWH015improvement of novel object recognition
regulation in microglia-mediated
neuroinflammation [92]
evidence that associates neutrophil-derived myeloperoxidase (MPO) in the pathogenesis of ADCB1R activation Murine model (male mice) induced with focal cerebral ischaemiaΔ9-THC and

SR141716
(CB1-R antagonist)/
AM630 (CB2R antagonist)
CB1R antagonist inhibited the neuroprotective effect of Δ9-THC
CB2R antagonist had no effect
  • reduction of the size of cerebral infarction given by MCA
  • affective, cognitive, sensory and somatic effects, neuroprotective properties [52,75]
mechanism of inhibition of myeloperoxidase independent of the cannabinoid receptorCBD(added space)
the neuroprotective effect CBD was not inhibited by both CB1R and CB2R antagonist → effects, independent of cannabinoid receptors
  • inhibition of MPO activity in neutrophils →anti-inflammatory mechanism
  • pre- and post-ischaemic neuroprotective effect [91]
β-amyloid (Aβ) plaques cause injuries in the pulvinar nucleus → disruption of thalamo-cortical circuits including disturbances in visual attention [93]CB1R
NAPE-PLD
FAAH
located in the thalamus
- pulvinar nucleus (lateral, medial and inferior)
- dorsal lateral geniculate nucleus
physiological connections withprefrontal cortex and amygdala [94,95,96,97]Coronal brain sections from Vervet monkey
  • lesioning of pulvinar nucleus → disturbance in perception of distracting stimuli [98]
  • modulation of visual and spatial perception and processing mechanisms
  • plasticity phenomena involving subcortical visual pathways
  • synchronization of neural activity [99]
(added space)
Patients with ADΔ9-THC
CBD
  • ↓of the lateral pulvinar nucleus ofAD patients [98]
  • regular consumtioninterferes with emotional face recognition [100]
Legend: CA1: dorsal hippocampus; BLA: basolateral amygdala; NAc: nucleus accumbens; PFC: prefrontal cortex; and CBD-DMH: cannabidiol dimethylheptyl; ↓: decrease; ↑ = increase.
Table 2. Implications of ECS in various pathological aspects of PD.
Table 2. Implications of ECS in various pathological aspects of PD.
Mechanisms of
PD Pathogenesis
Implications of the ECS in PDCannabinoid Receptor Ligands with Potential Benefits in
Therapeutic Management of PD
Target
Components
Physiological
Function
Disease Model and SpeciesCompoundBiological Effect
↓ tyrosine-hydroxylase-positive neurons in the substantia nigra pars compacta
down-regulation of CB2Rs in the substantia nigra
CB2Rneuroprotective effect are involved in neuroinflammation [25,105]MPTP-induced mouse model of Parkinson’s diseaseWIN 55,212 –2JW015
(CB2 receptor agonist)
protects neuron loss
reduces MPTP-induced microglial activation
reverses MPTP-associated motor deficits [37]
reduce inflammation in the brain of MPTP-treated mice [25]
side effects of current anti-parkinsonian therapies, especially L(3,4) dihydroxyphenylalamine L-DOPA-induced dyskinesia [108]CB1Rmodulation of neurotransmission and contribution to synaptic plasticity [108]WIN 55,212–2
HU210
protected nigrostriatal dopamine neurons
reduced microglia activation [109]
↓ dopaminergic neurons in the substantia nigra compacta and a significant reduction of striatal dopamine [104]CB2RCB2R-deficient mice showed an exacerbation of PD pathology [25]AM1241 (selective CB2R agonist)regenerated dopaminergic neurons
reversed the decreased CB2R level in the PD mouse brain [104]
currently therapy for PD is symptomatic whose efficacy is limited due to side effectsCB1Rneuroprotective properties against excitotoxicity
and oxidative stress neuroinflammation, which are also
associated with PD [104]
Clinical study on PD patientsCBDchange in patients’ lives possible neuroprotective effects assessed [106]
Table 3. Involvement of ECS in various pathological aspects of HD.
Table 3. Involvement of ECS in various pathological aspects of HD.
Mechanisms of
HD Pathogenesis
Implications of the ECS in HDCannabinoid Receptor Ligands with Potential Benefits in
Therapeutic Management of HD
Target
Components
Physiological
Function
Disease Model and SpeciesCompound/InterventionBiological Effect
mHTT - direct repressive effect on CB1R gene transcription→
Loss of CB1R binding in the striatum→
Reduction of CB1R [130,133]
CB1
Ractivation
CB1R are necessary to counteract neuronal degeneration [133]
Activation of CB1R pathway is associated with a protective effect [135]
CB1R activation protects neurons from NMDA-induced excitotoxicity and inhibits presynaptic release of glutamate [60]
R6/1 transgenic HD mouse modelWIN 55,212-2Antihyperkinetic activity
prevention of motor impairment [147]
Environment enrichmentUpregulation of CB1R binding → behavioral improvement [134]
R6/2 mouse model of HDΔ9-THC↓ of motor coordination deficits
improvement of motor and exploratory behavior
↓ of striatal atrophy and HTT aggregate accumulation [136]
3NP animal model of HDCBDReversibility or attenuation of alterations induced by 3NP [146]
CBGPrevention of striatal neuron death
Improvement of motor deficits
Reduction of inflammatory markers [153]
R6/2 model of HDGAT211
GAT228
GAT229
(positive allosteric modulators)
Improvement measures of health
GAT211 and GAT229 reduced psychoactivity, without tolerance or dependence [137,138,139]
N171-82Q transgenic modelCB1R gene inactivationEarlier and exacerbated motor alternations
Increased striatal aggregation frequency [136]
3NP animal model of HDCB1R are necessary to counteract neuronal degeneration [133]
Rat model of HDExcitotoxicity was increased through striatal injection of quinolinic acidWIN 55,212-2
CBD
Decreased bothglutamate levels and the effect of quinolinic acid on corticostriatal local field potential recordings [148]
Cell culture model of HD with mHTT expressive cellsCBD
Δ8-THC
Δ9-THC
51–84% protection against HTT-induced cell death [155]
Remark: Effects might be independent of CB1R and due to antioxidant mechanisms
Microglial CB2R → induced in HD patients and animal models
CB2R ablation exacerbates microglial activation and accelerates appearance of symptoms [140]
CB2RCB2R
activation → neuroprotective effect in HD models →
control of deleterious microglial activity [140]
Quinolinic-acid lesioned mice model of HDHU-308Reduction of neuronal damage in the striatum by attenuating glial activation [140]
Malonate-lesion rat model of HDAccelerated progression of the HD phenotype
Increased glial activation
Higher sensitivity to striatal neurodegeneration induced by excitotoxic processes [140]
R6/2 mice modelCB2R ablationFaster progression of the disease phenotype
Increased glial activation
Higher sensitivity to striatal neurodegeneration induced by excitotoxic processes [140]
CB1R/CB2R Human studies-patients with HDNabilone
Sativex®
Improvements in chorea
Improvements in the neuropsychiatric index
Trend for improvements in the Unified HD Rating scale motor score, dystonia subscore and behavior score [151]
Human studies-patients with early-onset HDNabilone
Sativex®
Dronabinol
Improvement of dystonia
Quality of life improvement
Behavior improvement [152]
TRPV1 HD rat model with bilateral striatal injection of 3NPAM404
(ECB reuptake inhibitor)
Reduction of hyperkinetic activity and restoration of neurochemical alterations [1,145]
Table 4. Implications of ECS in various pathological aspects of MS; ↓: decrease; ↑ = increase.
Table 4. Implications of ECS in various pathological aspects of MS; ↓: decrease; ↑ = increase.
Mechanisms of MS PathogenesisThe Endocannabinoid System and
Its Implications in MS
Cannabinoid Receptor Ligands with Potential Benefits in Therapeutic Management of MS
Target ComponentsPhysiological FunctionDisease Model and SpeciesCompoundBiological Effect
spasticity →
the mainly observed symptom in MS
is associated with spasms, pain and sleep disturbance [176,177]
CB1R and
CB2R
CB1R inhibits synaptic transmission →
main target for control of spasticity [204]
chronic relapsing EAEΔ9-THC
methanandamide (analogue of AEA)
(CB1R agonists)
WIN 55,212-2 (CB1R/CB2R agonist)
JWH-133
(CB2-R agonist)
amelioration of some motor symptoms such as limb spasticity, tremor and paralysis [203]
inflamamation→
recruitment of leukocytes from the blood into the CNS
adhesion to endothelial cells
(added space)
cerebrospinal fluid: increased glutamate level, differential expression of glutamate receptors
[232]
increased glutamate level → neurodegeneration due to excitotoxicity [210,216]
CB1R and
CB2R
CB2R have immunomodulatory properties [209]EAE induced
C57BL/6 mice immunized with MOG35–55 + pertussis toxin
WIN 55,212-2
SR 141716A
(CB1R antagonist)
SR144528
(CB2R antagonist)
CB1R antagonist → no influence on the protective effect

Biology 11 00440 i003
→ key role in the protective effect of WIN55212-2 [212]

Biology 11 00440 i004
stimulation →
attenuated EAE progression
potential target to inhibit leukocyte trafficking in EAE [209]
CB1Ractivation of cannabinoid receptors inhibits the release of glutamate presynaptically [217]Rat hippocampal neurons cultureAEA
Memantine
Δ9-THC
(added space)
antiglutamatergic effects by
↓ of Mg2+ concentration →↓ excitation level in the entire network of neurons in the culture
glutamatergic excitatory postsynaptic currents elicited by direct stimulation of the presynaptic neuron [217]
EAE induced
C57BL/6 mice—i.p. administration for 3 consecutive days
CBD(added space)
Low dose of CBD →
↓ inflammation → axonal damage
↓ spinal activation of glia
inhibition of T-cell migration in the spinal cord [220]

High dose of CBD →
↓microglial activity↓
↓cell infiltration and demyelination
↓axonal damage
↓levels of IL-6 [221]
CB1R and
CB2R
involvement in treating of neurodegenerative diseases driven by chronic neuro-inflammationEAE-induced
C57BL/6 mice immunized with MOG35–55 +
pertussis toxin
Δ9-THC + CBDΔ9-THC + CBD→attenuates the development of EAE [157]
CB1Rimmunosuppressive effects on astrocytesIn vitro method
TMEV-infected astrocytes
AEAdose-dependent potentiating of IL-6 [219]

inhibition of astrocytes activation →the production of IL-6 [231]

inhibition IL-1β, IL-6, IL-12 and IL-23 release in myeloid dendritic cells

inhibition of microglial activation [161]
Mouse model
TMEV-induced
demyelinating
disease
PEA↓ expression of IL-1, TNF-α↓ microglial activation in the spinal cord of mice [228]
UCM707
WIN 55,212-2
JWH-015
ACEA
↓ microglial activation
inhibition of MHC class II antigen expression
↓ of spinal cord infiltrating CD4T cells–
↓ the production of IL-1β, IL-6 and TNF-α [229,230]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vasincu, A.; Rusu, R.-N.; Ababei, D.-C.; Larion, M.; Bild, W.; Stanciu, G.D.; Solcan, C.; Bild, V. Endocannabinoid Modulation in Neurodegenerative Diseases: In Pursuit of Certainty. Biology 2022, 11, 440. https://doi.org/10.3390/biology11030440

AMA Style

Vasincu A, Rusu R-N, Ababei D-C, Larion M, Bild W, Stanciu GD, Solcan C, Bild V. Endocannabinoid Modulation in Neurodegenerative Diseases: In Pursuit of Certainty. Biology. 2022; 11(3):440. https://doi.org/10.3390/biology11030440

Chicago/Turabian Style

Vasincu, Alexandru, Răzvan-Nicolae Rusu, Daniela-Carmen Ababei, Mădălina Larion, Walther Bild, Gabriela Dumitrița Stanciu, Carmen Solcan, and Veronica Bild. 2022. "Endocannabinoid Modulation in Neurodegenerative Diseases: In Pursuit of Certainty" Biology 11, no. 3: 440. https://doi.org/10.3390/biology11030440

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