Open access peer-reviewed chapter

Cannabinoids and Motor Control of the Basal Ganglia: Therapeutic Potential in Movement Disorders

Written By

Teresa Morera-Herreras, Cristina Miguelez, Asier Aristieta, María Torrecilla, José Ángel Ruiz-Ortega and Luisa Ugedo

Submitted: 06 October 2015 Reviewed: 09 February 2016 Published: 15 June 2016

DOI: 10.5772/62438

From the Edited Volume

Cannabinoids in Health and Disease

Edited by Rosaria Meccariello and Rosanna Chianese

Chapter metrics overview

1,874 Chapter Downloads

View Full Metrics

Abstract

Cannabinoid receptors in the brain appear to be intimately involved in the motor control. Cannabinoid CB1 receptors are densely located in the basal ganglia (BG), a forebrain system that integrates cortical information to coordinate motor activity regulating signals. In fact, the administration of plant-derived, synthetic or endogenous cannabinoids produces several effects on motor function. These effects are paralleled to changes in the levels of different neurotransmitters in the BG, including GABA, dopamine and glutamate, all of which are important players in movement control.

Keywords

  • Basal Ganglia
  • endocannabinoids
  • Parkinson's disease
  • Huntington's disease
  • CB1 receptor
  • CB2 receptor
  • TRPV1

Abbreviation List

9-THC:

9-Tetrahydrocannabinol

9-THCV:

9-Tetrahydrocannabivarin

2-AG:

2-Arachidonoylglycerol

3-NP:

3-nitropropionic acid

6-OHDA:

6-hydroxidopamine

AEA:

Anandamide

BDNF:

Brain-derived neurotrophic factor

BG:

Basal Ganglia

CAG:

Cytosine-adenine-guanine

CBD:

Cannabidiol

CNS:

Central nervous system

ECS:

Endocannabinoid system

FAAH:

Fatty acid amide hydrolase

HD:

Huntington’s disease

HTT:

huntingtin protein

Internal/external globus pallidus:

GPi/GPe

LPS:

Lipopolysaccharide

MAGL:

Monoacylglycerol lipase

MPTP:

1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine

PD:

Parkinson’s disease

PEA:

Palmitoylethanolamide

SNc:

Substantia nigra pars compacta

SNr:

Substantia nigra pars reticulata

SPN:

Spiny projection neurons

STN:

Subthalamic nucleus

TRPV1:

Transient receptor potential vanilloid 1

VGLUT1:

vesicular glutamate transporter 1

VGLUT2:

vesicular glutamate transporter 2

Advertisement

1. Introduction

The identification of more than 60 oxygen-containing aromatic hydrocarbon compounds, known as cannabinoids in the plant Cannabis sativa, led to the discovery of the endocannabinoid system (ECS) [1]. This system consists of two well-characterized seven transmembrane G-protein coupled receptors, CB1 [2] and CB2 [3], the transient receptor potential vanilloid 1 (TRPV1) channel, a number of endogenous ligands and associated enzymes for biosynthesis and degradation (for review see [4,5]). The endocannabinoids are lipids localized in the central nervous system (CNS) and peripheral tissue, being anandamide (AEA) and 2-arachidonoylglycerol (2-AG) the best characterized ones [4]. These endocannabinoids are synthesized, on demand, after neuronal depolarization, released into the extracellular space and after CB receptors activation they are uptaken by a specific transport protein located on both neurons and glial cells to be subsequently degraded by intracellular enzymes [6].

The CB1 receptor is one of the most abundant cannabinoid receptor in the neurons from the CNS whereas the CB2 receptor is more predominant in glial cells and peripheral tissues [7]. CB receptors are coupled to Gi/o proteins, negatively to adenylyl cyclase and positively to mitogen-activated protein kinase and ion channels. Thereby activation of these receptors exerts diverse responses as inhibition of neurotransmitter release including GABA, glutamate, noradrenalin and dopamine [6,8,9], gene transcription and cell proliferation, differentiation and survival [10]. Commonly, the ECS is described as a neuromodulatory system which interacts with other neurotransmitter systems and may be implicated in (patho)physiological functions among others, those related to motor activity, neuron proliferation and inflammatory process. Consequently the ECS appears as a promising target for drug development. Cannabis sativa derivatives have been used medically for thousands of years, however the psychoactive effects together with tolerance and its potential abuse have limited the clinical application. Nowadays numerous efforts are being made to develop non-psychotropic cannabinoids that could be used in several CNS disorders.

This review will focus on recent studies clarifying the role of the ECS as a target to develop new therapeutic tools to treat movement disorders.

Advertisement

2. Neuroanatomical distribution of the endocannabinoid system in the BG

The ECS is highly expressed in the Basal Ganglia (BG), which is a highly organized network of subcortical nuclei composed of the striatum (caudate and putamen), subthalamic nucleus (STN), internal and external globus pallidus (or entopeduncular nucleus in rodents, GPi/EP and GPe) and substantia nigra (pars compacta, SNc, and pars reticulata, SNr). The BG connects the cortex with the thalamus, creating the cortico–basal ganglia–thalamo–cortical loop, and plays a crucial role on controlling movement activity.

High levels of CB1 receptors are found in the striatum and motor cortex where they are mainly present in projecting terminals (reviewed in [11]). Thus, CB1 receptors have been observed in glutamatergic corticostriatal afferences [12,13], striatal projections to the GPi/GPe and to the SNr [14,15] and also in subthalamonigral and subthalamopallidal terminals [15,16]. Moderate to dense CB2 immunoreactivity is also present in the cortex, striatum and SN [17] and is increased in several pathological conditions [18,19]. TRPV1 receptors have also been located in nigrostriatal terminals and on tyrosine hydroxylase-positive cells in the SNc [20,21].

The two most important endocannabinoid-synthesizing enzymes (N-acyl phosphatidylethanolamine phospholipase D and diacylglycerol lipase-alpha) as well as the endocannabinoid degradative enzymes (fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL)) are found in the BG, particularly in the striatum [2224]. Finally, both AEA and 2-AG are expressed within the BG and modulate the activity of the cortico–basal ganglia–thalamo–cortical loop leading to motor activity modulation (for review see [25]).

Advertisement

3. Motor effects of cannabinoids

Over the last three decades a wide range of experimental and clinical studies have demonstrated that the ECS plays a key role controlling motor function. Specifically, three lines of evidence support this idea. First, administration of plant-derived compounds, synthetic cannabinoids and endocannabinoids produces a variety of effects on motor activity in humans and laboratory animals. Second, endocannabinoids and their receptors are abundant in the BG and the cerebellum, two brain areas that exert direct control of motor function. Third, biochemical and functional alterations of the ECS have been related to the etiology of several movement disorders, since overactivity of the ECS signaling is associated with the progression of the nigral degeneration in patients with Parkinson’s disease (PD) [2628] and down-regulation of CB1 receptors has been reported in brain samples collected from patients with Huntington’s disease (HD) [2931].

Marijuana consumption affects psychomotor skills clearly reflected in poor performance of highly demanding tasks and potent impairment of motor coordination [3234], interfering with driving skills and increasing the risk of injuries [35]. In experimental animal models, the motor effects of the endogenous ligand AEA are similar to those induced by the plant-derived cannabinoid agonist Δ9-tetrahydrocannabinol (Δ9-THC), but less pronounced and of shorter duration. The main motor outcomes of cannabinoid agonists are hypoactivity and catalepsy, however biphasic effects have been described depending on the drug and dose used. While relatively low doses of Δ9-THC and AEA have been associated with a transient increase in motor activity, high doses produce motor inhibition and catalepsy [3645]. Additionally, both drugs AEA and Δ9-THC potentiate the catalepsy induced by a local administration of muscimol into the GP [46] and reduce stereotyped behaviors [4042]. Δ9-THC reduces the amphetamine-induced hyperlocomotion [47] and impairs fine motor control in rats [48].

Interestingly, cannabidiol (CBD), other plant-derived cannabinoid, has been proposed as a modulator of the effects induced by Δ9-THC (reviewed in [49]). Several studies conclude that CBD attenuates some Δ9-THC motor side effects such as catalepsy [5052], while others show that high doses of CBD fail to prevent or exacerbate the pharmacological effects of Δ9-THC, including hypoactivity [45,53]. Nevertheless, the influence of CBD on spontaneous locomotion seems to be limited [45,5355] and it may be unrelated to the CB receptors interaction [49].

Preclinical studies using synthetic cannabinoid agents have highlighted the role of the ECS on the control of motor function. Thus, systemic administration of the agonists WIN 55,212-2, CP55,940 and HU210 usually causes inhibition of motor activity [45,5661] and produces basal catalepsy [52,6264]. Activation of CB1 receptors by the agonist CP55,940 enhances the catalepsy induced by dopaminergic antagonists [65] and genetic inactivation of either dopamine D2 receptors or adenosine A2A receptors reduces the motor depression produced by CP55,940 [63]. In fact, the inhibitory motor effects caused by an intrastriatal injection of the CB1 agonist WIN 55,212-2 are mediated by a functional interaction between the adenosine A2A and the CB1 receptors [66], which are present as heteromers in the rat corticostriatal terminals [67]. In line with this, the specific adenosine A2A receptor antagonist MSX-3 blocks the inhibitory effect of the CB1 receptor agonist CP 55,940 on the hyperactivity induced by the dopamine D2 receptor agonist quinpirole [58]. Although, a strong functional interaction between striatal A2A and CB1 receptors is accepted, the mechanisms of this interaction are still not clear since it has been proposed a stimulatory role as well as an inhibitory role of A2A receptors on CB1 receptor-dependent effects [68].

Pharmacological agents that indirectly elevate endocannabinoid levels by inhibiting either their uptake or the intracellular degradation have been proposed as promising therapeutic agents for the treatment of diverse diseases including movement disorders (for review see [69]). These pharmacological agents are also referred as indirect agonists and they also elicit some cannabimimetic motor effects, including hypoactivity. On one hand, the selective and potent inhibition of the endocannabinoids transport by the arachidonic acid derivative, UCM707 potentiates the hypokinetic effects caused by a non-efficacious dose of AEA [70]. Similarly, the selective blockade of the AEA transport by AM404 elevates circulating levels of the endocannabinoid and causes hypoactivity [71,72], likely by activating TRPV1 [73]. On the other hand, selective inhibition of the endocannabinoids metabolizing enzymes has brought interesting results suggesting the differential role of the endogenous agents, AEA and 2-AG, in the unwanted motor effects induced by cannabinoid compounds. Specifically, inhibition of MAGL, but not FAAH, causes CB1 receptor-mediated hypoactivity. However, neither FAAH nor MAGL produce cataleptic effects usually elicited by direct agonists [74,75]. Thus, the FAAH inhibitor URB597 does not significantly alter motor function at doses that induces analgesia [74] and a single or repeated administration of PF-04457845, a specific and irreversible FAAH inhibitor, is well tolerated in healthy humans [76]. Conversely, the MAGL inhibitor JZL184 reduces locomotor activity [77,78], and dual inhibitors JZL195 and AS-57 induce also catalepsy [75,79].

From a general point of view, the motor effects of direct and indirect agonists are usually mediated by stimulation of CB1 receptor since its pharmacological blockade prevents the above mentioned effects [64,72,75,80,81]. In addition, a high dose of the CB1 receptor antagonist SR141716A causes hyperlocomotion by itself [82,83] while a lower dose does not affect motor activity [84]. Also the CB1 receptor antagonist AM251 administrated alone fails to alter basal locomotion but diminishes the psychomotor effects induced by the co-administration of amphetamine [85]. In agreement with these observations, mice lacking CB1 receptors exhibit several motor anomalies [86,87] and show less sensitivity to the psychomotor stimulants and sensitizing effects of the psychostimulants [88,89].

Although these findings provide evidence for the involvement of CB1 receptor-related mechanisms in motor control, other reports demonstrate that also the TRPV1 receptors can mediate effects of certain cannabinoids such as AEA [90]. The potential beneficial effects of cannabinoid agents in movement disorders are further reviewed here.

Advertisement

4. Alterations of the endocannabinoid system in movement disorders

4.1. Endocannabinoid system in Parkinson’s disease

PD is the second most common neurodegenerative disorder, after Alzheimer’s disease. It is a chronic and progressive neurodegenerative illness and its etiology is interpreted as a combination of genetic, environmental, and aging-related processes, although the exact cause of this degeneration is unknown. Clinical manifestation of the disease includes tremor at rest, worsening of voluntary movements, bradykinesia, muscle rigidity and postural instability [91,92]. All of these symptoms appear as a result of pathological processes, including neuroinflammation, mitochondrial dysfunction, oxidative stress, kinase pathways and calcium dysregulation, protein aggregation and prion-like processes, which result in a degeneration of dopamine neurons in the SNc and a subsequent dopaminergic depletion in the striatum, the main input nucleus of the BG neural circuit. [93].

Several components of the ECS are altered in movement disorders which have led to consider it as a possible target for developing new treatments for these disorders. The effects of cannabinoid compounds and their potential utility in PD has been tested, but inconsistent data have been produced, as there are many complex responses elicited by dopamine and its interaction with different cannabinoid mechanisms [94].

4.1.1. Endocannabinoid levels

The ECS in the BG becomes overactivated in PD. For instance, the cerebrospinal fluid of untreated patients with PD has at least two-fold higher levels of AEA, compared to controls, being this increase independent of either disease stage or progression [27]. Overall, results from preclinical studies support the clinical findings.

In 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-lesioned monkeys, the levels of both AEA and 2-AG were determined throughout the BG after induction of parkinsonism and L-DOPA treatment. Both AEA and 2-AG levels were elevated in the striatum of untreated parkinsonian monkeys and returned to those found in the striatum of normal monkeys after treatment with L-DOPA [95]. These results are similar to those reported in rodent models of PD, where AEA levels were estimated to be higher in the BG output regions, as compared to the striatum or GPe, while 2-AG levels were similar throughout the BG regions analyzed [96]. Similarly, in a rat model of PD induced by unilateral nigral lesion with 6-hydroxidopamine (6-OHDA), striatal levels of AEA, but not 2-AG levels, were increased. Moreover, decreased activity of the AEA membrane transporter and FAAH was observed, whereas the binding of AEA to CB receptors was unaffected [97]. In studies with reserpine-treated rats an increase of 2-AG levels in output BG regions, such as GP and SNr has been reported compared to other brain regions like the cortex, cerebellum, hippocampus, and striatum [98]. In mutant models of the disease, as the homozygous Pink1 knockout mice, the levels of 2-AG and AEA were not altered in striatum, and no change was observed as for the activity of FAAH and MAGL, responsible for endocannabinoids hydrolysis. Although the levels of 2-AG remains unchanged in PD, this endocannabinoid has revealed its neuroprotective role in the MPTP mouse model [99] as well as palmitoylethanolamide (PEA) [100].

4.1.2. Cannabinoid receptors

CB1 availability is unevenly modified in patients with PD, which is decreased in the SN but relatively increased in nigrostriatal, mesolimbic, and mesocortical dopaminergic projection areas [101]. In post-mortem samples from patients with PD increased CB1 receptor binding in the caudate-putamen was found and a higher stimulation by CB1 receptor agonists of [35S]GTPγS binding was measured in the same brain areas [26]. Other studies performed in PD human brain have shown that the expression of CB1 receptor mRNA was decreased in the caudate nucleus, anterior dorsal putamen and GPe, but remained unchanged in the other brain areas examined [102]. When the relationship between CB1 receptors and dopamine D2/D3 receptor densities was analyzed by receptor autoradiography, unchanged CB1 receptor density in the putamen and caudate of patients with PD (treated with L-DOPA) was found together with a parallel decreased in the density levels of dopamine D2/D3 receptors in the same nuclei [103].

Studies in animal models also show disparity of results concerning CB1 receptors. In MPTP-lesioned marmosets, increase in CB1 receptor binding was also confirmed in the caudate-putamen [26]. In contrast, in the reserpine-treated rat model of PD a topographically organized reduction in CB1 receptor mRNA expression in the striatum was found [98]. However, in other studies with 6-OHDA-lesioned rats, no significant change in CB1 receptor expression or binding was detected, although an increase in CB1 receptor mRNA levels in the striatum was found [104,105]. In Pink1 knockout mice, however, a significant reduction of binding ability of CB1 receptor agonists was found [106].

Regarding CB2 receptors, in neurotoxin rat models of PD, when 6-OHDA or lipopolysaccharide (LPS) is injected into the striatum, increased expression of the CB2 receptor and proportionally increased microglial activation was found [107]. Other CB receptors may have a role in PD treatment, thus the pharmacological blockade of TRPV1 receptors, which have been implicated in the modulation of dopamine transmission in the BG, attenuates the 6-OHDA induced hypokinesia [108]. TRPV1 receptors, where AEA binds, may play opposite roles to CB1 receptors in order to treat L-DOPA-induced dyskinesia [94,109]. Although it seem that the ECS suffers dramatic modifications in PD, so far clinical and preclinical studies show no conclusive results.

4.2. Endocannabinoid system in Huntington’s disease

HD is a rare autosomal-dominant progressive neurodegenerative disease that affects 4–10 per 100,000 people in Western countries [110]. It is caused by the mutation in the IT15 gene (HTT) that encodes the huntingtin protein(HTT) and consists in an expanded cytosine-adenine-guanine (CAG) repeat that yields an abnormally long polyglutamine sequence in HTT [111]. Healthy subjects have less than 35 CAG repeats, while genes with more than 39–40 CAG repeats (or more) encode pathogenic HTT. Although physiological HTT is widely found in the brain, the mutation leads to the expression of the aberrant protein harmful only to specific neuron populations. Indeed, the prominent degeneration of GABAergic spiny projecting neurons (SPN) in the striatum [112] and to a lesser extent, of glutamatergic cortical pyramidal neurons [113,114] are the hallmarks of the disease. In late stages of the disease other brain areas and neurons may also be affected, producing a complex disease. The clinical manifestation can appear early, at the ages of 30–50 years, and progressively worsen over the following years. On the course of the disease, the patient can suffer from movement disorders, cognitive disability and psychiatric symptoms such as personality changes, depression or memory loss, among others [115]. The early stages of the disease are characterized by choreic involuntary movements and cognitive impairment, which are related to mutant HTT aggregation and aberrant neurotransmission in the striatopallidal SPN and corticostriatal neurons, respectively (reviewed in [116]). In more advanced phases of the disease, when the direct pathway in the BG is also affected, the patients display parkinsonian-like symptoms, as bradykinesia and rigidity [117].

4.2.1. Endocannabinoid levels

As mentioned for PD, in HD The ECS also suffers notorious modifications on the course of the disease in the BG [30,31,118]. In post-mortem tissue, higher expression of FAAH was observed in the caudate-putamen from HD brains [119] while the opposite effect was found in the cortex [120].

The activity of the ECS has also been investigated in animal models of HD without consistent results, probably due to the different models and ages selected for the studies. In the R6/1 transgenic model of HD, significant decrease in AEA levels were detected in the hippocampus in contrast to increased 2-AG concentrations only found in the cortex at 12 weeks of age, prior to the motor disturbances [121]. Another study in R6/2 transgenic mice, which differ in CAG repeats number from R6/1 mice and show severe motor and cognitive defects at earlier ages, also reported decreased in cortical 2-AG during motor presymptomatic stages [122]. In symptomatic R6/2 mice, AEA, 2-AG and PEA were markedly decreased in the striatum and in a lesser extent in the hippocampus or cortex [122]. Other study performed also in R6/2 mice found increased 2-AG but unchanged AEA levels in the striatum [123]. As for the enzymes’ activity, Bari et al. found increased FAAH and decreased MAGL activity in the cortex together with reduced activity of AEA and 2-AG metabolic-enzymes at different ages in the striatum, which disagrees with the higher expression of FAAH reported by others [119]. The use of other animal models, such us the 3-nitropropionic acid (3-NP)-lesion rat, showed decreased AEA and 2-AG in the striatum but suggested increased AEA levels in the SN [124].

So far studies regarding enzyme activity and endocannabinoid levels in HD are limited and offer inconsistent results although overall endocannabinoid levels seem to be reduced in different brain areas. More homogeneous preclinical studies together with clinical investigations will be needed for clarifying these founds.

4.2.2. Cannabinoid receptors

Autoradiographical and immunohistochemical studies in human post-mortem tissue demonstrated that early stages of the disease are characterized by a loss of CB1, dopamine D2 and adenosine A2A receptors in the caudate nucleus, putamen and GPe, which corresponds to the deterioration of striatopallidal SPNs [30,118,125]. Later in the disease, when the striatonigral pathway also degenerates, even greater down-regulation of striatal and nigral CB1 receptors is observed [30]. In vivo imaging of CB1 in patients confirmed the down-expression of these receptors in cortical and subcortical structures, starting early in the development of the disease [126]. Despite the variety of animal models available for studying the disease, most preclinical studies back up the observations from human tissue. In transgenic mice, such as the HD94 or R6/2 transgenic models, loss of mRNA levels, binding, receptor expression and activation of GTP-binding proteins for cannabinoid CB1 receptors have been reported primarily in SPNs belonging to the indirect pathway in the early phases of the disease accompanied or not by cortical and hippocampal alterations [121,123,125,127129]. In models induced by the administration of 3-NP, where striatal neuronal loss is produced, additional loss of CB1 receptors is also observed [124,130,131]. Small animal PET studies performed in transgenic HD rat and mouse models, have also confirmed reduced CB1 receptor binding in the striatum and GPe in all stages, as well as alterations in the cortex, hippocampus, thalamic nuclei or cerebellum in more advanced phases of the disease [132,133]. Results from human brain and most animal models coincide in reporting down-regulation of striatal CB1 receptors in HD, which has been proposed to be induced by mutant HTT that controls gene promoter activity via repressor element 1 silencing transcription factor [119]. In this respect, genetic CB1 receptor deletion has been proven to worsen motor symptomatology and to exacerbate striatal degeneration by increasing excitotoxic damage and decreasing brain-derived neurotrophic factors [119,134,135]. Although rescuing CB1 receptors prevents the striatal morphological modifications observed in mice models of HD, it failed to improve motor deterioration [136].

Regarding CB2 receptors in the HD brain, studies performed in post-mortem human tissue suggest up-regulation of CB2 receptors on CD68-positive microglial cells [137] or slight increase on brain blood vessels [138] but not on astrocytic cells. These results agree in part with preclinical studies where up-regulation of CB2 receptors has been observed in the striatal microglial of R6/2 mice and striatal microglia and astrocytes in a rat model of HD induced by intrastriatal injection of malonate (mitochondrial complex II inhibitor) [137,139]. This compensatory increase of CB2 receptors may have a neuroprotective effect on the disease [137]. In this line, CB2 agonists have been proposed to decrease striatal neurodegeneration in malonate-lesioned rats [139]. In addition, CB2 depletion has been documented to accelerate the onset of the disease and aggravate the motor disabilities in a mouse model of HD [137,140].

Less attention has been paid to the TRPV1 channels, which also play an important role in the ECS. So far, only one publication has studied these receptors reporting no modification in TRPV1 binding in R6/2 mice at any age [123].

In general, preclinical and clinical studies have well characterized a down-regulation of CB1 receptors at different stages of HD, which may cause a role in the progression of the disease. Less is known about CB2 or TRPV1 receptors and further investigations are required to clarify their role in the neuropathology of the disease.

Advertisement

5. Potential therapeutic role of cannabinoids in movement disorders

5.1. Parkinson’s disease

Currently, pharmacological treatment for PD is focused on dopaminergic replacement, which controls the symptoms in both early and advanced stages [141]. However, its chronic use is limited due to the development of severe and disabling side effects, such as motor fluctuations or L-DOPA-induced dyskinesia. Besides, none of the available treatments are capable of slowing or stopping the disease progression. Cannabinoid-based therapies have been proposed as a promising approach for PD treatment, not only for their antiparkinsonian properties, but also for the neuroprotective and anti-inflammatory effects of these compounds.

5.1.1. Cannabinoids for the treatment of motor symptoms

Studies in animal models and patients with PD have indicated that both CB1 agonists and antagonists used alone or as coadjuvants, could be useful to treat different symptoms of this movement disorder. Thus, CB1 agonists have been shown to improve motor impairment in animal models of PD [142145] and to reduce tremor associated with the hyperactivity of the STN [145,146]. However, because of the hypokinetic profile of the cannabinoid agonists, it is unlikely that these drugs would be useful for alleviating bradykinesia in PD patients. In fact, the administration of agonists to humans or MTPT-lesioned primates enhanced motor disability (for review [34]).

On the other hand, behavioral changes in parkinsonian rodents also improved with the administration of CB1 antagonists [147149]. It has been suggested that blocking CB1 receptors could be useful in particular conditions, as when the patients do not respond to dopaminergic therapy or in advanced phases of the disease [147,148,150]. It is also important to consider the therapeutic benefits of TRPV1 receptor antagonists, given their role in regulating dopamine release from nigral neurons [151].

Cannabinoids also may be beneficial in treating L-DOPA-induced dyskinesia. Indeed, administration of CB1 agonists to parkinsonian rats chronically treated with L-DOPA reduced the occurrence of dyskinesia [109,152,153] without reducing the efficacy of L-DOPA to improve motor performance. In MPTP-lesioned monkeys and patients with PD the results are mixed. Indeed, both CB1 agonist and antagonists showed antidyskinetic effect [95,154157] while plant-derived cannabinoids failed to improve motor disability [158,159].

5.1.2. Cannabinoids as neuroprotective agents

In vivo as well as in vitro preclinical studies in animal models of PD have revealed that cannabinoid receptor modulation can be potentially useful for protecting dopaminergic neurons from progressive neurodegeneration. Although CB1 receptor-mediated effects cannot be excluded, some authors argue that CB1 receptors may have a minimal implication in neuroprotection [10,160162]. It seems plausible that neuroprotection is principally mediated by the antioxidant properties of the cannabinoids [160,163,164], given that oxidative stress is a major hallmark in the pathogenesis of PD. The effect of numerous phytocannabinoids, as ∆9-THC, CBD and ∆9-Tetrahydrocannabivarin (∆9-THCV), has been investigated in experimental models of PD (6-OHDA-lesioned rodents, MPTP or LPS-lesioned mice) [160,163,165]. These studies have showed that cannabinoids could be useful for developing novel neuroprotective therapies in PD due to their CB receptors-independent antioxidant actions. Double-blind trial carried out in patients with PD tried to evaluate the neuroprotective effect of CBD administration but definitive conclusions were not established [166].

Neuroprotection has also been provided by synthetic cannabinoids such as the endocannabinoid transporter inhibitor/vanilloid agonist AM404, or the CB1/CB2 receptors agonist CP55,940, which also produce antioxidant effects via cannabinoid receptor-independent mechanisms [163,167]. By contrast, selective CB1 or CB2 agonists failed to protect these neurons [163].

5.1.3. Cannabinoids as anti-inflammatory agents

Substantial evidence supports that inflammation plays a pivotal role in the death of neurons in the BG in PD. Activated microglia has been found in post-mortem brains from patients with PD [168] and elevated levels of inflammatory cytokines have been measured in the cerebrospinal fluid of patients [169171]. Cannabinoid compounds have demonstrated anti-inflammatory properties. Concretely, CB2 receptor agonists are the most promising cannabinoids to treat the inflammatory processes related to neurodegenerative disorders, although other cannabinoid agonists have also showed anti-inflammatory activity [172]. In fact, the selective CB1 receptor agonist arachidonoyl-2-chloroethylamide and other non-selective cannabinoid agonists have revealed strong anti-inflammatory properties [173].

CB2 receptor is expressed on microglia and it is strongly up-regulated when these cells are activated [174]. In vitro activation of microglial CB2 receptors leads to suppression of the release of pro-inflammatory cytokines reducing neurotoxicity [175]. The beneficial effect of CB2 receptor stimulation has also been proved in animal models of PD [161,165]. Moreover, an increase of sensitivity to the neurotoxin LPS in CB2 receptor knockout mice has been described [165] and the overexpression of this receptor subtype reduced the recruitment of glial cells to the lesion and decreased the level of various oxidative parameters [176].

Apart from the CB2 receptor, several studies have corroborated the anti-inflammatory potential of targeting CB1 receptor in PD. WIN 55,212-2 or HU 210 administration decreases the number of activated microglia and reduces the mRNA levels of the proinflammatory cytokine IL-6 in MPTP mice and LPS-treated rats [162,177,178]. In another study, 2 week pre-treatment with ∆9-THC and CBD, followed by 6-OHDA injection, decreased the loss of dopaminergic neurons in hemiparkinsonian rats [160].

All these studies suggest that the CB1 and CB2 receptors could be potential targets for anti-inflammatory intervention in PD.

5.2. Huntington’s disease

To date, the treatment of HD is merely symptomatic and focuses on the restoration of the neurotransmitter imbalance that occurs in the disease. For this reason, GABA agonists, dopamine depleting agents, neuroleptics, anti-glutamatergic agents, antidepressants or acetylcholinesterase inhibitors are often used for ameliorating the symptoms (for review see [179]), despite not being able to cure or stop the progression of the disease. In this situation cannabinoid compounds may play a promising therapeutic role as it has been proven in other neurodegenerative diseases [180,181]. Indeed, activation of cannabinoid receptors facilitates the activation of intracellular mechanisms related to cell homeostasis, repair and survival. Moreover, several pre-clinical and clinical studies have showed that cannabinoids can be useful in HD. The effects of cannabinoid compounds on HD can be divided in three different actions: improvement of motor symptoms, neuroprotection and anti-inflammatory activity.

5.2.1. Cannabinoids for the treatment of motor symptoms

As mentioned above, the most characteristic feature of HD is the motor impairment which varies along the progression of the disease. In this context, cannabinoid compounds have been tested for treating these motor symptoms in animal models of HD, some of them showing anti-hyperkinetic activity. The administration of the endocannabinoid re-uptake inhibitor, AM404, reduced hyperkinetic movements and restored the neurochemical alterations (dopamine and GABA reduction) in the HD rat model with bilateral striatal injection of 3-NP [130]. However, the anti-hyperkinetic effect showed by AM404 seems to be mainly consequence of TRPV1 receptor activation, and not CB1-dependent [130]. In the same rat model of HD, the administration of UCM707, another endocannabinoid uptake inhibitor, also showed anti-hyperkinetic activity that may be due to the recovery of GABAergic and glutamatergic function in the GP and SN, respectively [182]. In addition, Arvanil, a hybrid endocannabinoid and vanilloid compound, has also demonstrated anti-hyperkinetic activity, although it could not restore the neurochemical alterations in the 3-NP rat model of HD [183]. Similarly, the CB1 receptor agonist CP55,940 has showed anti-hyperkinetic activity, without changing dopamine and GABA levels [184]. Other compounds, VDM11 (endocannabinoid re-uptake inhibitor) and AM374 (endocannabinoid hydrolysis inhibitor), have not demonstrated any significant effect on hyperkinetic movements in the same HD model [184]. The chronic, but not the acute administration of the CB1 receptor agonist WIN 55,212-2 prevented the appearance of motor impairment in the R6/1 transgenic mice model of HD, without improving the social and cognitive deficits [185]. So far, the clinical use of plant-derived cannabinoids and synthetic analogues has shown disappointing results in the treatment of the motor symptoms. A case report and a pilot study of nabilone in patients with HD showed that the effect of this drug on HD motor symptoms is limited [186,187]. Other studies have demonstrated that the administration of nabilone or CBD did not improve the hyperkinetic alterations associated with HD [188,189]. The lack of therapeutic efficacy may lie on the mechanism of action of the cannabinoid agonists used. These cannabinoid drugs were unable to activate TRPV1 receptors, which have been proven to mediate anti-hyperkinetic in 3-NP-lesioned rats. Thus, the most appropriate compounds for future clinical trials, at least in the early stages of the disease, may be those able to activate both CB1 and TRPV1 receptors [190].

5.2.2. Cannabinoids as neuroprotective agents

The neuroprotective properties of the cannabinoids make them very interesting tools to stop the progression of HD. In fact, cannabinoid compounds may reduce cytotoxic processes that occur in neurons and glial cells in neurodegenerative diseases [181]. The protective effects are consequence of the intracellular signaling pathways activated by cannabinoid compounds [116], and are mediated by the activation of CB1, CB2 or TRPV1 receptors, but also due to CB receptor independent processes [172].

Cannabinoid CB1 receptors are located in neuronal glutamatergic terminals at presynaptic and postsynaptic levels. The activation of these receptors by cannabinoid agonists decreases glutamate release and limits the glutamatergic excitotoxicity in neurodegenerative processes [172]. Furthermore, CB1 receptor stimulation activates the phosphatidylinositol 3-kinase/Akt/mammalian target of rapamycin complex 1 pathway, protecting cells from excitotoxic damage, and facilitating brain-derived neurotrophic factor (BDNF) release [191]. The activation of CB1 receptors also diminishes the activity of voltage-sensitive calcium channels [192], reducing calcium dependent damaging pathways. Moreover, cannabinoid compounds activating CB1 receptors are also involved in some other mechanisms related to cell protection and survival, such as GABAergic signaling [193,194] or blood supply to lesioned brain areas [195]. On the other hand, CB2 receptors are mainly located in glial cells and their activation also supports cannabinoids-mediated neuroprotection. In fact, CB2 receptor activation reduces cytotoxic factors release and generation of reactive oxygen and nitric oxide-derived substances [10]. However, cannabinoids also induce CB receptor independent neuroprotective mechanisms. Some cannabinoid compounds can act on NMDA glutamatergic receptors, reducing high glutamate levels at postsynaptic levels and excitotoxicity. Moreover, cannabinoids restore the balance between oxidative and antioxidant mechanisms, mainly, by blocking reactive oxygen substances [196] but also facilitating endogenous antioxidant activity [163,197].

Several pre-clinical studies have been performed to clarify the role of cannabinoids in CB1 receptors mediated neuroprotection in HD. In PC12 cells expressing HTT, CB1 receptor stimulation showed different effects. Activation of CB1 receptors, coupled to Gi/o, induced cell protection by inhibiting cAMP and ERK phosphorylation. However, CB1 receptors could also couple to Gs, which stimulated cAMP and favored cell death in this HD model [198]. Similar results were also obtained in the in vitro model of striatal SPNs expressing wild-type (STHdhQ7/Q7) or mutant HTT (STHdhQ111/Q111) [199]. Studies in rodent models of HD have showed that cannabinoid compounds acting on CB1 receptors can reduce or delay the neurodegeneration associated with this disease [116]. The striatal injection of quinolinic acid in rats increases the glutamatergic transmission, and therefore, the excitotoxicity, mimicking some of the characteristic features of HD [200]. In this rat model of HD, the administration of the CB1 receptor agonist WIN 55,212-2 diminished the increment of glutamate levels induced by the quinolinic acid. In addition, WIN 55,212-2 also reduced the effect of quinolinic acid on corticostriatal local field potential recordings in vitro. The effects of WIN 55,212-2 were blocked by the CB1 receptor antagonist AM 251, demonstrating that the effects observed were CB1-dependent [200]. Thus, WIN 55,212-2 protected the striatum from the damage induced by quinolinic acid, although, this CB1 receptor agonist could not modify the motor abnormalities induced by quinolinic acid [200]. In R6/2 mice, the administration of the ∆9-THC improved the symptoms, the neuropathology and the molecular pathology related to this HD model, and induced BDNF expression in the striatum [119]. Moreover, the depletion of CB1 receptors in these transgenic mice impaired all the pathological features expressed by this HD model, and reduced BDNF expression [119]. Additionally, in R6/2 mice, the striatal injection of a recombinant adeno-associated viral vector encoding CB1 receptor induced the expression of CB1 receptors and BDNF and normalized the molecular pathological signs observed in this well established transgenic mice model of HD [191]. Interestingly, genetic rescue of CB1 receptors in the SPNs prevented the loss of vesicular glutamate transporter 1 and 2 (VGLUT1 and VGLUT2) and synaptophysin in the striatum, although it did not improve the motor phenotype expressed by R6/2 mice [136]. Moreover, enriched environment delayed the onset of motor alterations and the loss of CB1 receptors in R6/1 mice, slowing the progression of the disease [201,202]. All these studies support the role of CB1 receptors and CB1 agonists as useful therapeutic tools to reduce or delay the progression of HD. However, not all the studies have showed that CB1 receptor agonists induce protective effects. Indeed, chronic administration of HU210, ∆9-THC or URB597 did not modify the progressive impairment of motor activity in the transgenic R6/1 mice model of HD [203]. Additionally, in the malonate-lesioned rat model of HD, the administration of UCM707 did not delay the degeneration observed in this model [182]. Thus, further studies are needed to investigate and clarify the implication of CB1 receptors in HD and the potential therapeutic effects of CB1 receptor agonists.

Cannabinoid compounds that activate CB2 receptors have showed neuroprotective effects in HD [139]. The administration of the CB2 receptor agonist HU-308 in the quinolinic acid-lesioned mice attenuated glial activation and reduced the neuronal damage in the striatum [137]. Likewise, in the malonate-lesioned rat model of HD, the administration of the selective CB2 receptor agonist HU-308 protected striatal neurons from the apoptotic mechanisms activated by malonate [139]. This neuroprotection was blocked by the administration of the selective CB2 receptor antagonist SR144528, ratifying that the effect observed was CB2-dependent [139]. The neuroprotective effects of CB2 receptor agonists are associated with the reduction of the toxicity caused by reactive microglial cells [116,137]. The activation of CB2 receptors can also induce the expression of pro-survival substances [10,204]. In contrast, CB2 receptor-deficient R6/2 mice showed accelerated progression of the HD phenotype expressed by this model [137]. In fact, the ablation of CB2 receptors in these mice increased the glial activation and the sensitivity to striatal neurodegeneration induced by excitotoxic processes [137]. Thus, both in genetic and in toxin-lesioned rodent models of HD cannabinoids displayed CB2 receptor mediated protective effects [137,139].

Finally, cannabinoid compounds may also exert neuroprotective effects independent of CB1 or CB2 receptor activation in animal models of HD. This CB receptor independent protection displayed by some cannabinoids (i.e. ∆9-THC or CBC) is related to the blockage of reactive oxygen molecules [172]. The capacity to block reactive oxygen molecules may be due to the phenolic structures of these cannabinoid compounds. Indeed, in 3-NP-lesioned rat model of HD, the administration of CBC reduced the striatal atrophy produced by this toxin [197]. Thus, cannabinoid compounds may be useful to protect cells from the cytotoxicity associated with oxidative processes.

As mentioned above, clinical studies have been developed to examine the effect of cannabinoids on HD. However, most of these studies were designed to analyze the effect of specific cannabinoid compounds on specific symptoms, and not to study the neuroprotective effect of cannabinoids [116]. Moreover, results obtained from animal models of HD have showed that combination of cannabinoids can be more useful to protect neurons than single cannabinoids [116]. In fact, Sativex®, a combination of ∆9-THC and CBC, has showed protective effect on striatal neurons in some rodent models of HD [205,206]. A clinical trial in patients with HD concluded that Sativex® was safe and well tolerated, but it was not able to stop or slow the progression of the neurodegeneration (see https://clinicaltrials.gov/ct2/show/NCT01502046).

5.2.3. Cannabinoids as anti-inflammatory agents

As mentioned before, CB2 receptors are mainly located in glial cells, and are poorly expressed in the striatum in healthy condition. However, the striatal expression of these receptors is increased in patients with and animal models of HD. In HD, the reactive microglia may release inflammatory cytokines, reactive oxygen substances or nitric oxide [10,207,208]. Cannabinoid compounds activating CB2 receptors in glial cells can reduce the release of these factors, and promote the release of some anti-inflammatory cytokines (i.e. IL-10, IL-1ra) [209,210]. Moreover, it has been demonstrated that CB2 receptor agonists improve striatal inflammation in different rodent models of HD [137,139]. Interestingly, CB2 receptors are not involved in the psychotrophic effects induced by cannabinoid compounds. This property makes CB2 receptor agonists valuable compounds for future therapeutic approaches in HD.

Advertisement

6. Concluding remarks

To date, vast number of preclinical evidence have demonstrated that the ECS controls the motor activity in both physiological and pathological states. Cannabinoid compounds have been proven to ameliorate motor symptoms and drug-induced side effects. In addition, the anti-inflammatory and neuroprotective properties of these agents make them promising drugs to delay the progression of neurodegenerative diseases. For these reasons, further basic and clinical cannabis-based research could bring a new light into the treatment of movement disorders, such as PD and HD.

References

  1. 1. Mechoulam R, Shani A, Edery H, GrunfeldY. Chemical basis of hashish activity. Science. 1970;169(3945):611-2. doi: 10.1126/science.169.3945.611.
  2. 2. Matsuda LA, Lolait SJ, Brownstein MJ, Young AC, Bonner TI. Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature. 1990;346(6284):561-4. doi: 10.1038/346561a0.
  3. 3. Munro S, Thomas KL, Abu-Shaar M. Molecular characterization of a peripheral receptor for cannabinoids. Nature. 1993;365(6441):61-5. doi: 10.1038/365061a0.
  4. 4. Lu HC, Mackie K. An introduction to the endogenous cannabinoid system. Biol Psychiatry. 2015. doi: 10.1016/j.biopsych.2015.07.028.
  5. 5. Cacciola G, Chianese R, Chioccarelli T, Ciaramella V, Fasano S, Pierantoni R, et al. Cannabinoids and reproduction: a lasting and intriguing history. Pharmaceuticals. 2010;3(10):3275. doi: 10.3390/ph3103275.
  6. 6. Wilson RI, Nicoll RA. Endocannabinoid signaling in the brain. Science. 2002;296(5568):678-82. doi: 10.1126/science.1063545.
  7. 7. Svizenska I, Dubovy P, Sulcova A. Cannabinoid receptors 1 and 2 (CB1 and CB2), their distribution, ligands and functional involvement in nervous system structures–a short review. Pharmacol Biochem Behav. 2008;90(4):501-11. doi: 10.1016/j.pbb.2008.05.010.
  8. 8. Schlicker E, Timm J, Zentner J, Gothert M. Cannabinoid CB1 receptor-mediated inhibition of noradrenaline release in the human and guinea-pig hippocampus. Naunyn Schmiedebergs Arch Pharmacol. 1997;356(5):583-9. doi: 10.1007/PL00005155
  9. 9. Kuepper R, Ceccarini J, Lataster J, van Os J, van Kroonenburgh M, van Gerven JM, et al. Delta-9-tetrahydrocannabinol-induced dopamine release as a function of psychosis risk: 18F-fallypride positron emission tomography study. PLoS One. 2013;8(7):e70378. doi: 10.1371/journal.pone.0070378.
  10. 10. Fernandez-Ruiz J, Romero J, Velasco G, Tolon RM, Ramos JA, Guzman M. Cannabinoid CB2 receptor: a new target for controlling neural cell survival? Trends Pharmacol Sci. 2007;28(1):39-45. doi: 10.1016/j.tips.2006.11.001.
  11. 11. Hu SS, Mackie K. Distribution of the endocannabinoid system in the central nervous system. Handb Exp Pharmacol. 2015;231:59-93. doi: 10.1007/978-3-319-20825-1_3.
  12. 12. Gerdeman G, Lovinger DM. CB1 cannabinoid receptor inhibits synaptic release of glutamate in rat dorsolateral striatum. J Neurophysiol. 2001;85(1):468-71. doi: 10.1016/j.bbr.2014.05.029.
  13. 13. Kofalvi A, Rodrigues RJ, Ledent C, Mackie K, Vizi ES, Cunha RA, et al. Involvement of cannabinoid receptors in the regulation of neurotransmitter release in the rodent striatum: a combined immunochemical and pharmacological analysis. J Neurosci. 2005;25(11):2874-84. doi: 10.1523/JNEUROSCI.4232-04.2005.
  14. 14. Herkenham M, Lynn AB, de Costa BR, Richfield EK. Neuronal localization of cannabinoid receptors in the basal ganglia of the rat. Brain Res. 1991;547(2):267-74. doi: 10.1001/archpsyc.58.4.334.
  15. 15. Tsou K, Brown S, Sanudo-Pena MC, Mackie K, Walker JM. Immunohistochemical distribution of cannabinoid CB1 receptors in the rat central nervous system. Neuroscience. 1998;83(2):393-411. doi: 10.1016/S0306-4522(97)00436-3.
  16. 16. Mailleux P, Vanderhaeghen JJ. Localization of cannabinoid receptor in the human developing and adult basal ganglia. Higher levels in the striatonigral neurons. Neurosci Lett. 1992;148(1-2):173-6. doi: 10.1016/0304-3940(92)90832-R.
  17. 17. Gong JP, Onaivi ES, Ishiguro H, Liu QR, Tagliaferro PA, Brusco A, et al. Cannabinoid CB2 receptors: immunohistochemical localization in rat brain. Brain Res. 2006;1071(1):10-23. doi: 10.1016/j.brainres.2005.11.035.
  18. 18. Onaivi ES, Ishiguro H, Gong JP, Patel S, Perchuk A, Meozzi PA, et al. Discovery of the presence and functional expression of cannabinoid CB2 receptors in brain. Ann N Y Acad Sci. 2006;1074:514-36. doi: 10.1196/annals.1369.052.
  19. 19. Onaivi ES, Ishiguro H, Gu S, Liu QR. CNS effects of CB2 cannabinoid receptors: beyond neuro-immuno-cannabinoid activity. J Psychopharmacol. 2012;26(1):92-103. doi: 10.1177/0269881111400652.
  20. 20. Mezey E, Toth ZE, Cortright DN, Arzubi MK, Krause JE, Elde R, et al. Distribution of mRNA for vanilloid receptor subtype 1 (VR1), and VR1-like immunoreactivity, in the central nervous system of the rat and human. Proc Natl Acad Sci U S A. 2000;97(7):3655-60. doi: 10.1073/pnas.060496197.
  21. 21. Micale V, Cristino L, Tamburella A, Petrosino S, Leggio GM, Drago F, et al. Anxiolytic effects in mice of a dual blocker of fatty acid amide hydrolase and transient receptor potential vanilloid type-1 channels. Neuropsychopharmacology. 2009;34(3):593-606. doi: 10.1038/npp.2008.98.
  22. 22. Egertova M, Simon GM, Cravatt BF, Elphick MR. Localization of N-acyl phosphatidylethanolamine phospholipase D (NAPE-PLD) expression in mouse brain: a new perspective on N-acylethanolamines as neural signaling molecules. J Comp Neurol. 2008;506(4):604-15. doi: 10.1002/cne.21568.
  23. 23. Egertova M, Cravatt BF, Elphick MR. Comparative analysis of fatty acid amide hydrolase and cb(1) cannabinoid receptor expression in the mouse brain: evidence of a widespread role for fatty acid amide hydrolase in regulation of endocannabinoid signaling. Neuroscience. 2003;119(2):481-96. doi: 10.1016/S0306-4522(03)00145-3.
  24. 24. Yoshida T, Fukaya M, Uchigashima M, Miura E, Kamiya H, Kano M, et al. Localization of diacylglycerol lipase-alpha around postsynaptic spine suggests close proximity between production site of an endocannabinoid, 2-arachidonoyl-glycerol, and presynaptic cannabinoid CB1 receptor. J Neurosci. 2006;26(18):4740-51. doi: 10.1523/JNEUROSCI.0054-06.2006.
  25. 25. Morera-Herreras T, Miguelez C, Aristieta A, Ruiz-Ortega JA, Ugedo L. Endocannabinoid modulation of dopaminergic motor circuits. Front Pharmacol. 2012;3:110. doi: 10.3389/fphar.2012.00110.
  26. 26. Lastres-Becker I, Cebeira M, de Ceballos ML, Zeng BY, Jenner P, Ramos JA, et al. Increased cannabinoid CB1 receptor binding and activation of GTP-binding proteins in the basal ganglia of patients with Parkinson's syndrome and of MPTP-treated marmosets. Eur J Neurosci. 2001;14(11):1827-32. doi: 10.1046/j.0953-816x.2001.01812.x.
  27. 27. Pisani V, Moschella V, Bari M, Fezza F, Galati S, Bernardi G, et al. Dynamic changes of anandamide in the cerebrospinal fluid of Parkinson's disease patients. Mov Disord. 2010;25(7):920-4. doi: 10.1002/mds.23014.
  28. 28. Pisani A, Fezza F, Galati S, Battista N, Napolitano S, Finazzi-Agro A, et al. High endogenous cannabinoid levels in the cerebrospinal fluid of untreated Parkinson's disease patients. Ann Neurol. 2005;57(5):777-9. doi: 10.1002/ana.20462.
  29. 29. Glass M, Faull RL, Dragunow M. Loss of cannabinoid receptors in the substantia nigra in Huntington's disease. Neuroscience. 1993;56(3):523-7. doi: 10.1016/0306-4522(93)90352-G.
  30. 30. Glass M, Dragunow M, Faull RL. 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(3):505-19. doi: 10.1016/S0306-4522(00)00008-7.
  31. 31. Richfield EK, Herkenham M. Selective vulnerability in Huntington's disease: preferential loss of cannabinoid receptors in lateral globus pallidus. Ann Neurol. 1994;36(4):577-84. doi: 10.1002/ana.410360406.
  32. 32. Davis WM, Moreton JE, King WT, Pace HB. Marihuana on locomotor activity: biphasic effect and tolerance development. Res Commun Chem Pathol Pharmacol. 1972;3(1):29-35. doi: 10.1016/j.neures.2011.07.1163.
  33. 33. Dewey WL. Cannabinoid pharmacology. Pharmacol Rev. 1986;38(2):151-78. doi: 10.1016/0378-8741(87)90061-4.
  34. 34. Consroe P. Brain cannabinoid systems as targets for the therapy of neurological disorders. Neurobiol Dis. 1998;5(6 Pt B):534-51. doi: 10.1006/nbdi.1998.0220.
  35. 35. Volkow ND, Wang GJ, Telang F, Fowler JS, Alexoff D, Logan J, et al. Decreased dopamine brain reactivity in marijuana abusers is associated with negative emotionality and addiction severity. Proc Natl Acad Sci U S A. 2014;111(30):E3149-56. doi: 10.1073/pnas.1411228111.
  36. 36. Prescott WR, Gold LH, Martin BR. Evidence for separate neuronal mechanisms for the discriminative stimulus and catalepsy induced by delta 9-THC in the rat. Psychopharmacology (Berl). 1992;107(1):117-24. doi: 10.3389/fphar.2012.00110.
  37. 37. Pertwee RG. Pharmacological, physiological and clinical implications of the discovery of cannabinoid receptors. Biochem Soc Trans. 1998;26(2):267-72. doi: 10.1042/bst0260267.
  38. 38. Sulcova E, Mechoulam R, Fride E. Biphasic effects of anandamide. Pharmacol Biochem Behav. 1998;59(2):347-52. doi: 10.1016/S0091-3057(97)00422-X.
  39. 39. Sanudo-Pena MC, Romero J, Seale GE, Fernandez-Ruiz JJ, Walker JM. Activational role of cannabinoids on movement. Eur J Pharmacol. 2000;391(3):269-74. doi: 10.1016/S0014-2999(00)00044-3.
  40. 40. Jarbe TU, Andrzejewski ME, DiPatrizio NV. Interactions between the CB1 receptor agonist Delta 9-THC and the CB1 receptor antagonist SR-141716 in rats: open-field revisited. Pharmacol Biochem Behav. 2002;73(4):911-9. doi: 10.1016/S0091-3057(02)00938-3.
  41. 41. Navarro M, Fernandez-Ruiz JJ, de Miguel R, Hernandez ML, Cebeira M, Ramos JA. An acute dose of delta 9-tetrahydrocannabinol affects behavioral and neurochemical indices of mesolimbic dopaminergic activity. Behav Brain Res. 1993;57(1):37-46. doi: 10.1016/0166-4328(93)90059-Y.
  42. 42. Navarro M, Rubio P, de Fonseca FR. Behavioural consequences of maternal exposure to natural cannabinoids in rats. Psychopharmacology (Berl). 1995;122(1):1-14. doi: 10.1016/j.bbr.2013.09.022.
  43. 43. Romero J, Garcia-Palomero E, Lin SY, Ramos JA, Makriyannis A, Fernandez-Ruiz JJ. Extrapyramidal effects of methanandamide, an analog of anandamide, the endogenous CB1 receptor ligand. Life Sci. 1996;58(15):1249-57. doi: 10.1016/j.neures.2011.07.1163.
  44. 44. Katsidoni V, Kastellakis A, Panagis G. Biphasic effects of Delta9-tetrahydrocannabinol on brain stimulation reward and motor activity. Int J Neuropsychopharmacol. 2013;16(10):2273-84. doi: 10.1017/S1461145713000709.
  45. 45. Taffe MA, Creehan KM, Vandewater SA. Cannabidiol fails to reverse hypothermia or locomotor suppression induced by Delta(9) -tetrahydrocannabinol in Sprague-Dawley rats. Br J Pharmacol. 2015;172(7):1783-91. doi: 10.1111/bph.13024.
  46. 46. Wickens AP, Pertwee RG. delta 9-Tetrahydrocannabinol and anandamide enhance the ability of muscimol to induce catalepsy in the globus pallidus of rats. Eur J Pharmacol. 1993;250(1):205-8. doi: 10.1016/0014-2999(93)90646-Y.
  47. 47. Gorriti MA, Rodriguez de Fonseca F, Navarro M, Palomo T. Chronic (-)-delta9-tetrahydrocannabinol treatment induces sensitization to the psychomotor effects of amphetamine in rats. Eur J Pharmacol. 1999;365(2-3):133-42. doi: 10.1159/000317059.
  48. 48. McLaughlin PJ, Delevan CE, Carnicom S, Robinson JK, Brener J. Fine motor control in rats is disrupted by delta-9-tetrahydrocannabinol. Pharmacol Biochem Behav. 2000;66(4):803-9. doi: 10.1016/S0091-3057(00)00281-1.
  49. 49. McPartland JM, Duncan M, Di Marzo V, Pertwee RG. Are cannabidiol and Delta(9) -tetrahydrocannabivarin negative modulators of the endocannabinoid system? A systematic review. Br J Pharmacol. 2015;172(3):737-53. doi: 10.1111/bph.12944.
  50. 50. Karniol IG, Carlini EA. Pharmacological interaction between cannabidiol and delta 9-tetrahydrocannabinol. Psychopharmacologia. 1973;33(1):53-70. doi: 10.1007/BF00428793.
  51. 51. Formukong EA, Evans AT, Evans FJ. Inhibition of the cataleptic effect of tetrahydrocannabinol by other constituents of Cannabis sativa L. J Pharm Pharmacol. 1988;40(2):132-4. doi: 10.1111/j.2042-7158.1988.tb05198.
  52. 52. Gomes FV, Del Bel EA, Guimaraes FS. Cannabidiol attenuates catalepsy induced by distinct pharmacological mechanisms via 5-HT1A receptor activation in mice. Prog Neuropsychopharmacol Biol Psychiatry. 2013;46:43-7. doi: 10.1016/j.pnpbp.2013.06.005.
  53. 53. Hayakawa K, Mishima K, Hazekawa M, Sano K, Irie K, Orito K, et al. Cannabidiol potentiates pharmacological effects of Delta(9)-tetrahydrocannabinol via CB(1) receptor-dependent mechanism. Brain Res. 2008;1188:157-64. doi: 10.1016/j.brainres.2007.09.090.
  54. 54. Hiltunen AJ, Jarbe TU, Wangdahl K. Cannabinol and cannabidiol in combination: temperature, open-field activity, and vocalization. Pharmacol Biochem Behav. 1988;30(3):675-8. doi: 10.1016/S0304-3940(03)00569-X.
  55. 55. Espejo-Porras F, Fernandez-Ruiz J, Pertwee RG, Mechoulam R, Garcia C. Motor effects of the non-psychotropic phytocannabinoid cannabidiol that are mediated by 5-HT1A receptors. Neuropharmacology. 2013;75:155-63. doi: 10.1016/j.neuropharm.2013.07.024.
  56. 56. Arevalo C, de Miguel R, Hernandez-Tristan R. Cannabinoid effects on anxiety-related behaviours and hypothalamic neurotransmitters. Pharmacol Biochem Behav. 2001;70(1):123-31. doi: 10.1097/FBP.0000000000000154.
  57. 57. Drews E, Schneider M, Koch M. Effects of the cannabinoid receptor agonist WIN 55,212-2 on operant behavior and locomotor activity in rats. Pharmacol Biochem Behav. 2005;80(1):145-50. doi: 10.1016/j.pbb.2004.10.023.
  58. 58. Marcellino D, Carriba P, Filip M, Borgkvist A, Frankowska M, Bellido I, et al. Antagonistic cannabinoid CB1/dopamine D2 receptor interactions in striatal CB1/D2 heteromers. A combined neurochemical and behavioral analysis. Neuropharmacology. 2008;54(5):815-23. doi: 10.1016/j.neuropharm.2007.12.011.
  59. 59. Robinson L, Goonawardena AV, Pertwee RG, Hampson RE, Riedel G. The synthetic cannabinoid HU210 induces spatial memory deficits and suppresses hippocampal firing rate in rats. Br J Pharmacol. 2007;151(5):688-700. doi: 10.1038/sj.bjp.0707273.
  60. 60. Vlachou S, Stamatopoulou F, Nomikos GG, Panagis G. Enhancement of endocannabinoid neurotransmission through CB1 cannabinoid receptors counteracts the reinforcing and psychostimulant effects of cocaine. Int J Neuropsychopharmacol. 2008;11(7):905-23. doi: 10.1017/S1461145708008717.
  61. 61. Polissidis A, Chouliara O, Galanopoulos A, Rentesi G, Dosi M, Hyphantis T, et al. Individual differences in the effects of cannabinoids on motor activity, dopaminergic activity and DARPP-32 phosphorylation in distinct regions of the brain. Int J Neuropsychopharmacol. 2010;13(9):1175-91. doi: 10.1017/S1461145709991003.
  62. 62. Rodriguez de Fonseca F, Martin Calderon JL, Mechoulam R, Navarro M. Repeated stimulation of D1 dopamine receptors enhances (-)-11-hydroxy-delta 8-tetrahydrocannabinol-dimethyl-heptyl-induced catalepsy in male rats. Neuroreport. 1994;5(7):761-5. doi: 10.1097/00001756-199403000-00006.
  63. 63. Andersson M, Usiello A, Borgkvist A, Pozzi L, Dominguez C, Fienberg AA, et al. Cannabinoid action depends on phosphorylation of dopamine- and cAMP-regulated phosphoprotein of 32 kDa at the protein kinase A site in striatal projection neurons. J Neurosci. 2005;25(37):8432-8. doi: 10.1523/JNEUROSCI.1289-05.2005.
  64. 64. Rinaldi-Carmona M, Barth F, Heaulme M, Shire D, Calandra B, Congy C, et al. SR141716A, a potent and selective antagonist of the brain cannabinoid receptor. FEBS Lett. 1994;350(2-3):240-4. doi: 10.1016/0014-5793(94)00773-X.
  65. 65. Anderson JJ, Kask AM, Chase TN. Effects of cannabinoid receptor stimulation and blockade on catalepsy produced by dopamine receptor antagonists. Eur J Pharmacol. 1996;295(2-3):163-8. doi: 10.1016/0014-2999(95)00661-3.
  66. 66. Carriba P, Ortiz O, Patkar K, Justinova Z, Stroik J, Themann A, et al. Striatal adenosine A2A and cannabinoid CB1 receptors form functional heteromeric complexes that mediate the motor effects of cannabinoids. Neuropsychopharmacology. 2007;32(11):2249-59. doi: 10.1038/sj.npp.1301375.
  67. 67. Ferreira SG, Goncalves FQ, Marques JM, Tome AR, Rodrigues RJ, Nunes-Correia I, et al. Presynaptic adenosine A2A receptors dampen cannabinoid CB1 receptor-mediated inhibition of corticostriatal glutamatergic transmission. Br J Pharmacol. 2015;172(4):1074-86. doi: 10.1111/bph.12970.
  68. 68. Chiodi V, Ferrante A, Ferraro L, Potenza RL, Armida M, Beggiato S, et al. Striatal adenosine-cannabinoid receptor interactions in rats overexpressing adenosine A receptors. J Neurochem. 2015. doi: 10.1111/jnc.13421.
  69. 69. Pertwee RG. Elevating endocannabinoid levels: pharmacological strategies and potential therapeutic applications. Proc Nutr Soc. 2014;73(1):96-105. doi: 10.1017/S0029665113003649.
  70. 70. de Lago E, Fernandez-Ruiz J, Ortega-Gutierrez S, Viso A, Lopez-Rodriguez ML, Ramos JA. UCM707, a potent and selective inhibitor of endocannabinoid uptake, potentiates hypokinetic and antinociceptive effects of anandamide. Eur J Pharmacol. 2002;449(1-2):99-103. doi: 10.1111/j.1472-8206.2008.00626.
  71. 71. Gonzalez S, Romero J, de Miguel R, Lastres-Becker I, Villanua MA, Makriyannis A, et al. Extrapyramidal and neuroendocrine effects of AM404, an inhibitor of the carrier-mediated transport of anandamide. Life Sci. 1999;65(3):327-36. doi: 10.3389/fphar.2012.00110.
  72. 72. Giuffrida A, Rodriguez de Fonseca F, Nava F, Loubet-Lescoulie P, Piomelli D. Elevated circulating levels of anandamide after administration of the transport inhibitor, AM404. Eur J Pharmacol. 2000;408(2):161-8. doi: 10.1016/j.ejphar.2014.03.010.
  73. 73. Zygmunt PM, Chuang H, Movahed P, Julius D, Hogestatt ED. The anandamide transport inhibitor AM404 activates vanilloid receptors. Eur J Pharmacol. 2000;396(1):39-42. doi: 10.1016/S0014-2999(00)00207-7.
  74. 74. Jayamanne A, Greenwood R, Mitchell VA, Aslan S, Piomelli D, Vaughan CW. Actions of the FAAH inhibitor URB597 in neuropathic and inflammatory chronic pain models. Br J Pharmacol. 2006;147(3):281-8. doi: 10.1038/sj.bjp.0706510.
  75. 75. Long JZ, Nomura DK, Vann RE, Walentiny DM, Booker L, Jin X, et al. Dual blockade of FAAH and MAGL identifies behavioral processes regulated by endocannabinoid crosstalk in vivo. Proc Natl Acad Sci U S A. 2009;106(48):20270-5. doi: 10.1073/pnas.0909411106.
  76. 76. Li GL, Winter H, Arends R, Jay GW, Le V, Young T, et al. Assessment of the pharmacology and tolerability of PF-04457845, an irreversible inhibitor of fatty acid amide hydrolase-1, in healthy subjects. Br J Clin Pharmacol. 2012;73(5):706-16. doi: 10.1111/j.1365-2125.2011.04137.x.
  77. 77. Long JZ, Li W, Booker L, Burston JJ, Kinsey SG, Schlosburg JE, et al. Selective blockade of 2-arachidonoylglycerol hydrolysis produces cannabinoid behavioral effects. Nat Chem Biol. 2009;5(1):37-44. doi: 10.1038/nchembio.129.
  78. 78. Ignatowska-Jankowska B, Wilkerson JL, Mustafa M, Abdullah R, Niphakis M, Wiley JL, et al. Selective monoacylglycerol lipase inhibitors: antinociceptive versus cannabimimetic effects in mice. J Pharmacol Exp Ther. 2015;353(2):424-32. doi: 10.1124/jpet.114.222315.
  79. 79. Ramesh D, Gamage TF, Vanuytsel T, Owens RA, Abdullah RA, Niphakis MJ, et al. Dual inhibition of endocannabinoid catabolic enzymes produces enhanced antiwithdrawal effects in morphine-dependent mice. Neuropsychopharmacology. 2013;38(6):1039-49. doi: 10.1038/npp.2012.269.
  80. 80. Souilhac J, Poncelet M, Rinaldi-Carmona M, Le Fur G, Soubrie P. Intrastriatal injection of cannabinoid receptor agonists induced turning behavior in mice. Pharmacol Biochem Behav. 1995;51(1):3-7. doi: 10.1016/0091-3057(94)00396-Z.
  81. 81. Di Marzo V, Lastres-Becker I, Bisogno T, De Petrocellis L, Milone A, Davis JB, et al. Hypolocomotor effects in rats of capsaicin and two long chain capsaicin homologues. Eur J Pharmacol. 2001;420(2-3):123-31. doi: 10.1016/S0014-2999(01)01012-3.
  82. 82. Compton DR, Aceto MD, Lowe J, Martin BR. In vivo characterization of a specific cannabinoid receptor antagonist (SR141716A): inhibition of delta 9-tetrahydrocannabinol-induced responses and apparent agonist activity. J Pharmacol Exp Ther. 1996;277(2):586-94. doi: 10.1124/jpet.111.187815.
  83. 83. Bass CE, Griffin G, Grier M, Mahadevan A, Razdan RK, Martin BR. SR-141716A-induced stimulation of locomotor activity. A structure-activity relationship study. Pharmacol Biochem Behav. 2002;74(1):31-40. doi: 10.1016/S0091-3057(02)00945-0.
  84. 84. Navarro M, Hernandez E, Munoz RM, del Arco I, Villanua MA, Carrera MR, et al. Acute administration of the CB1 cannabinoid receptor antagonist SR 141716A induces anxiety-like responses in the rat. Neuroreport. 1997;8(2):491-6. doi: 10.1097/00001756-199701200-00023.
  85. 85. Thiemann G, van der Stelt M, Petrosino S, Molleman A, Di Marzo V, Hasenohrl RU. The role of the CB1 cannabinoid receptor and its endogenous ligands, anandamide and 2-arachidonoylglycerol, in amphetamine-induced behavioural sensitization. Behav Brain Res. 2008;187(2):289-96. doi: 10.1016/j.bbr.2007.09.022.
  86. 86. Zimmer A, Zimmer AM, Hohmann AG, Herkenham M, Bonner TI. Increased mortality, hypoactivity, and hypoalgesia in cannabinoid CB1 receptor knockout mice. Proc Natl Acad Sci U S A. 1999;96(10):5780-5. doi: 10.1073/pnas.96.10.5780.
  87. 87. Ledent C, Valverde O, Cossu G, Petitet F, Aubert JF, Beslot F, et al. Unresponsiveness to cannabinoids and reduced addictive effects of opiates in CB1 receptor knockout mice. Science. 1999;283(5400):401-4. doi: 10.1126/science.283.5400.401.
  88. 88. Corbille AG, Valjent E, Marsicano G, Ledent C, Lutz B, Herve D, et al. Role of cannabinoid type 1 receptors in locomotor activity and striatal signaling in response to psychostimulants. J Neurosci. 2007;27(26):6937-47. doi: 10.1523/JNEUROSCI.3936-06.2007.
  89. 89. Thiemann G, Di Marzo V, Molleman A, Hasenohrl RU. The CB(1) cannabinoid receptor antagonist AM251 attenuates amphetamine-induced behavioural sensitization while causing monoamine changes in nucleus accumbens and hippocampus. Pharmacol Biochem Behav. 2008;89(3):384-91. doi: 10.1016/j.pbb.2008.01.010.
  90. 90. Starowicz K, Nigam S, Di Marzo V. Biochemistry and pharmacology of endovanilloids. Pharmacol Ther. 2007;114(1):13-33. doi: 10.1016/j.pharmthera.2007.01.005.
  91. 91. Lang AE, Lozano AM. Parkinson's disease. First of two parts. N Engl J Med. 1998;339(15):1044-53. doi: 10.1056/NEJM199810083391506.
  92. 92. Lang AE, Lozano AM. Parkinson's disease. Second of two parts. N Engl J Med. 1998;339(16):1130-43. doi: 10.1056/NEJM199810153391607.
  93. 93. Kish SJ, Shannak K, Hornykiewicz O. Uneven pattern of dopamine loss in the striatum of patients with idiopathic Parkinson's disease. Pathophysiologic and clinical implications. N Engl J Med. 1988;318(14):876-80. doi: 10.1056/NEJM198804073181402.
  94. 94. Papa SM. The cannabinoid system in Parkinson's disease: multiple targets to motor effects. Exp Neurol. 2008;211(2):334-8. doi: 10.1016/j.expneurol.2008.03.009.
  95. 95. van der Stelt M, Fox SH, Hill M, Crossman AR, Petrosino S, Di Marzo V, et al. A role for endocannabinoids in the generation of parkinsonism and levodopa-induced dyskinesia in MPTP-lesioned non-human primate models of Parkinson's disease. FASEB J. 2005;19(9):1140-2. doi: 10.1096/fj.04-3010fje.
  96. 96. Di Marzo V, Hill MP, Bisogno T, Crossman AR, Brotchie JM. Enhanced levels of endogenous cannabinoids in the globus pallidus are associated with a reduction in movement in an animal model of Parkinson's disease. FASEB J. 2000;14(10):1432-8. doi: 10.1096/fj.14.10.1432.
  97. 97. Gubellini P, Picconi B, Bari M, Battista N, Calabresi P, Centonze D, et al. Experimental parkinsonism alters endocannabinoid degradation: implications for striatal glutamatergic transmission. J Neurosci. 2002;22(16):6900-7. doi: 20026732.
  98. 98. Silverdale MA, McGuire S, McInnes A, Crossman AR, Brotchie JM. Striatal cannabinoid CB1 receptor mRNA expression is decreased in the reserpine-treated rat model of Parkinson's disease. Exp Neurol. 2001;169(2):400-6. doi: 10.1006/exnr.2001.7649.
  99. 99. Mounsey RB, Mustafa S, Robinson L, Ross RA, Riedel G, Pertwee RG, et al. Increasing levels of the endocannabinoid 2-AG is neuroprotective in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson's disease. Exp Neurol. 2015;273:36-44. doi: 10.1016/j.expneurol.2015.07.024.
  100. 100. Esposito E, Impellizzeri D, Mazzon E, Paterniti I, Cuzzocrea S. Neuroprotective activities of palmitoylethanolamide in an animal model of Parkinson's disease. PLoS One. 2012;7(8):e41880. doi: 10.1371/journal.pone.0041880.
  101. 101. Van Laere K, Casteels C, Lunskens S, Goffin K, Grachev ID, Bormans G, et al. Regional changes in type 1 cannabinoid receptor availability in Parkinson's disease in vivo. Neurobiol Aging. 2012;33(3):620 e1-8. doi: 10.1016/j.neurobiolaging.2011.02.009.
  102. 102. Hurley MJ, Mash DC, Jenner P. Expression of cannabinoid CB1 receptor mRNA in basal ganglia of normal and parkinsonian human brain. J Neural Transm (Vienna). 2003;110(11):1279-88. doi: 10.1007/s00702-003-0033-7.
  103. 103. Farkas S, Nagy K, Jia Z, Harkany T, Palkovits M, Donohou SR, et al. The decrease of dopamine D(2)/D(3) receptor densities in the putamen and nucleus caudatus goes parallel with maintained levels of CB(1) cannabinoid receptors in Parkinson's disease: a preliminary autoradiographic study with the selective dopamine D(2)/D(3) antagonist [(3)H]raclopride and the novel CB(1) inverse agonist [(1)(2)(5)I]SD7015. Brain Res Bull. 2012;87(6):504-10. doi: 10.1016/j.brainresbull.2012.02.012.
  104. 104. Zeng BY, Dass B, Owen A, Rose S, Cannizzaro C, Tel BC, et al. Chronic L-DOPA treatment increases striatal cannabinoid CB1 receptor mRNA expression in 6-hydroxydopamine-lesioned rats. Neurosci Lett. 1999;276(2):71-4. doi: 10.1016/S0304-3940(99)00762-4.
  105. 105. Romero J, Berrendero F, Perez-Rosado A, Manzanares J, Rojo A, Fernandez-Ruiz JJ, et al. Unilateral 6-hydroxydopamine lesions of nigrostriatal dopaminergic neurons increased CB1 receptor mRNA levels in the caudate-putamen. Life Sci. 2000;66(6):485-94. doi: 10.1016/S0024320599006189.
  106. 106. Madeo G, Schirinzi T, Maltese M, Martella G, Rapino C, Fezza F, et al. Dopamine-dependent CB1 receptor dysfunction at corticostriatal synapses in homozygous PINK1 knockout mice. Neuropharmacology. 2016;101:460-70. doi:10.1016/j.neuropharm.2015.10.021.
  107. 107. Concannon RM, Okine BN, Finn DP, Dowd E. Differential upregulation of the cannabinoid CB(2) receptor in neurotoxic and inflammation-driven rat models of Parkinson's disease. Exp Neurol. 2015;269:133-41. doi: 10.1016/j.expneurol.2015.04.007.
  108. 108. Razavinasab M, Shamsizadeh A, Shabani M, Nazeri M, Allahtavakoli M, Asadi-Shekaari M, et al. Pharmacological blockade of TRPV1 receptors modulates the effects of 6-OHDA on motor and cognitive functions in a rat model of Parkinson's disease. Fundam Clin Pharmacol. 2013;27(6):632-40. doi: 10.1111/fcp.12015.
  109. 109. Morgese MG, Cassano T, Cuomo V, Giuffrida A. Anti-dyskinetic effects of cannabinoids in a rat model of Parkinson's disease: role of CB(1) and TRPV1 receptors. Exp Neurol. 2007;208(1):110-9. doi: 10.1016/j.expneurol.2007.07.021.
  110. 110. Pringsheim T, Wiltshire K, Day L, Dykeman J, Steeves T, Jette N. The incidence and prevalence of Huntington's disease: a systematic review and meta-analysis. Mov Disord. 2012;27(9):1083-91. doi: 10.1002/mds.25075.
  111. 111. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. The Huntington's Disease Collaborative Research Group. Cell. 1993;72(6):971-83. doi: 10.1016/0092-8674(93)90585-E.
  112. 112. Reiner A, Albin RL, Anderson KD, D'Amato CJ, Penney JB, Young AB. Differential loss of striatal projection neurons in Huntington disease. Proc Natl Acad Sci U S A. 1988;85(15):5733-7. doi: 10.1073/pnas.85.15.5733.
  113. 113. Heinsen H, Strik M, Bauer M, Luther K, Ulmar G, Gangnus D, et al. Cortical and striatal neurone number in Huntington's disease. Acta Neuropathol. 1994;88(4):320-33. doi: 10.1007/BF00310376.
  114. 114. Hedreen JC, Peyser CE, Folstein SE, Ross CA. Neuronal loss in layers V and VI of cerebral cortex in Huntington's disease. Neurosci Lett. 1991;133(2):257-61. doi: 10.1016/0304-3940(91)90583-F.
  115. 115. Walker FO. Huntington's disease. Lancet. 2007;369(9557):218-28. doi: 10.1016/S0140-6736(07)60111-1.
  116. 116. Sagredo O, Pazos MR, Valdeolivas S, Fernandez-Ruiz J. Cannabinoids: novel medicines for the treatment of Huntington's disease. Recent Pat CNS Drug Discov. 2012;7(1):41-8. doi: 10.2174/157488912798842278.
  117. 117. Berardelli A, Noth J, Thompson PD, Bollen EL, Curra A, Deuschl G, et al. Pathophysiology of chorea and bradykinesia in Huntington's disease. Mov Disord. 1999;14(3):398-403. doi: 0.1002/1531-8257(199905)14:3<398::AID-MDS1003>3.0.CO;2-F.
  118. 118. Allen KL, Waldvogel HJ, Glass M, Faull RL. Cannabinoid (CB(1)), GABA(A) and GABA(B) receptor subunit changes in the globus pallidus in Huntington's disease. J Chem Neuroanat. 2009;37(4):266-81. doi: 10.1016/S0891-0618(09)00020-9.
  119. 119. Blazquez C, Chiarlone A, Sagredo O, Aguado T, Pazos MR, Resel E, et al. Loss of striatal type 1 cannabinoid receptors is a key pathogenic factor in Huntington's disease. Brain. 2011;134(Pt 1):119-36. doi: 10.1093/brain/awq278.
  120. 120. Battista N, Bari M, Tarditi A, Mariotti C, Bachoud-Levi AC, Zuccato C, et al. Severe deficiency of the fatty acid amide hydrolase (FAAH) activity segregates with the Huntington's disease mutation in peripheral lymphocytes. Neurobiol Dis. 2007;27(1):108-16. doi: 10.1016/j.nbd.2007.04.012.
  121. 121. Dowie MJ, Bradshaw HB, Howard ML, Nicholson LF, Faull RL, Hannan AJ, et al. Altered CB1 receptor and endocannabinoid levels precede motor symptom onset in a transgenic mouse model of Huntington's disease. Neuroscience. 2009;163(1):456-65. doi: 10.1016/j.neuroscience.2009.06.014.
  122. 122. 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(1-2):307-13. doi: 10.1016/j.neuint.2007.06.031.
  123. 123. Bari M, Battista N, Valenza M, Mastrangelo N, Malaponti M, Catanzaro G, et al. In vitro and in vivo models of Huntington's disease show alterations in the endocannabinoid system. FEBS J. 2013;280(14):3376-88. doi: 10.1111/febs.12329.
  124. 124. Lastres-Becker I, Fezza F, Cebeira M, Bisogno T, Ramos JA, Milone A, et al. Changes in endocannabinoid transmission in the basal ganglia in a rat model of Huntington's disease. Neuroreport. 2001;12(10):2125-9. doi: 10.1097/00001756-200107200-00017.
  125. 125. Horne EA, Coy J, Swinney K, Fung S, Cherry AE, Marrs WR, et al. Downregulation of cannabinoid receptor 1 from neuropeptide Y interneurons in the basal ganglia of patients with Huntington's disease and mouse models. Eur J Neurosci. 2013;37(3):429-40. doi: 10.1111/ejn.12045.
  126. 126. Van Laere K, Casteels C, Dhollander I, Goffin K, Grachev I, Bormans G, et al. Widespread decrease of type 1 cannabinoid receptor availability in Huntington disease in vivo. J Nucl Med. 2010;51(9):1413-7. doi: 10.2967/jnumed.110.077156.
  127. 127. Denovan-Wright EM, Robertson HA. Cannabinoid receptor messenger RNA levels decrease in a subset of neurons of the lateral striatum, cortex and hippocampus of transgenic Huntington's disease mice. Neuroscience. 2000;98(4):705-13. doi: 10.1016/S0306-4522(00)00157-3.
  128. 128. Lastres-Becker I, Berrendero F, Lucas JJ, Martin-Aparicio E, Yamamoto A, Ramos JA, et al. Loss of mRNA levels, binding and activation of GTP-binding proteins for cannabinoid CB1 receptors in the basal ganglia of a transgenic model of Huntington's disease. Brain Res. 2002;929(2):236-42. doi: 10.1016/S0006-8993(01)03403-5.
  129. 129. Chiodi V, Uchigashima M, Beggiato S, Ferrante A, Armida M, Martire A, et al. Unbalance of CB1 receptors expressed in GABAergic and glutamatergic neurons in a transgenic mouse model of Huntington's disease. Neurobiol Dis. 2012;45(3):983-91. doi: 10.1016/j.nbd.2011.12.017.
  130. 130. Lastres-Becker I, Hansen HH, Berrendero F, De Miguel R, Perez-Rosado A, Manzanares J, et al. Alleviation of motor hyperactivity and neurochemical deficits by endocannabinoid uptake inhibition in a rat model of Huntington's disease. Synapse. 2002;44(1):23-35. doi: 10.1002/syn.10054.
  131. 131. Page KJ, Besret L, Jain M, Monaghan EM, Dunnett SB, Everitt BJ. Effects of systemic 3-nitropropionic acid-induced lesions of the dorsal striatum on cannabinoid and mu-opioid receptor binding in the basal ganglia. Exp Brain Res. 2000;130(2):142-50. doi: 10.1007/s002210050016.
  132. 132. Ooms M, Rietjens R, Rangarajan JR, Vunckx K, Valdeolivas S, Maes F, 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(12):2858-69. doi: 10.1016/j.neurobiolaging.2014.06.010.
  133. 133. Casteels C, Vandeputte C, Rangarajan JR, Dresselaers T, Riess O, Bormans G, et al. Metabolic and type 1 cannabinoid receptor imaging of a transgenic rat model in the early phase of Huntington disease. Exp Neurol. 2011;229(2):440-9. doi: 10.1016/j.expneurol.2011.03.014.
  134. 134. Chiarlone A, Bellocchio L, Blazquez C, Resel E, Soria-Gomez E, Cannich A, et al. A restricted population of CB1 cannabinoid receptors with neuroprotective activity. Proc Natl Acad Sci U S A. 2014;111(22):8257-62. doi: 10.1073/pnas.1400988111.
  135. 135. Mievis S, Blum D, Ledent C. Worsening of Huntington disease phenotype in CB1 receptor knockout mice. Neurobiol Dis. 2011;42(3):524-9. doi: 10.1016/j.nbd.2011.03.006.
  136. 136. Naydenov AV, Sepers MD, Swinney K, Raymond LA, Palmiter RD, Stella N. Genetic rescue of CB1 receptors on medium spiny neurons prevents loss of excitatory striatal synapses but not motor impairment in HD mice. Neurobiol Dis. 2014;71:140-50. doi: 10.1016/j.nbd.2014.08.009.
  137. 137. Palazuelos J, Aguado T, Pazos MR, Julien B, Carrasco C, Resel E, et al. Microglial CB2 cannabinoid receptors are neuroprotective in Huntington's disease excitotoxicity. Brain. 2009;132(Pt 11):3152-64. doi: 10.1093/brain/awp239.
  138. 138. Dowie MJ, Grimsey NL, Hoffman T, Faull RL, Glass M. Cannabinoid receptor CB2 is expressed on vascular cells, but not astroglial cells in the post-mortem human Huntington's disease brain. J Chem Neuroanat. 2014;59-60:62-71. doi: 10.1016/j.jchemneu.2014.06.004.
  139. 139. Sagredo O, Gonzalez S, Aroyo I, Pazos MR, Benito C, Lastres-Becker I, et al. Cannabinoid CB2 receptor agonists protect the striatum against malonate toxicity: relevance for Huntington's disease. Glia. 2009;57(11):1154-67. doi: 10.1002/glia.20838.
  140. 140. Bouchard J, Truong J, Bouchard K, Dunkelberger D, Desrayaud S, Moussaoui S, et al. Cannabinoid receptor 2 signaling in peripheral immune cells modulates disease onset and severity in mouse models of Huntington's disease. J Neurosci. 2012;32(50):18259-68. doi: 10.1523/JNEUROSCI.4008-12.2012.
  141. 141. Carlsson A. Treatment of Parkinson's with L-D OPA. The early discovery phase, and a comment on current problems. J Neural Transm (Vienna). 2002;109(5-6):777-87. doi: 10.1007/s007020200064.
  142. 142. Anderson LA, Anderson JJ, Chase TN, Walters JR. The cannabinoid agonists WIN 55,212-2 and CP 55,940 attenuate rotational behavior induced by a dopamine D1 but not a D2 agonist in rats with unilateral lesions of the nigrostriatal pathway. Brain Res. 1995;691(1-2):106-14. doi: 10.1016/0006-8993(95)00645-7.
  143. 143. Maneuf YP, Crossman AR, Brotchie JM. The cannabinoid receptor agonist WIN 55,212-2 reduces D2, but not D1, dopamine receptor-mediated alleviation of akinesia in the reserpine-treated rat model of Parkinson's disease. Exp Neurol. 1997;148(1):265-70. doi: 10.1006/exnr.1997.6645.
  144. 144. Brotchie JM. Adjuncts to dopamine replacement: a pragmatic approach to reducing the problem of dyskinesia in Parkinson's disease. Mov Disord. 1998;13(6):871-6. doi: 10.1002/mds.870130603.
  145. 145. Sanudo-Pena MC, Patrick SL, Khen S, Patrick RL, Tsou K, Walker JM. Cannabinoid effects in basal ganglia in a rat model of Parkinson's disease. Neurosci Lett. 1998;248(3):171-4. doi: 10.1016/S0304-3940(98)00368-1.
  146. 146. Sanudo-Pena MC, Tsou K, Walker JM. Motor actions of cannabinoids in the basal ganglia output nuclei. Life Sci. 1999;65(6-7):703-13. doi: 10.1016/S0024-3205(99)00293-3.
  147. 147. Fernandez-Espejo E, Caraballo I, de Fonseca FR, El Banoua F, Ferrer B, Flores JA, et al. Cannabinoid CB1 antagonists possess antiparkinsonian efficacy only in rats with very severe nigral lesion in experimental parkinsonism. Neurobiol Dis. 2005;18(3):591-601. doi: 10.1016/j.nbd.2004.10.015.
  148. 148. Gonzalez S, Scorticati C, Garcia-Arencibia M, de Miguel R, Ramos JA, Fernandez-Ruiz J. Effects of rimonabant, a selective cannabinoid CB1 receptor antagonist, in a rat model of Parkinson's disease. Brain Res. 2006;1073-1074:209-19. doi: 10.1016/j.brainres.2005.12.014.
  149. 149. Kelsey JE, Harris O, Cassin J. The CB(1) antagonist rimonabant is adjunctively therapeutic as well as monotherapeutic in an animal model of Parkinson's disease. Behav Brain Res. 2009;203(2):304-7. doi: 10.1016/j.bbr.2009.04.035.
  150. 150. Garcia-Arencibia M, Ferraro L, Tanganelli S, Fernandez-Ruiz J. Enhanced striatal glutamate release after the administration of rimonabant to 6-hydroxydopamine-lesioned rats. Neurosci Lett. 2008;438(1):10-3. doi: 10.1016/j.neulet.2008.04.041.
  151. 151. de Lago E, de Miguel R, Lastres-Becker I, Ramos JA, Fernandez-Ruiz J. Involvement of vanilloid-like receptors in the effects of anandamide on motor behavior and nigrostriatal dopaminergic activity: in vivo and in vitro evidence. Brain Res. 2004;1007(1-2):152-9. doi: 10.1016/j.brainres.2004.02.016.
  152. 152. Ferrer B, Asbrock N, Kathuria S, Piomelli D, Giuffrida A. Effects of levodopa on endocannabinoid levels in rat basal ganglia: implications for the treatment of levodopa-induced dyskinesias. Eur J Neurosci. 2003;18(6):1607-14. doi: 10.1046/j.1460-9568.2003.02896.
  153. 153. Martinez A, Macheda T, Morgese MG, Trabace L, Giuffrida A. The cannabinoid agonist WIN55212-2 decreases L-DOPA-induced PKA activation and dyskinetic behavior in 6-OHDA-treated rats. Neurosci Res. 2012;72(3):236-42. doi: 10.1016/j.neures.2011.12.006.
  154. 154. Fox SH, Henry B, Hill M, Crossman A, Brotchie J. Stimulation of cannabinoid receptors reduces levodopa-induced dyskinesia in the MPTP-lesioned nonhuman primate model of Parkinson's disease. Mov Disord. 2002;17(6):1180-7. doi: 10.1002/mds.10289.
  155. 155. Sieradzan KA, Fox SH, Hill M, Dick JP, Crossman AR, Brotchie JM. Cannabinoids reduce levodopa-induced dyskinesia in Parkinson's disease: a pilot study. Neurology. 2001;57(11):2108-11. doi: org/10.1212/WNL.57.11.2108.
  156. 156. Venderova K, Ruzicka E, Vorisek V, Visnovsky P. Survey on cannabis use in Parkinson's disease: subjective improvement of motor symptoms. Mov Disord. 2004;19(9):1102-6. doi: 10.1002/mds.20111.
  157. 157. Zuardi AW, Crippa JA, Hallak JE, Pinto JP, Chagas MH, Rodrigues GG, et al. Cannabidiol for the treatment of psychosis in Parkinson's disease. J Psychopharmacol. 2009;23(8):979-83. doi: 10.1177/0269881108096519.
  158. 158. Carroll CB, Bain PG, Teare L, Liu X, Joint C, Wroath C, et al. Cannabis for dyskinesia in Parkinson disease: a randomized double-blind crossover study. Neurology. 2004;63(7):1245-50. doi: 10.1212/01.WNL.0000140288.48796.8E.
  159. 159. Mesnage V, Houeto JL, Bonnet AM, Clavier I, Arnulf I, Cattelin F, et al. Neurokinin B, neurotensin, and cannabinoid receptor antagonists and Parkinson disease. Clin Neuropharmacol. 2004;27(3):108-10. doi: 10.1097/00002826-200405000-00003.
  160. 160. Lastres-Becker I, Molina-Holgado F, Ramos JA, Mechoulam R, Fernandez-Ruiz J. Cannabinoids provide neuroprotection against 6-hydroxydopamine toxicity in vivo and in vitro: relevance to Parkinson's disease. Neurobiol Dis. 2005;19(1-2):96-107. doi: 10.1016/j.nbd.2004.11.009.
  161. 161. Price DA, Martinez AA, Seillier A, Koek W, Acosta Y, Fernandez E, et al. WIN55,212-2, a cannabinoid receptor agonist, protects against nigrostriatal cell loss in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson's disease. Eur J Neurosci. 2009;29(11):2177-86. doi: 10.1111/j.1460-9568.2009.06764.x.
  162. 162. Chung YC, Bok E, Huh SH, Park JY, Yoon SH, Kim SR, et al. Cannabinoid receptor type 1 protects nigrostriatal dopaminergic neurons against MPTP neurotoxicity by inhibiting microglial activation. J Immunol. 2011;187(12):6508-17. doi: 10.4049/jimmunol.1102435.
  163. 163. Garcia-Arencibia M, Gonzalez S, de Lago E, Ramos JA, Mechoulam R, Fernandez-Ruiz J. Evaluation of the neuroprotective effect of cannabinoids in a rat model of Parkinson's disease: importance of antioxidant and cannabinoid receptor-independent properties. Brain Res. 2007;1134(1):162-70. doi: 10.1016/j.brainres.2006.11.063.
  164. 164. Carroll CB, Zeissler ML, Hanemann CO, Zajicek JP. Delta(9)-tetrahydrocannabinol (Delta(9)-THC) exerts a direct neuroprotective effect in a human cell culture model of Parkinson's disease. Neuropathol Appl Neurobiol. 2012;38(6):535-47. doi: 10.1111/j.1365-2990.2011.01248.x.
  165. 165. Garcia C, Palomo-Garo C, Garcia-Arencibia M, Ramos J, Pertwee R, Fernandez-Ruiz J. Symptom-relieving and neuroprotective effects of the phytocannabinoid Delta(9)-THCV in animal models of Parkinson's disease. Br J Pharmacol. 2011;163(7):1495-506. doi: 10.1111/j.1476-5381.2011.01278.x.
  166. 166. Chagas MH, Zuardi AW, Tumas V, Pena-Pereira MA, Sobreira ET, Bergamaschi MM, et al. Effects of cannabidiol in the treatment of patients with Parkinson's disease: an exploratory double-blind trial. J Psychopharmacol. 2014;28(11):1088-98. doi: 10.1177/0269881114550355.
  167. 167. Jimenez-Del-Rio M, Daza-Restrepo A, Velez-Pardo C. The cannabinoid CP55,940 prolongs survival and improves locomotor activity in Drosophila melanogaster against paraquat: implications in Parkinson's disease. Neurosci Res. 2008;61(4):404-11. doi: 10.1016/j.neures.2008.04.011.
  168. 168. McGeer PL, Itagaki S, Boyes BE, McGeer EG. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson's and Alzheimer's disease brains. Neurology. 1988;38(8):1285-91. doi: 10.3389/fncel.2015.00084.
  169. 169. Mogi M, Harada M, Kondo T, Riederer P, Inagaki H, Minami M, et al. Interleukin-1 beta, interleukin-6, epidermal growth factor and transforming growth factor-alpha are elevated in the brain from parkinsonian patients. Neurosci Lett. 1994;180(2):147-50. doi: 10.1016/0304-3940(94)90508-8.
  170. 170. Mogi M, Harada M, Kondo T, Riederer P, Nagatsu T. Brain beta 2-microglobulin levels are elevated in the striatum in Parkinson's disease. J Neural Transm Park Dis Dement Sect. 1995;9(1):87-92. doi: 10.1523/JNEUROSCI.3447-06.2006.
  171. 171. Muller T, Blum-Degen D, Przuntek H, Kuhn W. Interleukin-6 levels in cerebrospinal fluid inversely correlate to severity of Parkinson's disease. Acta Neurol Scand. 1998;98(2):142-4. doi: 10.1111/j.1600-0404.1998.tb01736.x.
  172. 172. Sagredo O, Garcia-Arencibia M, de Lago E, Finetti S, Decio A, Fernandez-Ruiz J. Cannabinoids and neuroprotection in basal ganglia disorders. Mol Neurobiol. 2007;36(1):82-91. doi: 10.1007/s12035-007-0004-3.
  173. 173. Fernandez-Ruiz J. Cannabinoids in neurodegeneration and neuroprotection. In: Mechoulam R, editor. Cannabinoids and Therapeutics. Berlin 2005. p.79-109. doi: 10.1007/3-7643-7358-X_5.
  174. 174. Benito C, Tolon RM, Pazos MR, Nunez E, Castillo AI, Romero J. Cannabinoid CB2 receptors in human brain inflammation. Br J Pharmacol. 2008;153(2):277-85. doi: 10.1038/sj.bjp.0707505.
  175. 175. Little JP, Villanueva EB, Klegeris A. Therapeutic potential of cannabinoids in the treatment of neuroinflammation associated with Parkinson's disease. Mini Rev Med Chem. 2011;11(7):582-90. doi: 10.2174/138955711795906905.
  176. 176. Ternianov A, Perez-Ortiz JM, Solesio ME, Garcia-Gutierrez MS, Ortega-Alvaro A, Navarrete F, et al. Overexpression of CB2 cannabinoid receptors results in neuroprotection against behavioral and neurochemical alterations induced by intracaudate administration of 6-hydroxydopamine. Neurobiol Aging. 2012;33(2):421 e1-16. doi: 10.1016/j.neurobiolaging.2010.09.012.
  177. 177. Marchalant Y, Brothers HM, Norman GJ, Karelina K, DeVries AC, Wenk GL. Cannabinoids attenuate the effects of aging upon neuroinflammation and neurogenesis. Neurobiol Dis. 2009;34(2):300-7. doi: 10.1016/j.nbd.2009.01.014.
  178. 178. Chung ES, Bok E, Chung YC, Baik HH, Jin BK. Cannabinoids prevent lipopolysaccharide-induced neurodegeneration in the rat substantia nigra in vivo through inhibition of microglial activation and NADPH oxidase. Brain Res. 2012;1451:110-6. doi: 10.1016/j.brainres.2012.02.058.
  179. 179. Pidgeon C, Rickards H. The pathophysiology and pharmacological treatment of Huntington disease. Behav Neurol. 2013;26(4):245-53. doi: 10.3233/BEN-2012-120267.
  180. 180. Fernandez-Ruiz J, Romero J, Ramos JA. Endocannabinoids and Neurodegenerative Disorders: Parkinson's Disease, Huntington's Chorea, Alzheimer's Disease, and Others. Handb Exp Pharmacol. 2015;231:233-59. doi: 10.1007/978-3-319-20825-1_8.
  181. 181. Kluger B, Triolo P, Jones W, Jankovic J. The therapeutic potential of cannabinoids for movement disorders. Mov Disord. 2015;30(3):313-27. doi: 10.1002/mds.26142.
  182. 182. de Lago E, Fernandez-Ruiz J, Ortega-Gutierrez S, Cabranes A, Pryce G, Baker D, et al. UCM707, an inhibitor of the anandamide uptake, behaves as a symptom control agent in models of Huntington's disease and multiple sclerosis, but fails to delay/arrest the progression of different motor-related disorders. Eur Neuropsychopharmacol. 2006;16(1):7-18. doi: 10.1016/j.euroneuro.2005.06.001.
  183. 183. de Lago E, Urbani P, Ramos JA, Di Marzo V, Fernandez-Ruiz J. Arvanil, a hybrid endocannabinoid and vanilloid compound, behaves as an antihyperkinetic agent in a rat model of Huntington's disease. Brain Res. 2005;1050(1-2):210-6. doi: 10.1016/j.brainres.2005.05.024.
  184. 184. Lastres-Becker I, de Miguel R, De Petrocellis L, Makriyannis A, Di Marzo V, Fernandez-Ruiz J. Compounds acting at the endocannabinoid and/or endovanilloid systems reduce hyperkinesia in a rat model of Huntington's disease. J Neurochem. 2003;84(5):1097-109. doi: 10.1046/j.1471-4159.2003.01595.x.
  185. 185. Pietropaolo S, Bellocchio L, Ruiz-Calvo A, Cabanas M, Du Z, Guzman M, et al. Chronic cannabinoid receptor stimulation selectively prevents motor impairments in a mouse model of Huntington's disease. Neuropharmacology. 2015;89:368-74. doi: 10.1016/j.neuropharm.2014.07.021.
  186. 186. Curtis A, Rickards H. Nabilone could treat chorea and irritability in Huntington's disease. J Neuropsychiatry Clin Neurosci. 2006;18(4):553-4. doi: 10.1176/appi.neuropsych.18.4.553.
  187. 187. 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(15):2254-9. doi: 10.1002/mds.22809.
  188. 188. Consroe P, Laguna J, Allender J, Snider S, Stern L, Sandyk R, et al. Controlled clinical trial of cannabidiol in Huntington's disease. Pharmacol Biochem Behav. 1991;40(3):701-8. doi: 0091-3057(91)90386-G.
  189. 189. Muller-Vahl KR, Schneider U, Emrich HM. Nabilone increases choreatic movements in Huntington's disease. Mov Disord. 1999;14(6):1038-40. doi: 10.1002/1531-8257(199911)14:63.0.CO;2-7.
  190. 190. Fernandez-Ruiz J. The endocannabinoid system as a target for the treatment of motor dysfunction. Br J Pharmacol. 2009;156(7):1029-40. doi: 10.1111/j.1476-5381.2008.00088.x.
  191. 191. Blazquez C, Chiarlone A, Bellocchio L, Resel E, Pruunsild P, Garcia-Rincon D, et al. The CB(1) cannabinoid receptor signals striatal neuroprotection via a PI3K/Akt/mTORC1/BDNF pathway. Cell Death Differ. 2015;22(10):1618-29. doi: 10.1038/cdd.2015.11.
  192. 192. Demuth DG, Molleman A. Cannabinoid signalling. Life Sci. 2006;78(6):549-63. doi: 10.1016/j.lfs.2005.05.055.
  193. 193. Monory K, Massa F, Egertova M, Eder M, Blaudzun H, Westenbroek R, et al. The endocannabinoid system controls key epileptogenic circuits in the hippocampus. Neuron. 2006;51(4):455-66. doi: 10.1016/j.neuron.2006.07.006.
  194. 194. Foldy C, Neu A, Jones MV, Soltesz I. Presynaptic, activity-dependent modulation of cannabinoid type 1 receptor-mediated inhibition of GABA release. J Neurosci. 2006;26(5):1465-9. doi: 10.1523/JNEUROSCI.4587-05.2006.
  195. 195. Randall MD, Harris D, Kendall DA, Ralevic V. Cardiovascular effects of cannabinoids. Pharmacol Ther. 2002;95(2):191-202. doi: 10.1016/S0163-7258(02)00258-9.
  196. 196. Marsicano G, Moosmann B, Hermann H, Lutz B, Behl C. Neuroprotective properties of cannabinoids against oxidative stress: role of the cannabinoid receptor CB1. J Neurochem. 2002;80(3):448-56. doi: 10.1046/j.0022-3042.2001.00716.x.
  197. 197. Sagredo O, Ramos JA, Decio A, Mechoulam R, Fernandez-Ruiz J. Cannabidiol reduced the striatal atrophy caused 3-nitropropionic acid in vivo by mechanisms independent of the activation of cannabinoid, vanilloid TRPV1 and adenosine A2A receptors. Eur J Neurosci. 2007;26(4):843-51. doi: 10.1111/j.1460-9568.2007.05717.x.
  198. 198. Scotter EL, Goodfellow CE, Graham ES, Dragunow M, Glass M. Neuroprotective potential of CB1 receptor agonists in an in vitro model of Huntington's disease. Br J Pharmacol. 2010;160(3):747-61. doi: 10.1111/j.1476-5381.2010.00773.x.
  199. 199. Laprairie RB, Bagher AM, Kelly ME, Denovan-Wright EM. Biased Type 1 Cannabinoid Receptor Signalling Influences Neuronal Viability in a Cell Culture Model of Huntington Disease. Mol Pharmacol. 2015. doi: 10.1124/mol.115.101980.
  200. 200. Pintor A, Tebano MT, Martire A, Grieco R, Galluzzo M, Scattoni ML, et al. The cannabinoid receptor agonist WIN 55,212-2 attenuates the effects induced by quinolinic acid in the rat striatum. Neuropharmacology. 2006;51(5):1004-12. doi: 10.1016/j.neuropharm.2006.06.013.
  201. 201. van Dellen A, Blakemore C, Deacon R, York D, Hannan AJ. Delaying the onset of Huntington's in mice. Nature. 2000;404(6779):721-2. doi: 10.1038/35008142.
  202. 202. Glass M, van Dellen A, Blakemore C, Hannan AJ, Faull RL. Delayed onset of Huntington's disease in mice in an enriched environment correlates with delayed loss of cannabinoid CB1 receptors. Neuroscience. 2004;123(1):207-12. doi: 10.1016/S0306-4522(03)00595-5.
  203. 203. Dowie MJ, Howard ML, Nicholson LF, Faull RL, Hannan AJ, Glass M. Behavioural and molecular consequences of chronic cannabinoid treatment in Huntington's disease transgenic mice. Neuroscience. 2010;170(1):324-36. doi: 10.1016/j.neuroscience.2010.06.056.
  204. 204. Fernandez-Ruiz J, Garcia C, Sagredo O, Gomez-Ruiz M, de Lago E. The endocannabinoid system as a target for the treatment of neuronal damage. Expert Opin Ther Targets. 2010;14(4):387-404. doi: 10.1517/14728221003709792.
  205. 205. Sagredo O, Pazos MR, Satta V, Ramos JA, Pertwee RG, Fernandez-Ruiz J. Neuroprotective effects of phytocannabinoid-based medicines in experimental models of Huntington's disease. J Neurosci Res. 2011;89(9):1509-18. doi: 10.1002/jnr.22682.
  206. 206. Valdeolivas S, Satta V, Pertwee RG, Fernandez-Ruiz J, Sagredo O. Sativex-like combination of phytocannabinoids is neuroprotective in malonate-lesioned rats, an inflammatory model of Huntington's disease: role of CB1 and CB2 receptors. ACS Chem Neurosci. 2012;3(5):400-6. doi: 10.1021/cn200114w.
  207. 207. Stella N. Cannabinoid signaling in glial cells. Glia. 2004;48(4):267-77. doi: 10.1002/glia.20084.
  208. 208. Walter L, Stella N. Cannabinoids and neuroinflammation. Br J Pharmacol. 2004;141(5):775-85. doi: 10.1038/sj.bjp.0705667.
  209. 209. Smith SR, Terminelli C, Denhardt G. Effects of cannabinoid receptor agonist and antagonist ligands on production of inflammatory cytokines and anti-inflammatory interleukin-10 in endotoxemic mice. J Pharmacol Exp Ther. 2000;293(1):136-50. doi: 10734163.
  210. 210. Molina-Holgado F, Pinteaux E, Moore JD, Molina-Holgado E, Guaza C, Gibson RM, et al. Endogenous interleukin-1 receptor antagonist mediates anti-inflammatory and neuroprotective actions of cannabinoids in neurons and glia. J Neurosci. 2003;23(16):6470-4. doi: 12878687.

Written By

Teresa Morera-Herreras, Cristina Miguelez, Asier Aristieta, María Torrecilla, José Ángel Ruiz-Ortega and Luisa Ugedo

Submitted: 06 October 2015 Reviewed: 09 February 2016 Published: 15 June 2016