The endocannabinoid system (ECS) is a highly versatile signaling system within the nervous system. Despite its widespread localization, its functions within the context of distinct neural processes are very well discernable and specific. This is remarkable, and the question remains as to how such specificity is achieved. One key player in the ECS is the cannabinoid type 1 receptor (CB1), a G protein–coupled receptor characterized by the complexity of its cell-specific expression, cellular and subcellular localization, and its adaptable regulation of intracellular signaling cascades. CB1 receptors are involved in different synaptic and cellular plasticity processes and in the brain’s bioenergetics in a context-specific manner. CB2 receptors are also important in several processes in neurons, glial cells, and immune cells of the brain. As polymorphisms in ECS components, as well as external impacts such as stress and metabolic challenges, can both lead to dysregulated ECS activity and subsequently to possible neuropsychiatric disorders, pharmacological intervention targeting the ECS is a promising therapeutic approach. Understanding the neurobiology of cannabinoid receptor signaling in depth will aid optimal design of therapeutic interventions, minimizing unwanted side effects.

Keywords: anandamide, 2-arachidonoylglycerol, behavior, cannabinoid receptor, neural communication

Introduction

The endocannabinoid system (ECS) comprises two cannabinoid receptors—CB 1 and CB 2 receptors—that belong to the family of seven transmembrane G protein-coupled receptors (GPCRs); the ligands for the cannabinoid receptors—the two major endocannabinoids (eCBs) anandamide (arachidonoylethanolamide, AEA) and 2-arachidonoylglycerol (2-AG), both being derivatives of the fatty acid arachidonic acid; and the machinery for the synthesis and degradation of eCBs. , The ECS is evolutionarily well conserved in vertebrates, is widely distributed in the body, and takes a central position in the regulation of a myriad of biological processes, both in neural and non-neural tissues. , It is intertwined with many neurotransmitter and lipid signaling systems, thereby integrated into broad functional networks. It appears that the ECS plays roles in the fine-tuning of physiological processes that keep the body in homeostatic set-points. In humans, several polymorphisms in ECS components with associated neuropathophysiological processes have been described, suggesting they promote susceptibility toward development of neuropsychiatric disorders. ECS dysregulation can also be induced by particular life factors, such as living under chronic stress, or by metabolic factors, such as with obesity. Pharmacological interventions targeting ECS activity aim to normalize such pathophysiological processes, thereby rescuing the subject from unfavorable allostatic set-points.

Since the discovery of the ECS in the 1990s, this signaling system has attracted intense attention, especially as it aided understanding of the effects of phytocannabinoids. Furthermore, detailing the various components of the ECS uncovered the fascinating complexity of how this signaling system acts in the functional network of the entire organism, in particular in the brain. This article presents a general overview of the neurobiology of the ECS. However, the vast literature on this subject makes it nearly impossible to touch on all aspects, and gaps are inevitable. Furthermore, due to space constraints, the discussion here focuses on the ECS of the brain in rodents and human, in particular the CB 1 receptor. However, the ECS is also widely involved in the regulation of peripheral immune, cardiovascular, metabolic, gastrointestinal, muscular, and peripheral nervous system processes, which in turn can influence central nervous system (CNS) functions. ,

Endocannabinoid mechanisms

The peculiarities of the cannabinoid receptors

CB 1 and CB 2 receptors feature the many typical characteristics of GPCRs, making these receptors highly versatile and adaptable, for example, regarding ligand binding, intracellular signaling coupling, homo- and heterodimerization, and subcellular localization. Here, the focus will be on the CB 1 receptor, the major cannabinoid receptor in the nervous system, but several aspects of CB 2 receptors will also be addressed. It is interesting to note that despite the ubiquitous occurrence of the CB 1 receptor in the nervous system, this signaling system appears to act in a highly specific manner in a given context. Several key features come into play and will be discussed below.

 

Receptor expression at the cellular level

Detailed expression analyses at the regional and cellular levels in the CNS revealed that the CB 1 receptor is present in virtually all brain regions and in all major cell types ( Figure 1 ), ie, in neurons, glial cells (astrocytes, oligodendrocytes), , and brain-resident immune cells (microglia). The CB 1 receptor is also present in the neurons of the major neurotransmitter systems (glutamate, γ-aminobutyric acid [GABA], serotonin, noradrenalin, acetylcholine), and possibly also in dopaminergic neurons. CB 1 receptor–expressing cells can also be classified according to the presence of peptidergic transmitters, such as corticotropin-releasing hormone, cholecystokinin, , and somatostatin. The expression levels can vary greatly, depending on the brain region and cell type. For example, the central amygdala contains very low levels of CB 1 receptor in only a few of the presumably GABAergic neurons. , On the other hand, very high levels of CB 1 receptor messenger RNA (mRNA) are detected in hippocampal and neocortical GABAergic interneurons. , Glutamatergic neurons generally contain rather low levels of CB 1 receptor. , Furthermore, the CB 1 receptor is barely detectable in astrocytes, , oligodendrocytes, ,, oligodendrocyte precursor cells (OPCs), , and adult neural stem cells (NSCs). , Specific functions of CB 1 receptor in these different populations and in many brain areas have been described in mouse by using conditional gene inactivation of CB 1 receptor in the respective cell types and/or brain regions, together with local pharmacological interventions. ,, Importantly, the relative abundance of CB 1 receptor does not indicate the importance of the receptor in a particular physiological process. ,

An external file that holds a picture, illustration, etc.
Object name is DCNS_22.3_Lutz_Figure1.jpg

Expression of cannabinoid type 1 and type 2 (CB1 and CB2) receptors in neural tissue. The endocannabinoid system is present in neurons, astrocytes, oligodendrocytes, oligodendrocyte precursor cells (OPCs), and microglia. Functional CB1 receptors are located on the plasma membrane, but also in mitochondria (mtCB1) of neurons and astrocytes. Presynaptic CB1 receptor suppresses neurotransmitter release, as shown here, at a glutamatergic synapse. For this process, postsynaptic increase of Ca2+ triggers the synthesis of endocannabinoids, which travel to the presynapse to activate CB1 receptor. Astrocytic CB1 receptor can regulate gliotransmitter release. AMPAR, AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid) type glutamate receptor; CB1/CB2, cannabinoid type 1/type 2 receptor; eCB, endocannabinoid; mGluR5, metabotropic glutamate receptor 5; mtCB1, mitochondrial CB1 receptor; NMDAR, NMDA (N-methyl-D-aspartate receptor) type glutamate receptor; OPC, oligodendrocyte precursor cell; TRPV1, transient receptor potential cation channel subfamily V member 1

CB 2 receptor expression has been predominantly described in peripheral immune cells and brain-resident immune cells, the macrophages, but was eventually also detected in neurons, a circumstance that has fueled many investigations on CB 2 receptor in neural functions. In the CNS, CB 2 receptor expression was reported in cells such as activated microglia, brain stem neurons, hippocampal glutamatergic neurons, and dopaminergic neurons of the ventral tegmental area. , CB 2 receptor transcripts are reported to be 100 to 200 times less abundant than CB 1 receptor mRNA, but are strongly upregulated in response to various insults, such as chronic pain, neuroinflammation, and stroke. , Genetic approaches, together with pharmacology and very specific and sensitive cellular detection methods of mRNA, paved the way for the recognition of CB 2 receptor in many neural functions.

In summary, the mRNA encoding the two major receptors for eCBs is very widely expressed in the brain in many cell types, allowing the involvement in numerous physiological and pathophysiological processes. In the next paragraph, the subcellular location of the proteins will be discussed, preparing the ground for detailing the involvement of the receptors in particular cell-signaling processes.

 

Receptor expression at the subcellular level

Presynaptic localization: The dominant site of CB 1 receptor protein location is the presynapse, where the activation of CB 1 receptor can suppress presynaptic neurotransmitter release. This mechanism is very well detailed for glutamatergic and GABAergic synapses, whereby the eCB 2-AG is generated in the postsynaptic site and travels retrogradely to the presynaptic site to stimulate the CB 1 receptor. This can result in a short-term decrease in Ca 2+ influx at the presynaptic terminal. ,,, This signaling can typically lead to processes called depolarization-induced suppression of excitation (DSE) at the glutamatergic synapse, and depolarization-induced suppression of inhibition (DSI) at the GABAergic synapse. Most of the investigations have substantiated 2-AG as the retrograde eCB, consistent with the neuroanatomical configuration with postsynaptic synthesizing and presynaptic degrading enzymes. For AEA, the situation is far from being understood. First, AEA synthesis machinery seems to be mainly located at the presynaptic site, whereas the degrading machinery is at the postsynaptic site. Secondly, besides the cannabinoid receptors, the postsynaptically acting transient receptor potential cation channel subfamily V member 1 (TRPV1) has to be considered as an AEA receptor as well. Altogether, DSE and DSI appear to be mediated by 2-AG but not by AEA, as evidenced via studies that used a genetically induced decrease in 2-AG and AEA signaling in hippocampal glutamatergic neurons by overexpression of 2-AG–degrading enzyme monoacylglycerol lipase (MAGL) and the AEA-degrading enzyme fatty acid amide hydrolase (FAAH), respectively.

Postsynaptic localization: Postsynaptic CB 1 receptor has been reported in electrophysiological experiments in neocortical GABAergic and glutamatergic neurons, and in hippocampal glutamatergic neurons involving potassium/sodium hyperpolarization-activated cyclic nucleotide-gated channel 1 (HCN1), possibly via a somatodendritic mechanism. Based on immunohistological analysis, CB 1 receptor has been described to be intracellular, within somatodendritic endosomes, and until now not yet on the cell membrane.

In summary, the neuronal CB 1 receptor is dominantly expressed at the presynapse; however, functional postsynaptic CB 1 receptor has also been reported. CB 2 receptor has been reported to be localized in the soma, but to date, no immunostaining has revealed a presynaptic or dendritic localization.

Intracellular localization: Receptor trafficking is central to activity regulation and function of GPCRs. Receptors are synthesized in the cell soma and transported to the destination—for CB 1 receptors, this is mainly the axonal terminal—and finally integrated into the cellular membrane. Distinct sequences of the CB 1 receptor are involved in axonal trafficking, , whereby alterations in such sequences can lead to eCB-mediated electrophysiological and behavioral alterations. Distribution of CB 1 receptor in the different neuronal compartments can be influenced by the activation state of the receptor. For example, pharmacological activation of CB 1 receptor increases its presence in the somatodendritic endosomes and decreases the receptor’s presence at the presynapse. This is explained by high internalization and retrograde transport of the CB 1 receptor. This process is very pronounced in hippocampus and neocortex, but absent in basal ganglia. On the other hand, treatment with CB 1 receptor antagonist leads to a decreased number of cell bodies with CB 1 receptor–containing endosomes. This observation suggests that under steady-state conditions, the CB 1 receptor is steadily activated and internalized into the soma. Using high-resolution microscopy, long-term treatment with the psychoactive phytocannabinoid Δ 9 -tetrahydrocannabinol (THC) was shown to lead to a decrease in CB 1 receptors located in the presynaptic membrane of hippocampal GABAergic interneurons. This internalization persists for many days after termination of THC treatment. Along these lines, genetic deletion of the 2-AG–degrading enzyme MAGL also leads to β-arrestin–mediated internalization and desensitization of the CB 1 receptor, an effect that is not observed in FAAH-deficient mice containing increased AEA levels.

Apart from intracellular localization in the endosomal compartment in the context of receptor trafficking, CB 1 receptor was also detected at mitochondrial membranes of neurons and named mtCB 1 receptor. Here it mediates the reduction in mitochondrial oxygen consumption upon stimulation by exogenous cannabinoids and eCBs, in the end decreasing the production of adenosine triphosphate (ATP). Subsequently, cannabinoids were shown to regulate complex I of the respiratory chain by modulation of mitochondrial protein kinase A (PKA) activity and downstream phosphorylation events. Thus, the ECS directly regulates mitochondrial energy production via mtCB 1 receptor. Moreover, mtCB 1 receptor was shown in the same study to mediate the memory-impairing effects of cannabinoids. Recently, activation of astrocytic mtCB 1 was shown to inhibit glucose metabolism and lactate production, altering neuronal functions and behavioral responses in a social-interaction test.

In summary, considering that eCBs are synthesized locally and that these lipid-signaling molecules diffuse and occupy a restricted three-dimensional space, generating a microdomain of eCB signaling, the precise localization of the cannabinoid receptors determines the downstream signaling. Research in recent years has established that functional CB 1 receptor is present both in the plasma membrane as well as in intracellular compartments, particularly in the outer mitochondrial membrane. Duration and intensity of eCB signaling is determined by the dynamics of eCB synthesis and degradation. Furthermore, it is thought that eCB signaling contains both tonic (constitutive) and phasic (short-term, “on-demand”) components. , Thus, important hallmarks of eCB signaling are its temporal and spatial restrictions, which of course is also a common feature of “classical” neurotransmitter receptor systems. Next, the diversity of the intracellular signal transduction upon cannabinoid receptor activation will be addressed.

Intracellular signaling: Among the GPCRs, the CB 1 receptor appears to be most highly expressed in the brain, with protein levels in the same range as for the major components of the excitatory and inhibitory neurotransmitter systems, ie, for N -methyl-d-aspartate (NMDA) and GABA A receptors, respectively. Intrinsically, GPCRs contain a rich repertoire for regulation of cellular processes upon ligand binding. , , Many aspects of CB 1 receptor signaling have been reported , ; however, we are far from understanding how CB 1 receptor signaling gains specificity depending on cellular context.

Dimerization: GPCRs contain the capacity to form homo- and heterodimers, ie, a particular GPCR interacts with the same type of receptor to form a homodimer, or a particular GPCR interacts with another type of receptor to form a heterodimer. Such dimerization processes are involved in signal integration upon receptor activation. Moreover, dimerization processes are also implicated in the emergence of mental disorders, and this might also be the case for cannabinoid receptors. Homodimers were reported for CB 1 receptors and heterodimers reported for CB 1 receptor with several other GPCRs, including CB 2 receptor, , dopamine D 2 receptor, , and serotonin 5-HT 2A receptor. Heterodimerization was reported to lead to alterations in signaling. For example, CB 1 receptor/dopamine D 2 –receptor heterodimers in cultured striatal neurons can change the intracellular signaling upon CB 1 receptor stimulation. Along the same line, a recent investigation showed the heterodimerization of CB 1 receptor with the adenosine A 2A receptor in striatum, whereby the costimulation of both receptors reduces intracellular signaling. CB 1 /CB 2 receptor heterodimers have also been reported to enhance ligand binding of the phytocannabinoid cannabigerol. Furthermore, heterodimerization was shown between cannabinoid receptor with non-GPCR. For example, heterodimerization between CB 2 receptor and the tyrosine kinase receptor HER2 is involved in the antitumor action of THC, whereby THC interrupts the dimer. Altogether, these features make up a powerful tool for cell-type–specific fine-tuning of intracellular signaling. Obviously, the two dimerizing receptors must be in tight proximity within the same cell compartment, providing a constraint for activation of this mechanism.

G protein coupling: CB 1 receptor is typically coupled to G i/o proteins, leading to a decreased production of cyclic adenosine monophosphate (cAMP) and an inhibition of N- and P/Q-type Ca 2+ channels, resulting in decreased Ca 2+ influx upon stimulation. , Under particular circumstances, G s coupling in striatal neurons was reported, in particular in concert with dopamine D 2 receptor signaling. G q coupling was reported in astrocytes. G αz was found as a CB 1 receptor–interacting G protein, detected from a biochemical pull-down proteomic experiment using hippocampal synaptosomes expressing tagged CB 1 receptor. Furthermore, biochemical analysis of G proteins interacting upon CB 1 receptor stimulation in neocortical extracts revealed the presence not only of G i/o , but also of G αz , G α12/13 , and G αq/11 . The coupling was dependent on the CB 1 receptor agonist used. Differences in the G protein coupling has been proposed to be a possible mechanism to explain the observation that low levels of CB 1 receptor proteins in glutamatergic hippocampal neurons show higher GTPγ-binding activity than the high CB 1 receptor content in GABAergic interneurons. However, these experiments were performed with whole hippocampal extracts without subcellular fractionation and represent an average over the entire tissue. Furthermore, the regulator of G protein–signaling (RGS) proteins constitute an important intracellular component in the control of GPCR signaling. For example, RGS proteins have been implicated in the interaction of CB 1 receptor with dopamine D 2 receptor regarding the regulation of eCB-mediated retrograde synaptic signaling of striatal neurons.

Altogether, these observations suggest that CB 1 receptor signaling can depend on the availability of the intracellular G protein pool and on the specific ligand that activates the CB 1 receptor. The former parameter is obviously also influenced by the presence of other GPCRs in the same subcellular domain, by competing for the same G protein pool and RGSs. Furthermore, as dysregulation of ECS activity has been reported for pathophysiological processes, alterations in these different constituents (G proteins, RGS, other GPCRs), can also lead to altered CB 1 receptor signaling.

Signaling via β-arrestin: Binding of CB 1 receptor to β-arrestins is important for the internalization of the receptor upon ligand stimulation, but β-arrestins are also signal transducers for GPCR intracellular pathways, such as extracellular-signal–regulated kinase (ERK), and c-jun terminal kinase (JNK). So-called biased β-arrestin signaling upon ligand activation of GPCRs has been recognized as therapeutically relevant also for CB 1 receptor signaling. , Deletion of the CB 1 receptor phosphorylation site that is involved in β-arrestin binding was reported to lead to resistance to cannabinoid tolerance and hypersensitivity to cannabinoids. CB 1 receptor can also activate the mammalian target of rapamycin (mTOR) pathway, eg, to regulate presynaptic protein synthesis in the context of long-term synaptic plasticity. On the other hand, mTOR mediates the amnesic effects of THC via the CB 1 receptor.

Receptor-interacting proteins: Furthermore, the presence of CB 1 receptor–interacting proteins can also contribute to the diversity of CB 1 receptor signaling. Here, the cannabinoid receptor interacting protein 1A (CRIP1A) is reported to influence CB 1 receptor agonist-induced regulation of excitatory neurotransmission, to modulate which G i/o subtypes interact with CB 1 receptor, and to attenuate CB 1 receptor internalization via β-arrestin.

Splice variants and posttranslational modifications: Lastly, for both CB 1 and CB 2 receptors, splice variants have been reported in rodents and human. , The in vivo significance of these variants has not yet been clarified, although differences in mRNA expression, receptor signaling, trafficking, and glycosylation have been reported. , , , Along these lines, posttranslational modifications, such as phosphorylation for the β-arrestin binding site and N-linked glycosylation, also have to be considered in the regulation of receptor activity. For example, reduction in glycosylation of the CB 1 receptor reduces the cell membrane expression of the receptor, but not ligand-binding affinity.

In summary, in the CNS, the ECS constitutes an important mechanism for the fine-tuned regulation of synaptic transmission in numerous projections and local networks in most, if not all, brain regions. Its proposed function as a so-called circuit breaker, due to the retrograde mechanism of the suppression of neurotransmitter release, is certainly a central aspect of the ECS function, but the ECS is acting beyond this and might also be seen as an integrator of synaptic processes, enabling encoding of information.

Cellular plasticity in the adult brain

 

Generation of new neurons

Besides the involvement of the ECS in synaptic plasticity, eCB-regulated plasticity is also present at the cellular level ( Figure 2 ). Adult neurogenesis, ie, the generation of newly born neurons from NSCs, is an important physiological process and is implicated in neuropsychiatric disorders. In humans and rodents, the major sites of neurogenesis are the subgranular zone (SGZ) of the dentate gyrus of the hippocampus and the subventricular zone (SVZ) in the lateral ventricle. In the adult, neurogenesis occurs continuously, but the rate of proliferation and subsequent differentiation are tightly regulated and can be influenced by different intrinsic and extrinsic factors. Proliferation is increased, eg, by enriched environment and physical activity, and decreased, eg, upon stress, during depression, and over the course of aging. , The role of the ECS in adult neurogenesis is well documented, as reviewed in detail elsewhere. Experiments performed in mice under home-cage conditions without behavioral challenges revealed that ubiquitous gene deficiency of CB 1 receptor and impaired 2-AG signaling in mice with deficiency of the 2-AG synthesizing enzyme diacylglycerol lipase-α (DAGLα) , lead to decreased NSC proliferation. On the contrary, under basal conditions, ubiquitous CB 2 receptor deficiency did not lead to impairments in NSC proliferation. Reduced proliferation was also observed in rodents treated with specific CB 1 – and CB 2 receptor antagonists. On the other hand, the enhancement of eCB signaling through the genetic inactivation of the AEA-degrading enzyme FAAH and the pharmacological blockade of the 2-AG–degrading enzyme MAGL led to increased NSC proliferation. Synthetic cannabinoid receptor agonist–stimulated neurogenesis appears to require both CB 1 and CB 2 receptors. Furthermore, the phytocannabinoid THC can also stimulate neurogenesis. Interestingly, the nonpsychotropic phytocannabinoid cannabidiol (CBD) was also reported to enhance neurogenesis, possibly through facilitating eCB signaling. Interestingly, CBD treatment was shown to increase AEA levels.

An external file that holds a picture, illustration, etc.
Object name is DCNS_22.3_Lutz_Figure2.jpg

Neurogenesis in the subgranular zone of the adult brain hippocampus. Both cannabinoid type 1 and type 2 (CB1 and CB2) receptors are expressed in neural stem cells and neural progenitor cells and participate in the proliferation of neural stem cells. CB1 receptor also acts later on in the differentiation of the newly generated neurons with regard to dendritic length and spine number. CB1/CB2, cannabinoid type 1/type 2 receptor

Exercise is an efficient intervention for increasing adult neurogenesis. Pharmacological blockade of the CB 1 receptor has been shown to blunt exercise-induced increase in proliferation in the SGZ. However, in another study, using CB 1 receptor–deficient mice, no genotype differences were observed in neurogenesis over a 6-week running period, but the CB 1 receptor–deficient mice showed reduced motivation to run. These divergent results might be explained by the differences in the pharmacological versus genetic blockade of CB 1 receptor. Results from an investigation in a mouse model of Down syndrome are very different. Here, interestingly, pharmacological blockade of CB 1 receptor led to the alleviation of impaired cognitive performance, synaptic plasticity, and neurogenesis, an effect possibly explained by the pathology of enhanced hippocampal CB 1 receptor expression and concomitant increased receptor function at excitatory terminals.

NSCs contain a functional ECS, including the expression of CB 1 and CB 2 receptors, and eCB-synthesizing and -degrading enzymes. In vitro experiments suggest the involvement of phosphoinositide 3-kinase/protein kinase B (PI3K/PKB), ERK, and mTOR complex 1 (mTORC) pathways in eCB-dependent stimulation of proliferation of NSCs. ,, For the CB 2 receptor, it has emerged that this receptor is mainly important under pathophysiological conditions, whereby reduction in neurogenesis induced by damaging conditions (epilepsy, alcohol, stroke, neurodegenerative processes) can be alleviated by CB 2 receptor activation.

It has to be recognized that the neurogenic niches are embedded into an environment that strongly affects proliferation and that also contains a functional ECS. Therefore, phenotypic outcomes after modulation of ECS activity depend on the cellular and temporal specificity of the targeting of the respective ECS components. A study using mice with ubiquitous loss of the CB 1 receptor and of FAAH revealed that eCB signaling controls neural progenitor differentiation in the adult brain by altering astroglial differentiation of newly born cells. The survival for the newly born cells was not changed. The question arises whether CB 1 receptor expressed in NSCs is directly involved in the promotion of newly born neurons from NSCs. To this end, specific genetic loss of the CB 1 receptor in NSCs revealed decreased proliferation. Furthermore, CB 1 receptor deficiency caused decreased dendritic branches and spine numbers in the differentiating neurons, reduced long-term potentiation (LTP) and short-term spatial memory, and increased depression-like behavior. Yet, no alteration in cell fate was observed, indicating that the effect on decreased astroglial differentiation observed in the ubiquitous CB 1 receptor–deficient mice was probably caused by the ECS changes around the NSC niche. The effect on dendritic branching observed on NSC-specific CB 1 receptor–deficient mice would also suggest a postsynaptic function of the CB 1 receptor that is required for maturation, but eventually, CB 1 receptor expression stops in terminally differentiated and integrated granule cells.

 

Generation of new oligodendrocytes

Another cell plasticity process is reported for the replenishment of oligodendrocytes from OPCs, which express the protein neural/glial antigen 2 (NG2). Upon challenges of the adult brain, such as damaging of the myelin, OPCs are capable of proliferating and differentiating into mature oligodendrocytes, alleviating or even repairing the damage. The ECS is functionally present in cultured OPCs and is required for maintaining proliferation, a process requiring PI3K/PKB/mTOR-signaling pathways, as inhibition of 2-AG synthesis and blockade of CB 1 /CB 2 receptors seem to induce cell-cycle arrest. The ECS is also present in mature oligodendrocytes. , , CB 1 /CB 2 -agonist stimulation leads to increased myelin basic protein (MBP) expression. Furthermore, in animal models of demyelination, stimulation of the ECS can alleviate various pathologies associated with demyelination. , Altogether, the ECS regulates cellular plasticity in the adult brain. Here, we focused on proliferation, but the ECS can also regulate various aspects of apoptosis and autophagy, two important cellular events for maintaining homeostasis in the body.

 

Neurogenesis in the embryo and early postnatal brain development

Finally, it is important to mention that the ECS has numerous functions during embryogenesis and at postnatal stages when the brain develops to its final maturation state. The understanding of these processes gives us essential mechanistic insights into the detrimental effects of cannabis use during these stages of development. As it is beyond the scope of this short presentation, the reader is referred to recent reviews on the ECS in neural development. ,

Dysregulated endocannabinoid system

Functions of the ECS help to maintain homeostasis in the body ( Figure 3A ), for example, through regulation of the stress response, feeding, and energy metabolism, and for ensuring the excitatory/inhibitory balance in the nervous system. Considering the temporal and spatial activity of the ECS, it is no surprise that a dysregulated ECS can lead to new set-points, called allostasis, that might then be implicated in distinct neuropsychiatric disorders ( Figure 3A, B ). Human polymorphisms in genes of the ECS can be linked to particular phenotypes, as summarized in detail in a recent review, suggesting a dysregulated ECS as origin of the observed altered behavior. A wealth of data is available for the phenotypes caused by the FAAH polymorphism rs324430 (C385A), which leads to decreased protein stability of FAAH in the A allele, leading to increased AEA levels. This point mutation was also introduced into the mouse genome, allowing for comparative studies in human and mouse. These investigations indicate that the A allele promotes fear extinction, reduces anxiety, and provides protection against stress-induced decreases in AEA. On the other hand, the A allele of FAAH constitutes a risk factor for developing anxiety and depression from repetitive childhood trauma, indicating the relevance of the developmental dynamics of ECS activity during childhood and adolescence. A recent case report describes a microdeletion in FAAH that led to increased AEA levels in blood and to insensitivity to pain. Several polymorphisms in the CB 1 receptor gene ( CNR1 ) have been described. None of them lead to overt biochemical changes in CB 1 receptor protein functions. However, these polymorphisms might contribute to the susceptibility for development of certain neuropsychiatric disorders, although the data are not as clear as for the FAAH C385A polymorphism; further studies are needed to substantiate previous observations with independent cohorts. Investigations on a CB 2 receptor polymorphism ( CNR2 ) recently summarized in a meta-analysis suggested an association of a particular SNP (rn2501432) with depression. Furthermore, a study revealed that the R allele of the CNR2 Q63R polymorphism together with the A allele of the FAAH C385A polymorphism are associated with enhanced vulnerability to childhood trauma and to a later anxious and depressive phenotype.

In humans, ECS activity can be monitored by measuring eCB levels in blood, saliva, hair, and cerebrospinal fluid. Expression of ECS genes can also be determined in the tissue post mortem. Numerous studies have reported alterations in eCB levels in patients suffering from various disorders, such as depression, posttraumatic stress disorder (PTSD), schizophrenia, anorexia nervosa, and Tourette syndrome. ,, Also, the course of therapy can be followed up by monitoring eCB levels, such as antidepressant treatment with electroconvulsive intervention. In investigations with patients, drug treatments for the respective diseases constitute a serious confounding for eCB measurements. eCB can also be monitored in healthy subjects who are undergoing psychological tests and other challenges, and it is hoped that such monitoring, in combination with other biomarkers, would make possible the evaluation of potential predispositions toward development of distinct disorders.

In humans, as recently reviewed, positron emission tomography (PET), a noninvasive method to determine ECS activity via radioactive tracers for cannabinoid receptors and ECS enzymes, has been applied to various disorders. Studies with alcohol-use disorders and schizophrenia are inconsistent, some reporting increased and others decreased CB 1 receptor binding. Data on aberrant CB 1 receptor binding in individuals with anorexia and PTSD are more coherent, but the data sets are still limited.

In summary, dysregulated ECS activity is reported in several neuropsychiatric disorders. This dysregulation may originate from genetic and/or epigenetic alterations in ECS components, or may be consequences of alterations of other signaling pathways, leading to changes in ECS activity.

The ECS as a therapeutic target

Given that dysregulated ECS activity can lead to allostatic set-points that could potentially represent pathological states, pharmacological treatment targeting ECS components is a promising strategy. A wealth of therapeutic applications has been investigated in preclinical animal models, but only in a few cases has translation to humans been achieved in clinical trials. The focus of the present discussion is on the modulation of the various components of the ECS ( Figure 3C, D ). Treatment options using phytocannabinoids are beyond the scope of this presentation, but information on this subject is found in other reviews.

An external file that holds a picture, illustration, etc.
Object name is DCNS_22.3_Lutz_Figure3.jpg

The endocannabinoid system is a homeostatic system. (A) It is proposed that various physiological processes need optimal ECS activity to maintain a homeostatic set-point. Aberrant ECS activity may lead to allostatic set-points, with the emergence of various diseases. (B) The shift from a homeostatic to an allostatic set-point may have different origins, such as genetic causes, but also life events and life history (eg, stress, trauma, metabolic challenges), acting, among others, via epigenetic mechanisms. (C) Pharmacological interventions aiming at regaining homeostasis by targeting different ECS components using different mechanistic approaches. (D) Pharmacological interventions targeting the different ECS components. Compounds acting on cannabinoid receptors with biased signaling effects toward G protein or β–arrestin pathways contain high potentials. 2-AG, 2-arachidonoylglycerol; AEA, anandamide (arachidonoylethanolamide); CB1/CB2, cannabinoid type 1/type 2 receptor; DAGL, diacylglycerol lipase; eCB, endocannabinoid; ET, eCB transporter; FAAH, fatty acid amide hydrolase; MAGL, monoacylglycerol lipase; NAM, negative allosteric receptor modulator; NAPE-PLD, 2-acyl phosphatidylethanolamide-specific phospholipase D; PAM, positive allosteric receptor modulator; PTSD, posttraumatic stress disorder

Complexity of eCB signaling molecules

Biosynthesis of eCBs occurs from membrane precursors, and eCB degradation products are precursors of eicosanoids. Thus, eCB signaling is integrated into a lipid metabolism and signaling network. In consequence, modification of activity of eCB synthesizing and degrading enzymes may also alter other lipid signaling systems. The synthesis and degradation machinery of AEA and 2-AG have different cellular and subcellular distributions, indicating differential functions. Furthermore, AEA and 2-AG have different pharmacological profiles with regard to interaction with their receptors, CB 1 and CB 2 receptors, but can also activate other receptors, such as TRPV1 and GABA A receptors, respectively. , Furthermore, endogenous peptides called pepcans or hemopressin were recently characterized ; these can act on CB 1 and CB 2 receptors, thereby modifying biological processes.

Cannabinoid receptors

For many years, the focus has been on CB 1 receptor antagonism/inverse agonism. However, the failure of rimonabant (Acomplia) because of CNS side effects stopped clinical applications. Nevertheless, convincing alternative strategies, in particular peripherally acting CB 1 receptor antagonists, have been developed in recent years and shown to be active without appreciable CNS side effects. , As peripheral organ systems interact with CNS functions, alleviation of dysregulated ECS activity in the periphery also has potential for beneficial therapeutic effects in dysregulated CNS functions. ,, As discussed above, the ECS is featured by its temporal and spatial specificity in signaling. Thus, both positive (PAM) and negative (NAM) allosteric receptor modulators constitute a promising alternative path compared with direct receptor agonism/antagonism. Indeed, preclinical research has shown very promising efficacy using such compounds. , Recently, pregnenolone and non-metabolizable derivatives thereof have emerged as a promising NAM of CB 1 receptor. , The development of receptor ligands with bias toward distinct intracellular pathways, mainly G protein versus β-arrestin, also represents a very promising strategy to increase the therapeutic effects and diminish unwanted signaling side effects. Pepcans (hemopressins) were reported to be a NAM for CB 1 receptor and a PAM for CB 2 receptor. A further strategy to enhance specificity is to generate dual-target drugs. For example, a mainly peripherally acting hybrid CB 1 /inducible nitric oxide synthase (iNOS) antagonist is effective in the treatment of experimental liver fibrosis, and the fusion peptide between hemopressin and neuropeptide VF shows potent antinociceptive effects with reduced cannabinoid-related side effects. Comparable approaches have been pursued for interfering with CB 2 receptor activity, although with slower progress. Particularly in the field of neuroinflammation and neurodegeneration, CB 2 receptor agonism seems to be promising. CB 2 receptor as a target is particularly attractive because of the lack of psychotropic activity, which is present in the case of CB 1 receptor agonism.

Catabolic and metabolic eCB enzymes

Knowing that eCBs cannot be stored in vesicles, that ECS activity is spatially and temporally regulated, and that AEA and 2-AG have different functions in the brain, the ability to pharmacologically interfere with biosynthesis and degradation of AEA and 2-AG, respectively, is useful. It also allows a possible increase in specificity at the site of action and may reduce side effects. Therefore, the inhibition of FAAH and MAGL separately, and the inhibition of both FAAH and MAGL together are the most investigated approaches in preclinical research and in clinical trials for various applications, such as pain, inflammation, anxiety, and depression-like behavior. As for modulators of cannabinoid receptors, peripherally acting compounds may help avoid CNS-derived side effects. , FAAH inhibitors also have preclinical applications. , Whereas hopes are high that such compounds will be useful in humans, this has yet to be achieved. Clinical trials must be conducted with care so that setbacks can be avoided, a lesson learned, for example, in the clinical trial using the FAAH inhibitor BIA 10-2474, which contained nonspecific reactions toward serine hydrolases other than FAAH, revealed from incidences of clinical neurotoxicity. Recently, new compounds inhibiting NAPE-PLD were reported. Interestingly, CBD can alter levels of AEA and other N -acylethanolamines, suggesting mechanisms via an indirect stimulation of the ECS, thereby possibly explaining the low level of side effects induced by CBD.

eCB membrane transporter

eCBs seem to be transported through a facilitated transporter across the plasma membrane. Despite the fact that to date no such protein has been cloned, drugs influencing the activity of such transporters have been found. The early generation of such compounds often did not have high efficacy and selectivity ; however, recently, a series of compounds have been developed that lack activity on cannabinoid receptor, eCB-degrading enzymes, and binding to fatty-acid binding protein. , Inhibition of such a transporter is expected to enhance the availability of extracellular eCBs, as eCBs cannot be transported into cells for degradation. Yet, it might also be argued that the export of eCBs is inhibited upon stimulated synthesis of eCBs, thereby increasing intracellular eCBs, leading, for example, to increased activation of mitochondrial CB 1 receptor (mtCB 1 ). Furthermore, because of the probable effect on both AEA and 2-AG levels, several different target receptors have to be considered (CB 1 /CB 2 receptor, TRPV1, peroxisome proliferator-activated receptor-γ [PPARγ]). Altogether, the inhibition of eCB membrane transporters would influence multiple cellular signaling systems, but represents a promising pharmacological target.

In summary, the diversity of ECS signaling molecules and their interactions with various receptors, together with signaling complexity of the receptor systems, makes pharmacological intervention of the ECS a challenging task, containing a considerable degree of unpredictability in the outcome of the biological effects in the whole organism.

Future directions and concluding remarks

Understanding of the brain’s ECS in its complexity at the mechanistic levels is very valuable for identifying promising disease states that can be optimally treated by modulating ECS activity. Given that eCB signaling is very widespread in the brain and intertwined with other signaling systems, the mechanistic insights will help minimize potentially unwanted side effects. Thus, it is important to understand ECS functions in the context of the entire organism. To this end, animal models, such as mice and rats, have been shown to be very suitable; owing to the evolutionary conservation of the ECS, these insights are expected to be transferrable to humans in many instances.

An important topic is the understanding of the integration of eCB signaling into the brain’s complex network. For example, excellent recent studies investigated pathways from amygdala to nucleus accumbens in the context of depressive-like behavior, and from amygdala to cortex regarding stress effects. Here, genetic manipulations of the CB 1 receptor functions were investigated in a pathway-specific manner, and in fact, very distinct functions of this receptor were uncovered. Thus, again, despite the very widespread presence of CB 1 receptor in the brain, its functions are amazingly specific. In further developments, high-resolution and specific genetic manipulations can be included, such as intersectional targeting of cells and circuits, and optogenetic-induced genetic manipulation. Furthermore, with the advent of the clustered regularly interspaced short palindromic sequences (CRISPR)/CRISPR-associated protein (Cas) technology, , the introduction of mutations into the mouse or rat genome, in particular point mutations that were characterized in in vitro structure-function analyses or are present in humans as small nucleotide polymorphisms (SNPs), opens the path to novel insights into the analysis of ECS components in the context of the entire organism.

Furthermore, SNPs in ECS components can also be investigated using human organoids generated from induced pluripotent stem cells, , with the additional potential of genomic manipulation using CRISPR/Cas technology and pharmacological interventions, for example, with cannabinoids, together with state-of-the-art analyses, such as single-cell RNA-seq. These approaches will further strengthen mechanistic insights into the roles of the ECS in psychiatric disorders.

Acknowledgments

There is no conflict of interest. This work was supported by the German Research Foundation DFG (CRC 1193 “Neurobiology of Resilience”; TRR-CRC 58 “Fear, Anxiety, Anxiety Disorders”). The author would like to thank Michael Plenikowski for his excellent work in the design of the illustrations.

REFERENCES

1. Piomelli D. The molecular logic of endocannabinoid signalling. Nat Rev Neurosci. 2003;4:873–884. [PubMed] []
2. Kano M, Ohno-Shosaku T, Hashimotodani Y, Uchigashima M, Watanabe M. Endocannabinoid-mediated control of synaptic transmission. Physiol Rev. 2009;89:309–380. [PubMed] []
3. Elphick MR. The evolution and comparative neurobiology of endocannabinoid signalling. Philos Trans R Soc Lond B Biol Sci. 2012;367:3201–3215. [PMC free article] [PubMed] []
4. Mechoulam R, Parker LA. The endocannabinoid system and the brain. Annu Rev Psychol. 2013;64:21–47. [PubMed] []
5. Ligresti A, De Petrocellis L, Di Marzo V. From phytocannabinoids to cannabinoid receptors and endocannabinoids: pleiotropic physiological and pathological roles through complex pharmacology. Physiol Rev. 2016;96:1593–1659. [PubMed] []
6. Cristino L, Bisogno T, Di Marzo V. Cannabinoids and the expanded endocannabinoid system in neurological disorders. Nat Rev Neurol. 2020;16:9–29. [PubMed] []
7. Lutz B, Marsicano G, Maldonado R, Hillard CJ. The endocannabinoid system in guarding against fear, anxiety and stress. Nat Rev Neurosci. 2015;16:705–718. [PMC free article] [PubMed] []
8. Ruiz de Azua I, Lutz B. Multiple endocannabinoid-mediated mechanisms in the regulation of energy homeostasis in brain and peripheral tissues. Cell Mol Life Sci. 2019;76:1341–1363. [PubMed] []
9. Muzik O, Diwadkar VA. Hierarchical control systems for the regulation of physiological homeostasis and affect: can their interactions modulate mood and anhedonia? Neurosci Biobehav Rev. 2019;105:251–261. [PubMed] []
10. Piazza PV, Cota D, Marsicano G. The CB1 Receptor as the cornerstone of exostasis. Neuron. 2017;93:1252–1274. [PubMed] []
11. Navarrete F, Garcia-Gutierrez MS, Jurado-Barba R, et al Endocannabinoid system components as potential biomarkers in psychiatry. Front Psychiatry. 2020;11:315. [PMC free article] [PubMed] []
12. Riebe CJ, Wotjak CT. Endocannabinoids and stress. Stress. 2011;14:384–397. [PubMed] []
13. Cinar R, Iyer MR, Kunos G. The therapeutic potential of second and third generation CB1R antagonists. Pharmacol Ther. 2020;208:107477. [PMC free article] [PubMed] []
14. Chicca A, Arena C, Manera C. Beyond the direct activation of cannabinoid receptors: new strategies to modulate the endocannabinoid system in CNS-related diseases. Recent Pat CNS Drug Discov. 2016;10:122–141. [PubMed] []
15. Tomaselli G, Vallee M. Stress and drug abuse-related disorders: the promising therapeutic value of neurosteroids focus on pregnenolone-progesterone-allopregnanolone pathway. Front Neuroendocrinol. 2019;55:100789. [PubMed] []
16. Maccarrone M, Bab I, Biro T, et al Endocannabinoid signaling at the periphery: 50 years after THC. Trends Pharmacol Sci. 2015;36:277–296. [PMC free article] [PubMed] []
17. Busquets-Garcia A, Gomis-Gonzalez M, Srivastava RK, et al Peripheral and central CB1 cannabinoid receptors control stress-induced impairment of memory consolidation. Proc Natl Acad Sci
U S A
.
2016;113:9904–9909. [PMC free article] [PubMed] []
18. Ruiz de Azua I, Mancini G, Srivastava RK, et al Adipocyte cannabinoid receptor CB1 regulates energy homeostasis and alternatively activated macrophages. J Clin Invest. 2017;127:4148–4162. [PMC free article] [PubMed] []
19. Bellocchio L, Soria-Gomez E, Quarta C, et al Activation of the sympathetic nervous system mediates hypophagic and anxiety-like effects of CB1 receptor blockade. Proc Natl Acad Sci U S A. 2013;110:4786–4791. [PMC free article] [PubMed] []
20. Suarez J, Rivera P, Aparisi RA, et al Adipocyte cannabinoid CB1 receptor deficiency alleviates high fat diet-induced memory deficit, depressive-like behavior, neuroinflammation and impairment in adult neurogenesis. Psychoneuroendocrinology. 2019;110:104418. [PubMed] []
21. Pertwee RG, Howlett AC, Abood ME, et al International union of basic and clinical pharmacology. LXXIX. Cannabinoid receptors and their ligands: beyond CB1 and CB2. Pharmacol Rev. 2010;62:588–631. [PMC free article] [PubMed] []
22. Wootten D, Christopoulos A, Marti-Solano M, Babu MM, Sexton PM. Mechanisms of signalling and biased agonism in G protein-coupled receptors. Nat Rev Mol Cell Biol. 2018;19:638–653. [PubMed] []
23. Han J, Kesner P, Metna-Laurent M, et al Acute cannabinoids impair working memory through astroglial CB1 receptor modulation of hippocampal LTD. Cell. 2012;148:1039–1050. [PubMed] []
24. Ilyasov AA, Milligan CE, Pharr EP, Howlett AC. The endocannabinoid system and oligodendrocytes in health and disease. Front Neurosci. 2018;12:733. [PMC free article] [PubMed] []
25. Araujo DJ, Tjoa K, Saijo K. The endocannabinoid system as a window into microglial biology and its relationship to autism. Front Cell Neurosci. 2019;13:424. [PMC free article] [PubMed] []
26. Cota D, Steiner MA, Marsicano G, et al Requirement of cannabinoid receptor type 1 for the basal modulation of hypothalamic-pituitary-adrenal axis function. Endocrinology. 2007;148:1574–1581. [PubMed] []
27. Marsicano G, Lutz B. Expression of the cannabinoid receptor CB1 in distinct neuronal subpopulations in the adult mouse forebrain. Eur J Neurosci. 1999;11:4213–4225. [PubMed] []
28. Katona I, Sperlágh B, Sík A, et al Presynaptically located CB1 cannabinoid receptors regulate GABA release from axon terminals of specific hippocampal interneurons. J Neurosci. 1999;19:4544–4558. [PMC free article] [PubMed] []
29. Zou S, Kumar U. Colocalization of cannabinoid receptor 1 with somatostatin and neuronal nitric oxide synthase in rat brain hippocampus. Brain Res. 2015;1622:114–126. [PubMed] []
30. Kamprath K, Romo-Parra H, Häring M, et al Short-term adaptation of conditioned fear responses through endocannabinoid signaling in the central amygdala. Neuropsychopharmacology. 2011;36:652–663. [PMC free article] [PubMed] []
31. Lange MD, Daldrup T, Remmers F, et al Cannabinoid CB1 receptors in distinct circuits of the extended amygdala determine fear responsiveness to unpredictable threat. Mol Psychiatry. 2017;22:1422–1430. [PubMed] []
32. Monory K, Massa F, Egertova M, et al The endocannabinoid system controls key epileptogenic circuits in the hippocampus. Neuron. 2006;51:455–466. [PMC free article] [PubMed] []
33. Katona I, Urban GM, Wallace M, et al Molecular composition of the endocannabinoid system at glutamatergic synapses. J Neurosci. 2006;26:5628–5637. [PMC free article] [PubMed] []
34. Gutierrez-Rodriguez A, Bonilla-Del RI, Puente N, et al Localization of the cannabinoid type-1 receptor in subcellular astrocyte compartments of mutant mouse hippocampus. Glia. 2018;66:1417–1431. [PubMed] []
35. Berrendero F, Sepe N, Ramos JA, Di Marzo V, Fernandez-Ruiz JJ. Analysis of cannabinoid receptor binding and mRNA expression and endogenous cannabinoid contents in the developing rat brain during late gestation and early postnatal period. Synapse. 1999;33:181–191. [PubMed] []
36. Molina-Holgado E, Vela JM, Arevalo-Martin A, et al Cannabinoids promote oligodendrocyte progenitor survival: involvement of cannabinoid receptors and phosphatidylinositol-3 kinase/Akt signaling. J Neurosci. 2002;22:9742–9753. [PMC free article] [PubMed] []
37. Gomez O, Sanchez-Rodriguez MA, Ortega-Gutierrez S, et al A basal tone of 2-arachidonoylglycerol contributes to early oligodendrocyte progenitor proliferation by activating phosphatidylinositol 3-kinase (PI3K)/AKT and the mammalian target of rapamycin (MTOR) pathways. J Neuroimmune Pharmacol. 2015;10:309–317. [PubMed] []
38. Aguado T, Palazuelos J, Monory K, et al The endocannabinoid system promotes astroglial differentiation by acting on neural progenitor cells. J Neurosci. 2006;26:1551–1561. [PMC free article] [PubMed] []
39. Zimmermann T, Maroso M, Beer A, et al Neural stem cell lineage-specific cannabinoid type-1 receptor regulates neurogenesis and plasticity in the adult mouse hippocampus. Cereb Cortex. 2018;28:4454–4471. [PMC free article] [PubMed] []
40. Lange MD, Daldrup T, Remmers F, et al Cannabinoid CB1 receptors in distinct circuits of the extended amygdala determine fear responsiveness to unpredictable threat. Mol Psychiatry. 2017;22:1422–1430. [PubMed] []
41. Hebert-Chatelain E, Desprez T, Serrat R, et al A cannabinoid link between mitochondria and memory. Nature. 2016;539:555–559. [PubMed] []
42. Robin LM, Oliveira da Cruz JF, Langlais VC, et al Astroglial CB1 receptors determine synaptic D-serine availability to enable recognition memory. Neuron. 2018;98:935–944. [PubMed] []
43. Tanaka M, Sackett S, Zhang Y. Endocannabinoid modulation of microglial phenotypes in neuropathology. Front Neurol. 2020;11:87. [PMC free article] [PubMed] []
44. Jordan CJ, Xi ZX. Progress in brain cannabinoid CB2 receptor research: from genes to behavior. Neurosci Biobehav Rev. 2019;98:208–220. [PMC free article] [PubMed] []
45. Van Sickle MD, Duncan M, Kingsley PJ, et al Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science. 2005;310:329–332. [PubMed] []
46. Stempel AV, Stumpf A, Zhang HY, et al Cannabinoid type 2 receptors mediate a cell type-specific plasticity in the hippocampus. Neuron. 2016;90:795–809. [PMC free article] [PubMed] []
47. Zhang HY, Gao M, Liu QR, et al Cannabinoid CB2 receptors modulate midbrain dopamine neuronal activity and dopamine-related behavior in mice. Proc Natl Acad Sci U S A. 2014;111(46):E5007–E5015. [PMC free article] [PubMed] []
48. Zhang HY, Gao M, Shen H, et al Expression of functional cannabinoid CB2 receptor in VTA dopamine neurons in rats. Addict Biol. 2017;22:752–765. [PMC free article] [PubMed] []
49. Yu SJ, Reiner D, Shen H, Wu KJ, Liu QR, Wang Y. Time-dependent protection of CB2 receptor agonist in stroke. PLoS One. 2015;10:e0132487. [PMC free article] [PubMed] []
50. Katona I, Freund TF. Multiple functions of endocannabinoid signaling in the brain. Annu Rev Neurosci. 2012;35:529–558. [PMC free article] [PubMed] []
51. Nyilas R, Dudok B, Urban GM, et al Enzymatic machinery for endocannabinoid biosynthesis associated with calcium stores in glutamatergic axon terminals. J Neurosci. 2008;28:1058–1063. [PMC free article] [PubMed] []
52. Egertova M, Cravatt BF, Elphick MR. Comparative analysis of fatty acid amide hydrolase and CB1 cannabinoid receptor expression in the mouse brain: evidence of a widespread role for fatty acid amide hydrolase in regulation of endocannabinoid signaling. Neuroscience. 2003;119:481–496. [PubMed] []
53. Gibson HE, Edwards JG, Page RS, Van Hook MJ, Kauer JA. TRPV1 channels mediate long-term depression at synapses on hippocampal interneurons. Neuron. 2008;57:746–759. [PMC free article] [PubMed] []
54. Guggenhuber S, Romo-Parra H, Bindila L, et al Impaired 2-AG signaling in hippocampal glutamatergic neurons: aggravation of anxiety-like behavior and unaltered seizure susceptibility. Int J Neuropsychopharmacol. 2015;19:pyv091. [PMC free article] [PubMed] []
55. Zimmermann T, Bartsch JC, Beer A, et al Impaired anandamide/palmitoylethanolamide signaling in hippocampal glutamatergic neurons alters synaptic plasticity, learning, and emotional responses. Neuropsychopharmacology. 2019;44:1377–1388. [PMC free article] [PubMed] []
56. Bacci A, Huguenard JR, Prince DA. Long-lasting self-inhibition of neocortical interneurons mediated by endocannabinoids. Nature. 2004;431:312–316. [PubMed] []
57. Marinelli S, Pacioni S, Cannich A, Marsicano G, Bacci A. Self-modulation of neocortical pyramidal neurons by endocannabinoids. Nat Neurosci. 2009;12:1488–1490. [PubMed] []
58. Maroso M, Szabo GG, Kim HK, et al Cannabinoid control of learning and memory through HCN channels. Neuron. 2016;89:1059–1073. [PMC free article] [PubMed] []
59. Thibault K, Carrel D, Bonnard D, et al Activation-dependent subcellular distribution patterns of CB1 cannabinoid receptors in the rat forebrain. Cereb Cortex. 2013;23:2581–2591. [PubMed] []
60. Fletcher-Jones A, Hildick KL, Evans AJ, Nakamura Y, Wilkinson KA, Henley JM. The C-terminal helix 9 motif in rat cannabinoid receptor type 1 regulates axonal trafficking and surface expression. Elife. 2019;8 [PMC free article] [PubMed] []
61. Wickert M, Hildick KL, Baillie GL, et al The F238L point mutation in the cannabinoid type 1 receptor enhances basal endocytosis via lipid rafts. Front Mol Neurosci. 2018;11:230. [PMC free article] [PubMed] []
62. Schneider M, Kasanetz F, Lynch DL, et al Enhanced functional activity of the cannabinoid type-1 receptor mediates adolescent behavior. J Neurosci. 2015;35:13975–13988. [PMC free article] [PubMed] []
63. Dudok B, Barna L, Ledri M, et al Cell-specific STORM super-resolution imaging reveals nanoscale organization of cannabinoid signaling. Nat Neurosci. 2015;18:75–86. [PMC free article] [PubMed] []
64. Imperatore R, Morello G, Luongo L, et al Genetic deletion of monoacylglycerol lipase leads to impaired cannabinoid receptor CB1R signaling and anxiety-like behavior. J Neurochem. 2015;135:799–813. [PubMed] []
65. Cravatt BF, Demarest K, Patricelli MP, et al Supersensitivity to anandamide and enhanced endogenous cannabinoid signaling in mice lacking fatty acid amide hydrolase. Proc Natl Acad Sci U S A. 2001;98:9371–9376. [PMC free article] [PubMed] []
66. Benard G, Massa F, Puente N, et al Mitochondrial CB1 receptors regulate neuronal energy metabolism. Nat Neurosci. 2012;15:558–564. [PubMed] []
67. Jimenez-Blasco D, Busquets-Garcia A, Hebert-Chatelain E, et al Glucose metabolism links astroglial mitochondria to cannabinoid effects. Nature. 2020;583(7817):603–608. [PubMed] []
68. Alger BE, Kim J. Supply and demand for endocannabinoids. Trends Neurosci. 2011;34:304–315. [PMC free article] [PubMed] []
69. Kenakin T. Biased receptor signaling in drug discovery. Pharmacol Rev. 2019;71:267–315. [PubMed] []
70. Zhao P, Furness SGB. The nature of efficacy at G protein-coupled receptors. Biochem Pharmacol. 2019;170:113647. [PubMed] []
71. Laprairie RB, Bagher AM, Denovan-Wright EM. Cannabinoid receptor ligand bias: implications in the central nervous system. Curr Opin Pharmacol. 2017;32:32–43. [PubMed] []
72. Al-Zoubi R, Morales P, Reggio PH. Structural Insights into CB1 receptor biased signaling. Int J Mol Sci. 2019;20(8):1837. [PMC free article] [PubMed] []
73. Borroto-Escuela DO, Fuxe K. Oligomeric receptor complexes and their allosteric receptor-receptor interactions in the plasma membrane represent a new biological principle for integration of signals in the CNS. Front Mol Neurosci. 2019;12:230. [PMC free article] [PubMed] []
74. Borroto-Escuela DO, Carlsson J, Ambrogini P, et al Understanding the role of GPCR heteroreceptor complexes in modulating the brain networks in health and disease. Front Cell Neurosci. 2017;11:37. [PMC free article] [PubMed] []
75. Wager-Miller J, Westenbroek R, Mackie K. Dimerization of G protein-coupled receptors: CB1 cannabinoid receptors as an example. Chem Phys Lipids. 2002;121:83–89. [PubMed] []
76. Callen L, Moreno E, Barroso-Chinea P, et al Cannabinoid receptors CB1 and CB2 form functional heteromers in brain. J Biol Chem. 2012;287:20851–20865. [PMC free article] [PubMed] []
77. Navarro G, Varani K, Reyes-Resina I, et al Cannabigerol action at cannabinoid CB1 and CB2 receptors and at CB1-CB2 heteroreceptor complexes. Front Pharmacol. 2018;9:632. [PMC free article] [PubMed] []
78. Bagher AM, Laprairie RB, Kelly ME, Denovan-Wright EM. Antagonism of dopamine receptor 2 long affects cannabinoid receptor 1 signaling in a cell culture model of striatal medium spiny projection neurons. Mol Pharmacol. 2016;89:652–666. [PubMed] []
79. Bagher AM, Laprairie RB, Toguri JT, Kelly MEM, Denovan-Wright EM. Bidirectional allosteric interactions between cannabinoid receptor 1 (CB1) and dopamine receptor 2 long (D2L) heterotetramers. Eur J Pharmacol. 2017;813:66–83. [PubMed] []
80. Vinals X, Moreno E, Lanfumey L, et al Cognitive impairment induced by delta9-tetrahydrocannabinol occurs through heteromers between cannabinoid CB1 and serotonin 5-HT2A receptors. PLoS Biol. 2015;13:e1002194. [PMC free article] [PubMed] []
81. Moreno E, Chiarlone A, Medrano M, et al Singular location and signaling profile of adenosine A2A-cannabinoid CB1 receptor heteromers in the dorsal striatum. Neuropsychopharmacology. 2018;43:964–977. [PMC free article] [PubMed] []
82. Navarro G, Varani K, Reyes-Resina I, et al Cannabigerol action at cannabinoid CB1 and CB2 receptors and at CB1-CB2 heteroreceptor complexes. Front Pharmacol. 2018;9:632. [PMC free article] [PubMed] []
83. Blasco-Benito S, Moreno E, Seijo-Vila M, et al Therapeutic targeting of HER2-CB2R heteromers in HER2-positive breast cancer. Proc Natl Acad Sci
U S A
.
2019;116:3863–3872. [PMC free article] [PubMed] []
84. Glass M, Felder CC. Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors augments cAMP accumulation in striatal neurons: evidence for a Gs linkage to the CB1 receptor. J Neurosci. 1997;17:5327–5333. [PMC free article] [PubMed] []
85. Kearn CS, Blake-Palmer K, Daniel E, Mackie K, Glass M. Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors enhances heterodimer formation: a mechanism for receptor crosstalk? Mol Pharmacol. 2005;67:1697–1704. [PubMed] []
86. Caballero-Floran RN, Conde-Rojas I, Oviedo CA, et al Cannabinoid-induced depression of synaptic transmission is switched to stimulation when dopaminergic tone is increased in the globus pallidus of the rodent. Neuropharmacology. 2016;110:407–418. [PubMed] []
87. Navarrete M, Araque A. Endocannabinoids mediate neuron-astrocyte communication. Neuron. 2008;57:883–893. [PubMed] []
88. Mattheus T, Kukla K, Zimmermann T, Tenzer S, Lutz B. Cell type-specific tandem affinity purification of the mouse hippocampal CB1 receptor-associated proteome. J Proteome Res. 2016;15:3585–3601. [PubMed] []
89. Diez-Alarcia R, Ibarra-Lecue I, Lopez-Cardona AP, et al Biased agonism of three different cannabinoid receptor agonists in mouse brain cortex. Front Pharmacol. 2016;7:415. [PMC free article] [PubMed] []
90. Steindel F, Lerner R, Häring M, et al Neuron-type specific cannabinoid-mediated G protein signalling in mouse hippocampus. J Neurochem. 2013;124:795–807. [PubMed] []
91. O’Brien JB, Wilkinson JC, Roman DL. Regulator of G-protein signaling (RGS) proteins as drug targets: progress and future potentials. J Biol Chem. 2019;294:18571–18585. [PMC free article] [PubMed] []
92. Song C, Anderson GR, Sutton LP, Dao M, Martemyanov KA. Selective role of RGS9-2 in regulating retrograde synaptic signaling of indirect pathway medium spiny neurons in dorsal striatum. J Neurosci. 2018;38:7120–7131. [PMC free article] [PubMed] []
93. Raehal KM, Bohn LM. The role of beta-arrestin2 in the severity of antinociceptive tolerance and physical dependence induced by different opioid pain therapeutics. Neuropharmacology. 2011;60:58–65. [PMC free article] [PubMed] []
94. Gurevich VV, Gurevich EV. GPCR signaling regulation: the role of GRKs and arrestins. Front Pharmacol. 2019;10:125. [PMC free article] [PubMed] []
95. Delgado-Peraza F, Ahn KH, Nogueras-Ortiz C, et al Mechanisms of biased b-arrestin-mediated signaling downstream from the cannabinoid 1 receptor. Mol Pharmacol. 2016;89:618–629. [PMC free article] [PubMed] []
96. Morgan DJ, Davis BJ, Kearn CS, et al Mutation of putative GRK phosphorylation sites in the cannabinoid receptor 1 (CB1R) confers resistance to cannabinoid tolerance and hypersensitivity to cannabinoids in mice. J Neurosci. 2014;34:5152–5163. [PMC free article] [PubMed] []
97. Younts TJ, Monday HR, Dudok B, et al Presynaptic protein synthesis is required for long-term plasticity of GABA release. Neuron. 2016;92:479–492. [PMC free article] [PubMed] []
98. Puighermanal E, Marsicano G, Busquets-Garcia A, Lutz B, Maldonado R, Ozaita A. Cannabinoid modulation of hippocampal long-term memory is mediated by mTOR signaling. Nat Neurosci. 2009;12:1152–1158. [PubMed] []
99. Guggenhuber S, Alpar A, Chen R, et al Cannabinoid receptor-interacting protein Crip1a modulates CB1 receptor signaling in mouse hippocampus. Brain Struct Funct. 2016;221:2061–2074. [PubMed] []
100. Booth WT, Walker NB, Lowther WT, Howlett AC. Cannabinoid receptor interacting protein 1a CRIP1a: function and structure. Molecules. 2019;24:3672. [PMC free article] [PubMed] []
101. Ruehle S, Wager-Miller J, Straiker A, et al Discovery and characterization of two novel CB1 receptor splice variants with modified N-termini in mouse. J Neurochem. 2017;142:521–533. [PMC free article] [PubMed] []
102. Liu QR, Huang NS, Qu H, et al Identification of novel mouse and rat CB1R isoforms and in silico modeling of human CB1R for peripheral cannabinoid therapeutics. Acta Pharmacol Sin. 2019;40:387–397. [PMC free article] [PubMed] []
103. Ryberg E, Vu HK, Larsson N, et al Identification and characterisation of a novel splice variant of the human CB1 receptor. FEBS Lett. 2005;579:259–264. [PubMed] []
104. Straiker A, Wager-Miller J, Hutchens J, Mackie K. Differential signalling in human cannabinoid CB1 receptors and their splice variants in autaptic hippocampal neurones. Br J Pharmacol. 2012;165:2660–2671. [PMC free article] [PubMed] []
105. Bagher AM, Laprairie RB, Kelly ME, Denovan-Wright EM. Co-expression of the human cannabinoid receptor coding region splice variants (hCB1) affects the function of hCB1 receptor complexes. Eur J Pharmacol. 2013;721:341–354. [PubMed] []
106. Rapino C, Castellucci A, Lizzi AR, et al Modulation of endocannabinoid-binding receptors in human neuroblastoma cells by tunicamycin. Molecules. 2019;24:1432. [PMC free article] [PubMed] []
107. Bond AM, Ming GL, Song H. Adult mammalian neural stem cells and neurogenesis: five decades later. Cell Stem Cell. 2015;17:385–395. [PMC free article] [PubMed] []
108. Toda T, Parylak SL, Linker SB, Gage FH. The role of adult hippocampal neurogenesis in brain health and disease. Mol Psychiatry. 2019;24:67–87. [PMC free article] [PubMed] []
109. Miller SM, Sahay A. Functions of adult-born neurons in hippocampal memory interference and indexing. Nat Neurosci. 2019;22:1565–1575. [PMC free article] [PubMed] []
110. Maccarrone M, Guzman M, Mackie K, Doherty P, Harkany T. Programming of neural cells by endocannabinoids: from physiological rules to emerging therapies. Nat Rev Neurosci. 2014;15:786–801. [PMC free article] [PubMed] []
111. Prenderville JA, Kelly AM, Downer EJ. The role of cannabinoids in adult neurogenesis. Br J Pharmacol. 2015;172:3950–3963. [PMC free article] [PubMed] []
112. Oddi S, Scipioni L, Maccarrone M. Endocannabinoid system and adult neurogenesis: a focused review. Curr Opin Pharmacol. 2019;50:25–32. [PubMed] []
113. Gao Y, Vasilyev DV, Goncalves MB, et al Loss of retrograde endocannabinoid signaling and reduced adult neurogenesis in diacylglycerol lipase knock-out mice. J Neurosci. 2010;30:2017–2024. [PMC free article] [PubMed] []
114. Jenniches I, Ternes S, Albayram O, et al Anxiety, stress, and fear response in mice with reduced endocannabinoid levels. Biol Psychiatry. 2016;79:858–868. [PubMed] []
115. Mensching L, Djogo N, Keller C, Rading S, Karsak M. Stable adult hippocampal neurogenesis in cannabinoid receptor CB2 deficient mice. Int J Mol Sci. 2019;20(15):3759. [PMC free article] [PubMed] []
116. Palazuelos J, Ortega Z, Diaz-Alonso J, Guzman M, Galve-Roperh I. CB2 cannabinoid receptors promote neural progenitor cell proliferation via mTORC1 signaling. J Biol Chem. 2012;287:1198–1209. [PMC free article] [PubMed] []
117. Zhang Z, Wang W, Zhong P, et al Blockade of 2-arachidonoylglycerol hydrolysis produces antidepressant-like effects and enhances adult hippocampal neurogenesis and synaptic plasticity. Hippocampus. 2015;25:16–26. [PMC free article] [PubMed] []
118. Rodrigues RS, Ribeiro FF, Ferreira F, Vaz SH, Sebastiao AM, Xapelli S. Interaction between cannabinoid type 1 and type 2 receptors in the modulation of subventricular zone and dentate gyrus neurogenesis. Front Pharmacol. 2017;8:516. [PMC free article] [PubMed] []
119. Jiang W, Zhang Y, Xiao L, et al Cannabinoids promote embryonic and adult hippocampus neurogenesis and produce anxiolytic- and antidepressant-like effects. J Clin Invest. 2005;115:3104–3116. [PMC free article] [PubMed] []
120. Wolf SA, Bick-Sander A, Fabel K, et al Cannabinoid receptor CB1 mediates baseline and activity-induced survival of new neurons in adult hippocampal neurogenesis. Cell Commun Signal. 2010;8:12. [PMC free article] [PubMed] []
121. Campos AC, Ortega Z, Palazuelos J, et al The anxiolytic effect of cannabidiol on chronically stressed mice depends on hippocampal neurogenesis: involvement of the endocannabinoid system. Int J Neuropsychopharmacol. 2013;16:1407–1419. [PubMed] []
122. Leishman E, Manchanda M, Thelen R, Miller S, Mackie K, Bradshaw HB. Cannabidiol’s upregulation of N-acyl ethanolamines in the central nervous system requires N-acyl phosphatidyl ethanolamine-specific phospholipase D. Cannabis Cannabinoid Res. 2018;3:228–241. [PMC free article] [PubMed] []
123. Hill MN, Titterness AK, Morrish AC, et al Endogenous cannabinoid signaling is required for voluntary exercise-induced enhancement of progenitor cell proliferation in the hippocampus. Hippocampus. 2010;20:513–523. [PMC free article] [PubMed] []
124. Dubreucq S, Koehl M, Abrous DN, Marsicano G, Chaouloff F. CB1 receptor deficiency decreases wheel-running activity: consequences on emotional behaviours and hippocampal neurogenesis. Exp Neurol. 2010;224:106–113. [PubMed] []
125. Navarro-Romero A, Vazquez-Oliver A, Gomis-Gonzalez M, et al Cannabinoid type-1 receptor blockade restores neurological phenotypes in two models for Down syndrome. Neurobiol Dis. 2019;125:92–106. [PubMed] []
126. Kirdajova D, Anderova M. NG2 cells and their neurogenic potential. Curr Opin Pharmacol. 2019;50:53–60. [PubMed] []
127. Kuhn S, Gritti L, Crooks D, Dombrowski Y. Oligodendrocytes in development, myelin generation and beyond. Cells. 2019;8:1424. [PMC free article] [PubMed] []
128. Gomez O, Sanchez-Rodriguez A, Le M, Sanchez-Caro C, Molina-Holgado F, Molina-Holgado E. Cannabinoid receptor agonists modulate oligodendrocyte differentiation by activating PI3K/Akt and the mammalian target of rapamycin mTOR. pathways. Br J Pharmacol. 2011;163:1520–1532. [PMC free article] [PubMed] []
129. Bernal-Chico A, Canedo M, Manterola A, et al Blockade of monoacylglycerol lipase inhibits oligodendrocyte excitotoxicity and prevents demyelination in vivo. Glia. 2015;63:163–176. [PubMed] []
130. Tomas-Roig J, Agbemenyah HY, Celarain N, Quintana E, Ramio-Torrenta L, Havemann-Reinecke U. Dose-dependent effect of cannabinoid WIN-55,212-2 on myelin repair following a demyelinating insult. Sci Rep. 2020;10:590. [PMC free article] [PubMed] []
131. Garcia-Arencibia M, Molina-Holgado E, Molina-Holgado F. Effect of endocannabinoid signalling on cell fate: life, death, differentiation and proliferation of brain cells. Br J Pharmacol. 2019;176:1361–1369. [PMC free article] [PubMed] []
132. Hurd YL, Manzoni OJ, Pletnikov MV, Lee FS, Bhattacharyya S, Melis M. Cannabis and the developing brain: insights into its long-lasting effects. J Neurosci. 2019;39:8250–8258. [PMC free article] [PubMed] []
133. Frau R, Miczan V, Traccis F, et al Prenatal THC exposure produces a hyperdopaminergic phenotype rescued by pregnenolone. Nat Neurosci. 2019;22:1975–1985. [PMC free article] [PubMed] []
134. de Salas-Quiroga A, Garcia-Rincon D, Gomez-Dominguez D, et al Long-term hippocampal interneuronopathy drives sex-dimorphic spatial memory impairment induced by prenatal THC exposure. Neuropsychopharmacology. 2020;45:877–886. [PMC free article] [PubMed] []
135. Rodrigues RS, Lourenco DM, Paulo SL, et al Cannabinoid actions on neural stem cells: implications for pathophysiology. Molecules. 2019;24:1350. [PMC free article] [PubMed] []
136. Sipe JC, Chiang K, Gerber AL, Beutler E, Cravatt BF. A missense mutation in human fatty acid amide hydrolase associated with problem drug use. Proc Natl Acad Sci U S A. 2002;99:8394–8399. [PMC free article] [PubMed] []
137. Dincheva I, Drysdale AT, Hartley CA, et al FAAH genetic variation enhances fronto-amygdala function in mouse and human. Nat Commun. 2015;6:6395. [PMC free article] [PubMed] []
138. Mayo LM, Asratian A, Linde J, et al Protective effects of elevated anandamide on stress and fear-related behaviors: translational evidence from humans and mice. Mol Psychiatry. 2020;25:993–1005. [PubMed] []
139. Gee DG, Fetcho RN, Jing D, et al Individual differences in frontolimbic circuitry and anxiety emerge with adolescent changes in endocannabinoid signaling across species. Proc Natl Acad Sci U S A. 2016;113:4500–4505. [PMC free article] [PubMed] []
140. Habib AM, Okorokov AL, Hill MN, et al Microdeletion in a FAAH pseudogene identified in a patient with high anandamide concentrations and pain insensitivity. Br J Anaesth. 2019;123:e249–e253. [PMC free article] [PubMed] []
141. Kong X, Miao Q, Lu X, et al The association of endocannabinoid receptor genes (CNR1 and CNR2) polymorphisms with depression: a meta-analysis. Medicine (Baltimore) 2019;98:e17403. [PMC free article] [PubMed] []
142. Lazary J, Eszlari N, Juhasz G, Bagdy G. A functional variant of CB2 receptor gene interacts with childhood trauma and FAAH gene on anxious and depressive phenotypes. J Affect Disord. 2019;257:716–722. [PubMed] []
143. Hillard CJ. Circulating endocannabinoids: from whence do they come and where are they going? Neuropsychopharmacology. 2018;43:155–172. [PMC free article] [PubMed] []
144. Muller-Vahl KR, Bindila L, Lutz B, et al Cerebrospinal fluid endocannabinoid levels in Gilles de la Tourette syndrome. Neuropsychopharmacology. 2020;45:1323–1329. [PMC free article] [PubMed] []
145. Kranaster L, Hoyer C, Mindt S, et al The novel seizure quality index for the antidepressant outcome prediction in electroconvulsive therapy: association with biomarkers in the cerebrospinal fluid. Eur Arch Psychiatry Clin Neurosci. 2019 Nov 23 doi: 10.1007/s00406-019-01086-x. Epub ahead of print. [PubMed] [CrossRef] []
146. Sloan ME, Grant CW, Gowin JL, Ramchandani VA, Le Foll B. Endocannabinoid signaling in psychiatric disorders: a review of positron emission tomography studies. Acta Pharmacol Sin. 2019;40:342–350. [PMC free article] [PubMed] []
147. Russo EB. Cannabis therapeutics and the future of neurology. Front Integr Neurosci. 2018;12:51. [PMC free article] [PubMed] []
148. Sarris J, Sinclair J, Karamacoska D, Davidson M, Firth J. Medicinal cannabis for psychiatric disorders: a clinically-focused systematic review. BMC Psychiatry. 2020;20:24. [PMC free article] [PubMed] []
149. Black N, Stockings E, Campbell G, et al Cannabinoids for the treatment of mental disorders and symptoms of mental disorders: a systematic review and meta-analysis. Lancet Psychiatry. 2019;6:995–1010. [PMC free article] [PubMed] []
150. Nomura DK, Morrison BE, Blankman JL, et al Endocannabinoid hydrolysis generates brain prostaglandins that promote neuroinflammation. Science. 2011;334:809–813. [PMC free article] [PubMed] []
151. Gomes I, Grushko JS, Golebiewska U, et al Novel endogenous peptide agonists of cannabinoid receptors. FASEB J. 2009;23:3020–3029. [PMC free article] [PubMed] []
152. Bauer M, Chicca A, Tamborrini M, et al Identification and quantification of a new family of peptide endocannabinoids (Pepcans) showing negative allosteric modulation at CB1 receptors. J Biol Chem. 2012;287:36944–36967. [PMC free article] [PubMed] []
153. Wei F, Zhao L, Jing Y. Signaling molecules targeting cannabinoid receptors: hemopressin and related peptides. Neuropeptides. 2020;79:101998. [PubMed] []
154. Christensen R, Kristensen PK, Bartels EM, Bliddal H, Astrup A. Efficacy and safety of the weight-loss drug rimonabant: a meta-analysis of randomised trials. Lancet. 2007;370:1706–1713. [PubMed] []
155. Amato G, Khan NS, Maitra R. A patent update on cannabinoid receptor 1 antagonists (2015-2018) Expert Opin Ther Pat. 2019;29:261–269. [PMC free article] [PubMed] []
156. Khurana L, Mackie K, Piomelli D, Kendall DA. Modulation of CB1 cannabinoid receptor by allosteric ligands: pharmacology and therapeutic opportunities. Neuropharmacology. 2017;124:3–12. [PMC free article] [PubMed] []
157. Slivicki RA, Xu Z, Kulkarni PM, et al Positive allosteric modulation of cannabinoid receptor type 1 suppresses pathological pain without producing tolerance or dependence. Biol Psychiatry. 2018;84:722–733. [PMC free article] [PubMed] []
158. Slivicki RA, Iyer V, Mali SS, et al Positive allosteric modulation of CB1 cannabinoid receptor signaling enhances morphine antinociception and attenuates morphine tolerance without enhancing morphine-induced dependence or reward. Front Mol Neurosci. 2020;13:54. [PMC free article] [PubMed] []
159. Tomaselli G, Vallee M. Stress and drug abuse-related disorders: the promising therapeutic value of neurosteroids focus on pregnenolone-progesterone-allopregnanolone pathway. Front Neuroendocrinol. 2019;55:100789. [PubMed] []
160. Petrucci V, Chicca A, Glasmacher S, et al Pepcan-12 (RVD-hemopressin) is a CB2 receptor positive allosteric modulator constitutively secreted by adrenals and in liver upon tissue damage. Sci Rep. 2017;7:9560. [PMC free article] [PubMed] []
161. Iyer MR, Cinar R, Katz A, et al Design, synthesis, and biological evaluation of novel, non-brain-penetrant, hybrid cannabinoid CB1R inverse agonist/inducible nitric oxide synthase (iNOS) inhibitors for the treatment of liver fibrosis. J Med Chem. 2017;60:1126–1141. [PubMed] []
162. Xu B, Xiao J, Xu K, et al VF-13, a chimeric peptide of VD-hemopressin(α) and neuropeptide VF, produces potent antinociception with reduced cannabinoid-related side effects. Neuropharmacology. 2020:108178. [PubMed] []
163. Morales P, Goya P, Jagerovic N. Emerging strategies targeting CB2 cannabinoid receptor: biased agonism and allosterism. Biochem Pharmacol. 2018;157:8–17. [PubMed] []
164. Fazio D, Criscuolo E, Piccoli A, Barboni B, Fezza F, Maccarrone M. Advances in the discovery of fatty acid amide hydrolase inhibitors: what does the future hold? Expert Opin Drug Discov. 2020:1–14. [PubMed] []
165. Gil-Ordonez A, Martin-Fontecha M, Ortega-
Gutierrez S, Lopez-Rodriguez ML. Monoacylglycerol lipase (MAGL) as a promising therapeutic target. Biochem Pharmacol. 2018;157:18–32. [PubMed] []
166. Bedse G, Hill MN, Patel S. 2-Arachidonoylglycerol modulation of anxiety and stress adaptation: from grass roots to novel therapeutics. Biol Psychiatry. 2020;S0006-3223(20):30049–4. [PMC free article] [PubMed] []
167. Ren SY, Wang ZZ, Zhang Y, Chen NH. Potential application of endocannabinoid system agents in neuropsychiatric and neurodegenerative diseases-focusing on FAAH/MAGL inhibitors. Acta Pharmacol Sin. 2020 Mar 18 doi: 10.1038/s41401-020-0385-7. Epub ahead of print. [PMC free article] [PubMed] [CrossRef] []
168. Vozella V, Ahmed F, Choobchian P, et al Pharmacokinetics, pharmacodynamics and safety studies on URB937, a peripherally restricted fatty acid amide hydrolase inhibitor, in rats. J Pharm Pharmacol. 2019;71:1762–1773. [PMC free article] [PubMed] []
169. D’Souza DC, Cortes-Briones J, Creatura G, et al Efficacy and safety of a fatty acid amide hydrolase inhibitor (PF-04457845) in the treatment of cannabis withdrawal and dependence in men: a double-blind, placebo-controlled, parallel group, phase 2a single-site randomised controlled trial. Lancet Psychiatry. 2019;6:35–45. [PubMed] []
170. Mayo LM, Asratian A, Linde J, et al Elevated anandamide, enhanced recall of fear extinction, and attenuated stress responses following inhibition of fatty acid amide hydrolase: a randomized, controlled experimental medicine trial. Biol Psychiatry. 2020;87:538–547. [PubMed] []
171. van Esbroeck ACM, Janssen APA, Cognetta AB III, et al Activity-based protein profiling reveals off-target proteins of the FAAH inhibitor BIA 10-2474. Science. 2017;356:1084–1087. [PMC free article] [PubMed] []
172. Mock ED, Mustafa M, Gunduz-Cinar O, et al Discovery of a NAPE-PLD inhibitor that modulates emotional behavior in mice. Nat Chem Biol. 2020;16:667–675. [PMC free article] [PubMed] []
173. Nicolussi S, Gertsch J. Endocannabinoid transport revisited. Vitam Horm. 2015;98:441–485. [PubMed] []
174. Chicca A, Nicolussi S, Bartholomaus R, et al Chemical probes to potently and selectively inhibit endocannabinoid cellular reuptake. Proc Natl Acad Sci U S A. 2017;114:E5006–E5015. [PMC free article] [PubMed] []
175. Reynoso-Moreno I, Chicca A, Flores-Soto ME, Viveros-Paredes JM, Gertsch J. The endocannabinoid reuptake inhibitor WOBE437 is orally bioavailable and exerts indirect polypharmacological effects via different endocannabinoid receptors. Front Mol Neurosci. 2018;11:180. [PMC free article] [PubMed] []
176. Shen CJ, Zheng D, Li KX, et al Cannabinoid CB1 receptors in the amygdalar cholecystokinin glutamatergic afferents to nucleus accumbens modulate depressive-like behavior. Nat Med. 2019;25:337–349. [PubMed] []
177. Marcus DJ, Bedse G, Gaulden AD, et al Endocannabinoid signaling collapse mediates stress-induced amygdalo-cortical strengthening. Neuron. 2020;105:1062–1076. [PMC free article] [PubMed] []
178. Hermann M, Stillhard P, Wildner H, et al Binary recombinase systems for high-resolution conditional mutagenesis. Nucleic Acids Res. 2014;42:3894–3907. [PMC free article] [PubMed] []
179. Madisen L, Garner AR, Shimaoka D, et al Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance. Neuron. 2015;85:942–958. [PMC free article] [PubMed] []
180. Karimova M, Baker O, Camgoz A, Naumann R, Buchholz F, Anastassiadis K. A single reporter mouse line for Vika, Flp, Dre, and Cre-recombination. Sci Rep. 2018;8:14453. [PMC free article] [PubMed] []
181. Yao S, Yuan P, Ouellette B, et al RecV recombinase system for in vivo targeted optogenomic modifications of single cells or cell populations. Nat Methods. 2020;17:422–429. [PMC free article] [PubMed] []
182. Shalem O, Sanjana NE, Zhang F. High-throughput functional genomics using CRISPR-Cas9. Nat Rev Genet. 2015;16:299–311. [PMC free article] [PubMed] []
183. Navabpour S, Kwapis JL, Jarome TJ. A neuroscientist’s guide to transgenic mice and other genetic tools. Neurosci Biobehav Rev. 2020;108:732–748. [PMC free article] [PubMed] []
184. Arlotta P, Pasca SP. Cell diversity in the human cerebral cortex: from the embryo to brain organoids. Curr Opin Neurobiol. 2019;56:194–198. [PubMed] []
185. Sidhaye J, Knoblich JA. Brain organoids: an ensemble of bioassays to investigate human neurodevelopment and disease. Cell Death Differ. 2020 Jun 1 doi: 10.1038/s41418-020-0566-4. Epub ahead of print. [PMC free article] [PubMed] [CrossRef] []
186. Fischer J, Heide M, Huttner WB. Genetic modification of brain organoids. . Front Cell Neurosci. 2019;13:558. [PMC free article] [PubMed] []
187. Ao Z, Cai H, Havert DJ, et al One-stop microfluidic assembly of human brain organoids to model prenatal cannabis exposure. Anal Chemi. 2020;92:4630–4638. [PubMed] []
188. Velasco S, Kedaigle AJ, Simmons SK, et al Individual brain organoids reproducibly form cell diversity of the human cerebral cortex. Nature. 2019;570:523–527. [PMC free article] [PubMed] []

Articles from Dialogues in Clinical Neuroscience are provided here courtesy of Taylor & Francis