Canna~Fangled Abstracts

CB1 Cannabinoid Receptors and their Associated Proteins

By August 12, 2013No Comments

pub med big
Logo of nihpa

Curr Med Chem. Author manuscript; available in PMC 2011 September 25.
Published in final edited form as:
PMCID:PMC3179980
NIHMSID:NIHMS314516

CB1 Cannabinoid Receptors and their Associated Proteins

The publisher’s final edited version of this article is available at Curr Med Chem
See other articles in PMC that cite the published article.

Abstract

CB1 receptors are G-protein coupled receptors (GPCRs) abundant in neurons, in which they modulate neurotransmission. The CB1receptor influence on memory and learning is well recognized, and disease states associated with CB1 receptors are observed in addiction disorders, motor dysfunction, schizophrenia, and in bipolar, depression, and anxiety disorders. Beyond the brain, CB1receptors also function in liver and adipose tissues, vascular as well as cardiac tissue, reproductive tissues and bone. Signal transduction by CB1 receptors occurs through interaction with Gi/o proteins to inhibit adenylyl cyclase, activate mitogen-activated protein kinases (MAPK), inhibit voltage-gated Ca2+channels, activate K+ currents (Kir), and influence Nitric Oxide (NO) signaling. CB1 receptors are observed in internal organelles as well as plasma membrane. β-Arrestins, adaptor protein AP-3, and G-protein receptor-associated sorting protein 1 (GASP1) modulate cellular trafficking. Cannabinoid Receptor Interacting Protein 1a (CRIP1a) is an accessory protein whose function has not been delineated. Factor Associated with Neutral sphingomyelinase (FAN) regulates ceramide signaling. Such diversity in cellular signaling and modulation by interacting proteins suggests that agonists and allosteric modulators could be developed to specifically regulate unique, cell type-specific responses.

Keywords: anandamide (arachidonylethanolamide), AP-3, 2-arachidonoylglycerol (2-AG), •-arrestin, CP55940, CRIP-1a, endocannabinoids, G-protein coupled receptors (GPCRs), GASP, rimonabant (SR141716), •9-tetrahydrocannabinol (THC), WIN55212-2

CB1 cannabinoid receptor physiology, pathology and pharmacology

CB1 receptors are G-protein coupled receptors (GPCRs) initially described as having such great abundance in brain tissue and neuronal cells, that the much lower levels in other tissues appeared to be of lesser significance [1]. In the early years of cannabinoid receptor characterization, pharmacological investigation of antinociception was emphasized because of the potential for a non-opioid analgesic, and pharmaceutical industry drug development programs were based upon this therapeutic opportunity [2,3]. This clinical application was thwarted as a result of the untoward side effects of memory impairment, cognitive dysfunction, and sedation [4,5], but has recently been reconsidered [6]. Although pharmaceutical development included treatment of nausea in cancer chemotherapy [7], the only CB1agonists to succeed to market have been dronabinol (also known as •9-tetrahydrocannabinol (THC)) and nabilone [810].

Current research has made significant progress in clarifying the role of CB1 receptors in such important aspects of cognition as reversal learning by which memory traces can be attenuated in the process of developing new patterns in response to novel relevant stimuli [11]. Highly coupled to memory and learning, CB1receptors play a role in addiction processes that extend from reinforcement of high-fat, sweet food intake [12] to dependence on reinforcing drugs and alcohol [1315]. CB1 receptors abundant in the basal ganglia are implicated in motor dysfunction including Huntington’s disease and amyotrophic lateral sclerosis [16]. Other central nervous system diseases that might involve a CB1 receptor component include schizophrenia, and bipolar, depression, and anxiety disorders [17,18]. To understand the mechanisms underlying these behavioral responses, we now know that CB1 receptors are not only involved in neuronal synaptic remodeling as learning takes place, but also in neurogenesis, neuronal migration and appropriate axonal targeting and synaptogenesis as the brain develops [1921].

The existence of CB1 receptors in liver and adipose tissues was not appreciated until clinical trials of the CB1 antagonist rimonabant (also known as SR141716A) for the treatment of obesity and dyslipidemias [22,23] uncovered the anomaly that decreases in adiposity in humans could not be entirely explained by a central nervous system effect of the drug to curtail food intake [24,25]. Current studies are focusing on both CB1 and CB2 receptors in hepatic metabolism and pathologies [26,27]. The interest in metabolic syndrome led to the expansion of studies of CB1receptors on vascular as well as cardiac tissue [28] [29,30]. CB1receptors are also relevant to many other organ systems including reproductive tissues [31], bone [32], and skin [33].

To summarize the relevant pharmacology, synthetic analogs of THC include classical cannabinoid agonists such as the highly potent and efficacious HU210, and the “non-classical” highly potent and efficacious agonists have a modified ring structure, such as CP55940 and CP55244 [34,35]. The aminoalkylindole CB1agonists include the full agonist WIN55212-2. The endogenous ligands for the CB1 receptor are a family of eicosanoid derivatives referred to as endocannabinoids, most notably 2-arachidonoylglycerol (2-AG) which behaves as a full agonist, and anandamide (arachidonylethanolamide), which behaves as a partial agonist. These CB1 agonists have been reviewed previously [3437], as well as in the present series (Mechoulam, 2010; Pertwee, 2010). CB1 antagonists include the high-affinity aryl pyrazole, rimonabant, formerly known as SR141716, as well as other CB1-selective antagonists of similar structure including AM251, taranabant, VCHSR and AM4113 [35,3842]. LY320135 is a CB1 antagonist having a structure that resembles the aminoalkylindole class of antagonists [43].

CB1 receptor cellular signaling via G-proteins

Domains of the CB1 receptor that selectively interact with Gi/o proteins have been identified [44]and appear to be coupled to Gαi/o proteins in the absence of exogenous agonists [4547]. The juxtamembrane domain, which extends from transmembrane (TM) 7 to the cys palmitoylation site, forms a helix referred to as H8 [4851]. This domain is able to trigger G-protein activation [44,48,52] via Gαo or Gαi3 [46,47,53]. CB1 receptor mutants truncated at the membrane surface of the TM7 were not able to inhibit Ca2+ channels in response to CB1 agonists [54]. The H8 amphipathic helix possesses cationic charges facing the G-protein surface that are important for agonist efficacy [48,50], with hydrophobic residues of the helix being critical for proper structural folding and positioning within the membrane bilayer [55]. A conformational change in the TM7 domain could displace the TM7-H8 elbow (see [51]. To test this, point mutations that replaced Leu at the elbow region with Phe or Ile were compared with the wild-type CB1 receptor [56]. Human Embryonic Kidney 293 (HEK293) cells heterologously expressing the mutant receptors exhibited reduced agonist-stimulated [35S]GTPγS binding to G-proteins, and the CB1 receptor mutants failed to interact with Gαi3 although they could still bind Gαi1 and Gαi2 [56]. These CB1 receptor mutants exhibited alterations in internalization in response to agonists [56]. The mutant receptors also exhibited significantly reduced inhibition of the Ca2+channels but a more rapid time course in response to WIN55212-2, which could be overcome by providing an abundance of the preferred G-protein, GαoA [56].

There is evidence to suggest that the CB1 receptor IL3 facilitates association with Gαi1 or Gαi2 [44,46,47,53]. Helical structures at the N-terminal side of IL3 possessing a BBXXB-like motif characterize the potential for this domain to interact with G-proteins [57]. The IL3 helical region near H6 is believed to interact with Gαi1 [53]. A double mutation of Leu-Ala in the IL3-H6 junction switched the helical domain to a single turn structure, thereby converting the CB1 receptor from interacting with Gi to interacting with Gs [53,58].

Signal transduction pathways utilized by CB1 receptors have often been associated with interaction with Gi/o class of G-proteins as demonstrated by their sensitivity to pertussis toxin [37,59]. Inhibition of adenylyl cyclase, activation of the mitogen activated protein kinase (MAPK) family of kinases, inhibition of voltage-gated Ca2+ channels, and activation of inwardly rectifying K+currents (Kir) are all signaling pathways that are pertussis toxin-sensitive (see reviews [34,37]. Regulation of ion channels by CB1receptors are the direct result of G•• release upon Gi/o activation [6064]. N-type voltage-gated Ca2+ currents were inhibited in differentiated neuroblastoma cells [62,63,6567] and P/Q-type Ca2+ currents were inhibited in rat cortical and cerebellar neurons and cultured AtT-20 pituitary cells heterologously expressing CB1receptors [68,69]. L-type Ca2+ currents in cat brain arterial smooth muscle cells were inhibited by CB1 agonists, although the mechanism may not directly involve G•• [70].

cAMP-Protein Kinase A (PKA) signaling

Inhibition of adenylyl cyclase isoforms 5 and 6 via G•i/o is the best characterized CB1 signaling pathway, as it results in reduced cyclic AMP-stimulated protein kinase A (PKA) activity. Reduced PKA activity impacts important cellular signaling events including voltage-dependent current flow at A-type K+ channels [71] and tyrosine phosphorylation of focal adhesion kinase (pp125FAK) and FAK-related non-kinase (FRNK) [7274]. In hippocampal neurons, the excitotoxic sequalae to NMDA-mediated Ca2+release was attenuated by CB1-mediated decreases in PKA phosphorylation of ryanodine channels [75].

If the ability of Gi/o proteins to interact withCB1 receptors is compromised by pertussis toxin treatment, interaction with Gs becomes possible [43,76,77]. CB1 receptor agonists increase cyclic AMP accumulation in globus pallidus [77,78], due to either augmentation of adenylyl cyclase 2/4/7 activation via Gβγ derived from Gi/o [79] or the sequestration of G• i/o proteins by other Gi/o-coupled receptors thereby facilitating coupling to Gs [61] [80]. A CB1-mediated activation of PKA has been shown to promote Ca2+ influx into neuroblastoma cells [81], phosphorylate DARPP-32 in corpus striatum neurons [82,83], and regulate K-type K+ channels in hippocampal neurons [84].

MAPK signaling

MAPK activation by CB1 agonists has been observed almost universally in cells expressing both endogenous and recombinant CB1 receptors [8588]. In CHO cells heterologously expressing CB1 receptors, extracellular signal-regulated kinase 1/2 (ERK1/2; also known as p42/p44 MAPK) required Gi/o as an initial step, presumably via the Gβγ subunits or else via CB1 agonist-stimulated association with •-arrestin [85]. There are multiple pathways that could result in CB1-mediated ERK1/2 phosphorylation and activation, and these may vary depending upon the cell type and the stimulus. Astrocytoma cells (and CHO cells expressing recombinant CB1 receptors) utilized a pathway involving phosphatidylinositol-3-kinase (PI3K) and protein kinase B (PKB, also known as Akt), which then activates Raf-1, MAP-ERK Kinase (MEK), and ERK1/2 [87,89,90]. In CHO cells heterologously expressing CB1 receptors, MAPK activation by insulin receptors or insulin-like growth factor (IGF) receptors could be blocked by SR141716-mediated sequestration of Gi/o proteins [91,92], suggesting some mechanism by which CB1receptors could be directly involved in the trans-activation of these receptor tyrosine kinases. Both glioblastoma or lung carcinoma cells employed cannabinoid-mediated trans-activation of epidermal growth factor (EGF) receptors in the PKB/Akt and ERK activation pathway [93].

N1E-115 neuroblastoma cells and hippocampal neurons utilized a pathway involving CB1-mediated attenuation of PKA activity and reduction of c-Raf phosphorylation, which facilitates MEK activation [94,95]. N18TG2 neuroblastoma cells appear to utilize a mechanism that involves PKC, a Ca2+-calmodulin-stimulated enzyme, a matrix metalloprotease and the activity of a vascular endothelial growth factor (VEGF)-like receptor. This ERK1/2 activation was associated with Ca2+ influx as determined by cellular uptake of radioisotopic Ca2+ [9698] In cultured cortical interneurons, endocannabinoids stimulated TrkB receptor tyrosine phosphorylation [99]. In PC12 cells heterologously expressing both TrkB receptors and CB1 receptors, co-immunoprecipitation studies implicated a complex formation in response to cannabinoid stimulation [99]. However, in nerve growth factor (NGF)-stimulated PC12 cells, anandamide was reported to inhibit TrkA receptor-induced Rap1/B-Raf/ERK activation [100]. CB1 receptors and Gi/o proteins could also regulateERK1/2 activation in PC12 cells via a non-receptor Src or Fyn tyrosine phosphorylation.

p38 MAPK was activated by cannabinoid receptor agonists in human vein endothelial cells [101], mouse hippocampal slices[102], and CHO cells heterologously expressing CB1receptors [103]. Jun N-terminal kinase (JNK1 and JNK2) was activated by CB1 agonists via Gi/o, PI3K and Ras in CHO cells heterologously expressing CB1 receptors [103]. Those studies implicated a transactivation of platelet-derived growth factor (PDGF) receptors in the cannabinoid agonist-mediated activation of JNK [103]. In cultured neuroblastoma cells, the CB1 receptor-mediated stimulation of JNK required Gi/o, activation of Rap1, Ral and Rac, and phosphorylation of Src tyrosine kinases [104106]. Src and JNK phosphorylation led to activation of the transcription factor Stat3, resulting in gene expression involved in neurite growth and elongation [104106].

Nitric Oxide (NO) signaling

CB1 receptor-mediated production and release of NO via neuronal NO synthase (nNOS) or endothelial (eNOS) has been demonstrated for a variety of tissues including saphenous vein segments [107], endothelial cells [108110], monocytes [111], brain slice preparations [112] and neuroblastoma cells [113,114]. Cannabinoid stimulation of NO-sensitive guanylyl cyclase and cyclic GMP production was detected in neuroblastoma cells [113] [115] [116]. The observation that CB1 agonists promote the translocation of NO-sensitive guanylyl cyclase from the cytosol to a membranous organelle compartment suggests that interactions between these proteins may occur in neuronal cells in the form of a receptor-effector complex [113].

In studies in which cannabinoid agonists reduced NO production in response to an excitotoxic stimulus in cerebellar granule cells, the CB1 receptor-mediated inhibition of voltage-gated Ca2+channels was shown to be responsible [116]. Similarly, in mouse cortical neurons undergoing excitotoxic degeneration in response to prolonged NMDA, CB1 receptor stimulation ameliorated NO generation by a mechanism that involved reduction in PKA activity and a change in the phosphorylation of nNOS [117119]. NO production by inducible NOS (iNOS) in response to an inflammatory stimulus in saphenous vein endothelial cells [107], RAW264.7 macrophages [120], microglial cells [121], and astrocytes [122124] was also reduced by cannabinoid receptor stimulation and the reduction in cyclic AMP [107,120,123,124].

CB1 receptor interaction with proteins other than G-proteins

The diverse intracellular localization of endogenously-expressed CB1 receptors was reported in studies using antibodies that recognize an N-terminal epitope of the receptor [125]. Immunocytochemical localization and sub-cellular fractionation studies indicated that a predominance of endogenously expressed CB1 receptor in neuroblastoma cells reside in internal organelles, particularly along the peri-nuclear membranes [125]. Constitutively synthesized CB1 receptors that had been pulse-chase labeled with [35S]-amino acids were degraded via at least two different mechanisms having elimination half-lives of approximately 5 hr (70% of receptors) and >24 hr (30% of receptors) in cells that had not been exposed to agonist ligands, suggesting the presence of multiple pools of functional receptors within neuronal cells [125]. Cellular sorting of CB1 receptors heterologously expressed in CHO-CB1 cells was recognized by FACS and confocal microscopy to transit from internal stores to the plasma membrane in response to the antagonist rimonabant and from plasma membrane to intracellular compartments in response to agonist CP55940 [126]. Pretreatment with CP 55940 resulted in rapid desensitization of the CB1 receptor agonist-mediated inhibition of forskolin-stimulated adenylyl cyclase or MAPK activation. Pretreatment of CHO-CB1 cells with rimonabant resulted in significant augmentation of CP55940-mediated activation of MAPK, but had no significant effect on CP55940-mediated inhibition of forskolin-stimulated adenylyl cyclase (Rinaldi-Carmona et al., 1998). Internalization of CB1receptors from neuronal plasma membranes at the soma has been proposed to be a mechanism by which these receptors can be diverted to plasma membrane loci along neuronal extensions [127129]. To add to the complexity of interpreting data, many studies have been performed using exogenously expressed CB1receptors that are fusion proteins with a fluorescent protein at the C-terminal tail, or else localization was determined using antibodies against epitopes in the C-terminal tail. Methods of detection that depend upon recognition of the C-terminal differ in their results when compared with endogenously expressed receptors having an accessible C-terminal for interaction with accessory proteins (see diversity of responses reported in [130]). Thus, it appears that domains along the CB1 receptor C-terminal have the potential to alter the dynamics of cellular trafficking and signal transduction (summarized in Table 1). CB1 receptors located at internal organelles may be of particular importance in autocrine regulation of cellular functions, particularly as sites of anandamide synthesis and degradation may be localized to intracellular membranous organelles (enzymes reviewed by [131]). Both cellular localization and function may be directed by CB1receptor association with a variety of interacting proteins.

Table 1

Role of CB1 Receptor-Associated Proteins in Protein Trafficking and Signal Transduction

Interaction of the CB1 Receptor with β-Arrestins

The level and duration of GPCR signaling activity is regulated by the two distinct processes of desensitization and endocytosis (see [132]). GPCR desensitization begins within seconds of agonist exposure and involves phosphorylation of agonist-activated receptors by GPCR kinases (GRKs). β-arrestin1 and β-arrestin2 immediately bind to the agonist-occupied, phosphorylated GPCR which prevents G-protein-mediated signal transduction, while the GPCR-arrestin complex associates with clathrin to initiate GPCR endocytosis. In addition to attenuating GPCR signaling and mediating GPCR internalization, β-arrestins function as scaffolds in GPCR-mediated endosome-based signaling pathways.

CB1 receptor desensitization by GRKs and β-arrestins was first demonstrated using Xenopus laevis oocytes and CB1 receptor mutants [133]. CB1 receptor-mediated activation of G-protein gated inwardly rectifying potassium channels (Kir3, also known as GIRK) was significantly desensitized in oocytes that co-expressed GRK3, β-arrestin 2, Kir3 channels, and CB1. A deletion CB1receptor mutant with 20 amino acids (residues 418–439) removed from the C-terminal tail of the receptor did not exhibit GRK3 and β-arrestin 2-mediated CB1 receptor desensitization in oocytes, which indicated that residues located in this region of the receptor are required for CB1 receptor desensitization by GRK3 and β-arrestin 2. Furthermore, the mutation of two possible GRK3 phosphorylation sites (S426A, S430A) significantly attenuated GRK3 and β-arrestin 2-mediated CB1 receptor desensitization in oocytes, which indicated that GRK3-mediated phosphorylation of S426 and S430 is necessary for CB1 receptor desensitization. Finally, these studies demonstrated S426A/S430A CB1 receptor mutants that were stably expressed in AtT20 cells retained their ability to be internalized. This finding suggested that distinct domains of the CB1 receptor are involved in GRK/β-arrestin-dependent CB1 receptor desensitization versus internalization. Later studies provided corroborative evidence supporting this seminal study that CB1 receptor desensitization is β-arrestin 2-dependent. The presynaptic expression of dominant negative GRK2 or β-arrestin 2 reduced desensitization of CB1 receptor-mediated presynaptic inhibition of glutamatergic neurotransmission in rat hippocampal neurons [134].

The S426A/S430A CB1 receptor desensitization-deficient mutant has also been used to demonstrate that CB1 receptor phosphorylation determines the time course of CB1 receptor agonist-mediated ERK1/2 activity [135]. The CB1 receptor agonist CP55940 transiently activated ERK1/2 in human embryonic kidney 293 (HEK293) cells stably expressing wild-type (WT) CB1receptors, with a peak of ERK1/2 phosphorylation at 5 min followed by a rapid dephosphorylation. In contrast, the duration of S426A/S430A CB1 receptor-mediated activation of ERK1/2 was significantly prolonged compared with WT, and was dynamically reversed by rimonabant in HEK293 cells. Thus, the time-course of CB1 receptor-mediated ERK1/2 and Kir3 activation both appear to be regulated by the same distinct domains of the CB1 receptor that are directly phosphorylated. During agonist treatment, β-arrestin was recruited to the plasma membrane in the S426A/S430A CB1 receptor mutant with the same kinetics as WT CB1 receptors, which suggests that phosphorylation of S426 and S430 is not required for β-arrestin recruitment. Nevertheless, a di-phosphorylated peptide created to mimic this domain was able to bind to •-arrestin 2 [136]. Based on the observation that the S426A/S430A CB1 receptor mutant could be internalized in HEK293 cells, Daigle and colleagues postulated that β-arrestin recruited to non-phosphorylated S426A/S430A CB1 receptors following agonist treatment retained its scaffolding functions [135].

CB1 receptor mutants have also been utilized to correlate CB1receptor internalization with β-arrestin recruitment [137]. The extreme carboxy terminal tail (amino acid residues 460–473) of the rat CB1 receptor contains a cluster of six serine (S) and threonine (T) residues that are potential sites for phosphorylation. HEK293 cells were stably transfected with CB1 receptors that were truncated at amino acid residue 460 (V460Z) or mutated at putative GRK phosphorylation sites (T461A/S463A, S465A/T466A, T468A/S469A, T461A-T466A, T461A-S469A) [137]. CB1 receptor internalization proceeded normally in HEK293 cells when the CB1receptor was truncated at residue 460 (V460Z) or when any two phosphorylation sites were mutated (T461A/S463A, S465A/T466A, T468A/S469A). However, CB1 receptor internalization was reduced when four sites were mutated (T461A through T466A), and was abolished when all six sites were mutated (T461A through S469A). In HEK293 cells in which the CB1 receptor internalized, β-arrestin was recruited to the plasma membrane and co-localized with CB1 receptors in response to agonist treatment, which suggests β-arrestin mediates CB1receptor internalization. β-arrestin was recruited to internalization-competent mutant CB1 receptors to the same maximal extent as to WT-CB1 receptors. However, the rate of β-arrestin recruitment to internalization-competent (V460Z, T461A/S463A, S465A/T466A, T468A/S469A) CB1 receptors was approximately 3-to 6-times slower compared with WT-CB1receptors. Moreover, β-arrestin was not recruited to internalization-incompetent mutant CB1 receptors (T461A through T466A, T461A through S469A). These results suggest that β-arrestin recruitment is coupled to the ability of the CB1 receptor to internalize, inasmuch as β-arrestin recruitment was attenuated when CB1 receptor internalization was reduced. Furthermore, β-arrestin recruitment is not solely a function of CB1 receptor phosphorylation, as β-arrestin was recruited to the V460Z CB1receptor mutant.

In vivo studies have also shown that β-arrestins can regulate CB1receptor activity. The role of β-arrestin 2 in cannabinoid-mediated behavioral effects was investigated in β-arrestin 2 (−/−) and β-arrestin 2 (+/+) mice [138]. No differences in CB1 receptor levels were observed in cerebellum, cortex or hippocampus of β-arrestin 2 (−/−) and β-arrestin 2 (+/+) mice. However, Δ9-THCproduced greater antinociception and hypothermia in β-arrestin2 (−/−) mice compared to β-arrestin2 (+/+) mice, while no differences were observed in either assay for other CB1 receptor agonists (e.g., CP55940, methanandamide). The finding that only THC activity was influenced by deletion of β-arrestin 2 suggests THC may activate CB1 receptors in a manner that leads to the recruitment of β-arrestin 2 and not β-arrestin 1. In another study, mice were chronically treated with THC to induce tolerance to the behavioral effects of THC and the relationship between THC-induced changes in CB1 receptor activity and the levels of GRKs and β-arrestins in specific mouse brain regions were investigated [139]. THC up-regulated GRK2, GRK4, and β-arrestin 1 levels in striatum, which suggested these proteins contribute to CB1receptor desensitization and endocytosis in this brain region. Chronic THC treatment increased GRK4 and β-arrestin 2 in cerebellum, whereas this treatment increased GRK2 and β-arrestin 2 levels in hippocampus. In the striatum and cerebellum, THC-induced up-regulation of GRKs and β-arrestins were ERK-dependent, because they were prevented in genetic (Ras-GRF1 knockout mice) and pharmacological (SL327-pretreated mice) models of ERK dysfunction. However, in the hippocampus, changes in GRKs and β-arrestins were ERK-independent. The findings in cerebellum and striatum suggest THC-induced ERK activation may play an important role in CB1 receptor desensitization and the expression of cannabinoid tolerance.

Interaction of the CB1 Receptor with adaptor protein AP-3 and GASP

Other interacting proteins are also involved in the regulation of CB1 receptor trafficking between plasma membranes and internal organelles. In a study of CB1 receptors endogenously expressed in Neuro2A neuroblastoma cells [130], a C-terminal antibody recognized CB1 receptors in intracellular compartments that merged with markers for late endosomes and lysosomes [130]. CB1 receptors co-localized with the adaptor protein AP-3• as detected in merged confocal images as well as co-immunoprecipitation studies [130]. AP-3 complex (comprised of • 3A and B (important for clathrin binding and sorting of LeuLeu signals), •, μ3A and B, and • subunits, with B isoforms being brain-specific) is a ubiquitous adaptor protein complex that responds to acidic Leu-Leu or Tyr-mediated sorting signals, and is intrinsic to vesicle budding from late endosomes [140]. No association of CB1 receptors was observed with AP-2 [130], which is associated with clathrin-dependent vesicle endocytosis from plasma membranes trafficking to early endosomes. Subcellular fractionation of Neuro2A cells by differential centrifugation as well as merged confocal micrographs identified CB1 receptors in the same compartment as endosomal markers AP-3• and Rab7. These data suggested that nascent CB1 receptors might bypass the plasma membrane and be delivered to the endosomal/lysosomal compartments under certain circumstances. Following depletion of cellular AP-3• by siRNA from Neuro2A or cultured hippocampal cells, a significantly greater fraction of CB1 receptors appeared on punctate domains along the plasma membrane surface, suggesting that one function of the AP-3 complex might be to divert CB1 receptors away from plasma membrane localization.

It is interesting to speculate on the role that intracellular CB1receptors might play in neuronal regulation. Subcellular fractionation of Neuro2A cells by differential centrifugation as well as merged confocal micrographs suggested that endosomal CB1 receptors co-localized with G• i and phosphoERK1/2, indicating their potential for functional signal transduction [130]. At presynaptic terminals, synaptic vesicle life cycle includes stages of birth by budding from endosomes, neurotransmitter filling of nascent vesicles, docking of filled vesicles at the synaptic active zone, priming, Ca2+ stimulated fusion and neurotransmitter release, followed by clathrin-dependent endocytosis of fused vesicle membrane and incorporation into early endosomes for recycling (reviewed by [140]). AP-3B is enriched in brain clathrin-coated vesicles and endosomes, and is associated with budding profiles on early-endosomes that traffic to late endocytic/lysosomal compartments. In neurons, AP-3B is found in neuronal soma and axonal terminals in both soluble and membrane-bound forms, where it might serve in selective sorting and transport of cargo from the trans-golgi network to the synapse. Using a PC12 pheochromocytoma cell model of synaptic-like micro-vesicle formation from endosomes, AP-3 functioned in a brefeldin-A-sensitive, ARF1 (ADP-ribosylation factor binding-protein 1)- dependent, and clathrin- and dynamin-independent manner to sort AP-3 cargo [141144]. Thus, one could speculate that the delivery of CB1 receptors via endosome-derived synaptic vesicles might serve as a mechanism to place functional CB1receptors at peri-synaptic membranes in correlation with the rate of synaptic vesicle release.

G-protein receptor-associated sorting protein 1 (GASP1) can regulate post-endocytic targeting of the CB1 receptor to the lysosome for degradation [145,146]. GASP1 could promote trafficking of those CB1 receptors that had been internalized in response to agonist ligands [145], suggesting that the plasma membrane receptors would have previously been coupled to heteromeric G-proteins as part of the G-protein signaling cycle. The association between GASP1 and the CB1 receptor was demonstrated by co-immunoprecipitation studies as well as pull-down assays using a glutathione-S-transferase (GST)-CB1-H8 peptide fusion protein [145]. Prolonged (>1 hr) agonist treatment (100 nM WIN55212-2) of HEK293 cells stably expressing EGFPN-term-CB1 receptors, resulted in confocal microscopic images of CB1-EGFP that merged with GASP1 as well as with Lyso-tracker dye, indicating the co-localization of these proteins at the lysosome [146]. When cultured neurons from neonatal rat spinal cord were treated with high concentrations of agonist (1.5 μM WIN55212-2), CB1 receptor elimination from soma and neuritic extensions was detectable at 6 hr, continued over the ensuing 24 hr, and was sustained at about 50% depletion for the next 24 hr of treatment [146]. This response was attenuated after the spinal neurons had been transduced by viral delivery of dominant-negative cGASP1-AAV, demonstrating a role for GASP1 in the down-regulation of CB1 receptors from their plasma membrane compartments in conditions of prolonged activation [146].

In in vivo studies in mice, the development of analgesic tolerance observed after four days of WIN55212-2-treatment could be attenuated by transduction with the dominant negative cGASP1-AAV, leading these researchers to suggest that GASP1 functions primarily to promote transit to the lysosome in conditions of tolerance [146]. A requirement for GASP1 for down-regulation of CB1 receptors in the dorsal horn of the mouse spinal cord was observed as a reduction from about 14% [3H]-CP55940 autoradiographic density after 7-day treatments with WIN55212-2 to only 5% reduction in the cGASP1-AAV-transduced mice [146]. Additional evidence from brain striatum has suggested that the role of GASP1 may be more extensive and complex than simply delivery of GPCRs to the lysosome for degradation. A GASP1-knock out mouse on a C57Bl/6 background was created, and the Bmax for D1-like, D2-like dopamine receptors and muscarinic-like receptors were determined after cocaine-sensitization or cocaine self-administration protocols that would increase exposure of these receptors to their endogenous agonists for long periods of time [147]. In contrast to what would be predicted if GASP1 regulated transit to lysosomes, the genetic deletion of GASP1 resulted in a decrease in receptor binding maxima in the self-administering animals, and no differences from WT in the control and cocaine-sensitized groups [147]. These data coupled with behavioral findings call into question the role that GASP1 may play in directing these receptors to be degraded, and suggest that GASP1 may have greater diversity in its cellular actions. GASP1 interacts with conserved FR residues within H8 in diverse GPCRs [148], many of which interact with G-proteins via the IL3 (as the CB1 receptor does with Gi1 and Gi2). Because Gi3 and Go utilize the CB1 receptor H8 domain for G-protein-mediated signal transduction, one would expect a competition to exist between GASP1 versus these G-proteins at the H8 domain of the CB1 receptor.

Interaction of the CB1 receptor with Cannabinoid Receptor Interacting Protein 1a (CRIP1a)

The CB1 receptor has recently been shown to interact with a novel protein, the cannabinoid receptor interacting-protein, CRIP1a/b, within its C-terminus domain. CRIP1a/b were discovered by the Lewis laboratory, which reported that deletion of the CB1 receptor C-terminal tail slowed the time to peak Ca2+ current inhibition, augmented the tonic inhibition of Ca2+ currents, and promoted the ability of the CB1 receptor to sequester G-proteins [54,149]. They hypothesized that the C-terminal tail could serve an auto-inhibitory function. In seeking an accessory protein to regulate this activity, they used the CB1 receptor distal C-terminal as bait in a yeast two-hybrid screen to identify a pair of splice variant proteins, CRIP1a and CRIP1b [150]. CRIP1a could bind to a GST-CB1-C-terminal tail fusion protein, and could be co-immunoprecipitated with the CB1 receptor [150].

The gene for CRIP1a and CRIP1b, is located on human chromosome two and consists of 3 coding exons. Both CRIP1a and CRIP1b have highly conserved sequence homology due to the fact that both proteins are encoded by exons 1 and 2. However, alternative splicing of exon 3 leads to the sequence variation and naming observed between these two isoforms; exon 3 encodes for amino acids 111–128 in CRIP1b and 111–164 in CRIP1a. Currently, both CRIP isoforms have been cloned and sequenced for human, rat and mouse, with significant homology seen among the rodent species. However, an alternative splicing variation has been found in mouse cerebellar tissue, which like CRIP1a/b is encoded for by exons 1 and 2, but also includes the noncoding regions of the 5′ end of exon 1 and the 3 prime end of exon 2. Genomic database searching has identified CRIP1a in all vertebrates, but CRIP1b is only found in primates indicating recent evolutionary processing of this gene [150]. In addition, CRIP1a distribution in mouse brain reveals co-expression with CB1 in excitatory glutamatergic neurons, but not in inhibitory GABAergic interneurons [151,152].

CB1 receptor mutagenesis studies show that the last 9 amino acids on the C-terminus of CB1 are the minimum residues required for CRIP1a/b binding [150]. Furthermore, both isoforms require amino acids 34–110 (comprising exons 1 and 2) to interact with the CB1. Because neither of the CRIP proteins interacts with the CB2 receptor, they possess a unique and specific means for modulating the physiological effects mediated through the CB1receptor. Alternative splicing isoforms that bind to CB2 have yet to be discovered, and if identified may serve to selectively and differentially alter CB1 and CB2 signaling.

To date, little is known about the functional relevance of the CRIP1 family of proteins due to the lack of a three-dimensional model and insufficient sequence homology to other known proteins. A major difference between CRIP1a and CRIP1b is that CRIP1a contains a palmitoylation site and a C-terminus PDZ class I ligand. Site specific palmitoylation is a known modification that many adaptor and scaffolding molecules undergo and serves to regulate the function and trafficking of G-proteins, GPCRs and Src family kinases [153]. Proteins that contain PDZ ligands are believed to have a role in determining the cellular distribution of proteins that possess PDZ domains that recognize them. The functional relevance of CRIP1a having a PDZ ligand is intriguing, as it may allow CRIP1a to 1) interact with other proteins and act as a scaffolding site to establish variations in signal transduction, 2) enable the formation of homo/heterodimerization between CB1and/or other receptors, and 3) modulate CB1 trafficking events such as localization, desensitization, or internalization.

Studies using superior cervical ganglion neurons (SCG) stably transfected with CB1 receptors have shown that the •• subunits of the trimeric G-protein complex can directly interact with N-type Ca2+ channels to inhibit Ca2+ influx [60]. This effect can be reversed by the antagonist/inverse agonist rimonabant, indicating that CB1 receptors release G•• subunits to suppress N-type Ca2+channel activity. CRIP1a, but not CRIP1b, attenuated the inhibitory effects of CB1 on N-type Ca2+ channels [150]. However, WIN55212-2 stimulation of CB1-mediated inhibition of N-type Ca2+ channels was unaltered by CRI P1a. These results taken together suggest that CRIP1a/b may functionally modulate CB1signal transduction in an agonist-independent manner. Additional research to define the effects of CRIP1a/b on other cannabinoid agonist and antagonist-regulated signal transduction will further delineate these mechanisms.

Many GPCRs can initiate agonist-independent regulation of signal transduction pathways, in which case ancillary proteins that internally bind to GPCRs can be key modulators of receptor-mediated events. There is also great variation in the C-termini of GPCRs, allowing for diverse and differential interactions, post-translational modifications, and trafficking seen within this superfamily of transmembrane proteins. Additionally, protein-protein interactions positively modulate GPCR signaling by influencing ligand-binding affinity and specificity, coupling between receptors, G-proteins and effectors, or targeting to specific subcellular locations. Receptor-interacting-proteins like Homers and dopamine receptor interacting proteins (DRIPs) regulate the intracellular activity of GPCRs using various mechanisms. For example, metabotropic glutamate receptors (mGluRs) are known to interact via their C-terminus with a family of proteins called Homers. Homer proteins have been implicated in a variety of roles including trafficking of type I mGluRs and receptor-mediated signal localization [154]. This class of proteins is subdivided into three families (1, 2, 3), and similar to what is seen in CRIP1a/b, families 1 and 2 are splice variations of the same gene. For the most part, Homer proteins are constitutively expressed and have two functional sites: an amino-terminal Enabled/Vasodilator-stimulated phosphoprotein (VASP) homology 1 (EVH1) domain and a coiled-coil leucine zipper carboxy-terminus. In addition to these two functional sites, Homers possess a PDZ domain that has been shown to bind with PSD-95, GKAP and Shank1 and 3 [155]. One important aspect on mGluR receptors is their ability to function as dimers, and therefore, the Homer family of proteins may participate in this process [156]. Homers are also known to form a complex between mGluRs and IP3 receptors [154]. Thus, if CRIPs exhibit functional homology to Homer proteins, then CRIPs may play a pivotal role in localization of the CB1 receptor to its signaling partners, such as voltage-gated Ca2+ channels and G-protein linked K+channels. However, unlike CB1 receptors, the majority of mGluRs are located at post synaptic density areas where Homers can be up-regulated by neuronal activity [154]. An example of this is that expression of Homer 1a is induced upon neuronal excitation [157], and due to the lack of a binding domain in its C-tail, can act as a dominate-negative protein by disrupting interactions between mGluR receptors and other proteins. In HEK293 cells, expression of Homer1a, but not its isoform Homer-1b, resulted in increased mGluR translocation to the plasma membrane. Studies suggest that Homer 1b influences mGluR 1 and 5 retention in intracellular stores until the up-regulation of Homer 1a evokes transit of receptors to the plasma membrane [158]. Because Homer 1a and CRIP1b both lack functional domains in their C-tail, they may perform similar functions as dominant negative regulators.

Like CB1 receptors, D2 dopamine receptors are GPCRs that are predominately expressed on presynaptic terminals, couple to Gi/o proteins, and initiate similar signal transduction pathways. D2receptors associate with interacting-proteins (DRIPs) that bind to the C-terminus of the receptor. The D2 receptor DRIP neuronal Ca2+ sensor 1 (NCS-1) inhibits D2 receptor desensitization in a Ca2+-dependent fashion by binding to GRK2 and preventing its phosphorylation of the D2 receptor [159]. Because CRIP1a/b bind to a domain near the CB1 receptor internalization site, CRIP may serve a similar function asNCS-1. To date, other dopamine receptor interacting-proteins (calcyon and DRIP78) have also been identified. The endoplasmic reticulum membrane-bound protein DRIP78 regulates the transport of many GPCRs (D1, M2, AT1), via their C-tail, to the plasma membrane [160]. Both sequestration and overexpression of DRIP78 results in D1 and •2AR localization in the endoplasmic reticulum, along with reduced ligand binding and receptor glycosylation [161].

The further identification of CRIP proteins as well as their functional importance will provide a greater understanding into the regulation and neurotransmission of cannabinoid receptors. Because CRIP1a colocalizes with CB1 signaling at excitatory synapses, it is a novel target for the treatment of disorders associated with excessive excitatory transmission, such as epilepsy [152]. Thus it will be important to continue research into the CRIP proteins to elucidate the mechanisms involved in CRIP1a/b modulation of CB1 receptors and the opportunity that this offers in the development of cannabinoid based drugs.

Interaction of the CB1 Receptor with Factor Associated with Neutral sphingomyelinase (FAN)

Signal transduction via regulation of the second messenger ceramide can be regulated by CB1 receptors either via sphingomyelin hydrolysis or by de novo synthesis of ceramide (see review [162,163]). To summarize, transient increases in intracellular ceramide are regulated by the interaction of the CB1receptor with neutral sphingomyelinase via the interacting protein FAN. Once released into the cell, this short-lived production of ceramide can regulate cellular metabolic processes. The enzymatic production of ceremide de novo is under the CB1 receptor regulation of serine palmitoyltransferase or ceramide synthase, and initiates signaling pathways leading to cyclooxygenase-2 expression and apoptosis in cells that are susceptible to this mediator [164166].

FAN directs the agonist-stimulated CB1 receptor to the regulation of sphingomyelin hydrolysis, leading to the release of ceramide [167]. In an astrocyte model, CB1 agonists promoted a complex of FAN with CB1 receptors, and this could be blocked by the CB1antagonist rimonabant, but not by pertussis toxin treatment, indicating that Gi/o proteins were not intrinsic to the process [167]. The requirement for FAN is based upon the attenuation of the ceramide generation in cells that express a dominant negative form of this protein[167]. CB1 receptors and FAN could co-immunprecipitate, demonstrating their interaction as a complex. Studies based upon FAN interactions with other proteins suggest that the C-terminal of the CB1 receptor is the site of a putative neutral sphingomyelinase activation domain [167]. Interestingly, FAN is a WD40 repeat protein, as is the G• subunit of G-proteins and many other adaptor proteins [163]).

Sphingomyelin hydrolysis in cultured astrocytes or glioma cells yielded an increase in intracellular ceramide within 15 min [87,168]. CB1-mediated release of ceramide activates the Raf-1-MEK-ERK pathway to regulate glucose metabolism [86]. The Guzman laboratory has speculated that this ceramide-signaling pathway may serve an important function of astrocytes to supply metabolic substrates to neurons for oxidative metabolism at synapses and other sites having rapid biosynthetic requirements [163].

Perspectives

The majority of the data regarding cellular mechanisms for CB1receptor actions have been based upon experiments performed in the brain or in neuronal models (Table 1). As we now are learning about CB1 receptors in many other cell types, it is important to recognize that cellular signal transduction for highly specialized cells may deviate from what we have learned regarding neurons. Neurons are excitable cells, and are structurally unique in their cell-cell interactions (i.e., they are polarized in the structure of axons and dendrites, and form unique sites of interaction, the synapse). Other cell types possess their unique abilities to respond to hormones and other mediators to regulate their specific functions within a tissue or organ. While we can use our present knowledge as a guide, we certainly cannot expect that all cell types will respond to CB1 receptor stimulation via activation of the same signaling pathways. This notion of diversity leads us to speculate that agonists and allosteric modulators might be discovered or developed which can trigger the regulation of unique, cell-specific responses.

Footnotes

Conflict of Interest: The authors have not received financial contributions to the work being reported, and do not report any conflict of interest. ACH has served on the speaker’s bureau for Sanofi-Aventis within the last three years.

 

References

1. Howlett AC, Bidaut-Russell M, Devane WA, Melvin LS, Johnson MR, Herkenham M. The cannabinoid receptor: biochemical, anatomical and behavioral characterization. Trends Neurosci.1990;13:420–423. [PubMed]
2. Johnson MR, Melvin LS, Althuis TH, Bindra JS, Harbert CA, Milne GM, Weissman A. Selective and potent analgetics derived from cannabinoids. J Clin Pharmacol. 1981;21:271S–282S.[PubMed]
3. Berlach DM, Shir Y, Ware MA. Experience with the synthetic cannabinoid nabilone in chronic noncancer pain. Pain Med.2006;7:25–29. [PubMed]
4. Jain AK, Ryan JR, McMahon FG, Smith G. Evaluation of intramuscular levonantradol and placebo in acute postoperative pain. J Clin Pharmacol. 1981;21:320S–326S. [PubMed]
5. Young AM, Katz JL, Woods JH. Behavioral effects of levonantradol and nantradol in the rhesus monkey. J Clin Pharmacol. 1981;21:348S–360S. [PubMed]
6. Clark AJ, Lynch ME, Ware M, Beaulieu P, McGilveray IJ, Gourlay D. Guidelines for the use of cannabinoid compounds in chronic pain. Pain Res Manag. 2005;10(Suppl A):44A–46A.[PubMed]
7. Laszlo J, Lucas VS, Jr, Hanson DC, Cronin CM, Sallan SE. Levonantradol for chemotherapy-induced emesis: phase I–II oral administration. J Clin Pharmacol. 1981;21:51S–56S. [PubMed]
8. Plasse TF, Gorter RW, Krasnow SH, Lane M, Shepard KV, Wadleigh RG. Recent clinical experience with dronabinol.Pharmacol Biochem Behav. 1991;40:695–700. [PubMed]
9. Slatkin NE. Cannabinoids in the treatment of chemotherapy-induced nausea and vomiting: beyond prevention of acute emesis.J Support Oncol. 2007;5:1–9. [PubMed]
10. Ware1 MA, Daeninck P, Maida V. A review of nabilone in the treatment of chemotherapy-induced nausea and vomiting. Ther Clin Risk Manag. 2008;4:99–107. [PMC free article] [PubMed]
11. Lutz B. The endocannabinoid system and extinction learning.Mol Neurobiol. 2007;36:92–101. [PubMed]
12. Kirkham TC. Cannabinoids and appetite: food craving and food pleasure. Int Rev Psychiatry. 2009;21:163–171. [PubMed]
13. Maldonado R, Valverde O, Berrendero F. Involvement of the endocannabinoid system in drug addiction. Trends Neurosci.2006;29:225–232. [PubMed]
14. Panagis G, Vlachou S, Nomikos GG. Behavioral pharmacology of cannabinoids with a focus on preclinical models for studying reinforcing and dependence-producing properties. Curr Drug Abuse Rev. 2008;1:350–374. [PubMed]
15. Beardsley PM, Thomas BF, McMahon LR. Cannabinoid CB1 receptor antagonists as potential pharmacotherapies for drug abuse disorders. Int Rev Psychiatry. 2009;21:134–142. [PubMed]
16. Fernandez-Ruiz J. The endocannabinoid system as a target for the treatment of motor dysfunction. Br J Pharmacol.2009;156:1029–1040. [PMC free article] [PubMed]
17. Leweke FM, Koethe D. Cannabis and psychiatric disorders: it is not only addiction. Addict Biol. 2008;13:264–275. [PubMed]
18. Moreira FA, Grieb M, Lutz B. Central side-effects of therapies based on CB1 cannabinoid receptor agonists and antagonists: focus on anxiety and depression. Best Pract Res Clin Endocrinol Metab. 2009;23:133–144. [PubMed]
19. Anavi-Goffer S, Mulder J. The polarised life of the endocannabinoid system in CNS development. Chem biochem.2009;10:1591–1598. [PubMed]
20. Harkany T, Mackie K, Doherty P. Wiring and firing neuronal networks: endocannabinoids take center stage. Curr Opin Neurobiol. 2008 [PubMed]
21. Galve-Roperh I, Aguado T, Palazuelos J, Guzman M. Mechanisms of control of neuron survival by the endocannabinoid system. Curr Pharm Des. 2008;14:2279–2288. [PubMed]
22. Van Gaal LF, Scheen AJ, Rissanen AM, Rossner S, Hanotin C, Ziegler O. Long-term effect of CB1 blockade with rimonabant on cardiometabolic risk factors: two year results from the RIO-Europe Study. Eur Heart J. 2008;29:1761–1771. [PubMed]
23. Despres JP. Pleiotropic effects of rimonabant: clinical implications. Curr Pharm Des. 2009;15:553–570. [PubMed]
24. de Kloet AD, Woods SC. Minireview: Endocannabinoids and their receptors as targets for obesity therapy. Endocrinology.2009;150:2531–2536. [PubMed]
25. Cota D, Sandoval DA, Olivieri M, Prodi E, D’Alessio DA, Woods SC, Seeley RJ, Obici S. Food intake-independent effects of CB1 antagonism on glucose and lipid metabolism. Obesity (Silver Spring) 2009;17:1641–1645. [PubMed]
26. Mallat A, Lotersztajn S. Endocannabinoids and liver disease. I. Endocannabinoids and their receptors in the liver. Am J Physiol Gastrointest Liver Physiol. 2008;294:G9–G12. [PubMed]
27. Pandey R, Hegde VL, Singh NP, Hofseth L, Singh U, Ray S, Nagarkatti M, Nagarkatti PS. Use of cannabinoids as a novel therapeutic modality against autoimmune hepatitis. Vitam Horm.2009;81:487–504. [PubMed]
28. Ralevic V, Kendall DA. Cannabinoid modulation of perivascular sympathetic and sensory neurotransmission. Curr Vasc Pharmacol. 2009;7:15–25. [PubMed]
29. Batkai S, Pacher P. Endocannabinoids and cardiac contractile function: pathophysiological implications. Pharmacol Res.2009;60:99–106. [PMC free article] [PubMed]
30. Hiley CR. Endocannabinoids and the heart. J Cardiovasc Pharmacol. 2009;53:267–276. [PMC free article] [PubMed]
31. Maccarrone M. Endocannabinoids: friends and foes of reproduction. Prog Lipid Res. 2009;48:344–354. [PubMed]
32. Idris AI. Role of cannabinoid receptors in bone disorders: alternatives for treatment. Drug News Perspect. 2008;21:533–540. [PubMed]
33. Biro T, Toth BI, Hasko G, Paus R, Pacher P. The endocannabinoid system of the skin in health and disease: novel perspectives and therapeutic opportunities. Trends Pharmacol Sci. 2009;30:411–420. [PMC free article] [PubMed]
34. Howlett AC, Barth F, Bonner TI, Cabral G, Casellas P, Devane WA, Felder CC, Herkenham M, Mackie K, Martin BR, Mechoulam R, Pertwee RG. International Union of Pharmacology. XXVII. Classification of Cannabinoid Receptors.Pharmacol Rev. 2002;54:161–202. [PubMed]
35. Pertwee RG. Pharmacological actions of cannabinoids. Handb Exp Pharmacol. 2005:1–51. [PubMed]
36. Di Marzo V, De Petrocellis L, Bisogno T. The biosynthesis, fate and pharmacological properties of endocannabinoids. Handb Exp Pharmacol. 2005:147–185. [PubMed]
37. Howlett AC. Cannabinoid receptor signaling. Handb Exp Pharmacol. 2005:53–79. [PubMed]
38. Barth F, Rinaldi-Carmona M. The development of cannabinoid antagonists. Curr Med Chem. 1999;6:745–755.[PubMed]
39. Thomas BF, Zhang Y, Brackeen M, Page KM, Mascarella SW, Seltzman HH. Conformational characteristics of the interaction of SR141716A with the CB1 cannabinoid receptor as determined through the use of conformationally constrained analogs. AAPS J.2006;8:E665–E671. [PMC free article] [PubMed]
40. Hagmann WK. The discovery of taranabant, a selective cannabinoid-1 receptor inverse agonist for the treatment of obesity. Arch Pharm (Weinheim) 2008;341:405–411. [PubMed]
41. Hurst DP, Lynch DL, Barnett-Norris J, Hyatt SM, Seltzman HH, Zhong M, Song ZH, Nie J, Lewis D, Reggio PH. N-(piperidin-1-yl)-5-(4-chlorophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-p yrazole-3-carboxamide (SR141716A) interaction with LYS 3.28(192) is crucial for its inverse agonism at the cannabinoid CB1 receptor. Mol Pharmacol. 2002;62:1274–1287. [PubMed]
42. Chambers AP, Vemuri VK, Peng Y, Wood JT, Olszewska T, Pittman QJ, Makriyannis A, Sharkey KA. A neutral CB1 receptor antagonist reduces weight gain in rat. Am J Physiol Regul Integr Comp Physiol. 2007;293:R2185–R2193. [PubMed]
43. Felder CC, Joyce KE, Briley EM, Glass M, Mackie KP, Fahey KJ, Cullinan GJ, Hunden DC, Johnson DW, Chaney MO, Koppel GA, Brownstein M. LY320135, a novel cannabinoid CB1 receptor antagonist, unmasks coupling of the CB1 receptor to stimulation of cAMP accumulation. J Pharmacol Exp Ther. 1998;284:291–297. [PubMed]
44. Howlett AC, Song C, Berglund BA, Wilken GH, Pigg JJ. Characterization of CB1 cannabinoid receptors using receptor peptide fragments and site-directed antibodies. Mol Pharmacol.1998;53:504–510. [PubMed]
45. Houston DB, Howlett AC. Differential receptor-G-protein coupling evoked by dissimilar cannabinoid receptor agonists. Cell Signal. 1998;10:667–674. [PubMed]
46. Mukhopadhyay S, McIntosh HH, Houston DB, Howlett AC. The CB(1) cannabinoid receptor juxtamembrane C-terminal peptide confers activation to specific G proteins in brain. Mol Pharmacol. 2000;57:162–170. [PubMed]
47. Mukhopadhyay S, Howlett AC. CB1 receptor-G protein association. Subtype selectivity is determined by distinct intracellular domains. Eur J Biochem. 2001;268:499–505.[PubMed]
48. Mukhopadhyay S, Cowsik SM, Lynn AM, Welsh WJ, Howlett AC. Regulation of Gi by the CB1 cannabinoid receptor C-terminal juxtamembrane region: structural requirements determined by peptide analysis. Biochemistry. 1999;38:3447–3455. [PubMed]
49. Choi G, Guo J, Makriyannis A. The conformation of the cytoplasmic helix 8 of the CB1 cannabinoid receptor using NMR and circular dichroism. Biochim Biophys Acta. 2005;1668:1–9.[PubMed]
50. Xie XQ, Chen JZ. NMR structural comparison of the cytoplasmic juxtamembrane domains of G-protein-coupled CB1 and CB2 receptors in membrane mimetic dodecylphosphocholine micelles. J Biol Chem. 2005;280:3605–3612. [PubMed]
51. Grace CR, Cowsik SM, Shim JY, Welsh WJ, Howlett AC. Unique helical conformation of the fourth cytoplasmic loop of the CB1 cannabinoid receptor in a negatively charged environment. J Struct Biol. 2007;159:359–368. [PMC free article] [PubMed]
52. Howlett AC, Mukhopadhyay S, Shim JY, Welsh WJ. Signal transduction of eicosanoid CB1 receptor ligands. Life Sci.1999;65:617–625. [PubMed]
53. Ulfers AL, McMurry JL, Miller A, Wang L, Kendall DA, Mierke DF. Cannabinoid receptor-G protein interactions: G(alphai1)-bound structures of IC3 and a mutant with altered G protein specificity. Protein Sci. 2002;11:2526–2531. [PMC free article][PubMed]
54. Nie J, Lewis DL. The proximal and distal C-terminal tail domains of the CB1 cannabinoid receptor mediate G protein coupling. Neuroscience. 2001;107:161–167. [PubMed]
55. Ahn KH, Pellegrini M, Tsomaia N, Yatawara AK, Kendall DA, Mierke DF. Structural analysis of the human cannabinoid receptor one carboxyl-terminus identifies two amphipathic helices. Biopolymers. 2009;91:565–573. [PMC free article][PubMed]
56. Anavi-Goffer S, Fleischer D, Hurst DP, Lynch DL, Barnett-Norris J, Shi S, Lewis DL, Mukhopadhyay S, Howlett AC, Reggio PH, Abood ME. Helix 8 Leu in the CB1 cannabinoid receptor contributes to selective signal transduction mechanisms. J Biol Chem. 2007;282:25100–25113. [PubMed]
57. Ulfers AL, McMurry JL, Kendall DA, Mierke DF. Structure of the third intracellular loop of the human cannabinoid 1 receptor.Biochemistry. 2002;41:11344–11350. [PubMed]
58. Abadji V, Lucas-Lenard JM, Chin C, Kendall DA. Involvement of the carboxyl terminus of the third intracellular loop of the cannabinoid CB1 receptor in constitutive activation of Gs. J Neurochem. 1999;72:2032–2038. [PubMed]
59. Howlett AC. Efficacy in CB1 receptor-mediated signal transduction. Br J Pharmacol. 2004;142:1209–1218.[PMC free article] [PubMed]
60. Pan X, Ikeda SR, Lewis DL. SR 141716A acts as an inverse agonist to increase neuronal voltage-dependent Ca2+ currents by reversal of tonic CB1 cannabinoid receptor activity. Mol Pharmacol. 1998;54:1064–1072. [PubMed]
61. Vasquez C, Lewis DL. The CB1 cannabinoid receptor can sequester G-proteins, making them unavailable to couple to other receptors. J Neurosci. 1999;19:9271–9280. [PubMed]
62. Priller J, Briley EM, Mansouri J, Devane WA, Mackie K, Felder CC. Mead ethanolamide, a novel eicosanoid, is an agonist for the central (CB1) and peripheral (CB2) cannabinoid receptors.Mol Pharmacol. 1995;48:288–292. [PubMed]
63. Pan X, Ikeda SR, Lewis DL. Rat brain cannabinoid receptor modulates N-type Ca2+ channels in a neuronal expression system. Mol Pharmacol. 1996;49:707–714. [PubMed]
64. Guo J, Ikeda SR. Endocannabinoids modulate N-type calcium channels and G-protein-coupled inwardly rectifying potassium channels via CB1 cannabinoid receptors heterologously expressed in mammalian neurons. Mol Pharmacol. 2004;65:665–674.[PubMed]
65. Caulfield MP, Brown DA. Cannabinoid receptor agonists inhibit Ca current in NG108-15 neuroblastoma cells via a pertussis toxin-sensitive mechanism. Br J Pharmacol. 1992;106:231–232.[PMC free article] [PubMed]
66. Mackie K, Hille B. Cannabinoids inhibit N-type calcium channels in neuroblastoma-glioma cells. Proc Natl Acad Sci USA.1992;89:3825–3829. [PMC free article] [PubMed]
67. Mackie K, Devane WA, Hille B. Anandamide, an endogenous cannabinoid, inhibits calcium currents as a partial agonist in N18 neuroblastoma cells. Mol Pharmacol. 1993;44:498–503.[PubMed]
68. Mackie K, Lai Y, Westenbroek R, Mitchell R. Cannabinoids activate an inwardly rectifying potassium conductance and inhibit Q-type calcium currents in AtT20 cells transfected with rat brain cannabinoid receptor. J Neurosci. 1995;15:6552–6561. [PubMed]
69. Hampson AJ, Bornheim LM, Scanziani M, Yost CS, Gray AT, Hansen BM, Leonoudakis DJ, Bickler PE. Dual effects of anandamide on NMDA receptor-mediated responses and neurotransmission. J Neurochem. 1998;70:671–676. [PubMed]
70. Gebremedhin D, Lange AR, Campbell WB, Hillard CJ, Harder DR. Cannabinoid CB1 receptor of cat cerebral arterial muscle functions to inhibit L-type Ca2+ channel current. Am J Physiol.1999;276:H2085–H2093. [PubMed]
71. Mu J, Zhuang SY, Hampson RE, Deadwyler SA. Protein kinase-dependent phosphorylation and cannabinoid receptor modulation of potassium A current (IA) in cultured rat hippocampal neurons. Pflugers Arch. 2000;439:541–546.[PubMed]
72. Zhou D, Song ZH. CB1 cannabinoid receptor-mediated tyrosine phosphorylation of focal adhesion kinase-related non-kinase. FEBS Lett. 2002;525:164–168. [PubMed]
73. Derkinderen P, Toutant M, Burgaya F, Le Bert M, Siciliano JC, de FV, Gelman M, Girault JA. Regulation of a neuronal form of focal adhesion kinase by anandamide. Science. 1996;273:1719–1722. [PubMed]
74. Zhou D, Song ZH. CB1 cannabinoid receptor-mediated neurite remodeling in mouse neuroblastoma N1E-115 cells. J Neurosci Res. 2001;65:346–353. [PubMed]
75. Zhuang SY, Bridges D, Grigorenko E, McCloud S, Boon A, Hampson RE, Deadwyler SA. Cannabinoids produce neuroprotection by reducing intracellular calcium release from ryanodine-sensitive stores. Neuropharmacology. 2005;48:1086–1096. [PubMed]
76. 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. [PubMed]
77. Bonhaus DW, Chang LK, Kwan J, Martin GR. Dual activation and inhibition of adenylyl cyclase by cannabinoid receptor agonists: evidence for agonist-specific trafficking of intracellular responses. J Pharmacol Exp Ther. 1998;287:884–888. [PubMed]
78. Maneuf YP, Brotchie JM. Paradoxical action of the cannabinoid WIN 55,212-2 in stimulated and basal cyclic AMP accumulation in rat globus pallidus slices. Br J Pharmacol.1997;120:1397–1398. [PMC free article] [PubMed]
79. Rhee MH, Bayewitch M, Avidor-Reiss T, Levy R, Vogel Z. Cannabinoid receptor activation differentially regulates the various adenylyl cyclase isozymes. J Neurochem. 1998;71:1525–1534. [PubMed]
80. Jarrahian A, Watts VJ, Barker EL. D2 dopamine receptors modulate Galpha-subunit coupling of the CB1 cannabinoid receptor. J Pharmacol Exp Ther. 2004;308:880–886. [PubMed]
81. Bash R, Rubovitch V, Gafni M, Sarne Y. The stimulatory effect of cannabinoids on calcium uptake is mediated by Gs GTP-binding proteins and cAMP formation. Neurosignals.2003;12:39–44. [PubMed]
82. Andersson M, Usiello A, Borgkvist A, Pozzi L, Dominguez C, Fienberg AA, Svenningsson P, Fredholm BB, Borrelli E, Greengard P, Fisone G. 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;25:8432–8438. [PubMed]
83. Borgkvist A, Marcellino D, Fuxe K, Greengard P, Fisone G. Regulation of DARPP-32 phosphorylation by Delta9-tetrahydrocannabinol. Neuropharmacology. 2008;54:31–35.[PubMed]
84. Hampson RE, Mu J, Deadwyler SA. Cannabinoid and kappa opioid receptors reduce potassium K current via activation of G(s) proteins in cultured hippocampal neurons. J Neurophysiol.2000;84:2356–2364. [PubMed]
85. Bouaboula M, Poinot-Chazel C, Bourrie B, Canat X, Calandra B, Rinaldi-Carmona M, Le Fur G, Casellas P. Activation of mitogen-activated protein kinases by stimulation of the central cannabinoid receptor CB1. Biochem J. 1995;312(Pt 2):637–641.[PMC free article] [PubMed]
86. Guzman M, Sanchez C. Effects of cannabinoids on energy metabolism. Life Sci. 1999;65:657–664. [PubMed]
87. Sanchez C, Galve-Roperh I, Rueda D, Guzman M. Involvement of sphingomyelin hydrolysis and the mitogen-activated protein kinase cascade in the Delta9-tetrahydrocannabinol-induced stimulation of glucose metabolism in primary astrocytes. Mol Pharmacol. 1998;54:834–843.[PubMed]
88. Wartmann M, Campbell D, Subramanian A, Burstein SH, Davis RJ. The MAP kinase signal transduction pathway is activated by the endogenous cannabinoid anandamide. FEBS Lett. 1995;359:133–136. [PubMed]
89. Galve-Roperh I, Rueda D, Gomez del Pulgar T, Velasco G, Guzman M. Mechanism of extracellular signal-regulated kinase activation by the CB(1) cannabinoid receptor. Mol Pharmacol.2002;62:1385–1392. [PubMed]
90. Gomez dP, Velasco G, Guzman M. The CB1 cannabinoid receptor is coupled to the activation of protein kinase B/Akt.Biochem J. 2000;347:369–373. [PMC free article] [PubMed]
91. Bouaboula M, Perrachon S, Milligan L, Canat X, Rinaldi-Carmona M, Portier M, Barth F, Calandra B, Pecceu F, Lupker J, Maffrand JP, Le Fur G, Casellas P. A selective inverse agonist for central cannabinoid receptor inhibits mitogen-activated protein kinase activation stimulated by insulin or insulin-like growth factor 1. Evidence for a new model of receptor/ligand interactions.J Biol Chem. 1997;272:22330–22339. [PubMed]
92. Bouaboula M, Desnoyer N, Carayon P, Combes T, Casellas P. Gi protein modulation induced by a selective inverse agonist for the peripheral cannabinoid receptor CB2: implication for intracellular signalization cross-regulation. Mol Pharmacol.1999;55:473–480. [PubMed]
93. Hart S, Fischer OM, Ullrich A. Cannabinoids induce cancer cell proliferation via tumor necrosis factor alpha-converting enzyme (TACE/ADAM17)-mediated transactivation of the epidermal growth factor receptor. Cancer Res. 2004;64:1943–1950. [PubMed]
94. Derkinderen P, Valjent E, Toutant M, Corvol JC, Enslen H, Ledent C, Trzaskos J, Caboche J, Girault JA. Regulation of extracellular signal-regulated kinase by cannabinoids in hippocampus. J Neurosci. 2003;23:2371–2382. [PubMed]
95. Davis MI, Ronesi J, Lovinger DM. A predominant role for inhibition of the adenylate cyclase/protein kinase A pathway in ERK activation by cannabinoid receptor 1 in N1E-115 neuroblastoma cells. J Biol Chem. 2003;278:48973–48980.[PubMed]
96. Korzh A, Keren O, Gafni M, Bar-Josef H, Sarne Y. Modulation of extracellular signal-regulated kinase (ERK) by opioid and cannabinoid receptors that are expressed in the same cell. Brain Res. 2008;1189:23–32. [PubMed]
97. Nabemoto M, Mashimo M, Someya A, Nakamura H, Hirabayashi T, Fujino H, Kaneko M, Okuma Y, Saito T, Yamaguchi N, Murayama T. Release of arachidonic acid by 2-arachidonoyl glycerol and HU210 in PC12 cells; roles of Src, phospholipase C and cytosolic phospholipase A(2)alpha. Eur J Pharmacol. 2008 [PubMed]
98. Rubovitch V, Gafni M, Sarne Y. The involvement of VEGF receptors and MAPK in the cannabinoid potentiation of Ca2+ flux into N18TG2 neuroblastoma cells. Brain Res Mol Brain Res.2004;120:138–144. [PubMed]
99. Berghuis P, Dobszay MB, Wang X, Spano S, Ledda F, Sousa KM, Schulte G, Ernfors P, Mackie K, Paratcha G, Hurd YL, Harkany T. Endocannabinoids regulate interneuron migration and morphogenesis by transactivating the TrkB receptor. Proc Natl Acad Sci USA. 2005;102:19115–19120. [PMC free article][PubMed]
100. Rueda D, Navarro B, Martinez-Serrano A, Guzman M, Galve-Roperh I. The endocannabinoid anandamide inhibits neuronal progenitor cell differentiation through attenuation of the Rap1/B-Raf/ERK pathway. J Biol Chem. 2002;277:46645–46650. [PubMed]
101. Liu J, Gao B, Mirshahi F, Sanyal AJ, Khanolkar AD, Makriyannis A, Kunos G. Functional CB1 cannabinoid receptors in human vascular endothelial cells. Biochem J. 2000;346(Pt 3):835–840. [PMC free article] [PubMed]
102. Derkinderen P, Ledent C, Parmentier M, Girault JA. Cannabinoids activate p38 mitogen-activated protein kinases through CB1 receptors in hippocampus. J Neurochem.2001;77:957–960. [PubMed]
103. Rueda D, Galve-Roperh I, Haro A, Guzman M. The CB(1) cannabinoid receptor is coupled to the activation of c-Jun N-terminal kinase. Mol Pharmacol. 2000;58:814–820. [PubMed]
104. He JC, Neves SR, Jordan JD, Iyengar R. Role of the Go/i signaling network in the regulation of neurite outgrowth. Can J Physiol Pharmacol. 2006;84:687–694. [PubMed]
105. Jordan JD, He JC, Eungdamrong NJ, Gomes I, Ali W, Nguyen T, Bivona TG, Philips MR, Devi LA, Iyengar R. Cannabinoid receptor-induced neurite outgrowth is mediated by Rap1 activation through G(alpha)o/i-triggered proteasomal degradation of Rap1GAPII. J Biol Chem. 2005;280:11413–11421.[PubMed]
106. He JC, Gomes I, Nguyen T, Jayaram G, Ram PT, Devi LA, Iyengar R. The G alpha(o/i)-coupled cannabinoid receptor-mediated neurite outgrowth involves Rap regulation of Src and Stat3. J Biol Chem. 2005;280:33426–33434. [PubMed]
107. Stefano GB, Salzet M, Magazine HI, Bilfinger TV. Antagonism of LPS and IFN-gamma induction of iNOS in human saphenous vein endothelium by morphine and anandamide by nitric oxide inhibition of adenylate cyclase. J Cardiovasc Pharmacol. 1998;31:813–820. [PubMed]
108. Fimiani C, Mattocks D, Cavani F, Salzet M, Deutsch DG, Pryor S, Bilfinger TV, Stefano GB. Morphine and anandamide stimulate intracellular calcium transients in human arterial endothelial cells: coupling to nitric oxide release. Cell Signal.1999;11:189–193. [PubMed]
109. Mombouli JV, Schaeffer G, Holzmann S, Kostner GM, Graier WF. Anandamide-induced mobilization of cytosolic Ca2+ in endothelial cells. Br J Pharmacol. 1999;126:1593–1600.[PMC free article] [PubMed]
110. Maccarrone M, Bari M, Lorenzon T, Bisogno T, Di MV, Finazzi-Agro A. Anandamide uptake by human endothelial cells and its regulation by nitric oxide. J Biol Chem. 2000;275:13484–13492. [PubMed]
111. Stefano GB, Liu Y, Goligorsky MS. Cannabinoid receptors are coupled to nitric oxide release in invertebrate immunocytes, microglia, and human monocytes. J Biol Chem. 1996;271:19238–19242. [PubMed]
112. Prevot V, Rialas CM, Croix D, Salzet M, Dupouy JP, Poulain P, Beauvillain JC, Stefano GB. Morphine and anandamide coupling to nitric oxide stimulates GnRH and CRF release from rat median eminence: neurovascular regulation. Brain Res.1998;790:236–244. [PubMed]
113. Jones JD, Carney ST, Vrana KE, Norford DC, Howlett AC. Cannabinoid receptor-mediated translocation of NO-sensitive guanylyl cyclase and production of cyclic GMP in neuronal cells.Neuropharmacology. 2008;54:23–30. [PMC free article][PubMed]
114. Carney ST, Lloyd ML, MacKinnon SE, Newton DC, Jones JD, Howlett AC, Norford DC. Cannabinoid regulation of nitric oxide synthase I (nNOS) in neuronal cells. J Neuroimmune Pharmacol.2009 [PMC free article] [PubMed]
115. Simmons ML, Murphy S. Induction of nitric oxide synthase in glial cells. J Neurochem. 1992;59:897–905. [PubMed]
116. Hillard CJ, Muthian S, Kearn CS. Effects of CB(1) cannabinoid receptor activation on cerebellar granule cell nitric oxide synthase activity. FEBS Lett. 1999;459:277–281. [PubMed]
117. Kim SH, Won SJ, Mao XO, Jin K, Greenberg DA. Molecular mechanisms of cannabinoid protection from neuronal excitotoxicity. Mol Pharmacol. 2006;69:691–696. [PubMed]
118. Rameau GA, Chiu LY, Ziff EB. Bidirectional regulation of neuronal nitric-oxide synthase phosphorylation at serine 847 by the N-methyl-D-aspartate receptor. J Biol Chem.2004;279:14307–14314. [PubMed]
119. Rameau GA, Tukey DS, Garcin-Hosfield ED, Titcombe RF, Misra C, Khatri L, Getzoff ED, Ziff EB. Biphasic coupling of neuronal nitric oxide synthase phosphorylation to the NMDA receptor regulates AMPA receptor trafficking and neuronal cell death. J Neurosci. 2007;27:3445–3455. [PubMed]
120. Jeon YJ, Yang KH, Pulaski JT, Kaminski NE. Attenuation of inducible nitric oxide synthase gene expression by delta 9-tetrahydrocannabinol is mediated through the inhibition of nuclear factor-kappa B/Rel activation. Mol Pharmacol.1996;50:334–341. [PubMed]
121. Cabral GA, Harmon KN, Carlisle SJ. Cannabinoid-mediated inhibition of inducible nitric oxide production by rat microglial cells: evidence for CB1 receptor participation. Adv Exp Med Biol.2001;493:207–214. [PubMed]
122. Molina-Holgado F, Lledo A, Guaza C. Anandamide suppresses nitric oxide and TNF-alpha responses to Theiler’s virus or endotoxin in astrocytes. Neuroreport. 1997;8:1929–1933.[PubMed]
123. Molina-Holgado F, Molina-Holgado E, Guaza C, Rothwell NJ. Role of CB1 and CB2 receptors in the inhibitory effects of cannabinoids on lipopolysaccharide-induced nitric oxide release in astrocyte cultures. J Neurosci Res. 2002;67:829–836.[PubMed]
124. Esposito G, Ligresti A, Izzo AA, Bisogno T, Ruvo M, Di Rosa M, Di MV, Iuvone T. The endocannabinoid system protects rat glioma cells against HIV-1 Tat protein-induced cytotoxicity. Mechanism and regulation. J Biol Chem. 2002;277:50348–50354. [PubMed]
125. McIntosh HH, Song C, Howlett AC. CB1 cannabinoid receptor: cellular regulation and distribution in N18TG2 neuroblastoma cells. Brain Res Mol Brain Res. 1998;53:163–173.[PubMed]
126. Rinaldi-Carmona M, Le Duigou A, Oustric D, Barth F, Bouaboula M, Carayon P, Casellas P, Le Fur G. Modulation of CB1 cannabinoid receptor functions after a long-term exposure to agonist or inverse agonist in the Chinese hamster ovary cell expression system. J Pharmacol Exp Ther. 1998;287:1038–1047.[PubMed]
127. Leterrier C, Laine J, Darmon M, Boudin H, Rossier J, Lenkei Z. Constitutive activation drives compartment-selective endocytosis and axonal targeting of type 1 cannabinoid receptors.J Neurosci. 2006;26:3141–3153. [PubMed]
128. Leterrier C, Bonnard D, Carrel D, Rossier J, Lenkei Z. Constitutive endocytic cycle of the CB1 cannabinoid receptor. J Biol Chem. 2004;279:36013–36021. [PubMed]
129. McDonald NA, Henstridge CM, Connolly CN, Irving AJ. An essential role for constitutive endocytosis, but not activity, in the axonal targeting of the CB1 cannabinoid receptor. Mol Pharmacol. 2007;71:976–984. [PubMed]
130. Rozenfeld R, Devi LA. Regulation of CB1 cannabinoid receptor trafficking by the adaptor protein AP-3. FASEB J.2008;22:2311–2322. [PMC free article] [PubMed]
131. Ahn K, McKinney MK, Cravatt BF. Enzymatic pathways that regulate endocannabinoid signaling in the nervous system. Chem Rev. 2008;108:1687–1707. [PMC free article] [PubMed]
132. Dromey JR, Pfleger KD. G protein coupled receptors as drug targets: the role of beta-arrestins. Endocr Metab Immune Disord Drug Targets. 2008;8:51–61. [PubMed]
133. Jin W, Brown S, Roche JP, Hsieh C, Celver JP, Kovoor A, Chavkin C, Mackie K. Distinct domains of the CB1 cannabinoid receptor mediate desensitization and internalization. J Neurosci.1999;19:3773–3780. [PubMed]
134. Kouznetsova M, Kelley B, Shen M, Thayer SA. Desensitization of cannabinoid-mediated presynaptic inhibition of neurotransmission between rat hippocampal neurons in culture.Mol Pharmacol. 2002;61:477–485. [PubMed]
135. Daigle TL, Kearn CS, Mackie K. Rapid CB1 cannabinoid receptor desensitization defines the time course of ERK1/2 MAP kinase signaling. Neuropharmacology. 2008;54:36–44.[PMC free article] [PubMed]
136. Bakshi K, Mercier RW, Pavlopoulos S. Interaction of a fragment of the cannabinoid CB1 receptor C-terminus with arrestin-2. FEBS Lett. 2007;581:5009–5016. [PMC free article][PubMed]
137. Daigle TL, Kwok ML, Mackie K. Regulation of CB1 cannabinoid receptor internalization by a promiscuous phosphorylation-dependent mechanism. J Neurochem.2008;106:70–82. [PMC free article] [PubMed]
138. Breivogel CS, Lambert JM, Gerfin S, Huffman JW, Razdan RK. Sensitivity to delta9-tetrahydrocannabinol is selectively enhanced in beta-arrestin2 −/− mice. Behav Pharmacol.2008;19:298–307. [PMC free article] [PubMed]
139. Rubino T, Vigano D, Premoli F, Castiglioni C, Bianchessi S, Zippel R, Parolaro D. Changes in the expression of G protein-coupled receptor kinases and beta-arrestins in mouse brain during cannabinoid tolerance: a role for RAS-ERK cascade. Mol Neurobiol. 2006;33:199–213. [PubMed]
140. Danglot L, Galli T. What is the function of neuronal AP-3?Biol Cell. 2007;99:349–361. [PubMed]
141. Faundez V, Horng JT, Kelly RB. ADP ribosylation factor 1 is required for synaptic vesicle budding in PC12 cells. J Cell Biol.1997;138:505–515. [PMC free article] [PubMed]
142. Faundez V, Horng JT, Kelly RB. A function for the AP3 coat complex in synaptic vesicle formation from endosomes. Cell.1998;93:423–432. [PubMed]
143. Shi G, Faundez V, Roos J, Dell’Angelica EC, Kelly RB. Neuroendocrine synaptic vesicles are formed in vitro by both clathrin-dependent and clathrin-independent pathways. J Cell Biol. 1998;143:947–955. [PMC free article] [PubMed]
144. Lichtenstein Y, Desnos C, Faundez V, Kelly RB, Clift-O’Grady L. Vesiculation and sorting from PC12-derived endosomes in vitro.Proc Natl Acad Sci USA. 1998;95:11223–11228. [PMC free article][PubMed]
145. Martini L, Waldhoer M, Pusch M, Kharazia V, Fong J, Lee JH, Freissmuth C, Whistler JL. Ligand-induced down-regulation of the cannabinoid 1 receptor is mediated by the G-protein-coupled receptor-associated sorting protein GASP1. FASEB J.2007;21:802–811. [PubMed]
146. Tappe-Theodor A, Agarwal N, Katona I, Rubino T, Martini L, Swiercz J, Mackie K, Monyer H, Parolaro D, Whistler J, Kuner T, Kuner R. A molecular basis of analgesic tolerance to cannabinoids. J Neurosci. 2007;27:4165–4177. [PubMed]
147. Boeuf J, Trigo JM, Moreau PH, Lecourtier L, Vogel E, Cassel JC, Mathis C, Klosen P, Maldonado R, Simonin F. Attenuated behavioural responses to acute and chronic cocaine in GASP-1-deficient mice. Eur J Neurosci. 2009;30:860–868. [PubMed]
148. Simonin F, Karcher P, Boeuf JJ, Matifas A, Kieffer BL. Identification of a novel family of G protein-coupled receptor associated sorting proteins. J Neurochem. 2004;89:766–775.[PubMed]
149. Nie J, Lewis DL. Structural domains of the CB1 cannabinoid receptor that contribute to constitutive activity and G-protein sequestration. J Neurosci. 2001;21:8758–8764. [PubMed]
150. Niehaus JL, Liu Y, Wallis KT, Egertova M, Bhartur SG, Mukhopadhyay S, Shi S, He H, Selley DE, Howlett AC, Elphick MR, Lewis DL. CB1 cannabinoid receptor activity is modulated by the cannabinoid receptor interacting protein CRIP 1a. Mol Pharmacol. 2007;72:1557–1566. [PMC free article] [PubMed]
151. Elphick MR, Wallis KT, Liu Y, Lewis DL. Egertova Localization of the CB1 cannabinoid receptor interacting protein (CRIP1a) in the brain. Annual Symposium on the Cannabinoids.2004;14:19.
152. Ludanyi A, Eross L, Czirjak S, Vajda J, Halasz P, Watanabe M, Palkovits M, Magloczky Z, Freund TF, Katona I. Downregulation of the CB1 cannabinoid receptor and related molecular elements of the endocannabinoid system in epileptic human hippocampus. J Neurosci. 2008;28:2976–2990.[PubMed]
153. Resh MD. Palmitoylation of ligands, receptors, and intracellular signaling molecules. Sci STKE. 2006;2006:re14.[PubMed]
154. Shiraishi-Yamaguchi Y, Furuichi T. The Homer family proteins. Genome Biol. 2007;8:206. [PMC free article] [PubMed]
155. Tu JC, Xiao B, Naisbitt S, Yuan JP, Petralia RS, Brakeman P, Doan A, Aakalu VK, Lanahan AA, Sheng M, Worley PF. Coupling of mGluR/Homer and PSD-95 complexes by the Shank family of postsynaptic density proteins. Neuron. 1999;23:583–592.[PubMed]
156. Xiao B, Tu JC, Petralia RS, Yuan JP, Doan A, Breder CD, Ruggiero A, Lanahan AA, Wenthold RJ, Worley PF. Homer regulates the association of group 1 metabotropic glutamate receptors with multivalent complexes of homer-related, synaptic proteins. Neuron. 1998;21:707–716. [PubMed]
157. Szumlinski KK, Kalivas PW, Worley PF. Homer proteins: implications for neuropsychiatric disorders. Curr Opin Neurobiol.2006;16:251–257. [PubMed]
158. Roche KW, Tu JC, Petralia RS, Xiao B, Wenthold RJ, Worley PF. Homer 1b regulates the trafficking of group I metabotropic glutamate receptors. J Biol Chem. 1999;274:25953–25957.[PubMed]
159. Bergson C, Levenson R, Goldman-Rakic PS, Lidow MS. Dopamine receptor-interacting proteins: the Ca(2+) connection in dopamine signaling. Trends Pharmacol Sci. 2003;24:486–492.[PubMed]
160. Bermak JC, Li M, Bullock C, Zhou QY. Regulation of transport of the dopamine D1 receptor by a new membrane-associated ER protein. Nat Cell Biol. 2001;3:492–498. [PubMed]
161. Dupre DJ, Hebert TE. Biosynthesis and trafficking of seven transmembrane receptor signalling complexes. Cell Signal.2006;18:1549–1559. [PubMed]
162. Guzman M, Sanchez C, Galve-Roperh I. Cannabinoids and cell fate. Pharmacol Ther. 2002;95:175–184. [PubMed]
163. Velasco G, Galve-Roperh I, Sanchez C, Blazquez C, Haro A, Guzman M. Cannabinoids and ceramide: two lipids acting hand-by-hand. Life Sci. 2005;77:1723–1731. [PubMed]
164. Ramer R, Weinzierl U, Schwind B, Brune K, Hinz B. Ceramide is involved in r(+)-methanandamide-induced cyclooxygenase-2 expression in human neuroglioma cells. Mol Pharmacol. 2003;64:1189–1198. [PubMed]
165. Galve-Roperh I, Sanchez C, Cortes ML, Gomez del Pulgar TG, Izquierdo M, Guzman M. Anti-tumoral action of cannabinoids: involvement of sustained ceramide accumulation and extracellular signal-regulated kinase activation. Nat Med.2000;6:313–319. [PubMed]
166. Blazquez C, Sanchez C, Velasco G, Guzman M. Role of carnitine palmitoyltransferase I in the control of ketogenesis in primary cultures of rat astrocytes. J Neurochem. 1998;71:1597–1606. [PubMed]
167. Sanchez C, Rueda D, Segui B, Galve-Roperh I, Levade T, Guzman M. The CB(1) cannabinoid receptor of astrocytes is coupled to sphingomyelin hydrolysis through the adaptor protein FAN. Mol Pharmacol. 2001;59:955–959. [PubMed]
168. Sanchez C, Galve-Roperh I, Canova C, Brachet P, Guzman M. Delta9-tetrahydrocannabinol induces apoptosis in C6 glioma cells. FEBS Lett. 1998;436:6–10. [PubMed]
potp font 1