Canna~Fangled Abstracts

International Union of Basic and Clinical Pharmacology. LXXIX. Cannabinoid Receptors and Their Ligands: Beyond CB1 and CB2

By December 5, 2010April 19th, 2021No Comments

Abstract

There are at least two types of cannabinoid receptors (CB1 and CB2). Ligands activating these G protein-coupled receptors (GPCRs) include the phytocannabinoid Δ9-tetrahydrocannabinol, numerous synthetic compounds, and endogenous compounds known as endocannabinoids. Cannabinoid receptor antagonists have also been developed. Some of these ligands activate or block one type of cannabinoid receptor more potently than the other type. This review summarizes current data indicating the extent to which cannabinoid receptor ligands undergo orthosteric or allosteric interactions with non-CB1, non-CB2 established GPCRs, deorphanized receptors such as GPR55, ligand-gated ion channels, transient receptor potential (TRP) channels, and other ion channels or peroxisome proliferator-activated nuclear receptors. From these data, it is clear that some ligands that interact similarly with CB1 and/or CB2 receptors are likely to display significantly different pharmacological profiles. The review also lists some criteria that any novel “CB3” cannabinoid receptor or channel should fulfil and concludes that these criteria are not currently met by any non-CB1, non-CB2 pharmacological receptor or channel. However, it does identify certain pharmacological targets that should be investigated further as potential CB3 receptors or channels. These include TRP vanilloid 1, which possibly functions as an ionotropic cannabinoid receptor under physiological and/or pathological conditions, and some deorphanized GPCRs. Also discussed are 1) the ability of CB1 receptors to form heteromeric complexes with certain other GPCRs, 2) phylogenetic relationships that exist between CB1/CB2 receptors and other GPCRs, 3) evidence for the existence of several as-yet-uncharacterized non-CB1, non-CB2 cannabinoid receptors; and 4) current cannabinoid receptor nomenclature.

I. Introduction

The main purpose of this review is to consider current knowledge about the extent to which established cannabinoid CB1 and CB2 receptor ligands target non-CB1, non-CB2 receptors or ion channels (section III). These considerations are preceded by a brief overview of the pharmacology of cannabinoid CB1 and CB2 receptors and their ligands and by a discussion of the evidence that CB1 receptors form heteromeric complexes with certain other receptors (section II). Also discussed in this review is the extent to which phylogenetic relationships exist between cannabinoid CB1 or CB2 receptors and other receptors (section IV). It ends by addressing the questions, first of whether cannabinoid CB1 and CB2 receptors should be renamed (section V), and second, of whether any non-CB1, non-CB2 receptor or channel should be reclassified as a cannabinoid “CB3” receptor or channel (section VI). The terms “CB1-selective” and “CB2-selective” have been used in this review to describe compounds that interact more potently with one cannabinoid receptor (CB1 or CB2) than with the other, irrespective of whether any of these compounds target CB1 or CB2 receptors more potently than a non-CB1, non-CB2 receptor or channel. Receptor nomenclature in this article complies with the recommendations of the International Union of Basic and Clinical Pharmacology nomenclature and also conforms to .

II. Cannabinoid CB1 and CB2 Receptors and their Ligands

A. CB1 and CB2 Receptors

The discovery in 1990 that an orphan G protein-coupled receptor (SKR6) derived from a rat cerebral cortex cDNA library mediates pharmacological effects of (−)-Δ9-tetrahydrocannabinol (Δ9-THC1), the main psychoactive constituent of cannabis, established the identity of the first cannabinoid receptor, which we now refer to as CB1 (). Three years later, in 1993, a G protein-coupled receptor (CX5) expressed in the human promyelocytic leukemic cell line HL60 was identified as a second cannabinoid receptor and named CB2 (). CB1 and CB2 receptors are members of the superfamily of G protein-coupled receptors (GPCRs). As discussed in greater detail elsewhere (), both these receptors inhibit adenylyl cyclase and activate mitogen-activated protein kinase by signaling through Gi/o proteins, which for the CB1 receptor can also mediate activation of A-type and inwardly rectifying potassium currents and inhibition of N- and P/Q-type calcium currents. In addition, CB1 receptors can signal through Gs proteins (). The ability of CB1 and CB2 receptors to signal through Gi/o proteins and, further downstream, through adenylyl cyclase is frequently exploited in two widely used in vitro bioassays: the [35S]GTPγS binding assay and the cAMP assay (). As well as orthosteric site(s), the CB1 receptor possesses one or more allosteric sites that can be targeted by ligands in a manner that enhances or inhibits the activation of this receptor by direct agonists ().

CB1 receptors are found mainly at the terminals of central and peripheral neurons, where they usually mediate inhibition of ongoing release of a number of different excitatory and inhibitory neurotransmitters (for review, see ). The distribution of these receptors within the central nervous system is such that their activation can affect processes such as cognition and memory, alter the control of motor function, and induce signs of analgesia. As to CB2 receptors, these are located predominantly in immune cells and, when activated, can modulate immune cell migration and cytokine release both outside and within the brain (for review, see ). There is also evidence that 1) some CB1 receptors are expressed by non-neuronal cells, including immune cells (), and 2) that CB2 receptors are expressed by some neurons, both within the brain and elsewhere (). The role of neuronal CB2 receptors remains to be established.

Finally, several polymorphisms in the genes of CB1 (CNR1) and CB2 (CNR2) receptors and in their proteins have been identified. Some of these have been linked to certain disorders that for CNR1 include 1) schizophrenia and 2) depression in Parkinson’s disease and for CNR2 include postmenopausal osteoporosis (for review, see ).

B. The Endocannabinoid System

The cloning of the CB1 receptor was followed by the discovery that mammalian tissues can both synthesize cannabinoid receptor agonists and release them onto cannabinoid receptors. The first of these “endocannabinoids” to be identified were N-arachidonoylethanolamine (anandamide) and 2-arachidonoyl glycerol (2-AG) (), both of which are synthesized on demand in response to elevations of intracellular calcium (for review, see ). Other compounds may also serve as endocannabinoids. These include N-dihomo-γ-linolenoylethanolamine, N-docosatetraenoylethanolamine, O-arachidonoylethanolamine (virodhamine), oleamide, N-arachidonoyl dopamine and N-oleoyl dopamine (for review, see ). Endocannabinoids and their receptors constitute the “endocannabinoid system.”

C. Cannabinoid CB1 and CB2 Receptor Ligands

 

1. Agonists that Target CB1 and CB2 Receptors with Similar Potency.

Several cannabinoid receptor agonists possess similar affinities for CB1 and CB2 receptors (Table 1). When classified according to their chemical structures (Fig. 1), these agonists fall essentially into four main groups: classical, nonclassical, aminoalkylindole, and eicosanoid (for review, see ,).

TABLE 1

Some Ki values of cannabinoid CB1/CB2 receptor ligands for the in vitro displacement of a tritiated compounda from specific binding sites on rat, mouse, or human CB1 and CB2 receptors

Unless otherwise indicated in the Reference column, see  for references. Structures of the compounds listed are shown in Figs. 1 to to55.

Cannabinoid Receptor Ligand Ki


Reference
CB1 CB2
nM
Section II.C.1
    (−)-Δ9-THC 5.05–80.3 3.13–75.3 See  for references
    HU-210 0.06–0.73 0.17–0.52
    CP55940 0.5–5.0 0.69–2.8
    R-(+)-WIN55212 1.89–123 0.28–16.2
    Anandamide 61–543 279–1940
    2-AG 58.3, 472 145, 1400
Section II.C.2
    Agonists with higher CB1 than CB2 affinity
        ACEA 1.4, 5.29 195, >2000
        Arachidonylcyclopropylamide 2.2 715
        R-(+)-methanandamide 17.9–28.3 815–868
        Noladin ether 21.2 >3000
    Agonists with higher CB2 than CB1 affinity
        JWH-133 677 3.4
        HU-308 >10000 22.7
        JWH-015 383 13.8
        AM1241 280 3.4
Section II.C.3
    Rimonabant (SR141716A) 1.8–12.3 514–13,200
    AM251 7.49 2290
    AM281 12 4200
    LY320135 141 14,900
    Taranabant 0.13, 0.27 170, 310
    NESS 0327 0.00035 21
    O-2050 2.5, 1.7 1.5 ; A. Thomas and R. G. Pertwee, unpublished data
Section II.C.4
    SR144528 50.3–>10,000 0.28–5.6
    AM630 5152 31.2
    JTE-907 2370 35.9
Section II.C.5
    11-OH-Δ8-THC 25.8 7.4
    Ajulemic acid 5.7, 32.3 56.1, 170.5 ; see also , for references
    Cannabinol 120–1130 96–301 See  for references
    Cannabigerol 81 2600
    Cannabidiol 4350–>10,000 2399–>10,000 See  for references
    N-Arachidonoyl dopamine 250 12,000
    Virodhamine 912 N.D.

N.D., no data.

aUsually [3H]CP55940, but sometimes [3H]SR141716A, [3H]R-(+)-WIN55212, or HU-243 (i.e., [3H]HU-210).

An external file that holds a picture, illustration, etc. Object name is zpg0041022940001.jpg

The structures of (−)-Δ9-tetrahydrocannabinol [(−)-Δ9-THC], HU-210, CP55940, R-(+)-WIN55212, anandamide, and 2-AG.

  • The classical group consists of dibenzopyran derivatives. It includes Δ9-THC, the main psychoactive constituent of cannabis, and (6aR)-trans-3-(1,1-dimethylheptyl)-6a, 7,10,10a-tetrahydro-1-hydroxy-6,6-dimethyl-6H-dibenzo[b,d]pyran-9-methanol (HU-210), a synthetic analog of (−)-Δ8-tetrahydrocannabinol. HU-210 displays high affinity for CB1 and CB2 receptors and also high potency and relative intrinsic activity as a cannabinoid receptor agonist. These properties are all thought to result mainly from the presence of its dimethylheptyl side chain. Δ9-THC possesses significantly lower CB1 and CB2 affinity than HU-210 and lower relative intrinsic activity at these receptors, an indication that Δ9-THC is a cannabinoid receptor partial agonist. Moreover, it displays even less relative intrinsic activity at CB2 than at CB1 receptors.

  • The nonclassical group contains bicyclic and tricyclic analogs of Δ9-THC that lack a pyran ring. A well known member of this group is (−)-cis-3-[2-hydroxy-4-(1,1-dimethylheptyl)phenyl]-trans-4-(3-hydroxypropyl)cyclohexanol (CP55940). This has been found to have slightly lower CB1 and CB2 affinities than HU-210 in some investigations but does seem to possess HU-210-like CB1 and CB2 receptor relative intrinsic activity.

  • Members of the aminoalkylindole group of cannabinoid CB1/CB2 receptor agonists have structures that differ markedly from those of both classical and nonclassical cannabinoids. The best known member of this group is R-(+)-[2,3-dihydro-5-methyl-3-(4-morpholinylmethyl)pyrrolo[1,2,3-de]-1,4-benzoxazin-6-yl]-1-naphthalenylmethanone mesylate [R-(+)-WIN55212]. This displays CP55940- and HU-210-like relative intrinsic activity at both CB1 and CB2 receptors. However, unlike HU-210 and CP55940, it has been found in some investigations to possess slightly higher CB2 than CB1 affinity.

  • Members of the eicosanoid group of cannabinoid CB1/CB2 receptor agonists have structures quite unlike those of classical, nonclassical, or aminoalkylindole cannabinoids. Two prominent members of this group are the endocannabinoids anandamide and 2-AG. Like Δ9-THC, anandamide behaves as a CB1 and CB2 receptor partial agonist and displays lower relative intrinsic activity for CB2 than for CB1. Its affinity for the CB1 receptor is also similar to that of Δ9-THC. This eicosanoid does, however, have slightly lower receptor affinity for CB2 than for CB1 and consequently displays less affinity for the CB2 receptor than Δ9-THC. 2-AG also has slightly less receptor affinity for CB2 than for CB1. It seems to have lower CB1 receptor potency than CP55940 but higher CB1 and CB2 receptor potency than anandamide and higher CB1 receptor relative intrinsic activity than anandamide or CP55940.

 

2. CB1– and CB2-Selective Cannabinoid Receptor Agonists.

Compounds that are significantly more potent at activating CB1 than CB2 receptors include three synthetic analogs of anandamide (Table 1 and Fig. 2): R-(+)-methanandamide, arachidonyl-2′-chloroethylamide (ACEA), and arachidonylcyclopropylamide (). Each of these compounds possesses significant potency and relative intrinsic activity as a CB1 receptor agonist. ACEA and arachidonylcyclopropylamide are both substrates for the anandamide-metabolizing enzyme fatty acid amide hydrolase, whereas this enzyme does not readily hydrolyze R-(+)-methanandamide. Noladin ether (2-arachidonyl glyceryl ether) () is also a CB1-selective agonist. It has been reported to possess CP55940-like CB1 receptor relative intrinsic activity but less potency as a CB1 receptor agonist than either CP55940 or 2-AG (). As to CB2-selective agonists (Table 1 and Fig. 2), those most frequently used as pharmacological tools are (6aR,10aR)-3-(1,1-dimethylbutyl)-6a,7,10,10a-tetrahydro-6,6,9-trimethyl-6H-dibenzo[b,d] pyran (JWH-133; a classical cannabinoid), {4-[4-(1,1-dimethylheptyl)-2,6-dimethoxy-phenyl]-6,6-dimethyl-bicyclo[3.1.1]hept-2-en-2-yl}-methanol (HU-308; a nonclassical cannabinoid), and (2-methyl-1-propyl-1H-indol-3-yl)-1-naphthalenylmethanone (JWH-015) and R-3-(2-iodo-5-nitrobenzoyl)-1-methyl-2-piperidinylmethyl)-1H-indole (AM1241) (aminoalkylindoles) (for review, see ).

An external file that holds a picture, illustration, etc. Object name is zpg0041022940002.jpg

The structures of the CB1-selective agonists ACEA, arachidonylcyclopropylamide (ACPA), methanandamide, and noladin ether and of the CB2-selective agonists JWH-133, HU-308, JWH-015, and AM1241.

 

3. CB1-Selective Competitive Antagonists.

As discussed in greater detail elsewhere (), the diarylpyrazole rimonabant (SR141716A), N-(piperidin-1-yl)-5-(4-iodophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboxamide (AM251), 1-(2,4-dichlorophenyl)-5-(4-iodophenyl)-4-methyl-N-4-morpholinyl-1H-pyrazole-3-carboxamide (AM281), 4-[[6-methoxy-2-(4-methoxyphenyl)-3-benzofuranyl]carbonyl]benzonitrile (LY320135), and taranabant (Fig. 3) can all block agonist-induced activation of cannabinoid CB1 receptors in a competitive manner and bind with significantly greater affinity to cannabinoid CB1 than cannabinoid CB2 receptors (Table 1). Although these compounds lack any ability to activate CB1 receptors when administered alone, there is evidence that in some CB1 receptor-containing tissues, they can induce responses opposite in direction from those elicited by a CB1 receptor agonist (). In some instances, at least, this may reflect an ability of these compounds to decrease the spontaneous coupling of CB1 receptors to their effector mechanisms that it is thought can occur in the absence of exogenously added or endogenously released CB1 agonists. There is also evidence that at least one of these compounds, rimonabant, can produce inverse cannabimimetic effects in a CB1 receptor-independent manner ().

An external file that holds a picture, illustration, etc. Object name is zpg0041022940003.jpg

The structures of the CB1-selective antagonists/inverse agonists, rimonabant, AM251, AM281, LY320135, and taranabant and of the CB2-selective antagonists/inverse agonists SR144528 and AM630.

Some CB1 receptor competitive antagonists have been developed that lack any detectable ability to induce signs of inverse agonism at the CB1 receptor when administered alone. One example of such a “neutral” antagonist (Table 1) is N-piperidinyl-[8-chloro-1-(2,4-dichlorophenyl)-1,4,5,6-tetrahydrobenzo[6,7]cyclohepta[1,2-c]pyrazole-3-carboxamide] (NESS O327) (Fig. 4), which is a structural analog of rimonabant and displays markedly higher affinity for CB1 than for CB2 receptors. This compound behaves as CB1 receptor antagonist both in vitro and in vivo and yet, by itself, does not affect [35S]GTPγS binding to rat cerebellar membranes (). Several other compounds have been reported to behave as neutral cannabinoid CB1 receptor antagonists (). These include (6aR,10aR)-3-(1-methanesulfonylamino-4-hexyn-6-yl)-6a,7,10,10a-tetrahydro-6,6,9-trimethyl-6H-dibenzo[b,d]pyran (O-2050) (Fig. 4), a sulfonamide analog of Δ8-tetrahydrocannabinol with an acetylenic side chain. It is noteworthy that the classification of this compound as a neutral antagonist is based on a very limited set of data, prompting a need for further research into its CB1 receptor pharmacology.

An external file that holds a picture, illustration, etc. Object name is zpg0041022940004.jpg

The structures of NESS O327 and O-2050.

 

4. CB2-Selective Competitive Antagonists.

[6-Iodo-2-methyl-1-[2-(4-morpholinyl)ethyl]-1H-indol-3-yl](4-methoxyphenyl)methanone (6-iodopravadoline) (AM630) and the diarylpyrazole N-[(1S)-endo-1,3,3-trimethyl bicyclo [2.2.1]heptan-2-yl]-5-(4-chloro-3-methylphenyl)-1-(4-methylbenzyl)-pyrazole-3-carboxamide (SR144528) (Fig. 3) are both more potent at blocking CB2 than CB1 receptor activation. They display much higher affinity for CB2 than for CB1 receptors (Table 1) and block agonist-induced CB2 receptor activation in a competitive manner (for review, see ). Both these compounds are thought to be CB2 receptor inverse agonists rather than neutral antagonists, because when administered by themselves, they can produce inverse cannabimimetic effects in CB2 receptor-expressing tissues (). Other notable examples of CB2-selective cannabinoid receptor antagonists/inverse agonists include N-(1,3-benzodioxol-5-ylmethyl)-1,2-dihydro-7-methoxy-2-oxo-8-(pentyloxy)-3-quinolinecarboxamide (JTE-907) () (Table 1) and the triaryl bis-sulfones N-[1(S)-[4-[[4-methoxy-2-[(4-methoxyphenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]methanesulfonamide) (Sch.225336), N-[1(S)-[4-[[4-chloro-2-[(2-fluorophenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]methanesulfonamide (Sch.356036), and N-[1(S)-[4-[[4-chloro-2-[(2-fluorophenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]-1,1,1-trifluoromethanesulfonamide (Sch.414319) (for review, see ). A neutral antagonist that selectively targets the CB2 receptor has not yet been developed.

 

5. Other Compounds.

Several other compounds that target cannabinoid CB1 and/or CB2 receptors with significant potency are mentioned in one or more subsequent sections of this review. The structures of these compounds and their affinities for CB1 and/or CB2 receptors are shown in Fig. 5 and Table 1, respectively. Two of these compounds, 11-hydroxy-Δ8-tetrahydrocannabinol and ajulemic acid (CT-3), are classical cannabinoids. 11-Hydroxy-Δ8-tetrahydrocannabinol possesses slightly greater potency than Δ9-THC as an inhibitor of adenylyl cyclase in murine neuroblastoma cells (). Compared with CP55940, ajulemic acid displays similar relative intrinsic activity but lower potency at both CB1 and CB2 receptors (). Three of the other compounds are plant cannabinoids (phytocannabinoids). They include cannabinol, which seems to be a CB1 receptor partial agonist (for review, see ). There have also been reports that cannabinol behaves as a reasonably potent CB2 receptor agonist in the cAMP assay but as a CB2 receptor inverse agonist in the [35S]GTPγS assay (for review, see ). The other two phytocannabinoids, cannabidiol and cannabigerol, seem to be CB1 receptor antagonists/inverse agonists (). In contrast, two structural analogs of cannabidiol, abnormal-cannabidiol and 5-methyl-4-[(1R,6R)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enyl]benzene-1,3-diol (O-1602) (Fig. 5) that are mentioned in sections III.A.6, III.A.7, III.A.8, and/or III.H.2, lack significant affinity for the CB1 receptor (for review, see ). Cannabidiol has also been reported to display significant potency in vitro as a CB2 receptor antagonist/inverse agonist (for review, see ).

An external file that holds a picture, illustration, etc. Object name is zpg0041022940005.jpg

The structures of (−)-11-hydroxy-Δ8-tetrahydrocannabinol, ajulemic acid, cannabinol, cannabidiol, abnormal-cannabidiol, O-1602, cannabigerol, virodhamine, and N-arachidonoyl dopamine.

Two other compounds listed in Table 1 are the endogenous eicosanoids virodhamine and N-arachidonoyl dopamine. In one investigation, virodhamine was found to activate CB2 receptors and to exhibit either partial agonist or antagonist activity at CB1 receptors (). However, in another investigation, it was found to behave as a CB1 receptor antagonist/inverse agonist (). As for N-arachidonoyl dopamine, there is evidence that this is a moderately potent CB1 receptor agonist (for review, see ). It is also noteworthy that N-oleoyl dopamine (sections III.B and III.E) possesses some affinity for the CB1 receptor (). However, four other pharmacologically active endogenous acylethanolamides mentioned in sections III and/or IV do not seem to display significant affinity for CB1 and/or CB2 receptors. These are linoleoyl ethanolamide, oleoyl ethanolamide, palmitoyl ethanolamide, and stearoyl ethanolamide (). There have been reports too by  and  that neither CB1 nor CB2 receptors are activated by the putative endogenous GPR55 agonist lysophosphatidyl inositol (section III.A).

D. CB1 Receptor Homomers and Heteromers: Nomenclature and Pharmacology

 

1. CB1 Receptor Homomers.

CB1 receptor homomers were originally identified as immunoreactive, high apparent molecular mass (180–220 kDa) bands on SDS-polyacrylamide gel electrophoresis (). The multimeric form from rat brain and mouse neuroblastoma cells was resistant to dissociation by conditions expected to disrupt disulfide bonds, or ionic or hydrophobic protein interactions in detergent solution (). However, the multimeric form from post mortem human brain seemed to be sensitive to sulfhydryl reagents when solubilized (). An antibody that selectively recognized the high molecular mass form was used to determine that CB1 multimers exhibited the same anatomical distribution as mixed forms in the brain, lending credence to the idea that CB1 receptors exist as homomers in vivo (). Insufficient evidence currently exists to allow any firm conclusions to be drawn about whether monomeric and homomeric forms exhibit differential signal transduction or intracellular trafficking patterns, or how interconversion is physiologically regulated.

 

2. CB1 Receptor Heteromers: A Brief Introduction.

CB1 receptors associate with other GPCRs to form heteromeric complexes (within 50–100 Å) as detected by fluorescence (FRET) or bioluminescence resonance energy transfer (). Guidelines for the nomenclature of associated GPCR proteins define receptor heteromers as “macromolecular complexes composed of functional receptor units with biochemical properties that are different from those of its individual components” (). A multimeric complex would be expected to influence agonist responses in an allosteric manner (). The guidelines defined allosteric interaction in the receptor heteromer as the “intermolecular interaction by which binding of a ligand to one of the receptor units in the receptor heteromer changes the binding properties of another receptor unit” (). Several “CB-X receptor heteromers” conform to the proposed conventions for structurally associated pairs in which the functional interactions influence ligand selectivity or relative intrinsic activity.

 

3. CB1-D2 Dopamine Receptor Heteromers.

CB1-D2 dopamine receptor heteromers were observed in FRET studies of D2-green fluorescent protein and CB1-yellow fluorescent protein fusion proteins expressed in HEK293 cells, with similar D2-CB1 receptor dimerization observed in the absence or presence of the CB1 agonist HU-210, D2 agonist quinpirole, or both together (). However, coimmunoprecipitatable complexes solubilized from a HEK293 cell heterologous expression system were promoted by the presence of agonists for both D2 and CB1 receptors (). Signal transduction in response to agonist stimulation of either CB1 or D2 receptors expressed alone in HEK293 cells is characterized by the Gi/o-dependent inhibition of forskolin-activated adenylyl cyclase (). However, coexpression of both CB1 and D2 receptors caused the effect of CP55940 on cAMP production to switch from inhibition to stimulation (). Combining agonists for both CB1 and D2 receptors also promoted a stimulation of cAMP accumulation when both receptors were expressed (). Synergistic activation of MAPK was also observed in response to simultaneous stimulation by both cannabinoid and dopaminergic agonists (). Neither the stimulation of cAMP production nor the activation of MAPK was pertussis toxin-sensitive, suggesting that Gi/o proteins were not required for the heterodimer responses to agonists in the HEK293 cell model (). However, Gαi1 overexpression inhibited cAMP production, suggesting that CB1-D2 receptor heteromers could interact with Gs only if the environment is not rich in Gi (). Desensitization of the D2-dopamine receptors by pretreatment with quinpirole reversed the ability of CP55940 to stimulate cAMP production (). Evidence can be found to support the hypothesis that CB1-D2 receptor heteromers function in vivo to convert the G protein preference from Gi to Gs. CP55940 decreased the high and low Kd affinities for dopamine (a function of the receptor-G protein interaction) in equilibrium binding assays of D2 receptors in rat striatal membranes (). Cannabinoid and D2 agonists converged to inhibit forskolin-activated adenylyl cyclase as a subadditive response in rodent and monkey striatal membranes (). In cultured striatal cells, costimulation by dopaminergic and cannabinoid agonists converted the response from a Gi-mediated inhibition to a Gs-mediated stimulation of cAMP production (). R-(+)-WIN55212-stimulation was reported to increase cAMP accumulation in globus pallidus slices (). Observations of cannabinoid-stimulated protein kinase A activation suggests that cAMP production is a viable signaling mechanism in basal ganglia (). However, caution should be observed in interpreting in vivo data, in that D1 dopamine () and A2A adenosine receptors () might contribute as components of a heteromeric complex. Immunocytochemical studies suggest that coexpression of CB1 and D2 receptors occurs on the output neurons of the olfactory tubercle, striatum, hippocampus, or neocortex (). Colocalization of CB1 and D2 receptors has been clearly identified in immunoelectron micrographs at the plasma membrane and endomembrane in dendritic spines in the nucleus accumbens (). However, it should be noted that these receptors can also be found individually distant from each other in the same soma or dendrite, or trans-synaptically ().

 

4. CB1-Opioid Receptor Heteromers.

CB1-opioid receptor heteromer formation was detected by an increased bioluminescence resonance energy transfer signal in HEK293 cells coexpressing recombinant yellow fluorescent protein-tagged CB1 and luciferase fused with μ-, δ-, or κ-opioid receptors (). However, functional interaction between CB1 and opioid receptors has thus far been reported only for the CB1-μ-opioid receptor pair. CB1-μ-opioid receptor association may be a factor in intracellular compartmentalization (). Morphine-stimulated [35S]GTPγS binding in HEK293 cells coexpressing CB1 receptors and μ-opioid receptors was attenuated by R-(+)-WIN55212 when calculated as a percentage of basal binding (). However, coexpression of CB1 receptors with μ-opioid receptors in HEK293 cells increased basal [35S]GTPγS binding such that subsequent stimulation by the μ-opioid receptor agonist [d-Ala2,N-Me-Phe4,Gly5-ol]-enkephalin (DAMGO) seemed to be reduced with respect to basal levels (). Basal [35S]GTPγS binding in the CB1-expressing cells was reversed by the CB1 antagonist LY320135, suggesting that the exogenously expressed CB1 receptors were able to constitutively activate a pool of G proteins (). CB1 receptor expression could constitutively reduce morphine- or DAMGO-stimulated MAPK activation in the absence of cannabinoid agonists, and this effect could be blocked by rimonabant but not by the putative neutral CB1 receptor antagonist, O-2050 (). Cannabinoid or opioid agonist actions in the coexpressed receptor model mutually reduced the ability of agonists at the heteroreceptor to activate MAPK (). In Neuro-2A cells expressing CB1 receptors and μ-opioid receptors, simultaneous application of agonists for both receptors suppressed Src and signal transducer and activator of transcription 3 phosphorylation and neurite outgrowth in a reciprocal manner. These findings, in total, would be consistent with mutual heterotropic allosterism. Nevertheless, caution must be exercised when interpreting signal transduction outcomes in heterologous expression systems, because the influence of membrane localization, protein stoichiometry, and accessory proteins may be missing (). Finally, CB1-μ-opioid receptor heteromers may function in cellular models that endogenously express both receptors. In SK-N-SH neuroblastoma and rat striatal membranes, stimulation of [35S]GTPγS binding by the CB1 agonist R-(+)-WIN55212 was reduced by the μ-opioid receptor agonist DAMGO, as was DAMGO-induced stimulation of [35S]GTPγS binding by R-(+)-WIN55212 (). Furthermore, immunoelectron microscopy studies demonstrated that CB1 receptors and μ-opioid receptors colocalize in dendritic spines of the medium spiny neurons of the striatum as well as in interneurons of the dorsal horn of the spinal cord (). However, in interpreting in vivo data, researchers should be cognizant that these receptors are also distributed individually and trans-synaptically ().

 

5. CB1-Orexin-1 Receptor Heteromers.

Evidence for CB1-orexin-1 (OX1) receptor heteromers comes from observations that when expressed in CHO cells, these receptors appear as clusters at the plasma membrane in immunoelectron micrographs (). FRET studies demonstrated close proximity of the CB1– and OX1-fluorescent fusion proteins expressed in HEK293 cells (). Although agonists activated MAPK in both of these receptors when expressed individually in CHO cells, coexpression resulted in a 100-fold increase in MAPK sensitivity to orexin A, a response that was reversed by the CB1 antagonist rimonabant or by pertussis toxin treatment (). On the other hand, coexpression had no appreciable effect on the potency of CP55940 to stimulate MAPK or to inhibit adenylyl cyclase (). CB1 and OX1 receptors were coexpressed predominantly in intracellular vesicles (). Treatment with antagonists for either receptor (CB1, rimonabant; OX1, SB674042) promoted trafficking of both receptors to the cell surface (), suggesting that CB1-OX1 receptor heteromerization influences cellular translocation of these receptors.

 

6. Other CB1 Receptor Heteromers.

Evidence that CB1 and GPCRs, in addition to D2, opioid, or OX1 receptors, may form receptor heteromers is based upon pharmacological cross-talk data, and until other kinds of data become available to support the existence of receptor heteromers, judgement on these pairs must be withheld. The GABAB antagonist phaclofen noncompetitively antagonized R-(+)-WIN55212-stimulated [35S]GTPγS binding in hippocampal membranes, and a CB1 antagonist competitively antagonized the response to 3-aminopropyl(methyl) phosphinic acid (SKF97541) (). Agonist-stimulation of CB1 and GABAB receptors, both endogenously expressed in cerebellar granule cells, resulted in a subadditive inhibition of adenylyl cyclase (.; ). The response to agonists for α2-adrenoceptor- or somatostatin receptor-mediated inhibition of N-type Ca2+ channels was reduced by expression of exogenous CB1 receptors in superior cervical ganglia neurons (). This effect of CB1 receptor expression could be reversed by overexpression of exogenous GαoB, Gβ1, and Gγ3 (), suggesting that these receptors may exist in a complex with shared Gi/o proteins required for Ca2+ channel regulation. The multiplicity of possible receptor “modules” that comprise functional units of signal transduction activity with other receptors, ion channels, and signal transducing effectors points to the complexity involved in interpreting data from in vivo studies (). Future investigations must determine the proximity of these receptors to each other, and provide definitive evidence for heterotropic allosteric interactions before these protein pairs can be advanced as receptor heteromer candidates.

III. The Extent to Which CB1 and CB2 Receptor Ligands Target Non-CB1, Non-CB2 Receptors and Ion Channels

A. The Deorphanized G Protein-Coupled Receptor, GPR55

Human GPR55 (hGPR55) was originally isolated in 1999 as an orphan GPCR with high levels of expression in human striatum () (GenBank accession no. NM_005683.3) and its gene mapped to human chromosome 2q37. GPR55 belongs to group δ of the rhodopsin-like (class A) receptors () and shows low sequence identity to both CB1 (13.5%) and CB2 (14.4%) receptors, which belong to group α of the class A GPCRs. A genetics study () has investigated the origins of the cannabinoid system and has concluded with respect to GPR55 that there is no significant sequence similarity between itself and CB1 or CB2. In particular, there is little sequence similarity in the areas responsible for ligand binding. Initial characterization of human GPR55 identified it as a potential member of either the purinergic or chemokine receptor family based on amino acid homology. However, it is most closely related (30% similarity) to GPR23/LPA4 (GenBank accession no. NM_005296.2), and GPR92/LPA5 (GenBank accession no. NM_020400.5) which have been shown to be lysophosphatidic acid (LPA) receptors () (sections III.B.4 and III.B.5). It also shares 29% identity with P2Y5/LPA6 (GenBank accession no. NM_005767.4), also shown to be a LPA receptor (), 27% identity with GPR35 (GenBank accession no. NM_005301.2), and 23% identity with the CCR4 chemokine receptor (GenBank accession no. NM_005508.4) ().

 

1. Reported Pharmacology of GPR55.

The current pharmacology of ligands at GPR55 is complicated and inconsistent. There are 11 reports containing data relating to ligand activity at GPR55. These reports use eight different cell backgrounds and six different assay endpoints that are all dependent on functional assays. So far, no binding data have been published. There are few examples of more than one laboratory repeating similar studies using equivalent cell background and assay technology. A range of assay strategies have been used to investigate the pharmacology of GPR55 and the mechanism of downstream signaling by this receptor remains uncertain. Using an approach that uses 12-amino acid peptides equivalent to the C-terminal sequences of the G proteins Gαi1/2, Gαi3, Gαs, and Gα13, as well as antibodies raised against those same peptides, it was demonstrated that the G protein preferentially coupling to GPR55 in [35S]GTPγS binding assays was Gα13 (). GPR55 has also been shown by other methods to use Gq, G12, or G13 for signal transduction, which results in downstream activation of RhoA and PLC (). This signaling mode is associated with temporal changes in cytoplasmic calcium, membrane-bound diacylglycerol, and plasma membrane topology. Involvement of the actin cytoskeleton has also been reported by . The reported activities of different ligands at GPR55 in various assays are summarized below and in Table 2.

TABLE 2

Reported activities of cannabinoid receptor ligands at recombinant GPR55 in various assays

The structures of all compounds listed are shown in Figures 115.

Ligand [35S]GTPγS Binding ERK1/2 Phosphorylation [Ca2+]i Mobilization β-Arrestin GPR55 Internalization RhoA Activation
LPI EC50 = 1 μMa EC50 = 200 nMa 30 nMa EC50 = 3.6 μMb 3 μMc 1 μMd,e
1 μMd,f 3 μMg,h EC50 = 1.2 μMc
3 μMc EC50 = 49 nMi
10 μMj
Anandamide EC50 = 18 nMk 10 μMj 5 μMg Very weak agonistb N.E.c 1 μMk
EC50 = 7.3 μMj N.E.c
2-AG EC50 = 3 nMk N.E.a,c N.E.h,i N.E.c N.E.c N.T.
Δ9-THC EC50 = 8 nMk N.E.a 5 μMg,h N.E.c N.E.c 5 μMg
Cannabidiol IC50 = 350 nMk 1 μM antagonistd N.E.g N.E.c N.E.c 1–10 μM antagonistd,k
Abn-CBD EC50 = 2.5 μMk N.E.a N.E.g N.E.b,c N.T. N.T.
O-1602 EC50 = 13 nMk 1 μMd 10 μMj N.E.c N.T. 1 μMd,e,f
EC50 = 2.5 nMl
CP55940 EC50 =5 nMk 10 μM antagonistc 10 μM antagonisti Ki ∼200 nMc Ki ∼200 nMc N.T.
Rimonabant N.T. N.E.a 1–2 μM antagonistg,h,j EC50 = 9.3 μMb 30 μMc N.T.
EC50 = 3.9 μMc
AM251 EC50 =39 nMk N.E.a N.T. 3 μMb 30 μMc N.T.
9.6 μMc
AM281 EC50 >30 μMk N.T. N.T. N.E.c N.E.c N.T.

N.E., no effect; N.T., not tested.

a: hGPR55 stably transfected in HEK293 cells with a tetracycline-inducible promoter.
b: hGPR55 transiently transfected in HEK293 cells.
c: hGPR55E stably transfected in HEK293 cells (β-arrestin assay) or U2OS cells (β-arrestin and GPR55 internalization assays).
d: human osteoclast primary cultures.
e: mouse osteoclast primary cultures.
f: BV-2 (mouse microglial cell line).
g: hGPR55 transiently transfected in HEK293 cells.
h: mouse DRG primary cultures.
i: hGPR55 stably transfected in HEK293 cells.
j: EA.hy926 (human umbilical vein derived endothelial cell line).
k: hGPR55 transiently transfected in HEK293s cells.
l: hGPR55 transiently transfected in HEK293T cells.

 

2. Anandamide.

The endocannabinoid ligand anandamide (section II.C.1) possesses significant affinity for both CB1 and CB2 receptors with slightly greater affinity for CB1 than for CB2 (Table 1). Using a [35S]GTPγS binding assay,  found that this ligand has an EC50 of 18 nM at GPR55 expressed in HEK293 cells and that it seems to have a higher potency for GPR55 than for either CB1 or CB2. Employing calcium mobilization assays, several other groups have demonstrated anandamide-induced GPR55 activation in HEK293 cells at a concentration of 5 μM and in EA.hy926 cells (EC50 = 7.3 μM), suggesting lower or similar potency to that reported for anandamide at CB1 and CB2 receptors ().  have also reported calcium oscillations in response to anandamide treatment in GPR55-expressing HEK293 cells. However, they could not demonstrate specificity for GPR55 because similar oscillations were seen in untransfected control cells. In contrast, three groups have reported that anandamide did not increase ERK1/2 phosphorylation via GPR55 in either HEK293 or U2OS cells (), whereas  did observe ERK1/2 phosphorylation at 10 μM in EA.hy926 cells. Several groups have also used β-arrestin and internalization assays to assess the properties of anandamide at GPR55. Although  reported weak agonist activity by anandamide,  found no evidence of anandamide dependent β-arrestin recruitment or of GPR55 receptor internalization. At 1 μM, anandamide was shown to activate RhoA in a GPR55-dependent manner in transfected HEK293 cells (), whereas  have demonstrated nuclear factor of activated T-cell activation using 10 μM anandamide in GPR55-expressing EA.hy926 cells.

 

3. 2-Arachidonoyl Glycerol.

The endocannabinoid ligand 2-AG (section II.C.1) binds to both CB1 and CB2 receptors with slightly greater affinity for CB1 than for CB2 (Table 1). 2-AG has been reported to be a 3 nM agonist of GPR55 in HEK293 cells using [35S]GTPγS binding as the assay (). However, in contrast to anandamide, no effect on calcium mobilization by 2-AG at 5 μM was seen in HEK293 cells ().  reported calcium oscillations in the presence of 3 to 30 μM 2-AG in GPR55-transfected HEK293 cells but could not demonstrate specific involvement of GPR55. 2-AG did not increase ERK1/2 phosphorylation in GPR55-expressing HEK293 or U2OS cells. In addition, 2-AG did not affect either β-arrestin recruitment or GPR55 receptor internalization ().

 

4. Lysophosphatidyl Inositol.

LPI has consistently been shown to be an agonist of GPR55. Thus, LPI (1 μM) can stimulate [35S]GTPγS binding to GPR55-expressing cell membranes () and has also been found to activate ERK1/2 in GPR55-transfected HEK293 cells with an EC50 of 200 nM as well as in GPR55-expressing U2OS cells and in human osteoclasts when tested at 10 and 1 μM, respectively (). Apparent GPR55-mediated calcium mobilization by LPI has been reported in EA.hy926 cells at 10 μM (), in mouse dorsal root ganglia at 3 μM (), and in HEK293 cells with EC50 values of 30 and 49 nM (). When using either β-arrestin recruitment or GPR55 receptor internalization assays, LPI was active as an apparent GPR55 agonist between 1 to 3 μM (). LPI has also been reported to induce GPR55-mediated activation of RhoA in transfected HEK293 cells (), as well as in human and mouse osteoclasts (). Evidence has also recently emerged that the endogenous compound, 2-arachidonoyl lysophosphatidyl inositol, activates GPR55 more potently than LPI in HEK293 cells and, hence, that this ligand may be the true intrinsic natural ligand for GPR55 ().

 

5. Δ9-Tetrahydrocannabinol.

The principal psychoactive component of the cannabis plant, Δ9-THC (section II.C.1), binds equally well to cannabinoid CB1 and CB2 receptors (Table 1). This cannabinoid has been reported to display significant potency as an agonist at GPR55 with an EC50 of 8 nM in a [35S]GTPγS binding assay performed with HEK293 cells (). Using 5 μM Δ9-THC,  reported a modest increase in intracellular calcium in both mouse and human GPR55-expressing HEK293 cells as well as in mouse dorsal root ganglia. In contrast, using GPR55 internalization and β-arrestin recruitment assays,  detected no sign of Δ9-THC-induced activation of GPR55, whereas  reported very weak GPR55-mediated β-arrestin recruitment in response to this ligand.

 

6. Abnormal-Cannabidiol.

Abnormal-cannabidiol (abn-CBD; section II.C.5 and Fig. 5) lacks significant affinity for CB1 and CB2 receptors but has been reported to have a number of in vivo effects through one or more as yet undefined receptors (section III.H.2). Both  and  have reported [35S]GTPγS binding data for this ligand at GPR55 expressed in HEK293 cells, albeit indicating 1000-fold different potencies (EC50 = 2.5 nM and 2.5 μM, respectively). No effect of this ligand on GPR55-mediated ERK1/2 activation was seen at a concentration of 1 μM () and no abn-CBD–induced GPR55-mediated mobilization of calcium was observed at 3 μM (). Furthermore, no GPR55-mediated activity of abn-CBD was seen when this ligand was tested in β-arrestin recruitment assays ().

 

7. Cannabidiol.

The phytocannabinoid cannabidiol (section II.C.5), which has therapeutic potential as an anti-inflammatory agent, displays relatively low affinity for CB1 and CB2 receptors (Table 1).  demonstrated that cannabidiol could antagonize the stimulation of [35S]GTPγS binding by anandamide, CP55940, and O-1602 in GPR55-transfected HEK293 cells with an IC50 of 350 nM. No confirmatory data from other laboratories are available using cannabidiol as an antagonist in transfected cells. However,  have reported that cannabidiol displays GPR55 antagonist activity in human osteoclasts at 1 μM using ERK1/2 phosphorylation and RhoA activation assays. Cannabidiol had no GPR55 agonist activity when assayed in calcium mobilization assays at 3 μM or when tested in β-arrestin recruitment assays (). Collectively, these data demonstrate that cannabidiol is an antagonist of GPR55.

 

8. O-1602.

O-1602 is an analog of abn-CBD in which the pentyl group has been replaced by a methyl group (section II.C.5 and Fig. 5). O-1602 has been reported to have activity at a non-CB1/CB2 receptor in the vasculature, the putative abnormal-cannabidiol receptor (section III.H.2). Using [35S]GTPγS binding assays, two independent groups have reported nanomolar activity of this compound at GPR55 expressed in HEK293 cells.  determined an EC50 of 13 nM, whereas  found an EC50 of 1.4 nM. O-1602 has been shown to promote apparent GPR55-mediated ERK1/2 phosphorylation in human osteoclasts as well as RhoA activation at 1 μM in HEK293 cells and in human and mouse osteoclasts.  reported no effect of O-1602 on calcium mobilization in GPR55-transfected HEK293 cells at 1 μM, whereas  did see a calcium signal in GPR55-expressing EA.hy926 cells in response to this compound at 10 μM. Collectively, these data support O-1602 as an agonist of GPR55, coupling signaling via G protein activation to RhoA. Using β-arrestin recruitment assays, however, neither  nor  observed any activity for O-1602 at GPR55.

 

9. CP55940.

The potent cannabinoid receptor agonist CP55940 (section II.C.1) has high affinity for both CB1 and CB2 receptors (Table 1) and is widely used in cannabinoid research as a pharmacological tool. Using a [35S]GTPγS binding assay, a potency (EC50 = 5 nM) similar to its potency as a CB1 and CB2 receptor agonist has been demonstrated for CP55940 at GPR55 (). Unfortunately, no confirmatory [35S]GTPγS binding data are available from other groups. However, this ligand has been investigated in other assays. Alone, CP55940 failed to demonstrate any agonist activity in either ERK1/2 activation () or calcium mobilization GPR55 assays (), although it did antagonize apparent GPR55-mediated ERK1/2 phosphorylation at a concentration of 10 μM ().

 

10. R-(+)-WIN55212.

R-(+)-WIN55212 (section II.C.1 and Table 1) is a potent, nonselective CB1 and CB2 receptor agonist that has been used in many studies of cannabinoid receptor function. The available data for R-(+)-WIN55212 activity at GPR55 are highly consistent, most laboratories finding no effect of this cannabinoid on any of the GPR55 assay end points used ().

 

11. Rimonabant.

Rimonabant (section II.C.3. and Table 1) is a potent CB1 receptor antagonist that was developed as an antiobesity agent. No [35S]GTPγS binding data on this compound at GPR55 have been published. However, signs of GPR55 antagonism have been detected at 1 μM in EA.hy926 cells and at 2 μM in HEK293 cells and mouse dorsal root ganglia using calcium mobilization assays (). In contrast,  reported agonist activity for rimonabant in the range of 100 nM to 3 μM, as indicated by elevations of intracellular calcium in GPR55-expressing HEK293 cells. GPR55 agonist activity by rimonabant is also reported for β-arrestin recruitment with EC50 values of 9.3 and 3.9 μM (). Likewise,  have reported receptor internalization at 30 μM rimonabant in GPR55-expressing HEK293 and U2OS cells, consistent with the findings of  that this compound can act as a GPR55 agonist. In addition,  demonstrated that rimonabant activates ERK1/2, cAMP response element-binding protein phosphorylation, and nuclear factor κ-light-chain-enhancer of activated B cells via GPR55 and also that it induces GPR55 internalization.

 

12. AM251.

Like rimonabant, AM251 (section II.C.3 and Table 1) is a potent CB1 receptor antagonist. It has been shown in a [35S]GTPγS binding assay using transfected HEK293 cells to be a high-potency agonist of GPR55 (EC50 = 39 nM) ().  have reported an EC50 of 612 nM for calcium mobilization in GPR55-expressing HEK293 cells. AM251 promotes β-arrestin recruitment with EC50 values of 3 and 9.6 μM () and GPR55 internalization ().

 

13. Other Ligands.

Using a [35S]GTPγS binding assay,  found that hGPR55 stably transfected into HEK293 cells was activated by nanomolar concentrations of the following lipids: noladin ether (section II.C.2 and Table 1) and virodhamine (section II.C.5 and Table 1), both of which are cannabinoid receptor agonists, and the non-CB1/CB2 receptor ligands, oleoyl ethanolamide and palmitoyl ethanolamide. Palmitoyl ethanolamide, which is of interest because of its potent anti-inflammatory, antiexcitotoxic, and antihyperalgesic properties (), also displays significant potency as a PPARα agonist (section III.G). It was originally thought to be an endogenous ligand for the CB2 receptor (). However, subsequent studies showed it to have little affinity for this receptor ().

 

14. Impact of Cell Lines and Expression Levels on GPR55 Data.

It is well known that cell lines present inconsistent phenotypes over time. For example,  have recently demonstrated that the androgen-insensitive PC-3 cell line exhibited two sublines that showed distinct receptor activation. The clonal background of HEK293 cells can differ markedly between laboratories. It is noteworthy, therefore, that GPR55 experiments were carried out by  with a HEK293s cell line and by  with a HEK293T cell line, and also, that these and other investigations into the ability of cannabinoid receptor ligands to target GPR55 () were each performed in a different laboratory. It is noteworthy, too, that although HEK293 cells are referred to as “human embryonic kidney” cells, a study on the origin of this cell line suggests that these cells may in fact have been derived by adenoviral transformation of a neuronal precursor present in the HEK cell cultures from which the original HEK293 cell line was obtained ().

Many in vitro studies of GPR55 have used transfected cells overexpressing this receptor (). If overexpression of the receptor induces constitutive activity, this can lead to altered ligand behavior (). Moreover, because of cell line and tissue heterogeneity, there may be accessory and other proteins in the various cell lines that modify the response of GPR55. The change in anandamide-induced CB1/GPR55 signaling that seems to occur in endothelial cells because of integrin clustering is one published example (). The manner in which GPR55 responds to its ligands may also be dependent on cell culture conditions. Moreover, HEK293 and other cells can synthesize lipid mediators, and this may alter the measured response (). The presence of endocannabinoids in serum has also been documented (), and other growth factors are present as well.

 

15. Conclusions.

Because of the large body of conflicting pharmacological data, no conclusive decision can yet be reached about whether GPR55 should be classified as a novel cannabinoid receptor. Particularly noteworthy are the mixed findings that have been obtained with the endocannabinoid anandamide. Thus, this compound has been found in GPR55 assays to stimulate [35S]GTPγS binding in the nanomolar range, to cause calcium mobilization in the micromolar range, but not to affect ERK1/2 phosphorylation or β-arrestin recruitment or to induce GPR55 internalization. These mixed findings may be the product of biased agonism at GPR55 or may have resulted simply from the use of different assay end points and cell systems. Therefore, although anandamide has been shown to be active at GPR55 in certain assays and cell types, the inconsistent manner with which it has been found to interact with this receptor prevents unequivocal designation of anandamide as a GPR55 ligand. Whether this inconsistency is the result of biased agonism or experimental variation remains to be determined; until it is, GPR55 cannot be considered an anandamide receptor.

The data for 2-AG are more consistent, albeit mainly negative. Thus, although it has been shown to display activity as a GPR55 agonist in a [35S]GTPγS binding assay, the majority of studies using calcium mobilization, ERK1/2 phosphorylation, or β-arrestin recruitment and receptor internalization have failed to demonstrate any effect of this endocannabinoid on GPR55. Consequently, there is no conclusive evidence at this time that this endocannabinoid is a ligand of GPR55. As for Δ9-THC, although it displays activity as a GPR55 agonist in [35S]GTPγS binding, calcium mobilization, and RhoA activation assays, it fails to stimulate ERK1/2 phosphorylation, β-arrestin recruitment, or GPR55 internalization. Whether this is a result of biased agonism by Δ9-THC or experimental variability remains to be determined.

A rare consensus among the articles published on GPR55 is that LPI is an agonist for this receptor. Another agreement among published reports is that the aminoalkylindole R-(+)-WIN55212, a potent CB1 and CB2 receptor agonist, does not target GPR55 as either an agonist or an antagonist. In contrast, both CP55940 and rimonabant have been found to behave as GPR55 agonists in some investigations but as GPR55 antagonists in others, possibly an indication that they possess low relative intrinsic activity as GPR55 agonists, although this remains to be established. Finally, the finding that 2-arachidonoyl lysophosphatidyl inositol is an endogenous agonist for GPR55 (section III.A.4) has revealed an interesting “parallel” between the chemical nature of GPR55 and CB1/CB2 receptor endogenous ligands.

B. Other Deorphanized G Protein-Coupled Receptors

 

1. GPR40, GPR41, GPR42, and GPR43.

In 2003, it was discovered by three different research groups that the receptor GPR40 can be activated by long- and medium-chain fatty acids (C6-C22; Table 3) (). The receptor has been renamed FFA1 because it is now thought to be a fatty acid receptor that is involved in the regulation of insulin release (). That GPR40 is indeed a fatty acid receptor has been confirmed in a recent investigation that used a new β-arrestin assay to deorphanize G protein-coupled receptors (). FFA1 can be activated by glitazone drugs that are activators of PPARγ (), and a number of other small-molecule agonists/antagonists for FFA1 have also been discovered (). However, FFA1 has not been reported to be activated or inhibited by any known cannabinoid CB1 or CB2 receptor agonist or antagonist.

TABLE 3

Deorphanized G protein-coupled receptors other than GPR55 that could possibly be activated by cannabinoid receptor ligands

In all cases, it was not reported whether the receptor was targeted by cannabinoid receptor ligands. See Section III.B for references and further details.

Receptor Recognized Agonist(s) EC50 Range
FFA1 (GPR40 Fatty acids (C6-C22) Micromolar
FFA2 (GPR43) Fatty acids (C1-C6) Micromolar to millimolar
FFA3 (GPR41) Fatty acids (C1-C6) Micromolar to millimolar
GPR84 Fatty acids (C9-C14) Micromolar
GPR120 Fatty acids (C14-C22) Micromolar
GPR3 Sphingosine-1-phosphate?
GPR6 Sphingosine-1-phosphate?
GPR12 Sphingosylphosphorylcholine?
GPR18 LPA Nanomolar to micromolar
Farnesyl pyrophosphate Nanomolar to micromolar
NAGly Nanomolar to micromolar
GPR23 LPA?
GPR119 OEA/OLDA Low micromolar

OEA, oleoyl ethanolamide; OLDA, N-oleoyl dopamine.

In the same year, two other orphan receptors, GPR41 and GPR43, were identified by three independent research groups as receptors for short-chain fatty acids (C1-C6; Table 3) () and these receptors have now been renamed FFA3 and FFA2, respectively (). FFA2 is found in adipose tissue, where its activation may increase leptin production and, in white blood cells, where it may stimulate chemotaxis. FFA3 is found on immune cells, in the gastrointestinal tract, and in adipose tissue. FFA2 knockout mice do not respond to acetate-induced reductions in plasma free fatty acid levels, indicating a role for this receptor in the stimulation of lipolysis. A series of small molecule phenylacetamides have been found to be more potent FFA2 agonists, and these may also act as allosteric ligands at an allosteric binding site on FFA2 (). FFA2 and FFA3 have not been reported to be activated or inhibited by any known agonists or antagonists for cannabinoid CB1 or CB2 receptors.

The GPR42 gene codes for a GPR42 receptor that is very similar to FFA3 but cannot be activated by short-chain fatty acids, prompting the suggestion that it is a pseudogene (). However, recent findings suggest that GPR42 could potentially be a functional gene in a fraction of the human population because of a polymorphism resulting in the presence of arginine at amino acid 174 of the receptor (). If that is the case, it is likely that an active GPR42 will have nearly the same properties as the FFA3 receptor.

 

2. GPR84 and GPR120.

The two orphan receptors, GPR84 and GPR120, seem to be receptors for medium-chain fatty acids (C9-C14; Table 3) () and for long-chain fatty acids (C14-C22) (), respectively.

GPR84 is highly expressed in bone marrow, and in splenic T cells and B cells, and results from studies with GPR84 knockout mice suggest that GPR84 is involved in regulating early IL-4 gene expression in activated T cells () and that it is expressed in activated microglial cells and macrophages (). Medium-chain fatty acids activate GPR84 as can be seen from their ability to decrease intracellular cAMP and to stimulate [35S]GTPγS binding to membranes from CHO cells stably expressing GPR84 (). Short- and long-chain fatty acids were inactive, and GPR84 has not been reported to be activated or inhibited by any known agonists or antagonists for cannabinoid CB1 or CB2 receptors.

GPR120 is found mainly in the intestinal tract, although it is also expressed by a number of other tissues (e.g., adipocytes, taste buds, and lung) (). More specifically, intestinal GPR120 is found in glucagon-like peptide-1 (GLP-1)-expressing endocrine cells in the large intestine () and in gastric inhibitory polypeptide-expressing K cells of the duodenum and jejunum (). Dietary fatty acids may promote intestinal CCK and GLP-1 release via activation of intestinal GPR120 (). Long-chain fatty acids (C14–C22; Table 3), especially unsaturated ones, activate GPR120 in cell lines stably expressing this receptor, as measured by an increase in intracellular Ca2+, whereas α-linolenic acid methyl ester lacks activity (). It is noteworthy that some plant-derived compounds, grifolin derivatives that do not contain a carboxylic group, can also activate GPR120 (). GPR120 has not been reported to be activated or inhibited by any known agonists or antagonists for cannabinoid CB1 or CB2 receptors.

 

3. GPR3, GPR6, and GPR12.

GPR3, GPR6, and GPR12 are constitutively active proteins that signal through Gαs to increase cAMP levels in cells expressing these receptors (). They are mainly expressed in the central nervous system, where they may contribute to the regulation of neuronal proliferation (), monoamine neurotransmission (), reward learning processes (), and energy expenditure (). They may also be involved in the regulation of meiosis in oocytes (). Their closest phylogenetic GPCR relatives are cannabinoid receptors, lysophospholipid receptors, and melanocortin receptors ().

It has been suggested that GPR3, GPR6, and GPR12 (Table 3) are all activated by sphingosine-1-phosphate and/or sphingosylphosphorylcholine at nanomolar concentrations (,). Results obtained in a recent investigation using β-arrestin recruitment instead of G protein activation as an assay for receptor agonism do, however, challenge this hypothesis as no sign of agonism was seen in response to sphingosine-1-phosphate or sphingosylphosphorylcholine at 8 and 42 μM, respectively (). A large number of endogenous lipids, including endocannabinoids, were screened in this investigation and none of these were found to activate GPR3, GPR6, or GPR12 (). However, anandamide and 2-AG did show weak agonist activity at the S1P1 receptor (edg1) at concentrations in the micromolar range ().

 

4. GPR18 and GPR92.

The chromosomal location of GPR18 has been determined as 13q32.3; it is a 331-amino acid GPCR. GPR18 (Table 3) is highly expressed in spleen, thymus, and peripheral lymphocyte subsets (). In GPR18-transfected cells, N-arachidonoyl glycine (NAGly) has been shown to induce intracellular Ca2+ mobilization at 10 μM (). Furthermore, the same study demonstrated inhibition of forskolin-stimulated cAMP production by NAGly with an EC50 of 20 nM; the effect was absent in untransfected cells and was pertussis toxin-sensitive, suggesting Gi-coupling. A more recent study used the β-arrestin PathHunter assay system to examine the pharmacological interactions of various lipids with a range of recently deorphanized GPCRs (). In this study, NAGly did not activate GPR18 but elicited a weak activation of GPR92, at concentrations above 10 μM. GPR92 mRNA is highly expressed in dorsal root ganglia, suggesting a role in sensory neuron transmission.  have also demonstrated that NAGly mobilizes intracellular Ca2+ and activates [35S]GTPγS binding in GPR92-expressing cells. However, the relative intrinsic activity of NAGly is significantly lower than that of the other putative endogenous GPR92 agonists, LPA and farnesyl pyrophosphate (). In addition, farnesyl pyrophosphate and LPA activate both Gq/11– and Gs-mediated signaling, whereas NAGly activates only Gq/11-mediated signaling. To date, there are no published data to indicate whether cannabinoid CB1 or CB2 receptor ligands can activate or block GPR18 or GPR92. It is noteworthy, however, that GPR18 may be a receptor for abnormal-cannabidiol (section III.H.2).

 

5. GPR23.

The orphan receptor GPR23/p2y9 is closely related to the purinergic P2Y receptor and mRNA for this receptor in the mouse is mainly found in ovary, uterus, and placenta (). It has been found (Table 3) that GPR23 is activated by LPA in the nanomolar range as indicated by intracellular calcium mobilization and cAMP formation (), probably through the activation of Gq/11 and Gs proteins (). Two other recent investigations have used β-arrestin recruitment to test the ability of LPA to activate GPR23. In one of these, activation was detected at 100 μM (), whereas in the other, no activation was induced by concentrations of up to 100 μM LPA (). GPR23 has not been reported to be activated or inhibited by any known agonists or antagonists for cannabinoid CB1 or CB2 receptors.

 

6. GPR119.

The human orphan receptor GPR119, identified by a basic local alignment search tool search of the genomic database, is an intronless GPCR belonging to the MECA (melanocortin, endothelial differentiation gene, cannabinoid, adenosine) cluster of receptors (). It is preferentially expressed in pancreatic and intestinal cells, where it is involved in the control of glucose-dependent insulin release and GLP-1 release, respectively (). Although GPR119 is phylogenetically related to cannabinoid receptors, only fatty acid amides interact with GPR119 (Table 3), the potency order of four of these being N-oleoyl dopamine > oleoyl ethanolamide > palmitoyl ethanolamide > anandamide (). Because only N-oleoyl dopamine and oleoyl ethanolamide have reasonably high (low micromolar) affinity for GPR119, and because neither of these lipids interacts with CB1 or CB2 receptors, GPR119 cannot be viewed as a cannabinoid receptor. Oleoyl ethanolamide also activates TRPV1 channels and PPARs (sections III.E and III.G).

 

7. Conclusions.

There is evidence that at least some cannabinoid receptor agonists do not activate GPR119, GPR3, GPR6, or GPR12 with significant potency (Table 3 and sections III.B3 and III.B.6). However, to our knowledge, cannabinoids have not been tested as ligands for most of the receptors mentioned in Table 3 or, in any case, no such data have been published. Clearly, therefore, there is a need for the receptors listed in Table 3 to be tested for their responsiveness to a broad spectrum of potential ligands, including a carefully selected range of cannabinoids, to help clarify their pharmacological profiles and physiological roles and so provide a conclusive deorphanization of these receptors.

C. Established G Protein-Coupled Receptors

At concentrations in the low micromolar range, some cannabinoid receptor agonists (Table 4) or antagonists seem to target certain G protein-coupled “noncannabinoid” receptors, in some instances probably by targeting allosteric sites on these receptors. These G protein-coupled receptors include muscarinic acetylcholine receptors and α2– and β-adrenoceptors and also opioid, adenosine, 5-HT, angiotensin, prostanoid, dopamine, melatonin, and tachykinin receptors.

TABLE 4

Effects of cannabinoid CB1 and/or CB2 receptor agonists on noncannabinoid established G protein-coupled receptors

See Section III.C for references and further details.

Receptor and Effect Endocannabinoid(s)? Effective Concentration Range Nonendogenous Cannabinoid Agonist(s)? Effective Concentration Range
Radioligand binding (↓)
    Opioid (μ- and δ-) N.D. N.D. Yesa Micromolar
    Acetylcholine (muscarinic) Yesa,b Micromolar Yesa,c Micromolar
    Adenosine A3 Yesa Micromolar No Micromolar
    5-HT1 or 5-HT2C Yesd Micromolar Yes Micromolar
Radioligand binding (↑)
    Acetylcholine (muscarinic) Yesa,b Micromolar Yesa,c Micromolar
    Adrenoceptors (β-) N.D. N.D. Yesa Micromolar
    5-HT2 No Micromolar Yes Micromolar

N.D., no data.

aMay target an allosteric site on this receptor.
bAlso R-(+)-methanandamide.
cOnly R-(+)-methanandamide.
dAlso 5-HT2A and 5-HT2B.

 

1. Opioid Receptors.

There is evidence that certain phytocannabinoids or synthetic cannabinoids can act allosterically at concentrations in the low micromolar range to accelerate the dissociation of ligands from the orthosteric sites on μ- and/or δ-opioid receptors. Thus, it has been found that the rate of dissociation of [3H]DAMGO ([3H][d-Ala2,N-Me-Phe4,Gly5-ol]-enkephalin), presumably from μ-opioid receptors, and of [3H] naltrindol, presumably from δ-opioid receptors, can be increased by Δ9-THC (EC50 = 21.4 and 10 μM, respectively) (). In contrast, rimonabant seems to displace [3H]DAMGO in a competitive manner (IC50 = 4.1 μM). Results obtained from equilibrium binding experiments with rat whole-brain membranes also suggest that Δ9-THC is a noncompetitive inhibitor of ligand binding to μ- and δ-opioid receptors (IC50 = 7 and 16 μM, respectively), although not to κ-opioid receptors or σ/phencyclidine receptors, and that inhibition of ligand binding to μ-opioid receptors can be induced by certain other cannabinoids (). In addition, it has been found by both  and  that rimonabant can induce radioligand displacement from μ-opioid receptors (IC50 = 3 and 5.7 μM, respectively), and by  that this CB1 receptor antagonist can induce radioligand displacement from κ-opioid receptors (IC50 = 3.9 μM).

 

2. Muscarinic Acetylcholine Receptors.

At concentrations in the micromolar range, both anandamide and R-(+)-methanandamide have been shown to modulate tritiated ligand binding to muscarinic acetylcholine receptors, probably by targeting allosteric sites on these receptors. Thus,  found that [3H]N-methylscopolamine and [3H]quinuclidinyl benzilate could be displaced from binding sites on adult human frontal cerebrocortical membranes in a noncompetitive manner by both anandamide (IC50 = 44 and 50 μM, respectively) and R-(+)-methanandamide (IC50 = 15 and 34 μM, respectively) but not R-(+)-WIN55212 (up to 5 μM). Both ethanolamides stimulated [3H]oxotremorine binding to these membranes at concentrations below 50 or 100 μM, although they did inhibit such binding at higher concentrations. It was concluded that these effects of anandamide did not require its conversion to arachidonic acid. It has also been found that anandamide and R-(+)-methanandamide but not R-(+)-WIN55212 can displace tritiated ligands from human M1 and M4 muscarinic acetylcholine receptors transfected into CHO cells (). IC50 values for the displacement of [3H]N-methylscopolamine and [3H]quinuclidinyl benzilate from M1 receptors were 2.8 and 13.5 μM, respectively, for anandamide and 1.45 and 6.5 μM, respectively, for R-(+)-methanandamide. Corresponding IC50 values for the displacement of these tritiated ligands from M4 receptors were 8.3 and 6.9 μM for anandamide and 9.8 and 3.1 μM for R-(+)-methanandamide. The effect of anandamide on tritiated ligand binding seemed to be noncompetitive and hence possibly allosteric in nature. It is noteworthy, however, that anandamide (10 μM) was subsequently found by the same research group not to affect the rate of dissociation of [3H]N-methylscopolamine from M1 binding sites ().

 

3. Other Established G Protein-Coupled Receptors.

It has been reported that at 10 μM, both rimonabant and AM251 can oppose the activation of adenosine A1 receptors in rat cerebellar membranes and that these CB1 receptor antagonists can inhibit basal [35S]GTPγS binding to these membranes, probably by blocking A1 receptor activation by endogenously released adenosine (). In addition, evidence has been obtained, first, that at 10 μM, both anandamide and 2-AG, but not AM251, R-(+)-WIN55212, or CP55940 can act as allosteric inhibitors at the human adenosine A3 receptor although not at the human adenosine A1 receptor (), and second, that both rimonabant (IC50 = 1.5 μM) and taranabant (IC50 = 3.4 μM) can induce radiolabeled ligand displacement from adenosine A3 receptors (). In contrast to taranabant (IC50 > 10 μM), rimonabant can also displace radiolabeled ligands from α2A– and α2C-adrenoceptors, from 5-HT6 and angiotensin AT1 receptors, and from prostanoid EP4, FP, and IP receptors with IC50 values ranging from 2 to 7.2 μM (). Taranabant, however, has been found to displace radiolabeled ligands from dopamine D1 and D3 receptors (Ki = 3.4 and 1.9 μM, respectively) and from melatonin MT1 receptors (Ki = 5.4 μM) (). In addition, both rimonabant (IC50 = 2 μM) and taranabant (IC50 = 0.5 μM) can induce radiolabeled ligand displacement from tachykinin NK2 receptors ().

There is also evidence that at 3 or 10 μM, but not higher or lower concentrations, both Δ9-THC and 11-hydroxy-Δ9-THC increase the affinity of [3H]dihydroalprenolol for β-adrenoceptors in mouse cerebral cortical membranes (). There is evidence too that [3H]5-HT binding to 5-HT1A, 5-HT1B, 5-HT1D, 5-HT1E, and/or 5-HT2C receptors in bovine cerebral cortical synaptic membranes can be reduced by 11-hydroxy-Δ8-THC and 11-oxo-Δ8-THC although not Δ8-THC at 10 μM, and by anandamide at 1 and 10 μM (). The same concentrations of these cannabinoids did not decrease [3H]ketanserin binding to 5-HT2A or 5-HT2B receptors, although evidence was obtained that anandamide can reduce radioligand binding to 5-HT1 and 5-HT2 receptors at 100 μM. In addition, there has been a report that [3H]ketanserin binding to 5-HT2 receptors in rat cerebral cortical membranes is enhanced by HU-210 at 500 nM ().

 

4. Conclusions.

There is evidence that at concentrations in the nanomolar or micromolar range, either or both of two CB1 receptor antagonists/inverse agonists, rimonabant and taranabant, can bind to some types of opioid, adrenergic, dopamine, 5-HT, adenosine, angiotensin, melatonin, tachykinin, and prostanoid receptors. There is also evidence that anandamide, 2-AG, and/or certain established nonendogenous CB1/CB2 receptor agonists can interact with at least some types of opioid, muscarinic acetylcholine, adrenergic, 5-HT, and adenosine receptors (Table 4). However, the potency with which these ligands target these receptors is significantly less than the potency with which they activate or block CB1 and/or CB2 receptors. Moreover, at least some of these interactions seem to be allosteric in nature. Consequently, no convincing case can be made for reclassifying any of the receptors mentioned in this section as a novel cannabinoid receptor.

D. Ligand-Gated Ion Channels

Several cannabinoid receptor agonists have been found to antagonize or enhance the activation of 5-HT3, nicotinic acetylcholine, glycine, and/or ionotropic glutamate (NMDA) receptors (Table 5).

TABLE 5

Effects of cannabinoid CB1 and/or CB2 receptor agonists on ligand-gated ion channels

See Section III.D for references and further details.

Gated Channel and Effect Endocannabinoid(s)? Effective Concentration Range Nonendogenous Cannabinoid Agonist(s)? Effective Concentration Range
Enhancement of activation
    Glycine (native) Yes Nanomolar Yes Nanomolar
    Glycine α1, α1β1 Yes Nanomolar Yes Nanomolar
    Glycine α1β Yes Nanomolar Yes Micromolar
    NMDA Yesa Nanomolar or micromolar Yesb Micromolar
Inhibition of activation
    5-HT3 Yesc Nanomolar Yesc Nanomolar
    Acetylcholine (nicotinic) Yesc Nanomolar Yesc Nanomolar or micromolar
    Glycine (native) Yes Nanomolar or micromolar No Nanomolar
    Glycine α1, α1β No Nanomolar Yes Micromolar
    Glycine α2, α3 No Micromolar Yes Nanomolar or micromolar
Radioligand binding (↓)
    Benzodiazepine No Micromolar Yes Micromolar
aAlso R-(+)-methanandamide.
bOnly R-(+)-methanandamide.
cMay target an allosteric site on this receptor.

 

1. 5-HT3 Receptors.

There has been one report that antagonism of the activation of 5-HT3 receptors by 5-HT can be induced by CP55940, R-(+)-WIN55212, and anandamide with IC50 values of 94, 310, and 190 nM, respectively (), and another report that these compounds induce antagonism of this kind with IC50 values of 648, 104, and 130 nM, respectively (). The second of these research groups also found 5-HT3 receptor activation to be potently antagonized by two other cannabinoid receptor agonists, Δ9-THC and JWH-015 (IC50 = 38 and 147 nM, respectively) and by the CB1 receptor antagonist/inverse agonist LY320135 (IC50 = 523 nM), although not by rimonabant at 1 μM. In other investigations, anandamide has been found to antagonize 5-HT3 receptor activation with IC50 values of 239 nM () or 3.7 μM (). There is evidence, at least for anandamide, CP55940, and R-(+)-WIN55212, that this antagonism is noncompetitive in nature and, indeed, that at least some of these cannabinoids may be targeting an allosteric site on the 5-HT3 receptor ().

 

2. Nicotinic Acetylcholine Receptors.

An allosteric mechanism may also underlie the antagonism of nicotinic acetylcholine receptors that is reportedly induced by anandamide, 2-AG, R-(+)-methanandamide, and CP55940 (IC50 = 230, 168, 183, and 3400 nM, respectively), although not by Δ9-THC or R-(+)-WIN55212 (). Evidence that anandamide can antagonize nicotinic acetylcholine receptors with significant potency has also been obtained by  (IC50 ∼ 300 nM) and  (IC50 = 900 nM). Anandamide behaved as a noncompetitive antagonist in the second of these investigations, in which further evidence that Δ9-THC (30 μM) is not a nicotinic acetylcholine receptor antagonist was also obtained. Arachidonic acid, a metabolite of anandamide, has also been found to antagonize nicotinic acetylcholine receptors, although the ability of anandamide to inhibit this ligand-gated ion channel does not seem to depend on its conversion to this unsaturated fatty acid ().

 

3. Glycine Receptors.

At least some cannabinoid receptor agonists have been found to modulate glycine-induced activation of subunits of glycine receptors with significant potency in a positive or negative manner. Thus, for example, results obtained from experiments with transfected cells () suggest the following:

  • Activation of human α1 subunits is enhanced by anandamide (IC50 = 38 or 319 nM), Δ9-THC (IC50 = 86 nM), and HU-210 (IC50 = 270 nM) but weakly inhibited by HU-308 and unaffected by R-(+)-WIN55212 at 30 μM.

  • Activation of neither human α2 nor rat α3 subunits is affected by anandamide at up to 30 μM, whereas activation of both is inhibited by HU-210 (IC50 = 90 and 50 nM, respectively), R-(+)-WIN55212 (IC50 = 220 and 86 nM, respectively), and HU-308 (IC50 = 1130 and 97 nM, respectively).

  • Activation of human α1β1 dimers is enhanced by anandamide and Δ9-THC (IC50 = 318 and 73 nM, respectively).

  • Activation of human α1β subunits is enhanced by anandamide (IC50 = 75 nM) and HU-210 (30 μM) but unaffected by R-(+)-WIN55212 and inhibited by HU-308 at 30 μM.

 

Δ9-THC (IC50 = 115 nM) and anandamide (IC50 = 230 nM) have also been found to enhance the activation of native glycine receptors in rat isolated ventral tegmental area neurons (). There is evidence too, from experiments with rat isolated hippocampal pyramidal neurons, that at 0.2 to 2 μM, anandamide and 2-AG can antagonize glycine receptor activation, and that this effect of anandamide is not TRPV1 channel-mediated ().

 

4. Other Ligand-Gated Ion Channels.

Evidence has also been obtained that anandamide (≥100 nM) and R-(+)-methanandamide (1 μM) but not Δ9-THC (≤10 μM) can enhance NMDA-induced activation of NMDA receptors (). In contrast, anandamide (1–100 μM) has been found not to affect binding of [3H]muscimol to GABAA receptors in bovine cerebral cortical synaptic membranes or, indeed, binding of the benzodiazepine [3H]flunitrazepam to these membranes (). Binding of [3H]flunitrazepam to synaptic membranes, however, has been found to be decreased by 11-hydroxy-Δ8-THC at 10 μM ().

 

5. Conclusions.

Some established endogenous and nonendogenous CB1/CB2 receptor agonists seem to block 5-HT3 and nicotinic acetylcholine receptors or enhance the activation of glycine and NMDA receptors (Table 5), the relative potencies displayed by these cannabinoids as blockers or enhancers of ligand-gated ion channel activation differing from those they display as CB1 or CB2 receptor agonists. One CB1 receptor antagonist/inverse agonist, LY320135, has also been found to block 5-HT3 receptors. Many of these effects on ligand-gated ion channel activation are induced by cannabinoid concentrations in the low to mid-nanomolar range and hence with significant potency. Even so, no convincing case can be made for reclassifying glycine, NMDA, 5-HT3, or nicotinic acetylcholine receptors as a novel cannabinoid receptor because 1) there is no evidence that glycine or NMDA receptors can be directly activated by any cannabinoid and 2) 5-HT3 and nicotinic acetylcholine receptors are blocked rather than activated by cannabinoids; this blockade seems to be noncompetitive/allosteric in nature. There is also some evidence that benzodiazepine receptors do not behave as cannabinoid receptors. Thus, there have been reports that benzodiazepine receptors are not targeted by anandamide and that although these receptors can be targeted by 11-hydroxy-Δ8-THC, this occurs at the rather high ligand concentration of 10 μM.

E. TRPV1 and other Transient Receptor Potential Channels

 

1. Transient Receptor Potential Channels: A Brief Introduction.

The transient receptor potential (TRP) superfamily of cation channels includes six subfamilies: “canonical,” “vanilloid” (TRPV), “melastatin” (TRPM) “polycystin,” “mucolipin,” and “ankyrin” (TRPA). TRP channels are six-transmembrane (TM) domain integral membrane proteins with cytosolic C- and N-terminal domains and a nonselective cation-permeable pore region between TMs 5 and 6 (). The various subfamilies differ particularly in the number of ankyrin repeats present in the N termini, which is null in TRPM and very high in TRPA channels. More than 50 members of the TRP family have been characterized in yeast, worms, insects, and fish and 28 in mammals so far (). They are involved in the transduction of a remarkable range of stimuli, including temperature, mechanical and osmotic stimuli, electrical charge, light, olfactive and taste stimuli, hypotonic cell swelling, and effects of xenobiotic substances and endogenous lipids (). It is noteworthy that mutations in different TRPs have been linked to human diseases, and their expression in tissues affected by pathological conditions is often increased (). To date, five types of TRP channels belonging to three subfamilies have been suggested to interact with phytocannabinoids, synthetic CB1 and CB2 receptor ligands, or endocannabinoids: TRPV1, TRPV2, TRPV4, TRPM8, and TRPA1. Data on these interactions are increasing and, importantly, have prompted two different research groups to propose that at least some TRP channels could be “ionotropic cannabinoid receptors” ().

 

2. TRPV1 Channels.

TRPV1 (initially known as VR1) was the first TRP channel to be cloned as a receptor for capsaicin, the natural product responsible for the pungency of hot chilli peppers (). It is also activated by noxious stimuli, such as heat (>43°C), protons (pH <6.9), and various other natural toxins. Consistent with the hypothesis that it is involved in pain, nociception, and temperature sensing (), TRPV1 is predominantly expressed in sensory neurons of unmyelinated axons (C fibers) and thin myelinated axons (Aδ fibers) of dorsal root and trigeminal ganglia (), where it can become up-regulated during nerve injury-induced thermal hyperalgesia () and diabetic neuropathy (). Evidence has accumulated for the presence of TRPV1 not only in sensory neurons but also in brain neurons and in non-neuronal cells, including epithelial, endothelial, glial, smooth muscle, mast and dendritic cells, lymphocytes, keratinocytes, osteoclasts, hepatocytes, myotubes, fibroblasts, and pancreatic β-cells (for review, see ). It is noteworthy that TRPV1 colocalizes with CB1 receptors in sensory (e.g., dorsal root ganglia and spinal cord) () and brain () neurons and with CB2 receptors in sensory neurons () and osteoclasts (). This colocalization makes possible several types of intracellular cross-talk (for review, see ) and might have important functional consequences for those endogenous and synthetic ligands that activate both cannabinoid and TRPV1 receptors ().

In fact, it is now well established from experimental work described in more than 300 articles that endocannabinoids, such as anandamide and N-arachidonoyl dopamine but not 2-AG, noladin ether, or virodhamine, bind to both human and rat TRPV1, upon which they act as full agonists (; for review, see ). Although the affinity of these compounds in binding assays with the human recombinant channel is similar to, or only slightly lower than, that of capsaicin (), their relative intrinsic activity and particularly their potency (Table 6) depends on the type of TRPV1 functional assay in which these pharmacological properties are assessed (i.e., enhancement of intracellular Ca2+ levels, induction of cation currents in neurons, or release of algogenic/vasodilatory peptides from sensory nervous tissue preparations) and on the experimental conditions used to carry out these assays. Furthermore, TRPV1 gating by its ligands can be markedly altered by several regulatory events and signals, including post-translational modifications (such as phosphorylation by several protein kinases), allosteric modulation by temperature, acid, membrane potential, membrane phospholipids (phosphatidylinositol bisphosphate in particular), metabotropic (including cannabinoid) receptor activation, neurotrophins, etc., all of which can be modulated by inflammatory and other pathological conditions. As a result, apparent anandamide-induced activation of TRPV1 has often been found to increase or decrease under such conditions. It has been proposed, for example, that endogenously released anandamide has higher efficacy and potency (). Although anandamide may be less potent as a TRPV1 agonist than as a CB1 receptor agonist, elevation of its endogenous levels with fatty acid amide hydrolase (FAAH) inhibitors [e.g., cyclohexylcarbamic acid 3′-carbamoyl-biphenyl-3-yl ester (URB597)] can lead to effects that are mediated by TRPV1 (). This, together with data indicating that exogenous anandamide can exert effects at TRPV1 in a way sensitive to inhibitors of anandamide cellular uptake in HEK293 cells (), trigeminal neurons (), and, ex vivo, in mesenteric arteries () represents strong indirect evidence that both endogenous and exogenous anandamide reaches its intracellular binding site on this molecular target.

TABLE 6

Cannabinoids, endogenous lipids, and synthetic CB1 and/or CB2 receptor ligands that interact with TRP receptors

Compound TRPV1


TRPV2 TRPM8 TRPA1
Calcium Assay or Mesenteric Artery Tension Currents by Patch Clamp
Anandamide EC50 = 1.15 μMa; EC50 = 0.63 μMb; EC50 = 0.16 μMc (rat TRPV1) EC50 = 4.9 μMd; EC50 = 2.5 μMe N.E. at 100 μMf IC50 = 0.15 μM (HEK293 cells) or 10 μM (DRG neurons) (vs. icilin); IC50 = 3.1 μM (vs. menthol in HEK293 cells)g,h EC50 = 4.9 μMi
2-AG N.E. at 10 μMb; EC50 = 8.4 μMf N.T. N.E. at 100 μMf N.E. N.T.
NADA EC50 = 26 nMb EC50 = 1.6 μM (DRG); 794 nM (TG)j N.T. IC50 = 0.7 μM (vs. icilin); IC50 = 2.0 μM (vs. menthol)g N.E. at 10 μMk
OEA EC50 = 3.7 μMl; EC50 = 0.4 μMc EC50 = 6.3 μMc N.T. N.T. N.T.
Δ9-THC N.E. at 100 μMf N.T. EC50 = 15.5 μMf IC50 = 0.16 μM (vs. icilin); IC50 = 0.15 μM (vs. menthol)h EC50 = 12 ± 2 μMm; EC50 = 32.3 μMf; EC50 = 0.23 μMh in HEK293 cells
HU-210 EC50 = 1.2 μM but very low maximal effectb; no effect at 100 μMf N.T. Slight agonism at 100 μMf N.T. N.E. at 100 μMf
CP55940 N.E. at 100 μMf N.T. Slight effect at 100 μMf N.T. Slight effect at 100 μMf
R-(+)-WIN55212 Slight effect at 100 μMf N.T. N.E. at 100 μMf N.T. EC50 = 9.8 μMf
Rimonabant Antagonist 1–5 μMn N.T. N.T. IC50 = 52.3 nM vs. icilin; no effect vs mentholg N.T.
JWH-015 EC50 = 41.2 μMf N.T. N.E. at 100 μMf N.T. Slight effect at 100 μMf
Cannabidiol EC50 = 3.5 μM (human)o; very little effect (rat)f N.T. EC50 = 3.7 μMf IC50 = 80 nM (vs. icilin); IC50 = 0.14 μM (vs. menthol)h >10 μMm; EC50 = 96 nMh in HEK293 cells; EC50 = 81.4 μMf
Cannabinol N.E. at 100 μMf N.T. EC50 = 77.7 μMf IC50 = 0.21 μM (vs. icilin)k >10 μMm; very little effect at 100 μMf; EC50 = 0.18 μM in HEK293 cellsk
Cannabichromene EC50 >50 μMp N.T. N.T. 40 μM (vs. icilin)k EC50 = 60 nMh
Cannabigerol EC50 = 11.5 μMp N.T. N.T. IC50 = 0.14 μM (vs. icilin); IC50 = 0.16 μM (vs. menthol)h EC50 = 3.4 μMh

N.E., no effect; N.T., not tested; NADA, N-arachidonoyl dopamine; OEA, oleoyl ethanolamide; TG, trigeminal ganglia.

kL. De Petrocellis and V. Di Marzo, unpublished observations.

It is noteworthy that anandamide and N-arachidonoyl dopamine seem to interact with TRPV1 at the same intracellular binding site as capsaicin.  observed that the TM3/4 region of the mammalian TRPV1 receptor is responsible for its sensitivity to capsaicin and anandamide, whereas the avian receptor, which is activated only by heat or protons, lacks part of this region. The binding site is located on the inner face of the plasma membrane, thus opening up the possibility that when anandamide is biosynthesized by cells that express TRPV1, it will activate this receptor before being released, thereby regulating Ca2+ homeostasis as an intracellular messenger (). A crucial role for Tyr511 and Ser512, located at the transition between the second intracellular loop and TM3, has been demonstrated, and other critical residues seem to be in the TM4 segment (). Tyr511 engages in a hydrophobic interaction with the hydrophobic tails of capsaicin, anandamide, and N-arachidonoyl dopamine, whereas the side-chain hydroxyl group of Thr550, as well as the aromatic region of Trp549, both present in the TM4 domain, might interact with the vanillyl, catecholamine, or ethanolamine moiety of TRPV1 ligands.

Synthetic CB1 and CB2 receptor ligands, such as HU-210, JWH-015, and rimonabant, have also been reported to interact with TRPV1, although usually with lower relative intrinsic activity and potency than anandamide, whereas N-arachidonoyl dopamine displays markedly greater potency than anandamide as a TRPV1 agonist (Table 6) (). In contrast, phytocannabinoids that do not activate CB1 receptors, such as cannabidiol and cannabigerol (section II.C.5), were shown to act as full agonists at TRPV1 receptors (). Cannabidiol, in particular, exhibits almost the same Ki as capsaicin at the human TRPV1 receptor (), but seems to be significantly less potent at rat TRPV1 (). It is noteworthy that anandamide and N-arachidonoyldopamine congeners with little or no activity at CB1/CB2 receptors, such as oleoyl ethanolamide, linoleoyl ethanolamide, and N-oleoyl dopamine (section II.C.5), are also potent TRPV1 agonists ().

 

3. Other TRPV Channels.

Apart from TRPV1, five other TRPV channels, all insensitive to capsaicin, have been identified and cloned to date. TRPV2, -3, and -4 are involved in high-temperature sensing and nociception, whereas TRPV5 and -6 intervene in Ca2+ absorption/reabsorption (for review, see ). TRPV2 was previously known as growth factor-regulated Ca2+ channel or vanilloid receptor-like 1 and is activated by higher temperatures than TRPV1, but not by protons (). Strong evidence (albeit obtained so far in only one laboratory) exists for the interaction of cannabidiol, Δ9-THC, cannabinol, and, to a smaller extent, 11-hydroxy-Δ9-THC, nabilone, CP55940, and HU-210, with TRPV2. These effects were observed by measuring elevation of intracellular Ca2+ in HEK293 cells transfected with rat recombinant TRPV2 cDNA. Apart from cannabidiol, these compounds exhibited EC50 values much higher, or Emax values much lower, than those observed in functional assays that measure cannabinoid receptor-mediated responses (Table 6). It is noteworthy that cannabidiol has also been found to evoke TRPV2-mediated cation currents and release of calcitonin gene-related peptide in cultured dorsal root ganglion (DRG) neurons (). Because TRPV2, like TRPV1, is immediately desensitized by its agonists, these findings might explain some of the anti-inflammatory and/or analgesic properties of cannabidiol.

TRPV4 is activated by a variety of physical and chemical stimuli, including heat and decreases in osmolarity (). TRPV4 was originally reported to be activated by anandamide and 2-AG, probably via the formation of cytochrome P450 metabolites of arachidonic acid, such as epoxyeicosatrienoic acids (). It is noteworthy that cytochrome P450 metabolites of anandamide activate cannabinoid receptors () but have never been tested on TRPV4.

 

4. Other TRP Channels: TRPM8 and TRPA1.

TRPM8 and TRPA1 belong to two subfamilies different from that of the capsaicin (TRPV1) receptor but are still involved in thermosensation. Indeed, TRPM8 and TRPA1 were suggested to be activated by cold temperatures as well as by natural compounds (such as menthol in the case of TRPM8) and by various irritants [(mustard oil isothiocyanates, acrolein, etc.) in the case of TRPA1] (for review, see ). Recent evidence has emerged that both anandamide and N-arachidonoyl dopamine, but not 2-AG, can efficaciously antagonize the stimulatory effect of two TRPM8 agonists, menthol and the synthetic compound icilin, on intracellular Ca2+ elevation in HEK293 cells transfected with rat recombinant TRPM8 (), as well as in DRG neurons (). It is noteworthy that they did this in a cannabinoid receptor-independent manner. The concentrations at which anandamide exerted this effect ranged from submicromolar to 10 μM, depending on the type of TRPM8 agonist and cellular system used for the Ca2+ assay. It is noteworthy that several nonpsychotropic phytocannabinoids exert a similar action at TRPM8, and at significantly lower concentrations. Cannabidiol, cannabigerol, Δ9-THC, and Δ9-THC-acid were almost equipotent (IC50 = 70–160 nM in transfected HEK293 cells), whereas cannabidiolic acid was the least potent among the tested compounds (IC50 = 0.9–1.6 μM) ().

Evidence that TRPA1 (ANKTM1) is activated by phytocannabinoid CB1/CB2 receptor agonists [i.e., Δ9-THC and cannabinol (sections II.C.1 and II.C.5)], was first described in the article reporting the cloning of this channel (). The authors of this article showed that concentrations of these two compounds ≥10 μM were necessary to induce TRPA1-mediated effects, including endothelium- and CB1/CB2-independent vasodilation of rat mesenteric arteries. Later,  showed that micromolar concentrations of R-(+)-WIN55212 as well as the CB2-selective agonist AM1241 1) elicited TRPA1-mediated elevation of Ca2+ and currents in transfected CHO cells and trigeminal neurons; 2) desensitized TRPA1- and TRPV1-expressing cells to the action of capsaicin; 3) inhibited capsaicin-evoked nocifensive behavior in vivo in wild-type mice and much less so in TRPA1-null mutant mice ().  recently confirmed that Δ9-THC and R-(+)-WIN55212 activate rat recombinant TRPA1 and also found that the cannabinoid receptor-inactive isomer of the latter compound, S-(−)-WIN55212, exerts a similar action, although at higher concentrations. Furthermore, the cannabinoid quinone 3S,4R-p-benzoquinone-3-hydroxy-2-p-mentha-(1,8)-dien-3-yl-5-pentyl (HU-331) was also quite potent. Almost at the same time,  found that in HEK293 cells transfected with rat TRPA1, Δ9-THC could be much more potent at eliciting TRPA1-mediated elevation of intracellular Ca2+ than previously suspected and that several nonpsychotropic phytocannabinoids were even more potent, cannabichromene (EC50 = 60 nM) being the most potent and cannabigerol and cannabidiolic acid (EC50 = 3.4–12.0 μM) the least potent. Cannabichromene also activated mustard oil-sensitive (TRPA1-expressing) DRG neurons, although with lower potency (EC50 = 34.3 μM). It is noteworthy that anandamide was also recently reported to weakly but efficaciously stimulate TRPA1-mediated elevation of intracellular Ca2+ in transfected HEK293 cells (EC50 = 4.9 μM) ().

 

5. Conclusions.

That some endocannabinoids, particularly anandamide, exert pharmacological effects in vivo by activating TRPV1 receptors has now been demonstrated in hundreds of investigations, including some in which supporting evidence was obtained from experiments with TRPV1-null mice. Furthermore, although Δ9-THC exerts only very weak effects at TRPV1, phytocannabinoids that do not activate CB1/CB2-receptors, on the one hand, and synthetic CB1/CB2 ligands, on the other hand, have been shown to interact with this protein with significant potency and/or relative intrinsic activity. Thus, a strong case can be made for classifying TRPV1 as an “ionotropic cannabinoid receptor,” although this channel is clearly targeted less selectively than cannabinoid CB1 or CB2 receptors because its activity can be directly modulated by several endogenous or xenobiotic compounds that are not CB1 or CB2 receptor ligands. The capability of endocannabinoids and cannabidiol to inhibit TRPM8 functional responses might be a consequence, in part, of their agonist activity at TRPV1, because it has been observed that most of the regulatory events and mediators that affect this latter channel in one direction often affect the TRPM8 channel in the opposite direction (). Further studies should be carried out to investigate fully the possibility that TRPA1 mediates some of the pharmacological effects of phytocannabinoids and/or synthetic Δ9-THC-mimetic compounds, although the fact that this TRP channel is also extremely promiscuous in its responsiveness to physiological and synthetic ligands should be kept in mind. Finally, too little evidence exists to date to allow any conclusions to be drawn about whether or not other TRP channels (e.g., TRPV2 and TRPV4) might be putative cannabinoid receptors.

F. Other Ion Channels

 

1. Calcium Channels.

Certain cannabinoid receptor agonists (Table 7) and antagonists have been found to antagonize T-type voltage-gated calcium channels at concentrations in the mid-nanomolar or low micromolar range. Thus, for example, evidence has been obtained () for CB1 receptor-independent inhibition of the following:

  • Cloned CaV3.1 channels by anandamide (IC50 = 4.2 μM), R-(+)-methanandamide (10 μM), N-arachidonoyl dopamine (IC50 = 513 nM), ACEA (10 μM), and Δ9-THC (IC50 = 1.6 μM).

  • Cloned CaV3.2 channels by anandamide (IC50 = 330 nM), N-arachidonoyl dopamine (IC50 = 1122 nM), R-(+)-methanandamide (1 μM), ACEA (10 μM), Δ9-THC (IC50 = 1.3 μM), HU-210 (10 but not 1 μM) and rimonabant (100 nM and 1 μM) but not by CP55940, R-(+)-WIN55212, or 2-AG at 10 μM.

  • Cloned CaV3.3 channels by anandamide (IC50 = 1.1 μM) and N-arachidonoyl dopamine (IC50 = 355 nM), by R-(+)-methanandamide and ACEA at 10 μM, and by Δ9-THC (4.3 μM).

  • Native T-type calcium channels by R-(+)-methanandamide and Δ9-THC at 1 μM and AM251 at 3 μM, but not by R-(+)-WIN55212 at 1 μM.

 

TABLE 7

Apparent cannabinoid CB1 receptor-independent effects of cannabinoid receptor agonists on ion channels other than TRP channels

See Section III.F for references and further discussion.

Ion Channel and Effect Endocannabinoid(s)? Effective Concentration Range Nonendogenous Cannabinoid Agonist(s)? Effective Concentration Range
Current enhancement/activation
    KCa (BK) K+ Yesa Nanomolar or micromolar Yesb Nanomolar or micromolar
    Voltage-gated Na+ No Micromolar Yes Nanomolar
Inhibition
    Ca2+ (native) Yesc Nanomolar or micromolar Yesb Micromolar
    T-type Ca2+ (CaV3.1) Yesc Micromolar Yes Micromolar
    T-type Ca2+ (CaV3.2) Yesa,c Nanomolar or micromolar Yes Micromolar
    T-type Ca2+ (CaV3.3) Yesc Micromolar Yes Micromolar
    T-type Ca2+ (native) N.D. N.D. Yes Micromolar
    N-, P-, or P/Q-type Ca2+ Yesc Micromolar Yes Micromolar
    KATP K+ Yesc,d Micromolar Yesb Micromolar
    TASK-1 K+ Yesa,c Nanomolar or micromolar Yes Nanomolarb or micromolar
    TASK-3 K+ Yesc Micromolar Yesb Micromolar
    Voltage-gated K+ (KV) Yesc Nanomolar or micromolar Yes Nanomolarb or micromolar
    Voltage-gated Na+ Yes Micromolar Yes Micromolare
Radioligand binding (↓)
    L-type Ca2+ Yesc,d Micromolar Yesb Micromolar
    KATP K+ Yesc,d Micromolar Yesb Micromolar
    Voltage-gated Na+ Yesd Micromolar Yesd Micromolar

N.D., no data.

aAnandamide but not 2-arachidonoyl glycerol.
bOnly R-(+)-methanandamide.
cAlso R-(+)-methanandamide.
dMay target an allosteric site on this ion channel.
eSlight enhancement of sodium currents has also been detected in response to 10 nM R-(+)-WIN55212 (section III.G.3.).

It is noteworthy that the phytocannabinoid cannabidiol has also been found to inhibit cloned CaV3.1, CaV3.2, and CaV3.3 calcium channels, its reported IC50 values for this inhibition being 813, 776, and 3631 nM, respectively (). In addition, a recent article reported that the endogenous lipo-amino acids, NAGly and N-arachidonoyl GABA, can potently inhibit CaV3.1, CaV3.2, and CaV3.3 channels (IC50 <1.0 μM) ().

In addition, there is evidence that anandamide, 2-AG, and R-(+)-methanandamide but not Δ9-THC, CP55940, or R-(+)-WIN55212 can bind to L-type voltage-gated calcium channels, possibly in a noncompetitive/allosteric manner, their reported IC50 or apparent Ki values ranging from 3.2 to 40 μM (). There have also been reports that 1) displacement of [n-methyl-3H]diltiazem from L-type calcium channels can be induced by both rimonabant (IC50 = 6.1 μM) and taranabant (IC50 = 300 nM) (); 2) that anandamide (100 nM) and rimonabant (1 μM) can inhibit potassium-evoked Ca2+ influx into neonatal rat cultured DRG sensory neurons (); and 3) that anandamide at 1 μM and 2-AG and R-(+)-methanandamide, but not Δ9-THC, CP55940, or R-(+)-WIN55212 at 10 μM, can inhibit depolarization-induced Ca2+ efflux in transverse tubule membrane vesicles (). The data obtained in these investigations suggest that these effects of anandamide, 2-AG, and R-(+)-methanandamide on calcium flux were not CB1 receptor-mediated.

There is also evidence that at least some cannabinoids can antagonize N-, P-, and P/Q-type voltage-gated calcium channels. Thus, there have been reports that N- or P/Q-type channels can be antagonized at 10 μM by R-(+)-WIN55212, S-(−)-WIN55212, and anandamide, although not by 2-AG or noladin ether (), and that P-type channels can be antagonized by R-(+)-WIN55212 (10 μM), anandamide (IC50 = 1.04 μM), R-(+)-methanandamide (2 μM), and 2-AG (10 μM) (). These effects on N-, P- and P/Q-type channels seemed not to be CB1 receptor- or TRPV1 receptor-mediated. Arachidonic acid can also target T-, P-, and L-type channels, but it is unlikely that the ability of anandamide to antagonize these channels depends on its metabolic conversion to this acid ().

 

2. Potassium Channels.

Results obtained in a number of investigations suggest that several types of potassium channel can be targeted by certain cannabinoid receptor agonists (Table 7) and antagonists in a cannabinoid CB1 receptor-independent manner at concentrations in the nanomolar or low micromolar range. These are members of the 2TM domain family of channels (KATP channels), the 4TM domain family of leak or background channels (TASK and TREK), and the 6TM domain family of voltage-gated channels (KV and calcium-activated KCa channels).

Turning first to ATP-sensitive inward-rectifier (KATP) channels, there is evidence that anandamide is a noncompetitive inhibitor of such channels (IC50 = 8.1 μM) and that this endocannabinoid also inhibits [3H]glibenclamide binding to KATP channels, again in a noncompetitive manner (IC50 = 6.3 μM). These channels can be inhibited by R-(+)-methanandamide at 10 μM but not by rimonabant or SR144528 at 1 μM ().

Moving on to the 4TM domain family, it has been found that human or rat TASK-1 channels can be inhibited by R-(+)-methanandamide (IC50 = 700 nM), by anandamide at 3 and 10 μM, and by rimonabant (slight inhibition), CP55940, and R-(+)-WIN55212 but not Δ9-THC, HU-210, or 2-AG at 10 μM (). There have also been reports that human, rat, or mouse TASK-3 channels can be inhibited by R-(+)-methanandamide and anandamide at 1 to 10 μM () and that bovine TREK-1 channels are inhibited by anandamide (IC50 = 5.1 μM) ().

As to KV channels, there is evidence for the following:

  • KV1.2 channels are inhibited by anandamide (IC50 = 2.7 μM), arachidonic acid (IC50 = 6.6 μM), and Δ9-THC (IC50 = 2.4 μM) ().

  • Human cardiac KV1.5 channels are inhibited by anandamide (IC50 = 0.9 μM), 2-AG (IC50 = 2.5 μM), and, with a similar potency, by both R-(+)-methanandamide and arachidonic acid ().

  • Human KV1.5 channels expressed on HEK293 cells are also blocked by anandamide, which displays significantly greater potency intracellularly (IC50 = 213 nM) than extracellularly (IC50 = 2.1 μM) ().

  • KV3.1 channels are inhibited by anandamide (0.1–3 μM) and arachidonic acid (3 μM) () and cardiac KV4.3 channels are inhibited by anandamide (IC50 = 400 nM), R-(+)-methanandamide (IC50 = 600 nM), and 2-AG (IC50 = 300 nM) ().

 

In addition, there has been a report that KV channels can be inhibited not only by anandamide (IC50 = 600 nM) and R-(+)-methanandamide (10 μM) but also by rimonabant (10 μM) and R-(+)-WIN55212 (20 μM), although not by arachidonic acid (). It has been found too that both rimonabant (IC50 = 2.5 μM) and taranabant (IC50 = 2.3 μM) can induce radiolabeled ligand displacement from rapid delayed rectifier KV channels (). Evidence has also recently been obtained that anandamide (1 μM) can strongly inhibit KV channel-mediated delayed rectifier outward potassium current (). A similar degree of inhibition was induced by R-(+)-methanandamide (1 μM). However, both R-(+)-WIN55212 and 2-AG induced less inhibition at 1 μM than anandamide.

Finally, there is evidence that anandamide (EC50 = 631 nM or 4.8 μM) and R-(+)-methanandamide (EC50 = 7.9 μM) but not 2-AG (up to 10 μM) can increase the activity of KCa (BK) channels (). Results obtained in experiments with rat isolated coronary arteries also suggest that such activation can be produced by both anandamide and R-(+)-methanandamide at concentrations of 0.3 to 3 μM, although not by JWH-133 at 1 μM ().

 

3. Sodium Channels.

There is evidence that voltage-gated sodium channels can be targeted by some cannabinoid receptor agonists at concentrations in the nanomolar or micromolar range (Table 7). Thus, for example, R-(+)-WIN55212 has been found by  to induce a slight enhancement (11.5%) of voltage-gated sodium currents in rat cultured trigeminal ganglion neurons at 10 nM but to inhibit these currents at higher concentrations (IC50 = 17.8 μM). In an earlier investigation,  also found that R-(+)-WIN55212 can inhibit voltage-gated sodium channels. More specifically, they obtained evidence that this aminoalkylindole can act in a CB1 receptor-independent manner to inhibit 1) depolarization of mouse brain synaptoneurosomes induced by the sodium channel-selective activator veratridine (IC50 = 21.1 μM); 2) veratridine-dependent release of l-glutamic acid (IC50 = 12.2 μM) and γ-aminobutyric acid (IC50 = 14.4 μM) from mouse purified whole-brain synaptosomes; and 3) the binding of the sodium channel site 2-selective ligand [3H]batrachotoxinin A 20-α-benzoate, to mouse brain synaptoneurosomal voltage-gated sodium channels (IC50 = 19.5 μM). Anandamide displayed similar inhibitory activity in these four bioassays (IC50 = 21.8, 5.1, 16.5, and 23.4 μM, respectively) () as did AM251 (IC50 = 8.9, 8.5, 9.2, and 11.2 μM, respectively) (). CP55940, N-arachidonoyl dopamine, noladin ether, and 2-AG have also been found to displace [3H]batrachotoxinin A 20-α-benzoate from mouse brain synaptoneurosomal voltage-gated sodium channels (IC50 = 22.3, 20.7, 51.2, and 90.4 μM, respectively) (,). The mechanism underlying the inhibition of binding induced in this assay by these four compounds, and by AM251 and anandamide but not by R-(+)-WIN55212, could well be allosteric in nature (,). It has been found too by  that the ability of R-(+)-WIN55212 and anandamide to inhibit veratridine-dependent depolarization of mouse synaptoneurosomes () extends to CP55940 (IC50 = 3.2 μM). This inhibition was produced by CP55940 in a noncompetitive manner. There have also been reports that voltage-gated sodium channels can be inhibited by anandamide in rat DRG sensory neurons (apparent Kd = 5.4 and 38.4 μM for tetrodotoxin-sensitive and tetrodotoxin-resistant sodium currents, respectively) () and by R-(+)-WIN55212 (10 μM) and noladin ether (50 μM) in frog parathyroid cells ().

 

4. Conclusions.

Several established endogenous and nonendogenous CB1/CB2 receptor agonists and some CB1 receptor antagonists can induce CB1 receptor-independent blockade of certain types of calcium, potassium, and sodium channels (Table 7), sometimes in an apparently noncompetitive/allosteric manner. Such blockade is induced by these agonists and antagonists at micromolar or nanomolar concentrations. There is also evidence that anandamide and R-(+)-methanandamide can activate KCa (BK) channels and that a subinhibitory concentration of R-(+)-WIN55212 (10 nM) can enhance voltage-gated sodium currents. Given the relatively high (nanomolar) potency with which CaV3.2, KV1.5, and KCa (BK) channels can apparently be targeted by anandamide, further research directed at seeking out CB1 receptor-independent physiological and/or pathological roles of these ion channels in the endocannabinoid system is clearly warranted.

G. Peroxisome Proliferator-Activated Receptors

Peroxisome proliferator-activated receptors (PPARs) are ligand-activated transcription factors that constitute part of the nuclear receptor family (). Activated PPARs form functional units as heterodimers with retinoid X receptors. Classical agonists at PPARs are fatty acids and their derivatives, ranging from oleic acid and arachidonic acid to leukotriene B4 and 15-deoxy-Δ12,14-prostaglandin J2. It is a widespread view that PPARs are not activated by a single endogenous ligand but are generalized lipid sensors, monitoring local changes in metabolism. There are three PPAR isoforms, PPARα, PPARβ, and PPARγ. PPARα is the target of the clinically employed antihyperlipidemic fibrates, including gemfibrozil and fenofibrate. PPARγ is a therapeutic target in type 2 diabetes. Its ligands include pioglitazone, rosiglitazone, and troglitazone, which are known collectively as thiazolidinediones. PPARβ is also known as PPARδ on the basis of differential naming of species orthologs and has yet to be targeted effectively in the clinic. All three isoforms are expressed in liver to some degree, although PPARα predominates in skeletal muscle and PPARγ in adipose tissue. Because they are nuclear receptors, signal transduction at PPARs is primarily directed through alterations in gene transcription. A number of PPAR target genes have been identified, many of which are associated with lipid turnover, such as long-chain fatty acyl-CoA dehydrogenase and lipoprotein lipase; these are often used as markers of PPAR activation both in vitro and in vivo.

 

1. Direct Evidence for Peroxisome Proliferator-Activated Receptor Activation or Occupancy by Cannabinoids and Related Molecules.

A number of cannabinoid CB1/CB2 agonists have also been reported to be PPAR agonists in in vitro experiments (Table 8). These include the archetypal endocannabinoids anandamide and 2-AG as well as the phytocannabinoids Δ9-THC and cannabidiol and the synthetic cannabinoids R-(+)-WIN55212 and ajulemic acid (Table 8). The potency of the majority of these agents is approximately 2 orders of magnitude lower at PPARs than at the conventional cannabinoid GPCRs (CB1 and CB2), although it is noteworthy that data from two investigations suggest that PPARγ can be activated by Δ9-THC () and R-(+)-WIN55212 () at 100 nM. There are also reports that rimonabant and AM251 can activate both PPARγ (at 10 μM) and PPARα (). However, it remains unclear whether this represents a direct or indirect phenomenon associated with high ligand concentrations. Two fatty acid ethanolamides (oleoyl ethanolamide and palmitoyl ethanolamide), which are essentially inactive at cannabinoid CB1 and CB2 receptors, are agonists with reasonable potency at PPARα (Table 8). This receptor seems to display some preference for the medium-chain mono- or diunsaturated fatty acid ethanolamides, such as oleoyl ethanolamide and linoleoyl ethanolamide, compared with the long-chain polyunsaturated fatty acid ethanolamides, such as anandamide, N-eicosapentaenoylethanolamine and N-docosohexaenoylethanolamine ().

TABLE 8

Potency of cannabinoids and endocannabinoid-like molecules at peroxisome proliferator-activated receptors (EC50 or IC50 values or effective concentrations)

Cannabinoid or Related Molecule PPARα PPARβ/δ PPARγ
μM
Palmitoyl ethanolamide 3a >30a >30a; >60b
Stearoyl ethanolamide >30a N.D. N.D.
Oleoyl ethanolamide 0.12c; 0.12d 1c >10c
Anandamide >10c; 10 to 30e N.D. 8b; 10f
2-AG N.D. >1g 10f; ∼30h
Noladin ether 10 to 30e N.D. ∼30h
Virodhamine 10 to 30e N.D. N.D.
R-(+)-WIN55212 20e N.D. N.D.
Δ9-THC >100e N.D. ∼0.3i
Cannabidiol N.D. N.D. 5j
Ajulemic acid >50j >50j 0.6j; 13k
Rimonabant Activationl N.D. 10l
AM251 Activationl N.D. 10l

N.D., no data.

aEC50 reporter gene assay in HeLa cells expressing recombinant receptors ().
bEC50 reporter gene assay in COS cells expressing recombinant receptors ().
cEC50 reporter gene assay in HeLa cells expressing recombinant receptors ().
dEC50 reporter gene assay in MCF7 cells expressing recombinant receptors ().
eIC50 competition for fluorescent ligand occupancy of ligand binding domain ().
fIC50 competition for fluorescent ligand occupancy of ligand binding domain ().
gEC50 reporter gene assay in human umbilical vein endothelial cells in the presence of a COX-2 inhibitor ().
hEC50 reporter gene assay in 3T3-L1 cells expressing recombinant receptors ().
iEC50 reporter gene assay in HEK293 cells expressing recombinant receptors ().
jIC50 competition for fluorescent ligand occupancy of ligand binding domain ().
kEC50 reporter gene assay in HEK293 cells expressing recombinant receptors ().
lEffective concentration in reporter gene assay in HEK293 cells expressing recombinant receptors; ligand concentration(s) for PPARα activation not published ().

Oxidative metabolism of endocannabinoids generates agents that are also active at PPARs. Thus, 15-lipoxygenase metabolism of 2-AG leads to the production of 15(S)-hydroxyeicosatetraenoic acid-glyceryl ester (). This agent was identified as a preferential agonist at PPARα in a reporter gene assay using NIH 3T3 cells as hosts, with an EC50 value of ∼3 μM, but was inactive up to 10 μM at PPARβ or PPARγ. It is noteworthy that the arachidonic acid metabolite 15(S)-hydroxyeicosatetraenoic acid at 10 μM is a preferential PPARγ agonist (). 2-(14,15-Epoxyeicosatrienoyl)glycerol (a product of 2-AG metabolism by epoxygenase) but not 2-AG itself also appeared to be a PPARα agonist (). However, mass spectrometry analysis suggested that the active entity was 14,15-dihydroxyeicosatrienoic acid, produced through sequential hydrolysis of the glyceryl ester and epoxide bonds. One of the problems this study highlights is that the extended periods necessary for identification of agonist action at PPARs (4 h or more) increases the potential for conversion of the added agent into an entity with altered activity at the receptor.

 

2. Indirect Evidence for Cannabinoid Activation of Peroxisome Proliferator-Activated Receptors.

Indirect means of identifying an action of cannabinoid-related molecules at PPARs include 1) the use of reporter genes in model cells where PPARs are either endogenously expressed or overexpressed and 2) the use of pharmacological or genetic inhibition of PPARs.

 

3. Reporter Gene Assays and Metabolism of Endocannabinoids.

COX-2-mediated metabolism of 2-AG, noladin ether, and anandamide led to greater activation of PPARβ compared with PPARγ, with no discernible PPARα activation, in human umbilical vein endothelial cells transfected with peroxisome proliferator-activated response element coupled to a reporter gene (). COX-2 inhibition alone did not alter PPAR activation, suggesting that conversion of endogenous COX-2 substrates was not sufficient to activate PPARs. 2-AG itself, in the presence of a COX-2 inhibitor, was ineffective, indicating that COX-2 metabolism was an obligatory step in the production of a PPAR ligand. siRNA knockdown of prostacyclin synthase (CYP8A1) inhibited PPAR activation, demonstrating that prostaglandin I2-glyceride was probably the endogenous ligand, although attempts to identify endogenous levels of this agent using mass spectrometry were unsuccessful ().

 

4. Antagonism.

In mouse primary splenocytes in vitro, anandamide evoked a concentration-dependent inhibition of interleukin-2 secretion that was not blocked by rimonabant or SR144528 but was inhibited by 2-chloro-5-nitro-N-4-pyridinyl-benzamide (T0070907), indicating a role for PPARγ in these effects (). Sequential metabolism in HeLa human cervical carcinoma cells by COX-2 and lipocalin type-prostaglandin D synthase led to activation of PPARγ by R-(+)-methanandamide, as evidenced by siRNA and 2-chloro-5-nitro-N-phenylbenzamide (GW9662) inhibition (). In vitro, the PPARγ antagonist GW9662 was able to block relaxations evoked by cannabidiol (), Δ9-THC (), N-arachidonoyl dopamine and anandamide () in rat isolated aorta. R-(+)-WIN55212-evoked apoptosis of HepG2 human hepatoma cells was inhibited by GW9662 (). Prolonged exposure of these cells to R-(+)-WIN55212 also induced increased expression of PPARγ, a response associated with PPARγ activation. Likewise, HU-210 exposure over several days increased PPARγ expression in 3T3-F442A mouse preadipocyte cells; however, this effect was blocked by rimonabant, indicating the involvement of cannabinoid CB1 receptors ().

 

5. Genetic Disruption.

An alternative method, although not without caveats, is to make use of animals in which the PPARs are genetically disrupted. These mice have been employed on a limited number of occasions in an attempt to study the effects of cannabinoid-related molecules, although, notably, not the archetypal endocannabinoids anandamide or 2-AG, or other ligands that are known to target CB1 or CB2 receptors. Thus, antinociceptive effects of palmitoyl ethanolamide, as well as the PPARα agonist 2-[[4-[2-[[(cyclohexylamino)carbonyl](4-cyclohexylbutyl)amino]ethyl]-phenyl]thio]-2-methylpropanoic acid (GW7647), were lost in mice in which the ppara gene was disrupted (). ppara gene disruption abrogated oleoyl ethanolamide-evoked feeding behaviors (), oleoyl ethanolamide-evoked lipolysis () and oleoyl ethanolamide-mediated neuroprotection (), but not oleoyl ethanolamide effects on visceral pain () or intestinal motility (). Genetic disruption of PPARγ in a homozygous manner results in embryonic lethality, so heterozygous mice have been generated and studied (), as have mice with conditional disruption of the pparg gene (). However, the use of these mice or those in which PPARβ is genetically disrupted () has not been reported in investigations of endocannabinoids and their analogs.

 

6. Amplification of Endocannabinoid Levels and Peroxisome Proliferator-Activated Receptors.

Nicotine-induced elevation of firing of ventral tegmental area neurons in anesthetized rats was inhibited by prior administration (60–120 min) of URB597, an inhibitor of the anandamide-metabolizing enzyme FAAH (). Rimonabant failed to alter the effects of URB597, whereas 1-[(4-chlorophenyl) methyl]-3-[(1,1-dimethylethyl)thio]-α,α-dimethyl-5-(1-methylethyl)-1H-indole-2-propanoic acid (MK886) was able to prevent the effects of URB597. The effects of URB597 were mimicked in vitro by administration of oleoyl ethanolamide, palmitoyl ethanolamide, or the PPARα agonist WY14643 in a manner sensitive to MK886 ().

Palmitoylallylamide, described as a low-potency inhibitor of FAAH (), evoked an inhibition of nociceptive behaviors in rat models of neuropathic pain 40 to 100 min after administration (). These effects were reduced after administration of rimonabant, SR144528, or MK886, dependent on the paradigm under study. In a passive avoidance paradigm, URB597 administration to rats 40 min before testing enhanced memory acquisition (), whereas Δ9-THC administration impaired memory. The URB597 enhancement was reduced by either rimonabant or MK886. Intraplantar administration of a low, but not high, dose of URB597 reduced pain behaviors in the carrageenan model of inflammatory pain () up to 210 min after injection. At this time, levels of anandamide and 2-AG, but not oleoyl ethanolamide or palmitoyl ethanolamide, were elevated in URB597 and carrageenan-treated paws compared with tissue from animals treated with carrageenan alone. The antinociceptive effects of URB597 were prevented by coadministration of the PPARα-selective antagonist [(2S)-2-[[(1Z)-1-methyl-3-oxo-3-[4-(trifluoromethyl)phenyl)-1-propenyl]amino]-3-[4-[2-(5-methyl-2-phenyl-4-oxazolyl)ethoxyphenylpropyl]-carbamic acid ethyl ester (GW6471) but not the PPARγ-selective antagonist GW9662, indicating a role for PPARα in these responses.

Blockade of endocannabinoid uptake by (5Z,8Z, 11Z,14Z)-N-(3-furanylmethyl)-5,8,11,14-eicosatetraenamide (UCM707) evoked neuroprotective effects in mouse mixed astrocytic neuronal cultures by activation of CB1, CB2, and PPARγ receptors (). Application of CB1 and CB2 receptor antagonists, but not the PPARγ antagonist GW9662, led to an exacerbation of the excitotoxic effects of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid application. Administration of N-(4-hydroxyphenyl)-5Z,8Z, 11Z,14Z-eicosatetraenamide (AM404), a mixed inhibitor of FAAH and endocannabinoid uptake, enhanced bacterial lipopolysaccharide-evoked tumor necrosis factor-α levels in rat plasma. Antagonists of CB1, CB2, and TRPV1 receptors attenuated elevations of tumor necrosis factor-α levels in response to LPS and the combination of LPS and AM404, whereas PPARγ antagonism interfered only with the AM404-evoked elevation (). Collectively, these data might be taken to indicate that PPARs are only activated significantly in vivo by endocannabinoids when endocannabinoid levels are pharmacologically amplified.

 

7. Regulation by Peroxisome Proliferator-Activated Receptors and Peroxisome Proliferator-Activated Receptor Ligands of the Endocannabinoid System.

There is some evidence that the functioning of the endocannabinoid system can be affected by PPARs and PPAR ligands. This has come primarily from experiments with adipocytes. For example, 24 h of activation of PPARγ by ciglitazone decreased levels of 2-AG, but not anandamide, in immature, but not mature, 3T3 F442A mouse adipocytes (). It remains to be determined whether this effect is due to a direct action on endocannabinoid synthesizing and/or metabolizing enzymes, or a consequence of cellular differentiation; interestingly, however,  have found that the PPARγ agonist rosiglitazone up-regulates FAAH in human adipocytes. In 3T3 L1 mouse preadipocyte cells, a related cell type, ppard gene silencing increased CB1 receptor gene expression (). Treatment of human adipocytes in vitro with the PPARγ agonist rosiglitazone markedly down-regulated CB1 receptor gene expression, whereas R-(+)-WIN55212 up-regulated PPARγ (). Treatment of human adipocytes in vitro with a different PPARγ agonist, ciglitazone, failed to alter endocannabinoids levels ().

In a distinct series of investigations, a number of PPARγ agonists, including ciglitazone, pioglitazone, rosiglitazone, and troglitazone, were shown to inhibit FAAH activity in rat brain membranes, as well as in intact C6 rat glioma and RBL-2H3 rat basophilic leukemia cells (). The low potency with which these inhibitions were induced (IC50 values ∼100 μM) suggests that this is an unlikely mechanism for influencing endocannabinoid levels.

 

8. Conclusions.

Although there has yet to be a systematic investigation of the activity of all the putative endocannabinoids at all three PPAR isoforms, anandamide and 2-AG have each already been reported to activate two PPAR receptors (Table 8): PPARγ and PPARα (anandamide) and PPARγ and PPARβ/δ (2-AG). Overall, the potencies of endocannabinoids and their metabolites as PPAR agonists or antagonists are relatively low compared with their potencies as agonists of canonical cannabinoid CB1/CB2 receptors. This might be taken as evidence that endocannabinoids are poor candidates as PPAR ligands in vivo. However, this fails to take into account background levels of established endogenous PPAR agonists. One estimate puts intracellular levels of long-chain fatty acids at 20 μM (), which is a level sufficient to occupy PPARs in cell-free systems. Although this background level may vary depending on the cell type and the active state of the cell, fluctuations in intracellular endocannabinoid levels may well prove sufficient to activate PPARs in vivo.

The best current evidence that endocannabinoids are endogenous agonists at PPARs in vivo derives from the use of a model of inflammatory pain. More specifically, local administration of a FAAH inhibitor has been found to induce local accumulation of both anandamide and 2-AG (but not oleoyl ethanolamide or palmitoyl ethanolamide) at a time when behavioral responses to the FAAH inhibitor could be blocked by local administration of a PPARα antagonist but not a PPARγ antagonist ().

H. Some Putative Receptors

 

1. Imidazoline-Like Receptors.

There is evidence that at least some cannabinoid receptor agonists and antagonists can interact with a non-I1, non-I2 subtype of the putative imidazoline receptor that may belong to a family of G protein-coupled sphingosine-1-phosphate/LPA receptors originally known as endothelial differentiation gene receptors. These putative receptors have been reported to be activated by CP55940 (300 nM) and R-(+)-WIN55212 (10 and 100 μM) but not by anandamide (1 μM) and to be blocked by rimonabant (1 μM) and LY320135 (0.1 to 10 μM) ().

 

2. The Putative Abnormal-Cannabidiol Receptor.

Since its original description as a receptor mediating the CB1/CB2-independent, endothelium-dependent vasodilator effect of anandamide and some atypical cannabinoids (), the putative abn-CBD receptor, also called the endothelial anandamide receptor (), has eluded molecular identification. The emergence of GPR55 as the first GPCR other than CB1 or CB2 receptors with high affinity for certain cannabinoid ligands has raised the question of its possible identity with the putative abn-CBD receptor.

There are a number of important parallels between the pharmacology of GPR55 and the putative abn-CBD receptor (see also section III.A.6). Thus, as summarized recently (), it has been shown that in at least some GPR55 and putative abn-CBD receptor bioassays, 1) the endocannabinoid, anandamide, is active; 2) the synthetic cannabinoid receptor agonist R-(+)-WIN55212 is inactive (in all GPR55 and abn-CBD bioassays); 3) abn-CBD and the compound O-1602 (section II.C.5 and Fig. 5) () act as agonists, whereas their parent compound, cannabidiol, acts as an antagonist; 4) the CB1 antagonist rimonabant, but not its close structural analog AM251, can inhibit both apparent GPR55- and apparent abn-CBD receptor-mediated responses, although with lower potency than it displays as a CB1 receptor antagonist; 5) the endogenous lipid N-arachidonoyl serine causes pertussis toxin- and 1,3-dimethoxy-5-methyl-2-[(1R,6R)-3-methyl-6-(1-methylethenyl)-2-cyclohexen-1-yl]benzene (O-1918)-sensitive, endothelium-dependent mesenteric vasodilation and thus may be an agonist of the abn-CBD-sensitive receptor (). It has also been reported recently that nanomolar concentrations of N-arachidonoyl serine promote angiogenesis and endothelial wound healing in human dermal microvascular endothelial cells and that these effects could be partly inhibited by siRNA-mediated knockdown of GPR55 and could also be antagonized by O-1918 but not by AM251 ().

On the other hand, there are some key differences: 1) vasodilator effects attributed to abn-CBD receptor activation are moderately sensitive to pertussis toxin, suggestive of Gi/o protein involvement, whereas GPR55 signals through Gα12, Gα13, or Gαq; 2) lysophosphatidyl inositol is a potent agonist of GPR55, but not of the putative abn-CBD receptor; and 3) the hypotensive/vasodilator actions of abn-CBD and its antagonism by the synthetic cannabinoid, O-1918 (), persist in GPR55 knockout mice (). However, the possibility exists that GPR55 and CB1 receptors may form heterodimers with pharmacological properties distinct from those of either one of the monomeric receptors, which may account for some or all of the above differences. Future experiments could address this question by exploring the presence of such heterodimers in various tissues under native conditions and/or by analyzing the pharmacology of cannabinoid-induced responses after the individual or joint transgenic expression of these two receptors in cells devoid of both.

Evidence has recently emerged that the putative abn-CBD receptor has features similar to those of GPR18 (). In BV-2 microglia and in GPR18-transfected HEK293 cells, but not in the parent HEK293 cells, abn-CBD and O-1602 potently stimulated cell migration and proliferation. Furthermore, the pertussis toxin-sensitive potent pro-migratory effect of NAGly, the putative endogenous ligand of GPR18 (section III.B.4), was inhibited by O-1918 or by the low-efficacy agonists N-arachidonoyl serine and cannabidiol. Although vascular effects were not tested in this study, it is noteworthy that in earlier experiments with human umbilical vein endothelial cells, abn-CBD-induced cell migration was found to be susceptible to inhibition by O-1918 or pertussis toxin (). These two compounds, O-1918 and pertussis toxin, have also been reported to inhibit NAGly-induced endothelium-dependent vasorelaxation of rat mesenteric artery segments (). However, the parallel between the putative abn-CBD receptor and GPR18 is not perfect. Rimonabant, which is a low-affinity antagonist of the vascular putative abn-CBD receptor (for review, see ), had no effect on GPR18-mediated microglial migration at concentrations up to 1 μM (). Clearly, the exciting possibility that the putative abn-CBD receptor is GPR18 should be followed up.

In conclusion, there is evidence to support the contention that both GPR55 and GPR18 have at least some of the features associated with the abn-CBD-sensitive receptor. This will need to be validated by further studies.

 

3. A Putative Receptor for Anandamide and R-(+)-WIN55212.

Investigations with C57BL/6 CB1(−/−) mice revealed that antinociception, locomotor hypoactivity, and catalepsy (immobility) responses could be elicited in these animals by anandamide but not Δ9-THC, leading  to postulate the existence of a non-CB1 receptor for this endocannabinoid agonist. Using CB1(−/−) brain membranes, they demonstrated that anandamide and R-(+)-WIN55212, but not Δ9-THC, CP55940, or HU-210, stimulated [35S]GTPγS binding, supporting the existence of a novel non-CB1, Gi/o-coupled receptor (). The potencies for anandamide (EC50 = 3.6 μM) and R-(+)-WIN55212 (EC50 = 1.8 μM) were low compared with the CB1 receptor (), and although rimonabant antagonized both these ligands, it did so in a manner that may not have been competitive in nature because it occurred only at concentrations (IC50 >1 μM) high enough to reduce basal [35S]GTPγS binding (). The anandamide/R-(+)-WIN55212-stimulated [35S]GTPγS binding was found in brain regions that were not abundant in CB1 receptors, such as the brain stem, midbrain, and spinal cord, as well as in some areas that are well populated with the CB1 receptor, such as the cerebral cortex and hippocampus (). The occurrence of non-CB1-mediated anandamide/R-(+)-WIN55212-stimulated [35S]GTPγS binding was later confirmed in another study carried out with CB1(−/−) mouse brain membranes (). A similar pharmacological profile is apparent in pertussis toxin-sensitive inhibition of cAMP accumulation in cultured mouse striatal astrocytes, which has been found to respond to anandamide and WIN55212, but poorly to CP55940, and to be resistant to antagonism by rimonabant (). In contrast to neurons, the mouse striatal astrocytes did not express immunoreactive CB1 receptors and did not exhibit [3H]SR141716A-binding activity ().

In conclusion, the current evidence that this target, and indeed the targets described in sections III.H.1 and III.H.2, are non-CB1, non-CB2 cannabinoid receptors is based upon pharmacological profiling only, with the additional caveats that only a limited number of agents were assessed and the two identified agonists also exhibit broad CB1/CB2 pharmacological activity. Moreover, the possibility that anandamide, or anandamide metabolic products, trigger responses via other bona fide targets is an equally plausible alternative potential mechanism that remains to be investigated. A radioligand binding assay for this putative receptor based upon [3H]R-(+)-WIN55212 would be plagued by the abundance of brain CB1 receptors, so a more selective radioligand probe would have to be developed to define the receptor binding profile. Thus, in the absence of a more comprehensive pharmacological profile and of a unique, identified protein, the naming of this pharmacological target as a definitive cannabinoid receptor cannot be justified.

 

4. The Putative CBsc Receptor for R-(+)-WIN55212.

A non-CB1R-(+)-WIN55212-sensitive receptor has been proposed to be present in the CA1 region of the hippocampus (). This putative cannabinoid receptor has been provisionally named CBsc because of its apparent location on hippocampal Schaffer collateral/commissural axon terminals (). Pharmacological evidence for the CBsc receptor is that 1 μM R-(+)-WIN55212 reduced amplitudes of evoked excitatory postsynaptic potentials in hippocampal slice preparations from CB1(−/−) (and wild-type) CD1 mice (). Although this response was observed in Wistar and Sprague-Dawley rats, the response to R-(+)-WIN55212 was absent in young/adult C56Bl/6 CB1(−/−) and wild-type C57BL/6 mice (). It was detectable, however, in neonatal and juvenile wild-type C57BL/6 mice (). Additional supporting evidence was the failure to identify immunoreactive CB1 receptors at glutamatergic presynaptic terminals (). However, such immunoreactivity has since been detected using more sensitive antibodies/methods (). In some experiments, the log concentration-response curve for R-(+)-WIN55212 exhibited an apparent EC50 of 2 μM but failed to develop an asymptotic maximum even at a concentration of 30 μM (). At such a high concentration, nonstereoselective R-(+)/S-(−)-WIN55212 effects on N-type voltage-gated Ca2+ channels could be responsible for the decrement in presynaptic neurotransmission (). In CA1 pyramidal cells and dentate gyrus granule cells, R-(+)-WIN55212 or CP55940 influenced glutamate release via a capsazepine-sensitive mechanism (). It is unlikely, however, that TRPV1 channels were responsible for this effect, because there is evidence that both capsaicin and capsazepine can reduce hippocampal glutamatergic neurotransmission in TRPV1(−/−) mice as well as in wild-type animals (). The putative CBsc receptor may be a TRPV1-like receptor. It is noteworthy that it is also unlikely that the proposed CBsc receptor is the same as the putative R-(+)-WIN55212 receptor first identified by , because evidence for the existence of this latter receptor came from experiments with C57BL/6 mouse tissue (section III.H.3).

In the absence of a more complete pharmacological characterization of the putative CBsc receptor, it remains possible that those effects on glutamate release that are induced in CB1-expressing Schaffer collaterals by cannabinoids with significant potency and that are sensitive to antagonism by either rimonabant or AM251, at least in rat and CD1 mouse tissue (), are in fact mediated by the CB1 receptor. Moreover, those effects on glutamate release that are induced by cannabinoids with low potency, and in a nonstereoselective, and CB1 antagonist-insensitive manner, could result from perturbations of ion channel functions.

In conclusion, given the positive results obtained with R-(+)-WIN55212 in CB1(−/−) CD1 mouse tissue and some negative results obtained with this cannabinoid in wild-type C57BL/6 mouse tissue, further experiments directed at establishing whether the proposed CBsc receptor truly is a novel cannabinoid receptor are clearly warranted.

IV. Phylogenetic Relationships

The sequence similarity (44% amino acid sequence identity) that the CX5 receptor (CB2) shares with the SKR6 receptor (CB1) provided crucial evidence that CX5 might be a cannabinoid receptor (). Thus, the identification of the cannabinoid CB2 receptor heralded a new approach to the discovery of putative cannabinoid receptors: homology-based searching of gene sequence databases. Furthermore, sequencing of the human genome () enabled genome-wide searches for receptors structurally and/or functionally related to CB1 and CB2.

An analysis of human gene/genome sequence data carried out in 2003 revealed more than 800 genes encoding GPCRs, which were grouped on the basis of sequence relationships, into five main families named after a prototypical receptor: 1) glutamate (G), 2) rhodopsin (R), 3) adhesion (A), 4) frizzled/taste2 (F), and 5) secretin (S), with the rhodopsin family further subdivided into groups α, β, γ, and δ (). This “GRAFS” classification system provides a useful framework both for analysis of relationships between CB1/CB2-type cannabinoid receptors and other G protein-coupled receptors and for investigation of the occurrence of putative non-CB1/CB2 G protein-coupled cannabinoid receptors in humans and other mammals.

A. CB1, CB2, and Other Rhodopsin α Group-Type G Protein-Coupled Receptors

Cannabinoid CB1 and CB2 receptors belong to the α group of rhodopsin-type GPCRs, which is composed largely of receptors for amine-type neurotransmitters and neuromodulators (e.g., serotonin, adrenaline, dopamine) (). Thus, CB1 and CB2 receptors are atypical of the α group in that they are activated endogenously by the lipid-type signaling molecules anandamide and 2-AG. CB1 and CB2 receptors are not, however, the only receptors in the α group that are activated by lipid ligands. The receptors in the α group that share the highest level of sequence similarity with CB1 and CB2 are a group of eight receptors that are activated by lysophospholipids. These include the sphingosine-1-phosphate (S1P) receptors (S1P1, S1P2, S1P3, S1P4, S1P5) and the LPA receptors LPA1, LPA2, and LPA3 (). It is noteworthy that S1P exerts CB1-independent cannabinoid-like effects, including thermal antinociception, hypothermia, catalepsy, and hypolocomotion (), whereas sphingosine and the sphingosine analog 2-amino-2-(2-[4-octylphenyl]ethyl)-1,3-propanediol (FTY720) inhibit binding of cannabinoid ligands to CB1 receptors in vitro (). Thus, there are some striking similarities in the pharmacological properties of the CB1 receptor and S1P receptors.

Cannabinoid CB1/CB2 receptors and the related lysophospholipid receptors belong to a distinct branch of α-type receptors (Fig. 6) that also includes receptors for melanocortin peptides (MC1-MC5), adenosine receptors (A1, A2A, A2B, A3), and the orphan receptors GPR3, GPR6, and GPR12 (). Thus, on the basis of analysis of human genome sequence data, it can be speculated that, as a consequence of multiple gene duplication events, an ancestral α-type GPCR gave rise to receptors that are activated by endocannabinoids, lysophospholipids, melanocortin peptides, and adenosine. Furthermore, comparative analysis of genome sequence data reveals that adenosine receptors have a wider phylogenetic distribution than CB1/CB2-type cannabinoid receptors, lysophospholipid receptors, melanocortin receptors, and GPR3, GPR6, and GPR12 (). Therefore, the common ancestor of this group of GPCRs may in fact have been an adenosine receptor. Furthermore, orthologs of the CB1/CB2-type cannabinoid receptors have only been found in the phylum Chordata (vertebrates, urochordates, and cephalochordates) (). Therefore, the duplication of a putative adenosine receptor gene that may have ultimately given rise to CB1/CB2-receptors probably occurred in a common ancestor of extant chordates.

An external file that holds a picture, illustration, etc. Object name is zpg0041022940006.jpg

Neighbor joining tree showing relationships between the human cannabinoid CB1 and CB2 receptors and other human rhodopsin α-group type G protein-coupled receptors. The tree was generated using the multiple sequence alignment program ClustalX, with bootstrapping (1000 bootstrap trials), and viewed using NJ plot. The tree shows receptors that are discussed in section IV.A of this review; the muscarinic acetylcholine receptors M1–M5 are included as an outgroup.

As highlighted above, among the receptors that are closely related (on the basis of sequence similarity) to CB1 and CB2 are three receptors known as GPR3, GPR6 and GPR12 (section III.B.3). It has been reported that these “orphan” receptors are activated in vitro by the lysophospholipid S1P (). However, a more recent screen using the β-arrestin PathHunter assay found that these receptors were not activated by S1P, even at concentrations as high as 8 μM (). Furthermore, GPR3, GPR6, and GPR12 are not activated by the endocannabinoids anandamide and 2-AG in this assay, suggesting that they may not be cannabinoid receptors. It is noteworthy, however, that the lysophospholipid receptor S1P1, was activated by anandamide and 2-AG, albeit only at concentrations in the micromolar range ().

Another rhodopsin α group-type GPCR that is of interest with respect to cannabinoid receptors is GPR119 (section III.B.6). This receptor is activated by oleoyl ethanolamide, an N-acylethanolamine with molecular properties similar to those of the endocannabinoid anandamide (). There is evidence that oleoyl ethanolamide has an important physiological role as a peripherally acting agent that reduces food intake () and this effect of oleoyl ethanolamide may be mediated, at least in part, by GPR119 (). Analysis of the sequence of GPR119 reveals that it shares structural similarity with adenosine receptors and amine receptors such as the 5-HT4 and 5-HT6 receptors, β1– and β2-adrenoceptors, and D1-type dopamine receptors. Furthermore, in a neighbor joining tree (Fig. 6) based on multiple sequence alignment, GPR119 is positioned within the branch of rhodopsin α group-type GPCRs that include cannabinoid CB1/CB2 receptors, lysophospholipid receptors, melanocortin receptors, and GPR3, GPR6, GPR12, and adenosine receptors. Thus, it seems that receptors for two different N-acylethanolamines, anandamide (CB1 and CB2) and oleoyl ethanolamide (GPR119), originated within this branch of GPCRs. This suggests that structural features characteristic of this branch of GPCRs may, perhaps, confer on them a propensity for ligand-binding associations with the N-acylethanolamine class of signaling molecules.

B. GPR55 and Other Rhodopsin δ Group-Type G Protein-Coupled Receptors

As discussed elsewhere in this review (section III.A), the concept that GPR55 is a “cannabinoid receptor” is controversial, and GPR55 may in fact be activated physiologically by lysophosphatidyl inositol (), which has been reported not to be a CB1 or CB2 receptor ligand (section II.C.5). The pharmacological properties of GPR55 will not be revisited here, but we will use this receptor as the starting point for a phylogenetic survey of “potential” cannabinoid receptors based on sequence similarity with GPR55. First, it is important to emphasize that GPR55 does not belong to the same group of rhodopsin-type GPCRs as CB1 and CB2; GPR55 falls within the δ group, whereas CB1 and CB2 receptors are in the α group, as discussed above. Therefore, if GPR55 is activated by endocannabinoids physiologically, then it must have acquired this property independently of the CB1/CB2-type cannabinoid receptors. The δ group of rhodopsin-type GPCRs comprises a heterogeneous collection of receptors, including MAS oncogene-related receptors, glycoprotein hormone receptors, purine receptors, and ∼460 odorant receptors ().

One of the receptors that shares a high level of sequence similarity with GPR55 is GPR35. However, there is no evidence that cannabinoids or other lipid signaling molecules activate this receptor. In fact it has been found that kynurenic acid, a product of tryptophan metabolism, acts as a GPR35 agonist (), although the physiological relevance of this finding remains unknown. It is noteworthy that among other receptors that share high levels of sequence similarity with GPR55 are receptors that have recently been identified as lysophospholipid receptors (Fig. 7). These include GPR23 and GPR92 (sections III.B.4 and III.B.5), which are activated by LPA and are now designated LPA4 () and LPA5 (), respectively, to distinguish them from the CB1/CB2-like LPA1-LPA3 receptors belonging to the α group of rhodopsin-type GPCRs. Thus, it seems that LPA receptors have evolved independently in both the α and δ branches of the rhodopsin family of GPCRs. GPR55, LPA4 (GPR23), and LPA5 (GPR92) belong to a group of closely related receptors that include P2-type purine receptors (e.g., P2Y1 and P2Y2) and a putative P2-like receptor originally designated P2Y5, which is in fact also activated by LPA and therefore has recently been designated as LPA6 (). Furthermore, the P2-like receptor originally designated P2Y10 has been identified as a lysophospholipid receptor that is activated by both S1P and LPA (). Another receptor that is closely related to P2Y1 is GPR174 (Fig. 7), but the ligand(s) that activate this receptor are not yet known.

An external file that holds a picture, illustration, etc. Object name is zpg0041022940007.jpg

Neighbor joining tree showing relationships between human GPR55 and other human rhodopsin δ-group type G protein-coupled receptors. The tree was generated using the multiple sequence alignment program ClustalX, with bootstrapping (1000 bootstrap trials), and viewed using NJ plot. The tree shows receptors that are discussed in section IV.B of this review.

It is interesting that GPR55 and the lysophospholipid receptors LPA4, LPA5, and LPA6 are closely related to P2-type purine receptors because, as discussed above, CB1/CB2-type cannabinoid receptors and LPA1-3 receptors are closely related to P1-type (adenosine) purine receptors. Thus, in different branches of the rhodopsin family of GPCRs (α and δ), lipid receptors that are activated by lysophospholipids or cannabinoids may have evolved independently from receptors that are activated by purines.

Other receptors in the δ group of rhodopsin-type GPCRs that are quite closely related to GPR55 and LPA4–6 are GPR17, GPR18 (section III.B.4), and GPR34 (Fig. 7). GPR17 is a P2Y-like receptor that is activated by both uracil nucleotides (e.g., UDP-glucose) and cysteinyl-leukotrienes (). GPR18, however, is of particular interest with respect to the endocannabinoid anandamide because it is activated by NAGly (), a lipoamino acid that also activates GPR92 (LPA5) (). Furthermore, GPR34 is activated by a different lipoamino acid: lysophosphatidyl-l-serine (). Finally, with lipid ligands as a recurring theme, there are GPR41 and GPR43 (section III.B.1 and Fig. 7), which are activated by short-chain fatty acids. Thus, within the branch of the δ group of rhodopsin-type GPCRs that include GPR55, there are a variety of related receptors that are activated by endocannabinoid/lysophospholipid-like molecules. However, it remains to be determined whether any of these receptors have pharmacological properties of the kind that could justify their classification as a “cannabinoid receptor.”

C. Conclusions

CB1/CB2-type cannabinoid receptors are phylogenetically restricted to the chordate branch of the animal kingdom. The lysophospholipid receptors S1P1, S1P2, S1P3, S1P4, S1P5, LPA1, LPA2, and LPA3 are the GPCRs that are most closely related to CB1/CB2-type receptors. Receptors for endocannabinoid/lysophospholipid-like molecules have evolved independently in different branches of the GPCR superfamily, but CB1 and CB2 are the only bona fide “cannabinoid receptors” that have been identified thus far.

V. Cannabinoid Receptor Nomenclature: CB or Not CB? That Is the Question

The terms “cannabinoid CB1” and “cannabinoid CB2” have been used throughout this review because this is the nomenclature that is currently recommended by the International Union of Basic and Clinical Pharmacology Committee on Receptor Nomenclature and Drug Classification and by its Subcommittee on Cannabinoid Receptors. It is noteworthy, however, that the adjective “cannabinoid” predates the discovery of cannabinoid receptors by many years and was originally coined to describe compounds, none of which is a structural analog of any known endocannabinoid. As a result, this term is widely used to describe not only all ligands that target CB1 or CB2 receptors but also other compounds with structures similar to the phytocannabinoid Δ9-THC, irrespective of whether they are or are not cannabinoid receptor agonists or antagonists. These additional cannabinoids include a number of nonpsychoactive compounds that are found in cannabis. The question arises, therefore, as to whether cannabinoid receptors should be renamed.

One possibility would be to rename these receptors after one of their endogenous agonists as is generally done for other receptors. However, selecting the most appropriate endocannabinoid for this purpose presents a formidable challenge. Thus, it is currently unclear whether anandamide, 2-AG, or some other established endocannabinoid should be selected for this purpose. Moreover, another as-yet-undiscovered endocannabinoid may emerge in the future as the most appropriate candidate for renaming cannabinoid receptors. Hence, for the present at least, this is probably not a sensible or viable option.

Another possibility would be to rename these receptors “endocannabinoid” receptors (for example, endocannabinoid CB1 and endocannabinoid CB2 receptors). Such nomenclature is of course tautologous, because all receptors are expected to have endogenous ligands. Nonetheless, it would remove the current confusion created by the term “cannabinoid” and yet, by retaining the terms “CB1” and “CB2,” would most likely not be the cause of any significant new confusion.

These and other options will be regularly considered by the International Union of Basic and Clinical Pharmacology Committee on Receptor Nomenclature and Drug Classification Subcommittee on Cannabinoid Receptors. In the meantime, it would be prudent to retain the present nomenclature for these receptors.

VI. Overall Conclusions and Future Directions

In conclusion, the endocannabinoid system seems to interact in a significant manner with several other endogenous systems. Thus, there is good evidence that CB1 receptors form heteromers with certain other GPCRs and that this heteromerization affects the manner in which the CB1 receptor responds to agonists (section II.D). In addition, it is generally accepted that at least some endocannabinoids, as well as Δ9-THC and several synthetic CB1/CB2 receptor agonists and antagonists, can interact with a number of established non-CB1, non-CB2 GPCRs, ligand-gated ion channels, ion channels, and nuclear receptors (PPARs) (section III). Of particular interest are channels or non-CB1, non-CB2 receptors that seem to be activated or blocked by some CB1/CB2 receptor ligands with potencies that differ little from those with which they target CB1 and/or CB2 receptors as agonists or antagonists. Examples include 1) enhancement of the activation of glycine receptors by anandamide, Δ9-THC, HU-210, and R-(+)-WIN55212 (section III.D.3); 2) the enhancement of NMDA-induced activation of NMDA receptors by anandamide (section III.D.4); 3) the inhibition of T-type voltage-gated calcium channels by anandamide and rimonabant (section III.F.1); and 4) the inhibition of voltage-gated KV3.1 and KV4.3 potassium channels and the activation of calcium-activated potassium (BK) channels by anandamide, 2-AG, and/or R-(+)-methanandamide (section III.F.2). Some channels may be targeted by CB1/CB2 receptor ligands allosterically. These are 5-HT3 and nicotinic acetylcholine receptors that can be potently activated by anandamide and certain other CB1/CB2 receptor ligands (sections III.D.1 and III.D.2). For some GPCRs (section III.C.3) and ligand-gated ion channels (section III.D), evidence that they can interact with CB1/CB2 receptor ligands has come solely from data obtained in radiolabeled ligand displacement experiments, prompting a need for further research directed at establishing whether such binding leads to receptor/channel activation or blockade. It should be noted, however, that CB1/CB2 receptor ligands such as anandamide, Δ9-THC, HU-210, 11-hydroxy-Δ8-THC, rimonabant, and taranabant did not display particularly high potency in these displacement experiments.

There seems to be no correlation between the ability of compounds to activate or block CB1 and/or CB2 receptors and their ability to target other receptors or channels. Moreover, some receptors and channels have been found to be activated by CB1/CB2 receptor antagonists or antagonized or inhibited by CB1/CB2 receptor agonists in a CB1/CB2 receptor independent manner. Examples include the antagonism of GPR55 and inhibition of 5-HT3 receptors and certain ion channels that has been observed in some experiments in response to CB1/CB2 receptor agonists (sections III.A.9., III.D.1, and III.F) and the activation of GPR55 and PPARγ (sections III.A.11., III.A.12 and III.G.1) that can apparently be induced by certain CB1 receptor antagonists in some bioassay systems. It should be borne in mind, therefore, that 1) some ligands that interact similarly with CB1 or CB2 receptors are likely to display significantly different pharmacological profiles in both some in vitro and in vivo bioassay systems and in the clinic and 2) a cannabinoid receptor antagonist might modulate the actions of a cannabinoid receptor agonist not only directly, through competitive antagonism or by inducing an inverse agonist effect at the cannabinoid receptor (sections II.C.3 and II.C.4), but also indirectly, by producing one or more CB1/CB2 receptor-independent effects. Clearly, therefore, a number of CB1/CB2 receptor agonists and antagonists together constitute a library of compounds, each with its own distinct “pharmacological fingerprint.” It is noteworthy, too, that the terms “CB1-selective” and “CB2-selective” are often used to indicate that a particular ligand activates or blocks one of these two receptors more potently than the other. However, because at least some ligands that can be described in this way interact no more potently with a CB1 or CB2 receptor than with one or more other type of receptor or channel (section III), any use of these terms should be accompanied by an appropriate definition or caveat (section I). It is also worth noting that anandamide seems to be the first endogenous molecule to have been found capable of activating certain GPCRs, ligand-gated ion channels, ion channels, and nuclear receptors and hence members of superfamilies that display negligible structural homology.

An important question to arise from the data presented in this review is whether any known mammalian channel or non-CB1/CB2 receptor should be classified as a novel cannabinoid “CB3” receptor or channel. We propose that any such receptor or channel should meet at least some of the following criteria:

  1. It should be activated at its orthosteric site and with significant potency by an established CB1/CB2 receptor ligand.

  2. It should be activated by at least one established endogenous CB1/CB2 receptor agonist at “physiologically relevant” concentrations.

  3. If it is a GPCR, it should display significant amino acid sequence similarity with the CB1 or the CB2 receptor, which are members of the α group of rhodopsin-type GPCRs.

  4. It should not be a “well established” non-CB1/CB2 receptor or channel, especially if there is already strong evidence that 1) this is activated endogenously by a non-CB1/CB2 receptor ligand with appropriate potency and relative intrinsic activity and 2) this is not activated endogenously by any endocannabinoid with appropriate potency and relative intrinsic activity.

  5. It should be expressed by mammalian cells that are known to be exposed to concentrations of endogenously released endocannabinoid molecules capable of eliciting a response.

Criterion 1 is not met by any receptors or channels that seem only to be potently targeted by CB1/CB2 receptor ligands in an allosteric manner. These probably include 5-HT3 and nicotinic acetylcholine receptors (section III.D). Criteria 1 and 2 are not met by established GPCRs or deorphanized GPCRs other than GPR55, because there is, at least currently, no evidence that an endogenous or synthetic CB1/CB2 receptor agonist activates any of these with significant potency (sections III.B.7 and III.C). In addition, criterion 2 does not seem to be met by PPARs, although it should be borne in mind that intracellular levels of established endogenous PPAR agonists can be very high (section III.G.8).

Criterion 3 is not met by GPR55 but is met by GPR3, GPR6, GPR12, and other non-CB1/CB2 members of the α group of rhodopsin-type GPCRs (section IV). However, there is no pharmacological evidence that any of these non-GPR55, deorphanized receptors behaves as a cannabinoid receptor (sections III.B.3 and IV.A). Moreover, many α group rhodopsin-type GPCRs (for example, α1-, α2-, and β-adrenoceptors and established 5-HT, dopamine, adenosine, melanocortin, sphingosine 1-phosphate, and lysophosphatidic acid receptors) are each excluded by criterion 4 from being a novel cannabinoid receptor. The case for considering GPR55 as a non-CB1/CB2 cannabinoid receptor is further weakened 1) by the finding that it can be potently activated by an endogenous, non-CB1/CB2 receptor agonist, LPI, and 2) by the inconsistent nature of much of the pharmacological data that have been generated to-date in GPR55 experiments with CB1/CB2 receptor ligands (section III.A). Criterion 4 is also not met by calcium, potassium, or sodium channels, ligand-gated ion channels, or nuclear receptors that are targeted by CB1/CB2 receptor ligands (sections III.D., III.F., and III.G). As to criterion 5, this is not met, at least at the present time, by the putative receptors discussed in section III.H. It is also noteworthy that at least some of these putative receptors seem to be activated by endogenous or synthetic CB1 or CB2 receptor agonists with rather low potency.

It is concluded that according to the five criteria listed in this section, no channel, non-CB1/CB2 established receptor or deorphanized receptor should currently be classified or reclassified as a novel cannabinoid receptor. It is noteworthy, however, that the TRPV1 channel does seem to meet criteria 1, 4, and 5, at least in part. Thus, 1) it is activated by anandamide at the capsaicin binding site, albeit with lower potency than the CB1 receptor, and also by N-arachidonoyl dopamine; 2) it is colocalized in some neurons with CB1 receptors and hence presumably exposed to endogenously produced endocannabinoids; and 3) anandamide was the first TRPV1 endogenous agonist to be identified (section III.E.2). It is also noteworthy that there is already evidence that TRPV1 channels are activated by endogenously released anandamide, at least in the presence of a FAAH inhibitor (section III.E.2). Clearly, there is a need for further research directed at exploring the possibility that the TRPV1 channel should be regarded as being either an “ionotropic cannabinoid CB3 receptor” or a dual TRPV1/CB3 channels. Such research should also investigate the hypothesis that the extent to which endogenously released anandamide activates TRPV1 channels increases under certain pathological conditions (section III.E.2). Increased activation of this kind might mean that the TRPV1 channel would also meet criterion 2 in disease if not in health. Thus, it might well be considered acceptable for the term “physiologically relevant” in criterion 2 to encompass concentrations of an endogenously released established endocannabinoid that activate a putative cannabinoid CB3 receptor only under pathological conditions. An alternative possibility (that anandamide should be classified as a dual endocannabinoid/endovanilloid) also merits further investigation. So too does the question of whether any other non-CB1/CB2 receptor/channel that displays significant sensitivity to an endocannabinoid is ever exposed to active concentrations of this endocannabinoid when it is released endogenously in the absence or presence of drugs that inhibit its metabolism or enhance its biosynthesis. Such research should perhaps focus initially on endocannabinoid-sensitive receptors or channels that are colocalized with CB1 or CB2 receptors. It will also be important to explore both the pharmacology of GPR55 more fully and the ability of CB1/CB2 receptor ligands to target other deorphanized receptors. The search for a cannabinoid CB3 receptor should and will continue.

Acknowledgments.

This work was supported in part by the Intramural Research Program of the National Institutes of Health National Institute on Alcohol Abuse and Alcoholism (to G.K.); by the National Institutes of Health National Institute on Drug Abuse [Grants DA009789 (to R.G.P., V.D., R.A.R., R.M.), DA021696 (to K.M.), DA023204 (to M.E.A.), DA03672 (to R.G.P., R.A.R.), DA03690 (to A.C.H.), DA03934 (to R.G.P., R.A.R.), DA05274 (to M.E.A.)]; and by funding from the UNIK program “Food, Fitness and Pharma” (to H.S.H.).

This article is available online at http://pharmrev.aspetjournals.org.

doi:10.1124/pr.110.003004.

 

1Abbreviations:

Δ9-THC
(−)-Δ9-tetrahydrocannabinol
2-AG
2-arachidonoyl glycerol
abn-CBD
abnormal-cannabidiol
ACEA
arachidonyl-2′-chloroethylamide
AM1241
R-3-(2-iodo-5-nitrobenzoyl)-1-methyl-2-piperidinylmethyl)-1H-indole
AM251
N-(piperidin-1-yl)-5-(4-iodophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboxamide
AM281
1-(2,4-dichlorophenyl)-5-(4-iodophenyl)-4-methyl-N-4-morpholinyl-1H-pyrazole-3-carboxamide
AM404
N-(4-hydroxyphenyl)-5Z,8Z,11Z,14Z-eicosatetraenamide
AM630
[6-iodo-2-methyl-1-[2-(4-morpholinyl)ethyl]-1H-indol-3-yl](4-methoxyphenyl)methanone (6-iodopravadoline)
CHO
Chinese hamster ovary
COX
cyclooxygenase
CP55940
(−)-cis-3-[2-hydroxy-4-(1,1-dimethylheptyl)phenyl]-trans-4-(3-hydroxypropyl)cyclohexanol
CT-3
ajulemic acid
DAMGO
[d-Ala2,N-Me-Phe4,Gly5-ol]-enkephalin
DRG
dorsal root ganglion
ERK
extracellular-signal regulated kinase
FAAH
fatty acid amide hydrolase
FRET
fluorescence resonance energy transfer
FTY720
2-amino-2-(2-[4-octylphenyl]ethyl)-1,3-propanediol
GLP-1
glucagon-like peptide-1
GPCR
G protein-coupled receptor
GTPγS
guanosine 5′-O-(3-thio)triphosphate
GW6471
[(2S)-2-[[(1Z)-1-methyl-3-oxo-3-[4-(trifluoromethyl)phenyl)-1-propenyl]amino]-3-[4-[2-(5-methyl-2-phenyl-4-oxazolyl)ethoxyphenylpropyl]-carbamic acid ethyl ester
GW7647
2-[[4-[2-[[(cyclohexylamino)carbonyl](4-cyclohexylbutyl)amino] ethyl]-phenyl]thio]-2-methylpropanoic acid
GW9662
2-chloro-5-nitro-N-phenylbenzamide
HEK
human embryonic kidney
HU-210
(6aR)-trans-3-(1,1-dimethylheptyl)-6a,7,10,10a-tetrahydro-1-hydroxy-6,6-dimethyl-6H-dibenzo[b,d]pyran-9-methanol
HU-243
[3H](6aR)-trans-3-(1,1-dimethylheptyl)-6a,7,10,10a-tetrahydro-1-hydroxy-6,6-dimethyl-6H-dibenzo[b,d]pyran-9-methanol
HU-308
{4-[4-(1,1-dimethylheptyl)-2,6-dimethoxy-phenyl]-6,6-dimethyl-bicyclo[3.1.1] hept-2-en-2-yl}-methanol
HU-331
3S,4R-p-benzoquinone-3-hydroxy-2-p-mentha-(1,8)-dien-3-yl-5-pentyl
JTE-907
N-(1,3-benzodioxol-5-ylmethyl)-1,2-dihydro-7-methoxy-2-oxo-8-(pentyloxy)-3-quinolinecarboxamide
JWH-015
(2-methyl-1-propyl-1H-indol-3-yl)-1-naphthalenylmethanone
JWH-133
(6aR,10aR)-3-(1,1-dimethylbutyl)-6a,7,10,10a-tetrahydro-6,6,9-trimethyl-6H-dibenzo [b,d]pyran
LOX
lipoxygenase
LPA
lysophosphatidic acid
LPI
lysophosphatidyl inositol
LY320135
4-[[6-methoxy-2-(4-methoxyphenyl)-3-benzofuranyl]carbonyl]benzonitrile
MAPK
mitogen-activated protein kinase
MK886
1-[(4-chlorophenyl)methyl]-3-[(1,1-dimethylethyl)thio]-α,α-dimethyl-5-(1-methylethyl)-1H-indole-2-propanoic acid
NAGly
N-arachidonoyl glycine
NESS O327
N-piperidinyl-[8-chloro-1-(2,4-dichlorophenyl)-1,4,5,6-tetrahydrobenzo[6,7]cyclohepta[1,2-c]pyrazole-3-carboxamide]
NMDA
N-methyl-d-aspartate
noladin ether
2-arachidonyl glyceryl ether
O-1602
5-methyl-4-[(1R,6R)-3-methyl-6-(prop-1-en-2-yl)cyclohex-2-enyl]benzene-1,3-diol
O-1918
1,3-dimethoxy-5-methyl-2-[(1R,6R)-3-methyl-6-(1-methylethenyl)-2-cyclohexen-1-yl]benzene
O-2050
(6aR,10aR)-3-(1-methanesulfonylamino-4-hexyn-6-yl)-6a,7,10,10a-tetrahydro-6,6,9-trimethyl-6H-dibenzo[b,d]pyran
OX
orexin
PPAR
peroxisome proliferator-activated receptor
R-(+)-WIN55212
R-(+)-[2,3-dihydro-5-methyl-3-(4-morpholinylmethyl)pyrrolo[1,2,3-de]-1,4-benzoxazin-6-yl]-1-naphthalenylmethanone mesylate
RhoA
Ras homolog gene family, member A
Sch.225336
N-[1(S)-[4-[[4-methoxy-2-[(4-methoxyphenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]methanesulfonamide)
Sch.356036
N-[1(S)-[4-[[4-chloro-2-[(2-fluorophenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]methanesulfonamide
Sch.414319
N-[1(S)-[4-[[4-chloro-2-[(2-fluorophenyl)sulfonyl]phenyl]sulfonyl]phenyl]ethyl]-1,1,1-trifluromethanesulfonamide
siRNA
small interfering RNA
SKF97541
3-aminopropyl(methyl)phosphinic acid
SR141716A
rimonabant
SR144528
N-((1S)-endo-1,3,3-trimethyl bicyclo heptan-2-yl]-5-(4-chloro-3-methylphenyl)-1-(4-methylbenzyl)-pyrazole-3-carboxamide)
T0070907
2-chloro-5-nitro-N-4-pyridinyl-benzamide
TM
transmembrane
TRP
transient receptor potential
TRPA
transient receptor potential ankyrin
TRPM
transient receptor potential melastatin
UCM707
(5Z,8Z,11Z,14Z)-N-(3-furanylmethyl)-5,8,11,14-eicosatetraenamide
URB597
cyclohexylcarbamic acid 3′-carbamoyl-biphenyl-3-yl ester.

 

References

  • Abadji V, Lin S, Taha G, Griffin G, Stevenson LA, Pertwee RG, Makriyannis A. (1994) (R)-methanandamide: a chiral novel anandamide possessing higher potency and metabolic stabilityJ Med Chem 37:1889–1893 [PubMed[]
  • Adam L, Salois D, Rihakova L, Lapointe S, St-Onge S, Labrecque J, Payza K. (2007) Positive allosteric modulators of CB1 receptors, in 17th Annual Symposium of the Cannabinoids; 2007 Jun 26–Jul 1; St-Sauveur, Canada p. 86 International Cannabinoid Research Society, Burlington, Vermont []
  • Ahluwalia J, Urban L, Bevan S, Nagy I. (2003a) Anandamide regulates neuropeptide release from capsaicin-sensitive primary sensory neurons by activating both the cannabinoid 1 receptor and the vanilloid receptor 1 in vitroEur J Neurosci 17:2611–2618 [PubMed[]
  • Ahluwalia J, Yaqoob M, Urban L, Bevan S, Nagy I. (2003b) Activation of capsaicin-sensitive primary sensory neurones induces anandamide production and releaseJ Neurochem 84:585–591 [PubMed[]
  • Akiyama TE, Sakai S, Lambert G, Nicol CJ, Matsusue K, Pimprale S, Lee YH, Ricote M, Glass CK, Brewer HB, Jr, et al. (2002) Conditional disruption of the peroxisome proliferator-activated receptor gamma gene in mice results in lowered expression of ABCA1, ABCG1, and apoE in macrophages and reduced cholesterol effluxMol Cell Biol 22:2607–2619 [PMC free article] [PubMed[]
  • Akopian AN, Ruparel NB, Jeske NA, Patwardhan A, Hargreaves KM. (2009) Role of ionotropic cannabinoid receptors in peripheral antinociception and antihyperalgesiaTrends Pharmacol Sci 30:79–84 [PMC free article] [PubMed[]
  • Akopian AN, Ruparel NB, Patwardhan A, Hargreaves KM. (2008) Cannabinoids desensitize capsaicin and mustard oil responses in sensory neurons via TRPA1 activationJ Neurosci 28:1064–1075 [PMC free article] [PubMed[]
  • Alexander SP, Mathie A, Peters JA. (2009) Guide to Receptors and Channels (GRAC), 4th EditionBr J Pharmacol 158 (Suppl 1):S1–S254 [PMC free article] [PubMed[]
  • Alvarado M, Goya P, Macías-González M, Pavón FJ, Serrano A, Jagerovic N, Elguero J, Gutiérrez-Rodríguez A, García-Granda S, Suardíaz M, et al. (2008) Antiobesity designed multiple ligands: synthesis of pyrazole fatty acid amides and evaluation as hypophagic agentsBioorg Med Chem 16:10098–10105 [PubMed[]
  • Amorós I, Barana A, Caballero R, Gómez R, Osuna L, Lillo MP, Tamargo J, Delpón E. (2010) Endocannabinoids and cannabinoid analogues block human cardiac KV4.3 channels in a receptor-independent mannerJ Mol Cell Cardiol 48:201–210 [PubMed[]
  • Anand U, Otto WR, Sanchez-Herrera D, Facer P, Yiangou Y, Korchev Y, Birch R, Benham C, Bountra C, Chessell IP, et al. (2008) Cannabinoid receptor CB2 localisation and agonist-mediated inhibition of capsaicin responses in human sensory neuronsPain 138:667–680 [PubMed[]
  • Andersson DA, Adner M, Högestätt ED, Zygmunt PM. (2002) Mechanisms underlying tissue selectivity of anandamide and other vanilloid receptor agonistsMol Pharmacol 62:705–713 [PubMed[]
  • Andersson M, Usiello A, Borgkvist A, Pozzi L, Dominguez C, Fienberg AA, Svenningsson P, Fredholm BB, Borrelli E, Greengard P, et al. (2005) Cannabinoid action depends on phosphorylation of dopamine- and cAMP-regulated phosphoprotein of 32 kDa at the protein kinase A site in striatal projection neuronsJ Neurosci 25:8432–8438 [PMC free article] [PubMed[]
  • Appendino G, Ligresti A, Minassi A, Cascio MG, Allarà M, Taglialatela-Scafati O, Pertwee RG, De Petrocellis L, Di Marzo V. (2009) Conformationally constrained fatty acid ethanolamides as cannabinoid and vanilloid receptor probesJ Med Chem 52:3001–3009 [PubMed[]
  • Artmann A, Petersen G, Hellgren LI, Boberg J, Skonberg C, Nellemann C, Hansen SH, Hansen HS. (2008) Influence of dietary fatty acids on endocannabinoid and N-acylethanolamine levels in rat brain, liver and small intestineBiochim Biophys Acta 1781:200–212 [PubMed[]
  • Bab I, Zimmer A, Melamed E. (2009) Cannabinoids and the skeleton: from marijuana to reversal of bone lossAnn Med 41:560–567 [PubMed[]
  • Baek JH, Zheng Y, Darlington CL, Smith PF. (2008) Cannabinoid CB2 receptor expression in the rat brainstem cochlear and vestibular nucleiActa Otolaryngol 128:961–967 [PubMed[]
  • Barana A, Amorós I, Caballero R, Gómez R, Osuna L, Lillo MP, Blázquez C, Guzmán M, Delpón E, Tamargo J. (2010) Endocannabinoids and cannabinoid analogues block cardiac hKV1.5 channels in a cannabinoid receptor-independent mannerCardiovasc Res 85:56–67 [PubMed[]
  • Barann M, Molderings G, Brüss M, Bönisch H, Urban BW, Göthert M. (2002) Direct inhibition by cannabinoids of human 5-HT3A receptors: probable involvement of an allosteric modulatory siteBr J Pharmacol 137:589–596 [PMC free article] [PubMed[]
  • Barbara G, Alloui A, Nargeot J, Lory P, Eschalier A, Bourinet E, Chemin J. (2009) T-type calcium channel inhibition underlies the analgesic effects of the endogenous lipoamino acidsJ Neurosci 29:13106–13114 [PMC free article] [PubMed[]
  • Beltramo M, Bernardini N, Bertorelli R, Campanella M, Nicolussi E, Fredduzzi S, Reggiani A. (2006) CB2 receptor-mediated antihyperalgesia: possible direct involvement of neural mechanismsEur J Neurosci 23:1530–1538 [PubMed[]
  • Benninger F, Freund TF, Hájos N. (2008) Control of excitatory synaptic transmission by capsaicin is unaltered in TRPV1 vanilloid receptor knockout miceNeurochem Int 52:89–94 [PMC free article] [PubMed[]
  • Berg AP, Talley EM, Manger JP, Bayliss DA. (2004) Motoneurons express heteromeric TWIK-related acid-sensitive K+ (TASK) channels containing TASK-1 (KCNK3) and TASK-3 (KCNK9) subunitsJ Neurosci 24:6693–6702 [PMC free article] [PubMed[]
  • Bharate SB, Nemmani KV, Vishwakarma RA. (2009) Progress in the discovery and development of small-molecule modulators of G-protein-coupled receptor 40 (GPR40/FFA1/FFAR1): an emerging target for type 2 diabetesExpert Opin Ther Pat 19:237–264 [PubMed[]
  • Bidaut-Russell M, Howlett AC. (1991) Cannabinoid receptor-regulated cyclic AMP accumulation in the rat striatumJ Neurochem 57:1769–1773 [PubMed[]
  • Bisogno T, Hanus L, De Petrocellis L, Tchilibon S, Ponde DE, Brandi I, Moriello AS, Davis JB, Mechoulam R, Di Marzo V. (2001) Molecular targets for cannabidiol and its synthetic analogues: effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of anandamideBr J Pharmacol 134:845–852 [PMC free article] [PubMed[]
  • Bisogno T, Melck D, Bobrov MYu, Gretskaya NM, Bezuglov VV, De Petrocellis L, Di Marzo V. (2000) N-Acyl-dopamines: novel synthetic CB1 cannabinoid-receptor ligands and inhibitors of anandamide inactivation with cannabimimetic activity in vitro and in vivoBiochem J 351:817–824 [PMC free article] [PubMed[]
  • Bjursell M, Gerdin AK, Jönsson M, Surve VV, Svensson L, Huang XF, Törnell J, Bohlooly-Y M. (2006) G protein-coupled receptor 12 deficiency results in dyslipidemia and obesity in miceBiochem Biophys Res Commun 348:359–366 [PubMed[]
  • Borgkvist A, Marcellino D, Fuxe K, Greengard P, Fisone G. (2008) Regulation of DARPP-32 phosphorylation by Δ9-tetrahydrocannabinolNeuropharmacology 54:31–35 [PubMed[]
  • Bouaboula M, Hilairet S, Marchand J, Fajas L, Le Fur G, Casellas P. (2005) Anandamide induced PPARγ transcriptional activation and 3T3–L1 preadipocyte differentiationEur J Pharmacol 517:174–181 [PubMed[]
  • Bouchard C, Pagé J, Bédard A, Tremblay P, Vallières L. (2007) G protein-coupled receptor 84, a microglia-associated protein expressed in neuroinflammatory conditionsGlia 55:790–800 [PubMed[]
  • Bradshaw HB, Walker JM. (2005) The expanding field of cannabimimetic and related lipid mediatorsBr J Pharmacol 144:459–465 [PMC free article] [PubMed[]
  • Breivogel CS, Childers SR. (2000) Cannabinoid agonist signal transduction in rat brain: comparison of cannabinoid agonists in receptor binding, G-protein activation, and adenylyl cyclase inhibitionJ Pharmacol Exp Ther 295:328–336 [PubMed[]
  • Breivogel CS, Griffin G, Di Marzo V, Martin BR. (2001) Evidence for a new G protein-coupled cannabinoid receptor in mouse brainMol Pharmacol 60:155–163 [PubMed[]
  • Briscoe CP, Tadayyon M, Andrews JL, Benson WG, Chambers JK, Eilert MM, Ellis C, Elshourbagy NA, Goetz AS, Minnick DT, et al. (2003) The orphan G protein-coupled receptor GPR40 is activated by medium and long chain fatty acidsJ Biol Chem 278:11303–11311 [PubMed[]
  • Brown AJ, Goldsworthy SM, Barnes AA, Eilert MM, Tcheang L, Daniels D, Muir AI, Wigglesworth MJ, Kinghorn I, Fraser NJ, et al. (2003) The orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acidsJ Biol Chem 278:11312–11319 [PubMed[]
  • Butt C, Alptekin A, Shippenberg T, Oz M. (2008) Endogenous cannabinoid anandamide inhibits nicotinic acetylcholine receptor function in mouse thalamic synaptosomesJ Neurochem 105:1235–1243 [PubMed[]
  • Cabral GA, Staab A. (2005) Effects on the immune systemHandb Exp Pharmacol 168:385–423 [PubMed[]
  • Calandra B, Portier M, Kernéis A, Delpech M, Carillon C, Le Fur G, Ferrara P, Shire D. (1999) Dual intracellular signaling pathways mediated by the human cannabinoid CB1 receptorEur J Pharmacol 374:445–455 [PubMed[]
  • Canals M, Milligan G. (2008) Constitutive activity of the cannabinoid CB1 receptor regulates the function of co-expressed mu opioid receptorsJ Biol Chem 283:11424–11434 [PubMed[]
  • Carriba P, Navarro G, Ciruela F, Ferré S, Casadó V, Agnati L, Cortés A, Mallol J, Fuxe K, Canela EI, et al. (2008) Detection of heteromerization of more than two proteins by sequential BRET-FRETNat Methods 5:727–733 [PubMed[]
  • Carriba P, Ortiz O, Patkar K, Justinova Z, Stroik J, Themann A, Müller C, Woods AS, Hope BT, Ciruela F, et al. (2007) Striatal adenosine A2A and cannabinoid CB1 receptors form functional heteromeric complexes that mediate the motor effects of cannabinoidsNeuropsychopharmacology 32:2249–2259 [PubMed[]
  • Cascio MG, Gauson LA, Stevenson LA, Ross RA, Pertwee RG. (2010) Evidence that the plant cannabinoid cannabigerol is a highly potent α2-adrenoceptor agonist and moderately potent 5HT1A receptor antagonistBr J Pharmacol 159:129–141 [PMC free article] [PubMed[]
  • Caterina MJ, Leffler A, Malmberg AB, Martin WJ, Trafton J, Petersen-Zeitz KR, Koltzenburg M, Basbaum AI, Julius D. (2000) Impaired nociception and pain sensation in mice lacking the capsaicin receptorScience 288:306–313 [PubMed[]
  • Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. (1997) The capsaicin receptor: a heat-activated ion channel in the pain pathwayNature 389:816–824 [PubMed[]
  • Cheer JF, Cadogan AK, Marsden CA, Fone KC, Kendall DA. (1999) Modification of 5-HT2 receptor mediated behaviour in the rat by oleamide and the role of cannabinoid receptorsNeuropharmacology 38:533–541 [PubMed[]
  • Chemin J, Monteil A, Perez-Reyes E, Nargeot J, Lory P. (2001) Direct inhibition of T-type calcium channels by the endogenous cannabinoid anandamideEMBO J 20:7033–7040 [PMC free article] [PubMed[]
  • Chemin J, Nargeot J, Lory P. (2007) Chemical determinants involved in anandamide-induced inhibition of T-type calcium channelsJ Biol Chem 282:2314–2323 [PubMed[]
  • Chen JK, Chen J, Imig JD, Wei S, Hachey DL, Guthi JS, Falck JR, Capdevila JH, Harris RC. (2008) Identification of novel endogenous cytochrome p450 arachidonate metabolites with high affinity for cannabinoid receptorsJ Biol Chem 283:24514–24524 [PMC free article] [PubMed[]
  • Childers SR, Pacheco MA, Bennett BA, Edwards TA, Hampson RE, Mu J, Deadwyler SA. (1993) Cannabinoid receptors: G-protein-mediated signal transduction mechanismsBiochem Soc Symp 59:27–50 [PubMed[]
  • Christopoulos A, Wilson K. (2001) Interaction of anandamide with the M1 and M4 muscarinic acetylcholine receptorsBrain Res 915:70–78 [PubMed[]
  • Chu ZL, Carroll C, Chen R, Alfonso J, Gutierrez V, He H, Lucman A, Xing C, Sebring K, Zhou J, et al. (2010) N-oleoyldopamine enhances glucose homeostasis through the activation of GPR119Mol Endocrinol 24:161–170 [PMC free article] [PubMed[]
  • Chu ZL, Jones RM, He H, Carroll C, Gutierrez V, Lucman A, Moloney M, Gao H, Mondala H, Bagnol D, et al. (2007) A role for beta-cell-expressed G protein-coupled receptor 119 in glycemic control by enhancing glucose-dependent insulin releaseEndocrinology 148:2601–2609 [PubMed[]
  • Chun J, Goetzl EJ, Hla T, Igarashi Y, Lynch KR, Moolenaar W, Pyne S, Tigyi G. (2002) International Union of Pharmacology. XXXIV. Lysophospholipid receptor nomenclaturePharmacol Rev 54:265–269 [PubMed[]
  • Ciana P, Fumagalli M, Trincavelli ML, Verderio C, Rosa P, Lecca D, Ferrario S, Parravicini C, Capra V, Gelosa P, et al. (2006) The orphan receptor GPR17 identified as a new dual uracil nucleotides/cysteinyl-leukotrienes receptorEMBO J 25:4615–4627 [PMC free article] [PubMed[]
  • Cinar R, Freund TF, Katona I, Mackie K, Szucs M. (2008) Reciprocal inhibition of G-protein signaling is induced by CB1 cannabinoid and GABAB receptor interactions in rat hippocampal membranesNeurochem Int 52:1402–1409 [PubMed[]
  • Cinar R, Szücs M. (2009) CB1 receptor-independent actions of SR141716 on G-protein signaling: coapplication with the μ-opioid agonist Tyr-D-AlaGly-(NMe)Phe-Gly-ol unmasks novel, pertussis toxin-insensitive opioid signaling in μ-opioid receptor-Chinese hamster ovary cellsJ Pharmacol Exp Ther 330:567–574 [PubMed[]
  • Cluny NL, Keenan CM, Lutz B, Piomelli D, Sharkey KA. (2009) The identification of peroxisome proliferator-activated receptor alpha-independent effects of oleoylethanolamide on intestinal transit in miceNeurogastroenterol Motil 21:420–429 [PubMed[]
  • Cristino L, de Petrocellis L, Pryce G, Baker D, Guglielmotti V, Di Marzo V. (2006) Immunohistochemical localization of cannabinoid type 1 and vanilloid transient receptor potential vanilloid type 1 receptors in the mouse brainNeuroscience 139:1405–1415 [PubMed[]
  • D’Agostino G, La Rana G, Russo R, Sasso O, Iacono A, Esposito E, Raso GM, Cuzzocrea S, Lo Verme J, Piomelli D, et al. (2007) Acute intracerebroventricular administration of palmitoylethanolamide, an endogenous PPAR-α agonist, modulates carrageenan-Induced paw edema in miceJ Pharmacol Exp Ther 322:1137–1143 [PubMed[]
  • Davis JB, Gray J, Gunthorpe MJ, Hatcher JP, Davey PT, Overend P, Harries MH, Latcham J, Clapham C, Atkinson K, et al. (2000) Vanilloid receptor-1 is essential for inflammatory thermal hyperalgesiaNature 405:183–187 [PubMed[]
  • De Jesús ML, Sallés J, Meana JJ, Callado LF. (2006) Characterization of CB1 cannabinoid receptor immunoreactivity in postmortem human brain homogenatesNeuroscience 140:635–643 [PubMed[]
  • De Petrocellis L, Bisogno T, Davis JB, Pertwee RG, Di Marzo V. (2000) Overlap between the ligand recognition properties of the anandamide transporter and the VR1 vanilloid receptor: inhibitors of anandamide uptake with negligible capsaicin-like activityFEBS Lett 483:52–56 [PubMed[]
  • De Petrocellis L, Bisogno T, Maccarrone M, Davis JB, Finazzi-Agro A, Di Marzo V. (2001) The activity of anandamide at vanilloid VR1 receptors requires facilitated transport across the cell membrane and is limited by intracellular metabolismJ Biol Chem 276:12856–12863 [PubMed[]
  • De Petrocellis L, Di Marzo V. (2009) Role of endocannabinoids and endovanilloids in Ca2+ signallingCell Calcium 45:611–624 [PubMed[]
  • De Petrocellis L, Starowicz K, Moriello AS, Vivese M, Orlando P, Di Marzo V. (2007) Regulation of transient receptor potential channels of melastatin type 8 (TRPM8): effect of cAMP, cannabinoid CB1 receptors and endovanilloidsExp Cell Res 313:1911–1920 [PubMed[]
  • De Petrocellis L, Vellani V, Schiano-Moriello A, Marini P, Magherini PC, Orlando P, Di Marzo V. (2008) Plant-derived cannabinoids modulate the activity of transient receptor potential channels of ankyrin type-1 and melastatin type-8J Pharmacol Exp Ther 325:1007–1015 [PubMed[]
  • Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D, Mandelbaum A, Etinger A, Mechoulam R. (1992) Isolation and structure of a brain constituent that binds to the cannabinoid receptorScience 258:1946–1949 [PubMed[]
  • Di Marzo V, Breivogel CS, Tao Q, Bridgen DT, Razdan RK, Zimmer AM, Zimmer A, Martin BR. (2000) Levels, metabolism, and pharmacological activity of anandamide in CB1 cannabinoid receptor knockout mice: evidence for non-CB1, non-CB2 receptor-mediated actions of anandamide in mouse brainJ Neurochem 75:2434–2444 [PubMed[]
  • Di Marzo V, Cristino L. (2008) Why endocannabinoids are not all alikeNat Neurosci 11:124–126 [PubMed[]
  • Di Marzo V, De Petrocellis L, Bisogno T. (2005) The biosynthesis, fate and pharmacological properties of endocannabinoidsHandb Exp Pharmacol 168:147–185 [PubMed[]
  • Di Marzo V, De Petrocellis L, Fezza F, Ligresti A, Bisogno T. (2002) Anandamide receptorsProstaglandins Leukot Essent Fatty Acids 66:377–391 [PubMed[]
  • Dolezelova E, Nothacker HP, Civelli O, Bryant PJ, Zurovec M. (2007) Drosophila adenosine receptor activates cAMP and calcium signalingInsect Biochem Mol Biol 37:318–329 [PubMed[]
  • Duan Y, Liao C, Jain S, Nicholson RA. (2008a) The cannabinoid receptor agonist CP-55,940 and ethyl arachidonate interfere with [3H]batrachotoxinin A 20 α-benzoate binding to sodium channels and inhibit sodium channel functionComp Biochem Physiol C Toxicol Pharmacol 148:244–249 [PubMed[]
  • Duan Y, Zheng J, Nicholson RA. (2008b) Inhibition of [3H]batrachotoxinin A-20α-benzoate binding to sodium channels and sodium channel function by endocannabinoidsNeurochem Int 52:438–446 [PubMed[]
  • Dubi N, Gheber L, Fishman D, Sekler I, Hershfinkel M. (2008) Extracellular zinc and zinc-citrate, acting through a putative zinc-sensing receptor, regulate growth and survival of prostate cancer cellsCarcinogenesis 29:1692–1700 [PubMed[]
  • Dyson A, Peacock M, Chen A, Courade JP, Yaqoob M, Groarke A, Brain C, Loong Y, Fox A. (2005) Antihyperalgesic properties of the cannabinoid CT-3 in chronic neuropathic and inflammatory pain states in the ratPain 116:129–137 [PubMed[]
  • Eichele K, Ramer R, Hinz B. (2009) R(+)-Methanandamide-induced apoptosis of human cervical carcinoma cells involves a cyclooxygenase-2-dependent pathwayPharm Res 26:346–355 [PubMed[]
  • Ellis J, Pediani JD, Canals M, Milasta S, Milligan G. (2006) Orexin-1 receptor-cannabinoid CB1 receptor heterodimerization results in both ligand-dependent and -independent coordinated alterations of receptor localization and functionJ Biol Chem 281:38812–38824 [PubMed[]
  • Elphick MR. (2002) Evolution of cannabinoid receptors in vertebrates: identification of a CB2 gene in the puffer fish Fugu rubripesBiol Bull 202:104–107 [PubMed[]
  • Elphick MR. (2007) BfCBR: a cannabinoid receptor ortholog in the cephalochordate Branchiostoma floridae (Amphioxus)Gene 399:65–71 [PubMed[]
  • Elphick MR, Egertová M. (2001) The neurobiology and evolution of cannabinoid signallingPhilos Trans R Soc Lond B Biol Sci 356:381–408 [PMC free article] [PubMed[]
  • Elphick MR, Egertová M. (2005) The phylogenetic distribution and evolutionary origins of endocannabinoid signallingHandb Exp Pharmacol 168:283–297 [PubMed[]
  • Elphick MR, Satou Y, Satoh N. (2003) The invertebrate ancestry of endocannabinoid signalling: an orthologue of vertebrate cannabinoid receptors in the urochordate Ciona intestinalisGene 302:95–101 [PubMed[]
  • Evans RM, Scott RH, Ross RA. (2004) Multiple actions of anandamide on neonatal rat cultured sensory neuronesBr J Pharmacol 141:1223–1233 [PMC free article] [PubMed[]
  • Everaerts W, Nilius B, Owsianik G. (2010) The vanilloid transient receptor potential channel TRPV4: from structure to diseaseProg Biophys Mol Biol 103:2–17 [PubMed[]
  • Facci L, Dal Toso R, Romanello S, Buriani A, Skaper SD, Leon A. (1995) Mast cells express a peripheral cannabinoid receptor with differential sensitivity to anandamide and palmitoylethanolamideProc Natl Acad Sci USA 92:3376–3380 [PMC free article] [PubMed[]
  • Fan P. (1995) Cannabinoid agonists inhibit the activation of 5-HT3 receptors in rat nodose ganglion neuronsJ Neurophysiol 73:907–910 [PubMed[]
  • Fang X, Hu S, Xu B, Snyder GD, Harmon S, Yao J, Liu Y, Sangras B, Falck JR, Weintraub NL, et al. (2006) 4,15-Dihydroxyeicosatrienoic acid activates peroxisome proliferator-activated receptor-αAm J Physiol Heart Circ Physiol 290:H55–H63 [PubMed[]
  • Ferré S, Baler R, Bouvier M, Caron MG, Devi LA, Durroux T, Fuxe K, George SR, Javitch JA, Lohse MJ, et al. (2009a) Building a new conceptual framework for receptor heteromersNat Chem Biol 5:131–134 [PMC free article] [PubMed[]
  • Ferré S, Goldberg SR, Lluis C, Franco R. (2009b) Looking for the role of cannabinoid receptor heteromers in striatal functionNeuropharmacology 56 (Suppl 1):226–234 [PMC free article] [PubMed[]
  • Fioravanti B, De Felice M, Stucky CL, Medler KA, Luo MC, Gardell LR, Ibrahim M, Malan TP, Jr, Yamamura HI, Ossipov MH, et al. (2008) Constitutive activity at the cannabinoid CB1 receptor is required for behavioral response to noxious chemical stimulation of TRPV1: antinociceptive actions of CB1 inverse agonistsJ Neurosci 28:11593–11602 [PMC free article] [PubMed[]
  • Fisyunov A, Tsintsadze V, Min R, Burnashev N, Lozovaya N. (2006) Cannabinoids modulate the P-type high-voltage-activated calcium currents in Purkinje neuronsJ Neurophysiol 96:1267–1277 [PubMed[]
  • Fong TM, Guan XM, Marsh DJ, Shen CP, Stribling DS, Rosko KM, Lao J, Yu H, Feng Y, Xiao JC, et al. (2007) Antiobesity efficacy of a novel cannabinoid-1 receptor inverse agonist, N-[(1S,2S)-3-(4-chlorophenyl)-2-(3-cyanophenyl)-1-methylpropyl]-2-methyl-2-{[5-(trifluoromethyl)pyridin-2-yl]oxy}propanamide (MK-0364), in rodentsJ Pharmacol Exp Ther 321:1013–1022 [PubMed[]
  • Fong TM, Shearman LP, Stribling DS, Shu J, Lao J, Huang CR, Xiao JC, Shen CP, Tyszkiewicz J, Strack AM, et al. (2009) Pharmacological efficacy and safety profile of taranabant in preclinical speciesDrug Dev Res 70:349–362 []
  • Forman BM, Chen J, Evans RM. (1997) Hypolipidemic drugs, polyunsaturated fatty acids, and eicosanoids are ligands for peroxisome proliferator-activated receptors α and δProc Natl Acad Sci USA 94:4312–4317 [PMC free article] [PubMed[]
  • Fredriksson R, Höglund PJ, Gloriam DE, Lagerström MC, Schiöth HB. (2003a) Seven evolutionarily conserved human rhodopsin G protein-coupled receptors lacking close relativesFEBS Lett 554:381–388 [PubMed[]
  • Fredriksson R, Lagerström MC, Lundin LG, Schiöth HB. (2003b) The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprintsMol Pharmacol 63:1256–1272 [PubMed[]
  • Fu H, Xiao JM, Cao XH, Ming ZY, Liu LJ. (2008) Effects of WIN55,212–2 on voltage-gated sodium channels in trigeminal ganglion neurons of ratsNeurol Res 30:85–91 [PubMed[]
  • Fu J, Gaetani S, Oveisi F, Lo Verme J, Serrano A, Rodríguez De Fonseca F, Rosengarth A, Luecke H, Di Giacomo B, Tarzia G, et al. (2003) Oleylethanolamide regulates feeding and body weight through activation of the nuclear receptor PPAR-αNature 425:90–93 [PubMed[]
  • Fuxe K, Marcellino D, Rivera A, Diaz-Cabiale Z, Filip M, Gago B, Roberts DC, Langel U, Genedani S, Ferraro L, et al. (2008) Receptor-receptor interactions within receptor mosaics. Impact on neuropsychopharmacologyBrain Res Rev 58:415–452 [PubMed[]
  • Gantz I, Muraoka A, Yang YK, Samuelson LC, Zimmerman EM, Cook H, Yamada T. (1997) Cloning and chromosomal localization of a gene (GPR18) encoding a novel seven transmembrane receptor highly expressed in spleen and testisGenomics 42:462–466 [PubMed[]
  • Ghosh M, Wang H, Ai Y, Romeo E, Luyendyk JP, Peters JM, Mackman N, Dey SK, Hla T. (2007) COX-2 suppresses tissue factor expression via endocannabinoid-directed PPARδ activationJ Exp Med 204:2053–2061 [PMC free article] [PubMed[]
  • Giuliano M, Pellerito O, Portanova P, Calvaruso G, Santulli A, De Blasio A, Vento R, Tesoriere G. (2009) Apoptosis induced in HepG2 cells by the synthetic cannabinoid WIN: involvement of the transcription factor PPARγBiochimie 91:457–465 [PubMed[]
  • Glass M, Felder CC. (1997) Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors augments cAMP accumulation in striatal neurons: evidence for a Gs linkage to the CB1 receptorJ Neurosci 17:5327–5333 [PMC free article] [PubMed[]
  • Godlewski G, Offertáler L, Osei-Hyiaman D, Mo FM, Harvey-White J, Liu J, Davis MI, Zhang L, Razdan RK, Milman G, et al. (2009) The endogenous brain constituent N-arachidonoyl L-serine is an activator of large conductance Ca2+-activated K+ channelsJ Pharmacol Exp Ther 328:351–361 [PMC free article] [PubMed[]
  • Gong JP, Onaivi ES, Ishiguro H, Liu QR, Tagliaferro PA, Brusco A, Uhl GR. (2006) Cannabinoid CB2 receptors: immunohistochemical localization in rat brainBrain Res 1071:10–23 [PubMed[]
  • Gonthier MP, Hoareau L, Festy F, Matias I, Valenti M, Bès-Houtmann S, Rouch C, Robert-Da Silva C, Chesne S, Lefebvre d’Hellencourt C, et al. (2007) Identification of endocannabinoids and related compounds in human fat cellsObesity 15:837–845 [PubMed[]
  • Gras D, Chanez P, Urbach V, Vachier I, Godard P, Bonnans C. (2009) Thiazolidinediones induce proliferation of human bronchial epithelial cells through the GPR40 receptorAm J Physiol Lung Cell Mol Physiol 296:L970–L978 [PubMed[]
  • Griffin G, Tao Q, Abood ME. (2000) Cloning and pharmacological characterization of the rat CB2 cannabinoid receptorJ Pharmacol Exp Ther 292:886–894 [PubMed[]
  • Guo J, Ikeda SR. (2004) Endocannabinoids modulate N-type calcium channels and G-protein-coupled inwardly rectifying potassium channels via CB1 cannabinoid receptors heterologously expressed in mammalian neuronsMol Pharmacol 65:665–674 [PubMed[]
  • Guzmán M, Lo Verme J, Fu J, Oveisi F, Blázquez C, Piomelli D. (2004) Oleoylethanolamide stimulates lipolysis by activating the nuclear receptor peroxisome proliferator-activated receptor α (PPAR-α)J Biol Chem 279:27849–27854 [PubMed[]
  • Hájos N, Freund TF. (2002a) Distinct cannabinoid sensitive receptors regulate hippocampal excitation and inhibitionChem Phys Lipids 121:73–82 [PubMed[]
  • Hájos N, Freund TF. (2002b) Pharmacological separation of cannabinoid sensitive receptors on hippocampal excitatory and inhibitory fibersNeuropharmacology 43:503–510 [PubMed[]
  • Hájos N, Katona I, Naiem SS, MacKie K, Ledent C, Mody I, Freund TF. (2000) Cannabinoids inhibit hippocampal GABAergic transmission and network oscillationsEur J Neurosci 12:3239–3249 [PubMed[]
  • Hájos N, Ledent C, Freund TF. (2001) Novel cannabinoid-sensitive receptor mediates inhibition of glutamatergic synaptic transmission in the hippocampusNeuroscience 106:1–4 [PubMed[]
  • Hampson AJ, Bornheim LM, Scanziani M, Yost CS, Gray AT, Hansen BM, Leonoudakis DJ, Bickler PE. (1998) Dual effects of anandamide on NMDA receptor-mediated responses and neurotransmissionJ Neurochem 70:671–676 [PubMed[]
  • Hanus L, Abu-Lafi S, Fride E, Breuer A, Vogel Z, Shalev DE, Kustanovich I, Mechoulam R. (2001) 2-Arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1 receptorProc Natl Acad Sci USA 98:3662–3665 [PMC free article] [PubMed[]
  • Hara T, Hirasawa A, Sun Q, Sadakane K, Itsubo C, Iga T, Adachi T, Koshimizu TA, Hashimoto T, Asakawa Y, et al. (2009) Novel selective ligands for free fatty acid receptors GPR120 and GPR40Naunyn Schmiedebergs Arch Pharmacol 380:247–255 [PubMed[]
  • Hejazi N, Zhou C, Oz M, Sun H, Ye JH, Zhang L. (2006) Δ9-Tetrahydrocannabinol and endogenous cannabinoid anandamide directly potentiate the function of glycine receptorsMol Pharmacol 69:991–997 [PubMed[]
  • Henquet C, Di Forti M, Morrison P, Kuepper R, Murray RM. (2008) Gene-environment interplay between cannabis and psychosisSchizophr Bull 34:1111–1121 [PMC free article] [PubMed[]
  • Henstridge CM, Balenga NA, Ford LA, Ross RA, Waldhoer M, Irving AJ. (2009) The GPR55 ligand L-α-lysophosphatidylinositol promotes RhoA-dependent Ca2+ signaling and NFAT activationFASEB J 23:183–193 [PubMed[]
  • Henstridge CM, Balenga NA, Schröder R, Kargl JK, Platzer W, Martini L, Arthur S, Penman J, Whistler JL, Kostenis E, et al. (2010) GPR55 ligands promote receptor coupling to multiple signalling pathwaysBr J Pharmacol 160:604–614 [PMC free article] [PubMed[]
  • Hermann H, De Petrocellis L, Bisogno T, Schiano Moriello A, Lutz B, Di Marzo V. (2003) Dual effect of cannabinoid CB1 receptor stimulation on a vanilloid VR1 receptor-mediated responseCell Mol Life Sci 60:607–616 [PubMed[]
  • Hermann H, Marsicano G, Lutz B. (2002) Coexpression of the cannabinoid receptor type 1 with dopamine and serotonin receptors in distinct neuronal subpopulations of the adult mouse forebrainNeuroscience 109:451–460 [PubMed[]
  • Hilairet S, Bouaboula M, Carrière D, Le Fur G, Casellas P. (2003) Hypersensitization of the orexin 1 receptor by the CB1 receptor: evidence for cross-talk blocked by the specific CB1 antagonist, SR141716J Biol Chem 278:23731–23737 [PubMed[]
  • Hillard CJ, Bloom AS. (1982) Δ9-Tetrahydrocannabinol-induced changes in β-adrenergic receptor binding in mouse cerebral cortexBrain Res 235:370–377 [PubMed[]
  • Hillard CJ, Manna S, Greenberg MJ, DiCamelli R, Ross RA, Stevenson LA, Murphy V, Pertwee RG, Campbell WB. (1999) Synthesis and characterization of potent and selective agonists of the neuronal cannabinoid receptor (CB1)J Pharmacol Exp Ther 289:1427–1433 [PubMed[]
  • Hinckley M, Vaccari S, Horner K, Chen R, Conti M. (2005) The G-protein-coupled receptors GPR3 and GPR12 are involved in cAMP signaling and maintenance of meiotic arrest in rodent oocytesDev Biol 287:249–261 [PubMed[]
  • Hirasawa A, Tsumaya K, Awaji T, Katsuma S, Adachi T, Yamada M, Sugimoto Y, Miyazaki S, Tsujimoto G. (2005) Free fatty acids regulate gut incretin glucagon-like peptide-1 secretion through GPR120Nature Med 11:90–94 [PubMed[]
  • Hoffman AF, Macgill AM, Smith D, Oz M, Lupica CR. (2005) Species and strain differences in the expression of a novel glutamate-modulating cannabinoid receptor in the rodent hippocampusEur J Neurosci 22:2387–2391 [PMC free article] [PubMed[]
  • Holzer P. (1988) Local effector functions of capsaicin-sensitive sensory nerve endings: involvement of tachykinins, calcitonin gene-related peptide and other neuropeptidesNeuroscience 24:739–768 [PubMed[]
  • Horswill JG, Bali U, Shaaban S, Keily JF, Jeevaratnam P, Babbs AJ, Reynet C, Wong Kai In P. (2007) PSNCBAM-1, a novel allosteric antagonist at cannabinoid CB1 receptors with hypophagic effects in ratsBr J Pharmacol 152:805–814 [PMC free article] [PubMed[]
  • Howlett AC. (1987) Cannabinoid inhibition of adenylate cyclase: relative activity of constituents and metabolites of marihuanaNeuropharmacology 26:507–512 [PubMed[]
  • Howlett AC. (2005) Cannabinoid receptor signalingHandb Exp Pharmacol 168:53–79 [PubMed[]
  • Howlett AC, Barth F, Bonner TI, Cabral G, Casellas P, Devane WA, Felder CC, Herkenham M, Mackie K, Martin BR, et al. (2002) International Union of Pharmacology. XXVII. Classification of cannabinoid receptorsPharmacol Rev 54:161–202 [PubMed[]
  • Hu H, He LY, Gong Z, Li N, Lu YN, Zhai QW, Liu H, Jiang HL, Zhu WL, Wang HY. (2009) A novel class of antagonists for the FFAs receptor GPR40Biochem Biophys Res Commun 390:557–563 [PubMed[]
  • Ichimura A, Hirasawa A, Hara T, Tsujimoto G. (2009) Free fatty acid receptors act as nutrient sensors to regulate energy homeostasisProstaglandins Other Lipid Mediat 89:82–88 [PubMed[]
  • Ignatov A, Lintzel J, Hermans-Borgmeyer I, Kreienkamp HJ, Joost P, Thomsen S, Methner A, Schaller HC. (2003a) Role of the G-protein-coupled receptor GPR12 as high-affinity receptor for sphingosylphosphorylcholine and its expression and function in brain developmentJ Neurosci 23:907–914 [PMC free article] [PubMed[]
  • Ignatov A, Lintzel J, Kreienkamp HJ, Schaller HC. (2003b) Sphingosine-1-phosphate is a high-affinity ligand for the G protein-coupled receptor GPR6 from mouse and induces intracellular Ca2+ release by activating the sphingosine-kinase pathwayBiochem Biophys Res Commun 311:329–336 [PubMed[]
  • Ishii S, Noguchi K, Yanagida K. (2009) Non-Edg family lysophosphatidic acid (LPA) receptorsProstaglandins Other Lipid Mediat 89:57–65 [PubMed[]
  • Itoh Y, Kawamata Y, Harada M, Kobayashi M, Fujii R, Fukusumi S, Ogi K, Hosoya M, Tanaka Y, Uejima H, et al. (2003) Free fatty acids regulate insulin secretion from pancreatic β cells through GPR40Nature 422:173–176 [PubMed[]
  • Iwamura H, Suzuki H, Ueda Y, Kaya T, Inaba T. (2001) In vitro and in vivo pharmacological characterization of JTE-907, a novel selective ligand for cannabinoid CB2 receptorJ Pharmacol Exp Ther 296:420–425 [PubMed[]
  • Jaggar SI, Hasnie FS, Sellaturay S, Rice AS. (1998) The anti-hyperalgesic actions of the cannabinoid anandamide and the putative CB2 receptor agonist palmitoylethanolamide in visceral and somatic inflammatory painPain 76:189–199 [PubMed[]
  • Járai Z, Wagner JA, Varga K, Lake KD, Compton DR, Martin BR, Zimmer AM, Bonner TI, Buckley NE, Mezey E, et al. (1999) Cannabinoid-induced mesenteric vasodilation through an endothelial site distinct from CB1 or CB2 receptorsProc Natl Acad Sci USA 96:14136–14141 [PMC free article] [PubMed[]
  • Jarrahian A, Hillard CJ. (1997) Arachidonylethanolamide (anandamide) binds with low affinity to dihydropyridine binding sites in brain membranesProstaglandins Leukot Essent Fatty Acids 57:551–554 [PubMed[]
  • Jarrahian A, Watts VJ, Barker EL. (2004) D2 dopamine receptors modulate Gα-subunit coupling of the CB1 cannabinoid receptorJ Pharmacol Exp Ther 308:880–886 [PubMed[]
  • Jhaveri MD, Richardson D, Robinson I, Garle MJ, Patel A, Sun Y, Sagar DR, Bennett AJ, Alexander SP, Kendall DA, et al. (2008) Inhibition of fatty acid amide hydrolase and cyclooxygenase-2 increases levels of endocannabinoid related molecules and produces analgesia via peroxisome proliferator-activated receptor-alpha in a model of inflammatory painNeuropharmacology 55:85–93 [PubMed[]
  • Johns DG, Behm DJ, Walker DJ, Ao Z, Shapland EM, Daniels DA, Riddick M, Dowell S, Staton PC, Green P, et al. (2007) The novel endocannabinoid receptor GPR55 is activated by atypical cannabinoids but does not mediate their vasodilator effectsBr J Pharmacol 152:825–831 [PMC free article] [PubMed[]
  • Johnson DE, Heald SL, Dally RD, Janis RA. (1993) Isolation identification and synthesis of an endogenous arachidonic amide that inhibits calcium channel antagonist 1,4-dihydropyridine bindingProstaglandins Leukot Essent Fatty Acids 48:429–437 [PubMed[]
  • Johnson DM, Garrett EM, Rutter R, Bonnert TP, Gao YD, Middleton RE, Sutton KG. (2006) Functional mapping of the transient receptor potential vanilloid 1 intracellular binding siteMol Pharmacol 70:1005–1012 [PubMed[]
  • Jordt SE, Bautista DM, Chuang HH, McKemy DD, Zygmunt PM, Högestätt ED, Meng ID, Julius D. (2004) Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1Nature 427:260–265 [PubMed[]
  • Jordt SE, Julius D. (2002) Molecular basis for species-specific sensitivity to “hot” chili peppersCell 108:421–430 [PubMed[]
  • Kapur A, Zhao P, Sharir H, Bai Y, Caron MG, Barak LS, Abood ME. (2009) Atypical responsiveness of the orphan receptor GPR55 to cannabinoid ligandsJ Biol Chem 284:29817–29827 [PMC free article] [PubMed[]
  • Kathmann M, Flau K, Redmer A, Tränkle C, Schlicker E. (2006) Cannabidiol is an allosteric modulator at mu- and delta-opioid receptorsNaunyn-Schmiedebergs Arch Pharmacol 372:354–361 [PubMed[]
  • Katona I, Urbán GM, Wallace M, Ledent C, Jung KM, Piomelli D, Mackie K, Freund TF. (2006) Molecular composition of the endocannabinoid system at glutamatergic synapsesJ Neurosci 26:5628–5637 [PMC free article] [PubMed[]
  • Katsuma S, Hatae N, Yano T, Ruike Y, Kimura M, Hirasawa A, Tsujimoto G. (2005) Free fatty acids inhibit serum deprivation-induced apoptosis through GPR120 in a murine enteroendocrine cell line STC-1J Biol Chem 280:19507–19515 [PubMed[]
  • Kawamura Y, Fukaya M, Maejima T, Yoshida T, Miura E, Watanabe M, Ohno-Shosaku T, Kano M. (2006) The CB1 cannabinoid receptor is the major cannabinoid receptor at excitatory presynaptic sites in the hippocampus and cerebellumJ Neurosci 26:2991–3001 [PMC free article] [PubMed[]
  • Kearn CS, Blake-Palmer K, Daniel E, Mackie K, Glass M. (2005) Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors enhances heterodimer formation: a mechanism for receptor cross-talk? Mol Pharmacol 67:1697–1704 [PubMed[]
  • Kenakin TP. (2001) Quantitation in receptor pharmacologyReceptors Channels 7:371–385 [PubMed[]
  • Kim HI, Kim TH, Shin YK, Lee CS, Park M, Song JH. (2005) Anandamide suppression of Na+ currents in rat dorsal root ganglion neuronsBrain Res 1062:39–47 [PubMed[]
  • Kimura T, Ohta T, Watanabe K, Yoshimura H, Yamamoto I. (1998) Anandamide, an endogenous cannabinoid receptor ligand, also interacts with 5-hydroxytryptamine (5-HT) receptorBiol Pharm Bull 21:224–226 [PubMed[]
  • Kimura T, Yamamoto I, Ohta T, Yoshida H, Watanabe K, Ho IK, Yoshimura H. (1996) Changes in 5-HT receptor binding induced by tetrahydrocannabinol metabolites in bovine cerebral cortexRes Commun Alcohol Subst Abuse 17:57–69 []
  • Kishida K, Shimomura I, Nishizawa H, Maeda N, Kuriyama H, Kondo H, Matsuda M, Nagaretani H, Ouchi N, Hotta K, et al. (2001) Enhancement of the aquaporin adipose gene expression by a peroxisome proliferator-activated receptor gammaJ Biol Chem 276:48572–48579 [PubMed[]
  • Kohno M, Hasegawa H, Inoue A, Muraoka M, Miyazaki T, Oka K, Yasukawa M. (2006) Identification of N-arachidonylglycine as the endogenous ligand for orphan G-protein-coupled receptor GPR18Biochem Biophys Res Commun 347:827–832 [PubMed[]
  • Kotarsky K, Boketoft A, Bristulf J, Nilsson NE, Norberg A, Hansson S, Owman C, Sillard R, Leeb-Lundberg LM, Olde B. (2006) Lysophosphatidic acid binds to and activates GPR92, a G protein-coupled receptor highly expressed in gastrointestinal lymphocytesJ Pharmacol Exp Ther 318:619–628 [PubMed[]
  • Kotarsky K, Nilsson NE, Flodgren E, Owman C, Olde B. (2003) A human cell surface receptor activated by free fatty acids and thiazolidinedione drugsBiochem Biophys Res Commun 301:406–410 [PubMed[]
  • Kozak KR, Gupta RA, Moody JS, Ji C, Boeglin WE, DuBois RN, Brash AR, Marnett LJ. (2002) 15-Lipoxygenase metabolism of 2-arachidonylglycerol: generation of a peroxisome proliferator-activated receptor α agonistJ Biol Chem 277:23278–23286 [PubMed[]
  • Lagalwar S, Bordayo EZ, Hoffmann KL, Fawcett JR, Frey WH., 2nd (1999) Anandamides inhibit binding to the muscarinic acetylcholine receptorJ Mol Neurosci 13:55–61 [PubMed[]
  • Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, et al. (2001) Initial sequencing and analysis of the human genomeNature 409:860–921 [PubMed[]
  • Lane JR, Beukers MW, Mulder-Krieger T, Ijzerman AP. (2010) The endocannabinoid 2-arachidonylglycerol is a negative allosteric modulator of the human A3 adenosine receptorBiochem Pharmacol 79:48–56 [PubMed[]
  • Lanzafame AA, Guida E, Christopoulos A. (2004) Effects of anandamide on the binding and signaling properties of M1 muscarinic acetylcholine receptorsBiochem Pharmacol 68:2207–2219 [PubMed[]
  • Lattin JE, Schroder K, Su AI, Walker JR, Zhang J, Wiltshire T, Saijo K, Glass CK, Hume DA, Kellie S, et al. (2008) Expression analysis of G protein-coupled receptors in mouse macrophagesImmunome Res 4:5. [PMC free article] [PubMed[]
  • Lauckner JE, Jensen JB, Chen HY, Lu HC, Hille B, Mackie K. (2008) GPR55 is a cannabinoid receptor that increases intracellular calcium and inhibits M currentProc Natl Acad Sci USA 105:2699–2704 [PMC free article] [PubMed[]
  • Lauffer L, Iakoubov R, Brubaker PL. (2008) GPR119: “double-dipping” for better glycemic controlEndocrinology 149:2035–2037 [PubMed[]
  • Le Poul E, Loison C, Struyf S, Springael JY, Lannoy V, Decobecq ME, Brezillon S, Dupriez V, Vassart G, Van Damme J, et al. (2003) Functional characterization of human receptors for short chain fatty acids and their role in polymorphonuclear cell activationJ Biol Chem 278:25481–25489 [PubMed[]
  • Lee CW, Rivera R, Dubin AE, Chun J. (2007) LPA4/GPR23 is a lysophosphatidic acid (LPA) receptor utilizing Gs-, Gq/Gi-mediated calcium signaling and G12/13-mediated Rho activationJ Biol Chem 282:4310–4317 [PubMed[]
  • Lee CW, Rivera R, Gardell S, Dubin AE, Chun J. (2006) GPR92 as a new G12/13– and Gq-coupled lysophosphatidic acid receptor that increases cAMP, LPA5J Biol Chem 281:23589–23597 [PubMed[]
  • Lee T, Schwandner R, Swaminath G, Weiszmann J, Cardozo M, Greenberg J, Jaeckel P, Ge H, Wang Y, Jiao X, et al. (2008) Identification and functional characterization of allosteric agonists for the G protein-coupled receptor FFA2Mol Pharmacol 74:1599–1609 [PubMed[]
  • Lenman A, Fowler CJ. (2007) Interaction of ligands for the peroxisome proliferator-activated receptor γ with the endocannabinoid systemBr J Pharmacol 151:1343–1351 [PMC free article] [PubMed[]
  • Liao C, Zheng J, David LS, Nicholson RA. (2004) Inhibition of voltage-sensitive sodium channels by the cannabinoid 1 receptor antagonist AM 251 in mammalian brainBasic Clin Pharmacol Toxicol 94:73–78 [PubMed[]
  • Liaw CW, Connolly DT. (2009) Sequence polymorphisms provide a common consensus sequence for GPR41 and GPR42DNA Cell Biol 28:555–560 [PubMed[]
  • Ligresti A, Moriello AS, Starowicz K, Matias I, Pisanti S, De Petrocellis L, Laezza C, Portella G, Bifulco M, Di Marzo V. (2006) Antitumor activity of plant cannabinoids with emphasis on the effect of cannabidiol on human breast carcinomaJ Pharmacol Exp Ther 318:1375–1387 [PubMed[]
  • Lin S, Khanolkar AD, Fan P, Goutopoulos A, Qin C, Papahadjis D, Makriyannis A. (1998) Novel analogues of arachidonylethanolamide (anandamide): affinities for the CB1 and CB2 cannabinoid receptors and metabolic stabilityJ Med Chem 41:5353–5361 [PubMed[]
  • Liu H, Enyeart JA, Enyeart JJ. (2007) Potent inhibition of native TREK-1 K+ channels by selected dihydropyridine Ca2+ channel antagonistsJ Pharmacol Exp Ther 323:39–48 [PubMed[]
  • Liu J, Li H, Burstein SH, Zurier RB, Chen JD. (2003) Activation and binding of peroxisome proliferator-activated receptor γ by synthetic cannabinoid ajulemic acidMol Pharmacol 63:983–992 [PubMed[]
  • Lobo MK, Cui Y, Ostlund SB, Balleine BW, Yang XW. (2007) Genetic control of instrumental conditioning by striatopallidal neuron-specific S1P receptor GPR6Nat Neurosci 10:1395–1397 [PubMed[]
  • Loría F, Petrosino S, Hernangómez M, Mestre L, Spagnolo A, Correa F, Di Marzo V, Docagne F, Guaza C. (2010) An endocannabinoid tone limits excitotoxicity in vitro and in a model of multiple sclerosisNeurobiol Dis 37:166–176 [PubMed[]
  • Lo Verme J, Fu J, Astarita G, La Rana G, Russo R, Calignano A, Piomelli D. (2005a) The nuclear receptor peroxisome proliferator-activated receptor-α mediates the antiinflammatory actions of palmitoylethanolamideMol Pharmacol 67:15–19 [PubMed[]
  • Lo Verme J, Gaetani S, Fu J, Oveisi F, Burton K, Piomelli D. (2005b) Regulation of food intake by oleoylethanolamideCell Mol Life Sci 62:708–716 [PubMed[]
  • LoVerme J, Russo R, La Rana G, Fu J, Farthing J, Mattace-Raso G, Meli R, Hohmann A, Calignano A, Piomelli D. (2006) Rapid broad-spectrum analgesia through activation of peroxisome proliferator-activated receptor-αJ Pharmacol Exp Ther 319:1051–1061 [PubMed[]
  • Lozovaya N, Yatsenko N, Beketov A, Tsintsadze T, Burnashev N. (2005) Glycine receptors in CNS neurons as a target for nonretrograde action of cannabinoidsJ Neurosci 25:7499–7506 [PMC free article] [PubMed[]
  • Lunn CA, Reich EP, Fine JS, Lavey B, Kozlowski JA, Hipkin RW, Lundell DJ, Bober L. (2008) Biology and therapeutic potential of cannabinoid CB2 receptor inverse agonistsBr J Pharmacol 153:226–239 [PMC free article] [PubMed[]
  • Maccarrone M, Pauselli R, Di Rienzo M, Finazzi-Agrò A. (2002) Binding, degradation and apoptotic activity of stearoylethanolamide in rat C6 glioma cellsBiochem J 366:137–144 [PMC free article] [PubMed[]
  • Mackie K. (2005) Cannabinoid receptor homo- and heterodimerizationLife Sci 77:1667–1673 [PubMed[]
  • Maingret F, Patel AJ, Lazdunski M, Honoré E. (2001) The endocannabinoid anandamide is a direct and selective blocker of the background K+ channel TASK-1EMBO J 20:47–54 [PMC free article] [PubMed[]
  • Maione S, Bisogno T, de Novellis V, Palazzo E, Cristino L, Valenti M, Petrosino S, Guglielmotti V, Rossi F, Di Marzo V. (2006) Elevation of endocannabinoid levels in the ventrolateral periaqueductal grey through inhibition of fatty acid amide hydrolase affects descending nociceptive pathways via both cannabinoid receptor type 1 and transient receptor potential vanilloid type-1 receptorsJ Pharmacol Exp Ther 316:969–982 [PubMed[]
  • Maneuf YP, Brotchie JM. (1997) Paradoxical action of the cannabinoid WIN 55,212–2 in stimulated and basal cyclic AMP accumulation in rat globus pallidus slicesBr J Pharmacol 120:1397–1398 [PMC free article] [PubMed[]
  • Marcellino D, Carriba P, Filip M, Borgkvist A, Frankowska M, Bellido I, Tanganelli S, Müller CE, Fisone G, Lluis C, et al. (2008) Antagonistic cannabinoid CB1/dopamine D2 receptor interactions in striatal CB1/D2 heteromers. A combined neurochemical and behavioral analysisNeuropharmacology 54:815–823 [PubMed[]
  • Martin BR, Stevenson LA, Pertwee RG, Breivogel CS, Williams W, Mahadevan A, Razdan RK. (2002) Agonists and silent antagonists in a series of cannabinoid sulfonamides, in 12th Annual Symposium of the Cannabinoids; 2002 Jul 10–14; Pacific Grove, CA p. 2 International Cannabinoid Research Society, Burlington, Vermont []
  • Matias I, Gonthier MP, Orlando P, Martiadis V, De Petrocellis L, Cervino C, Petrosino S, Hoareau L, Festy F, Pasquali R, et al. (2006) Regulation, function, and dysregulation of endocannabinoids in models of adipose and β-pancreatic cells and in obesity and hyperglycemiaJ Clin Endocrinol Metab 91:3171–3180 [PubMed[]
  • Matsuda LA, Lolait SJ, Brownstein MJ, Young AC, Bonner TI. (1990) Structure of a cannabinoid receptor and functional expression of the cloned cDNANature 346:561–564 [PubMed[]
  • Mazzola C, Medalie J, Scherma M, Panlilio LV, Solinas M, Tanda G, Drago F, Cadet JL, Goldberg SR, Yasar S. (2009) Fatty acid amide hydrolase (FAAH) inhibition enhances memory acquisition through activation of PPAR-α nuclear receptorsLearn Mem 16:332–337 [PMC free article] [PubMed[]
  • McDonald HA, Neelands TR, Kort M, Han P, Vos MH, Faltynek CR, Moreland RB, Puttfarcken PS. (2008) Characterization of A-425619 at native TRPV1 receptors: a comparison between dorsal root ganglia and trigeminal gangliaEur J Pharmacol 596:62–69 [PubMed[]
  • McHugh D, Hu SS, Rimmerman N, Juknat A, Vogel Z, Walker JM, Bradshaw HB. (2010) N-arachidonoyl glycine, an abundant endogenous lipid, potently drives directed cellular migration through GPR18, the putative abnormal cannabidiol receptorBMC Neurosci. 11:44. [PMC free article] [PubMed[]
  • McKemy DD. (2005) How cold is it? TRPM8 and TRPA1 in the molecular logic of cold sensationMol Pain 1:16. [PMC free article] [PubMed[]
  • McPartland JM, Matias I, Di Marzo V, Glass M. (2006) Evolutionary origins of the endocannabinoid systemGene 370:64–74 [PubMed[]
  • Mechoulam R, Ben-Shabat S, Hanus L, Ligumsky M, Kaminski NE, Schatz AR, Gopher A, Almog S, Martin BR, Compton DR. (1995) Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptorsBiochem Pharmacol 50:83–90 [PubMed[]
  • Melis M, Pillolla G, Luchicchi A, Muntoni AL, Yasar S, Goldberg SR, Pistis M. (2008) Endogenous fatty acid ethanolamides suppress nicotine-induced activation of mesolimbic dopamine neurons through nuclear receptorsJ Neurosci 28:13985–13994 [PMC free article] [PubMed[]
  • Meschler JP, Howlett AC. (2001) Signal transduction interactions between CB1 cannabinoid and dopamine receptors in the rat and monkey striatumNeuropharmacology 40:918–926 [PubMed[]
  • Michalik L, Auwerx J, Berger JP, Chatterjee VK, Glass CK, Gonzalez FJ, Grimaldi PA, Kadowaki T, Lazar MA, O’Rahilly S, et al. (2006) International Union of Pharmacology. LXI. Peroxisome proliferator-activated receptorsPharmacol Rev 58:726–741 [PubMed[]
  • Milligan G, Smith NJ. (2007) Allosteric modulation of heterodimeric G-protein-coupled receptorsTrends Pharmacol Sci 28:615–620 [PubMed[]
  • Milman G, Maor Y, Abu-Lafi S, Horowitz M, Gallily R, Batkai S, Mo FM, Offertaler L, Pacher P, Kunos G, et al. (2006) N-arachidonoyl L-serine, an endocannabinoid-like brain constituent with vasodilatory propertiesProc Natl Acad Sci USA 103:2428–2433 [PMC free article] [PubMed[]
  • Miyauchi S, Hirasawa A, Iga T, Liu N, Itsubo C, Sadakane K, Hara T, Tsujimoto G. (2009) Distribution and regulation of protein expression of the free fatty acid receptor GPR120Naunyn Schmiedebergs Arch Pharmacol 379:427–434 [PubMed[]
  • Mo FM, Offertáler L, Kunos G. (2004) Atypical cannabinoid stimulates endothelial cell migration via a Gi/Go-coupled receptor distinct from CB1, CB2 or EDG-1Eur J Pharmacol 489:21–27 [PubMed[]
  • Molderings GJ, Bönisch H, Hammermann R, Göthert M, Brüss M. (2002) Noradrenaline release-inhibiting receptors on PC12 cells devoid of α2– and CB1 receptors: similarities to presynaptic imidazoline and edg receptorsNeurochem Int 40:157–167 [PubMed[]
  • Molderings GJ, Likungu J, Göthert M. (1999) Presynaptic cannabinoid and imidazoline receptors in the human heart and their potential relationshipNaunyn-Schmiedebergs Arch Pharmacol 360:157–164 [PubMed[]
  • Monory K, Tzavara ET, Lexime J, Ledent C, Parmentier M, Borsodi A, Hanoune J. (2002) Novel, not adenylyl cyclase-coupled cannabinoid binding site in cerebellum of miceBiochem Biophys Res Commun 292:231–235 [PubMed[]
  • Moreno-Galindo EG, Barrio-Echavarría GF, Vásquez JC, Decher N, Sachse FB, Tristani-Firouzi M, Sánchez-Chapula JA, Navarro-Polanco RA. (2010) Molecular basis for a high-potency open-channel block of Kv1.5 channel by the endocannabinoid anandamideMol Pharmacol 77:751–758 [PubMed[]
  • Morgese MG, Cassano T, Cuomo V, Giuffrida A. (2007) Anti-dyskinetic effects of cannabinoids in a rat model of Parkinson’s disease: role of CB1 and TRPV1 receptorsExp Neurol 208:110–119 [PMC free article] [PubMed[]
  • Movahed P, Jönsson BA, Birnir B, Wingstrand JA, Jørgensen TD, Ermund A, Sterner O, Zygmunt PM, Högestätt ED. (2005) Endogenous unsaturated C18 N-acylethanolamines are vanilloid receptor (TRPV1) agonistsJ Biol Chem 280:38496–38504 [PubMed[]
  • Mukhopadhyay S, Chapnick BM, Howlett AC. (2002) Anandamide-induced vasorelaxation in rabbit aortic rings has two components: G protein dependent and independentAm J Physiol Heart Circ Physiol 282:H2046–H2054 [PubMed[]
  • Mukhopadhyay S, McIntosh HH, Houston DB, Howlett AC. (2000) The CB1 cannabinoid receptor juxtamembrane C-terminal peptide confers activation to specific G proteins in brainMol Pharmacol 57:162–170 [PubMed[]
  • Munro S, Thomas KL, Abu-Shaar M. (1993) Molecular characterization of a peripheral receptor for cannabinoidsNature 365:61–65 [PubMed[]
  • Murakami M, Shiraishi A, Tabata K, Fujita N. (2008) Identification of the orphan GPCR, P2Y10 receptor as the sphingosine-1-phosphate and lysophosphatidic acid receptorBiochem Biophys Res Commun 371:707–712 [PubMed[]
  • Navarro HA, Howard JL, Pollard GT, Carroll FI. (2009) Positive allosteric modulation of the human cannabinoid (CB1) receptor by RTI-371, a selective inhibitor of the dopamine transporterBr J Pharmacol 156:1178–1184 [PMC free article] [PubMed[]
  • Németh B, Ledent C, Freund TF, Hájos N. (2008) CB1 receptor-dependent and -independent inhibition of excitatory postsynaptic currents in the hippocampus by WIN 55,212–2Neuropharmacology 54:51–57 [PMC free article] [PubMed[]
  • Nicholson RA, Liao C, Zheng J, David LS, Coyne L, Errington AC, Singh G, Lees G. (2003) Sodium channel inhibition by anandamide and synthetic cannabimimetics in brainBrain Res 978:194–204 [PubMed[]
  • Nilius B, Owsianik G, Voets T, Peters JA. (2007) Transient receptor potential cation channels in diseasePhysiol Rev 87:165–217 [PubMed[]
  • Nilius B, Voets T. (2005) TRP channels: a TR(I)P through a world of multifunctional cation channelsPflugers Arch 451:1–10 [PubMed[]
  • Nilsson NE, Kotarsky K, Owman C, Olde B. (2003) Identification of a free fatty acid receptor, FFA2R, expressed on leukocytes and activated by short-chain fatty acidsBiochem Biophys Res Commun 303:1047–1052 [PubMed[]
  • Noguchi K, Ishii S, Shimizu T. (2003) Identification of P2y9/GPR23 as a novel G protein-coupled receptor for lysophosphatidic acid, structurally distant from the edg familyJ Biol Chem 278:25600–25606 [PubMed[]
  • Norrod AG, Puffenbarger RA. (2007) Genetic polymorphisms of the endocannabinoid systemChem Biodivers 4:1926–1932 [PubMed[]
  • Offertáler L, Mo FM, Bátkai S, Liu J, Begg M, Razdan RK, Martin BR, Bukoski RD, Kunos G. (2003) Selective ligands and cellular effectors of a G protein-coupled endothelial cannabinoid receptorMol Pharmacol 63:699–705 [PubMed[]
  • Oh DY, Yoon JM, Moon MJ, Hwang JI, Choe H, Lee JY, Kim JI, Kim S, Rhim H, O’Dell DK, et al. (2008) Identification of farnesyl pyrophosphate and N-arachidonylglycine as endogenous ligands for GPR92J Biol Chem 283:21054–21064 [PMC free article] [PubMed[]
  • Ohno-Shosaku T, Tsubokawa H, Mizushima I, Yoneda N, Zimmer A, Kano M. (2002) Presynaptic cannabinoid sensitivity is a major determinant of depolarization-induced retrograde suppression at hippocampal synapsesJ Neurosci 22:3864–3872 [PMC free article] [PubMed[]
  • Oka S, Nakajima K, Yamashita A, Kishimoto S, Sugiura T. (2007) Identification of GPR55 as a lysophosphatidylinositol receptorBiochem Biophys Res Commun 362:928–934 [PubMed[]
  • Oka S, Toshida T, Maruyama K, Nakajima K, Yamashita A, Sugiura T. (2009) 2-Arachidonoyl-sn-glycero-3-phosphoinositol: a possible natural ligand for GPR55J Biochem 145:13–20 [PubMed[]
  • Okada Y, Imendra KG, Miyazaki T, Hotokezaka H, Fujiyama R, Zeredo JL, Miyamoto T, Toda K. (2005) Biophysical properties of voltage-gated Na+ channels in frog parathyroid cells and their modulation by cannabinoidsJ Exp Biol 208:4747–4756 [PubMed[]
  • Oliver D, Lien CC, Soom M, Baukrowitz T, Jonas P, Fakler B. (2004) Functional conversion between A-type and delayed rectifier K+ channels by membrane lipidsScience 304:265–270 [PubMed[]
  • O’Sullivan SE, Bennett AJ, Kendall DA, Randall MD. (2006) Cannabinoid ligands as activators of peroxisome proliferator-activated receptor gamma (PPARγ), in 16th Annual Symposium of the Cannabinoids; 2006 Jun 25–28; Budapest, Hungary p. 59 International Cannabinoid Research Society, Burlington, Vermont []
  • O’Sullivan SE, Kendall DA, Randall MD. (2009a) Time-dependent vascular effects of endocannabinoids mediated by peroxisome proliferator-activated receptor gamma (PPARγ)PPAR Res 2009:425289. [PMC free article] [PubMed[]
  • O’Sullivan SE, Sun Y, Bennett AJ, Randall MD, Kendall DA. (2009b) Time-dependent vascular actions of cannabidiol in the rat aortaEur J Pharmacol 612:61–68 [PubMed[]
  • O’Sullivan SE, Tarling EJ, Bennett AJ, Kendall DA, Randall MD. (2005) Novel time-dependent vascular actions of Δ9-tetrahydrocannabinol mediated by peroxisome proliferator-activated receptor γBiochem Biophys Res Commun 337:824–831 [PubMed[]
  • Overton HA, Babbs AJ, Doel SM, Fyfe MC, Gardner LS, Griffin G, Jackson HC, Procter MJ, Rasamison CM, Tang-Christensen M, et al. (2006) Deorphanization of a G protein-coupled receptor for oleoylethanolamide and its use in the discovery of small-molecule hypophagic agentsCell Metab 3:167–175 [PubMed[]
  • Owsianik G, D’hoedt D, Voets T, Nilius B. (2006) Structure-function relationship of the TRP channel superfamilyRev Physiol Biochem Pharmacol 156:61–90 [PubMed[]
  • Oz M, Jackson SN, Woods AS, Morales M, Zhang L. (2005) Additive effects of endogenous cannabinoid anandamide and ethanol on α7-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytesJ Pharmacol Exp Ther 313:1272–1280 [PubMed[]
  • Oz M, Ravindran A, Diaz-Ruiz O, Zhang L, Morales M. (2003) The endogenous cannabinoid anandamide inhibits α7 nicotinic acetylcholine receptor-mediated responses in Xenopus oocytesJ Pharmacol Exp Ther 306:1003–1010 [PubMed[]
  • Oz M, Tchugunova Y, Dinc M. (2004b) Differential effects of endogenous and synthetic cannabinoids on voltage-dependent calcium fluxes in rabbit T-tubule membranes: comparison with fatty acidsEur J Pharmacol 502:47–58 [PubMed[]
  • Oz M, Tchugunova YB, Dunn SM. (2000) Endogenous cannabinoid anandamide directly inhibits voltage-dependent Ca2+ fluxes in rabbit T-tubule membranesEur J Pharmacol 404:13–20 [PubMed[]
  • Oz M, Yang KH, Dinc M, Shippenberg TS. (2007) The endogenous cannabinoid anandamide inhibits cromakalim-activated K+ currents in follicle-enclosed Xenopus oocytesJ Pharmacol Exp Ther 323:547–554 [PubMed[]
  • Oz M, Zhang L, Morales M. (2002) Endogenous cannabinoid, anandamide, acts as a noncompetitive inhibitor on 5-HT3 receptor-mediated responses in Xenopus oocytesSynapse 46:150–156 [PubMed[]
  • Oz M, Zhang L, Ravindran A, Morales M, Lupica CR. (2004a) Differential effects of endogenous and synthetic cannabinoids on α7-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytesJ Pharmacol Exp Ther 310:1152–1160 [PubMed[]
  • Pacheco MA, Ward SJ, Childers SR. (1993) Identification of cannabinoid receptors in cultures of rat cerebellar granule cellsBrain Res 603:102–110 [PubMed[]
  • Pagano C, Pilon C, Calcagno A, Urbanet R, Rossato M, Milan G, Bianchi K, Rizzuto R, Bernante P, Federspil G, et al. (2007) The endogenous cannabinoid system stimulates glucose uptake in human fat cells via phosphatidylinositol 3-kinase and calcium-dependent mechanismsJ Clin Endocrinol Metab 92:4810–4819 [PubMed[]
  • Pan X, Ikeda SR, Lewis DL. (1998) SR 141716A acts as an inverse agonist to increase neuronal voltage-dependent Ca2+ currents by reversal of tonic CB1 cannabinoid receptor activityMol Pharmacol 54:1064–1072 [PubMed[]
  • Parker HE, Habib AM, Rogers GJ, Gribble FM, Reimann F. (2009) Nutrient-dependent secretion of glucose-dependent insulinotropic polypeptide from primary murine K cellsDiabetologia 52:289–298 [PMC free article] [PubMed[]
  • Parmar N, Ho WS. (2010) N-arachidonoyl glycine, an endogenous lipid that acts as a vasorelaxant via nitric oxide and large conductance calcium-activated potassium channelsBr J Pharmacol 160:594–603 [PMC free article] [PubMed[]
  • Pasternack SM, von Kügelgen I, Aboud KA, Lee YA, Rüschendorf F, Voss K, Hillmer AM, Molderings GJ, Franz T, Ramirez A, et al. (2008) G protein-coupled receptor P2Y5 and its ligand LPA are involved in maintenance of human hair growthNat Genet 40:329–334 [PubMed[]
  • Paugh SW, Cassidy MP, He H, Milstien S, Sim-Selley LJ, Spiegel S, Selley DE. (2006) Sphingosine and its analog, the immunosuppressant 2-amino-2-(2-[4-octylphenyl]ethyl)-1,3-propanediol, interact with the CB1 cannabinoid receptorMol Pharmacol 70:41–50 [PubMed[]
  • Pertwee RG. (1999) Pharmacology of cannabinoid receptor ligandsCurr Med Chem 6:635–664 [PubMed[]
  • Pertwee RG. (2004) Novel pharmacological targets for cannabinoidsCurr Neuropharmacol 2:9–29 []
  • Pertwee RG. (2005a) Pharmacological actions of cannabinoidsHandb Exp Pharmacol 168:1–51 [PubMed[]
  • Pertwee RG. (2005b) The therapeutic potential of drugs that target cannabinoid receptors or modulate the tissue levels or actions of endocannabinoidsAAPS J 7:E625–E654 [PMC free article] [PubMed[]
  • Pertwee RG. (2005c) Inverse agonism and neutral antagonism at cannabinoid CB1 receptorsLife Sci 76:1307–1324 [PubMed[]
  • Pertwee RG. (2008a) The diverse CB1 and CB2 receptor pharmacology of three plant cannabinoids: Δ9-tetrahydrocannabinol, cannabidiol and Δ9-tetrahydrocannabivarinBr J Pharmacol 153:199–215 [PMC free article] [PubMed[]
  • Pertwee RG. (2008b) Ligands that target cannabinoid receptors in the brain: from THC to anandamide and beyondAddict Biol 13:147–159 [PubMed[]
  • Pertwee RG, Ross RA. (2002) Cannabinoid receptors and their ligandsProstaglandins Leukot Essent Fatty Acids 66:101–121 [PubMed[]
  • Peters JM, Lee SS, Li W, Ward JM, Gavrilova O, Everett C, Reitman ML, Hudson LD, Gonzalez FJ. (2000) Growth, adipose, brain, and skin alterations resulting from targeted disruption of the mouse peroxisome proliferator-activated receptor β(δ)Mol Cell Biol 20:5119–5128 [PMC free article] [PubMed[]
  • Pickel VM, Chan J, Kash TL, Rodríguez JJ, MacKie K. (2004) Compartment-specific localization of cannabinoid 1 (CB1) and μ-opioid receptors in rat nucleus accumbensNeuroscience 127:101–112 [PubMed[]
  • Pickel VM, Chan J, Kearn CS, Mackie K. (2006) Targeting dopamine D2 and cannabinoid-1 (CB1) receptors in rat nucleus accumbensJ Comp Neurol 495:299–313 [PMC free article] [PubMed[]
  • Pietr M, Kozela E, Levy R, Rimmerman N, Lin YH, Stella N, Vogel Z, Juknat A. (2009) Differential changes in GPR55 during microglial cell activationFEBS Lett 583:2071–2076 [PubMed[]
  • Poling JS, Rogawski MA, Salem N, Jr, Vicini S. (1996) Anandamide, an endogenous cannabinoid, inhibits Shaker-related voltage-gated K+ channelsNeuropharmacology 35:983–991 [PubMed[]
  • Porter AC, Sauer JM, Knierman MD, Becker GW, Berna MJ, Bao J, Nomikos GG, Carter P, Bymaster FP, Leese AB, et al. (2002) Characterization of a novel endocannabinoid, virodhamine, with antagonist activity at the CB1 receptorJ Pharmacol Exp Ther 301:1020–1024 [PubMed[]
  • Price MR, Baillie GL, Thomas A, Stevenson LA, Easson M, Goodwin R, McLean A, McIntosh L, Goodwin G, Walker G, et al. (2005a) Allosteric modulation of the cannabinoid CB1 receptorMol Pharmacol 68:1484–1495 [PubMed[]
  • Price TJ, Patwardhan AM, Flores CM, Hargreaves KM. (2005b) A role for the anandamide membrane transporter in TRPV1-mediated neurosecretion from trigeminal sensory neuronsNeuropharmacology 49:25–39 [PMC free article] [PubMed[]
  • Price TJ, Patwardhan A, Akopian AN, Hargreaves KM, Flores CM. (2004) Modulation of trigeminal sensory neuron activity by the dual cannabinoid-vanilloid agonists anandamide, N-arachidonoyl-dopamine and arachidonyl-2-chloroethylamideBr J Pharmacol 141:1118–1130 [PMC free article] [PubMed[]
  • Qin N, Neeper MP, Liu Y, Hutchinson TL, Lubin ML, Flores CM. (2008) TRPV2 is activated by cannabidiol and mediates CGRP release in cultured rat dorsal root ganglion neuronsJ Neurosci 28:6231–6238 [PMC free article] [PubMed[]
  • Rashid MH, Inoue M, Bakoshi S, Ueda H. (2003a) Increased expression of vanilloid receptor 1 on myelinated primary afferent neurons contributes to the antihyperalgesic effect of capsaicin cream in diabetic neuropathic pain in miceJ Pharmacol Exp Ther 306:709–717 [PubMed[]
  • Rashid MH, Inoue M, Kondo S, Kawashima T, Bakoshi S, Ueda H. (2003b) Novel expression of vanilloid receptor 1 on capsaicin-insensitive fibers accounts for the analgesic effect of capsaicin cream in neuropathic painJ Pharmacol Exp Ther 304:940–948 [PubMed[]
  • Rinaldi-Carmona M, Barth F, Millan J, Derocq JM, Casellas P, Congy C, Oustric D, Sarran M, Bouaboula M, Calandra B, et al. (1998) SR 144528, the first potent and selective antagonist of the CB2 cannabinoid receptorJ Pharmacol Exp Ther 284:644–650 [PubMed[]
  • Rios C, Gomes I, Devi LA. (2006) Mu opioid and CB1 cannabinoid receptor interactions: reciprocal inhibition of receptor signaling and neuritogenesisBr J Pharmacol 148:387–395 [PMC free article] [PubMed[]
  • Roberts LA, Christie MJ, Connor M. (2002) Anandamide is a partial agonist at native vanilloid receptors in acutely isolated mouse trigeminal sensory neuronsBr J Pharmacol 137:421–428 [PMC free article] [PubMed[]
  • Roche M, Kelly JP, O’Driscoll M, Finn DP. (2008) Augmentation of endogenous cannabinoid tone modulates lipopolysaccharide-induced alterations in circulating cytokine levels in ratsImmunology 125:263–271 [PMC free article] [PubMed[]
  • Rockwell CE, Kaminski NE. (2004) A cyclooxygenase metabolite of anandamide causes inhibition of interleukin-2 secretion in murine splenocytesJ Pharmacol Exp Ther 311:683–690 [PubMed[]
  • Rockwell CE, Snider NT, Thompson JT, Vanden Heuvel JP, Kaminski NE. (2006) Interleukin-2 suppression by 2-arachidonyl glycerol is mediated through peroxisome proliferator activated receptorγ independently of cannabinoid receptors 1 and 2Mol Pharmacol 70:101–111 [PubMed[]
  • Rodriguez JJ, Mackie K, Pickel VM. (2001) Ultrastructural localization of the CB1 cannabinoid receptor in μ-opioid receptor patches of the rat caudate putamen nucleusJ Neurosci 21:823–833 [PMC free article] [PubMed[]
  • Ross HR, Gilmore AJ, Connor M. (2009) Inhibition of human recombinant T-type calcium channels by the endocannabinoid N-arachidonoyl dopamineBr J Pharmacol 156:740–750 [PMC free article] [PubMed[]
  • Ross HR, Napier I, Connor M. (2008) Inhibition of recombinant human T-type calcium channels by Δ9-tetrahydrocannabinol and cannabidiolJ Biol Chem 283:16124–16134 [PMC free article] [PubMed[]
  • Ross RA. (2009) The enigmatic pharmacology of GPR55Trends Pharmacol Sci 30:156–163 [PubMed[]
  • Ross RA, Brockie HC, Stevenson LA, Murphy VL, Templeton F, Makriyannis A, Pertwee RG. (1999) Agonist-inverse agonist characterization at CB1 and CB2 cannabinoid receptors of L759633, L759656 and AM630Br J Pharmacol 126:665–672 [PMC free article] [PubMed[]
  • Ross RA, Coutts AA, McFarlane SM, Anavi-Goffer S, Irving AJ, Pertwee RG, MacEwan DJ, Scott RH. (2001a) Actions of cannabinoid receptor ligands on rat cultured sensory neurones: implications for antinociceptionNeuropharmacology 40:221–232 [PubMed[]
  • Ross RA, Gibson TM, Brockie HC, Leslie M, Pashmi G, Craib SJ, Di Marzo V, Pertwee RG. (2001b) Structure-activity relationship for the endogenous cannabinoid, anandamide, and certain of its analogues at vanilloid receptors in transfected cells and vas deferensBr J Pharmacol 132:631–640 [PMC free article] [PubMed[]
  • Rossi F, Siniscalco D, Luongo L, De Petrocellis L, Bellini G, Petrosino S, Torella M, Santoro C, Nobili B, Perrotta S, et al. (2009) The endovanilloid/endocannabinoid system in human osteoclasts: possible involvement in bone formation and resorptionBone 44:476–484 [PubMed[]
  • Rubino T, Realini N, Castiglioni C, Guidali C, Viganó D, Marras E, Petrosino S, Perletti G, Maccarrone M, Di Marzo V, et al. (2008) Role in anxiety behavior of the endocannabinoid system in the prefrontal cortexCereb Cortex 18:1292–1301 [PubMed[]
  • Ruiu S, Pinna GA, Marchese G, Mussinu JM, Saba P, Tambaro S, Casti P, Vargiu R, Pani L. (2003) Synthesis and characterization of NESS 0327: a novel putative antagonist of the CB1 cannabinoid receptorJ Pharmacol Exp Ther 306:363–370 [PubMed[]
  • Ryberg E, Larsson N, Sjögren S, Hjorth S, Hermansson NO, Leonova J, Elebring T, Nilsson K, Drmota T, Greasley PJ. (2007) The orphan receptor GPR55 is a novel cannabinoid receptorBr J Pharmacol 152:1092–1101 [PMC free article] [PubMed[]
  • Sade H, Muraki K, Ohya S, Hatano N, Imaizumi Y. (2006) Activation of large-conductance, Ca2+-activated K+ channels by cannabinoidsAm J Physiol Cell Physiol 290:C77–C86 [PubMed[]
  • Sagan S, Venance L, Torrens Y, Cordier J, Glowinski J, Giaume C. (1999) Anandamide and WIN 55212–2 inhibit cyclic AMP formation through G-protein-coupled receptors distinct from CB1 cannabinoid receptors in cultured astrocytesEur J Neurosci 11:691–699 [PubMed[]
  • Salio C, Fischer J, Franzoni MF, Mackie K, Kaneko T, Conrath M. (2001) CB1-cannabinoid and μ-opioid receptor co-localization on postsynaptic target in the rat dorsal hornNeuroreport 12:3689–3692 [PubMed[]
  • Sasso O, La Rana G, Vitiello S, Russo R, D’Agostino G, Iacono A, Russo E, Citraro R, Cuzzocrea S, Piazza PV, et al. (2010) Palmitoylethanolamide modulates pentobarbital-evoked hypnotic effect in mice. Involvement of allopregnanolone biosynthesisEur Neuropsychopharmacol 20:195–206 [PubMed[]
  • Savinainen JR, Järvinen T, Laine K, Laitinen JT. (2001) Despite substantial degradation, 2-arachidonoylglycerol is a potent full efficacy agonist mediating CB1 receptor-dependent G-protein activation in rat cerebellar membranesBr J Pharmacol 134:664–672 [PMC free article] [PubMed[]
  • Savinainen JR, Saario SM, Niemi R, Järvinen T, Laitinen JT. (2003) An optimized approach to study endocannabinoid signaling: evidence against constitutive activity of rat brain adenosine A1 and cannabinoid CB1 receptorsBr J Pharmacol 140:1451–1459 [PMC free article] [PubMed[]
  • Sawzdargo M, Nguyen T, Lee DK, Lynch KR, Cheng R, Heng HH, George SR, O’Dowd BF. (1999) Identification and cloning of three novel human G protein-coupled receptor genes GPR52, Psi GPR53 and GPR55: GPR55 is extensively expressed in human brainMol Brain Res 64:193–198 [PubMed[]
  • Shapira M, Gafni M, Sarne Y. (1998) Independence of, and interactions between, cannabinoid and opioid signal transduction pathways in N18TG2 cellsBrain Res 806:26–35 [PubMed[]
  • Shapira M, Vogel Z, Sarne Y. (2000) Opioid and cannabinoid receptors share a common pool of GTP-binding proteins in cotransfected cells, but not in cells which endogenously coexpress the receptorsCell Mol Neurobiol 20:291–304 [PubMed[]
  • Shaw G, Morse S, Ararat M, Graham FL. (2002) Preferential transformation of human neuronal cells by human adenoviruses and the origin of HEK 293 cellsFASEB J 16:869–871 [PubMed[]
  • Shen M, Thayer SA. (1998) The cannabinoid agonist Win55,212–2 inhibits calcium channels by receptor-mediated and direct pathways in cultured rat hippocampal neuronsBrain Res 783:77–84 [PubMed[]
  • Shimasue K, Urushidani T, Hagiwara M, Nagao T. (1996) Effects of anandamide and arachidonic acid on specific binding of (+)-PN200–110, diltiazem and (−)- desmethoxyverapamil to L-type Ca2+ channelEur J Pharmacol 296:347–350 [PubMed[]
  • Showalter VM, Compton DR, Martin BR, Abood ME. (1996) Evaluation of binding in a transfected cell line expressing a peripheral cannabinoid receptor (CB2): identification of cannabinoid receptor subtype selective ligandsJ Pharmacol Exp Ther 278:989–999 [PubMed[]
  • Sim-Selley LJ, Goforth PB, Mba MU, Macdonald TL, Lynch KR, Milstien S, Spiegel S, Satin LS, Welch SP, Selley DE. (2009) Sphingosine-1-phosphate receptors mediate neuromodulatory functions in the CNSJ Neurochem 110:1191–1202 [PMC free article] [PubMed[]
  • Skaper SD, Buriani A, Dal Toso R, Petrelli L, Romanello S, Facci L, Leon A. (1996) The ALIAmide palmitoylethanolamide and cannabinoids, but not anandamide, are protective in a delayed postglutamate paradigm of excitotoxic death in cerebellar granule neuronsProc Natl Acad Sci USA 93:3984–3989 [PMC free article] [PubMed[]
  • Smart D, Gunthorpe MJ, Jerman JC, Nasir S, Gray J, Muir AI, Chambers JK, Randall AD, Davis JB. (2000) The endogenous lipid anandamide is a full agonist at the human vanilloid receptor (hVR1)Br J Pharmacol 129:227–230 [PMC free article] [PubMed[]
  • Smith NJ, Stoddart LA, Devine NM, Jenkins L, Milligan G. (2009) The action and mode of binding of thiazolidinedione ligands at free fatty acid receptor 1J Biol Chem 284:17527–17539 [PMC free article] [PubMed[]
  • Soga T, Ohishi T, Matsui T, Saito T, Matsumoto M, Takasaki J, Matsumoto S, Kamohara M, Hiyama H, Yoshida S, et al. (2005) Lysophosphatidylcholine enhances glucose-dependent insulin secretion via an orphan G-protein-coupled receptorBiochem Biophys Res Commun 326:744–751 [PubMed[]
  • Spivak CE, Lupica CR, Oz M. (2007) The endocannabinoid anandamide inhibits the function of α4β2 nicotinic acetylcholine receptorsMol Pharmacol 72:1024–1032 [PubMed[]
  • Starowicz K, Nigam S, Di Marzo V. (2007) Biochemistry and pharmacology of endovanilloidsPharmacol Ther 114:13–33 [PubMed[]
  • Steffens M, Zentner J, Honegger J, Feuerstein TJ. (2005) Binding affinity and agonist activity of putative endogenous cannabinoids at the human neocortical CB1 receptorBiochem Pharmacol 69:169–178 [PubMed[]
  • Stoddart LA, Smith NJ, Milligan G. (2008) International Union of Pharmacology. LXXI. Free fatty acid receptors FFA1, -2, and -3: pharmacology and pathophysiological functionsPharmacol Rev 60:405–417 [PubMed[]
  • Suardíaz M, Estivill-Torrús G, Goicoechea C, Bilbao A, Rodríguez de Fonseca F. (2007) Analgesic properties of oleoylethanolamide (OEA) in visceral and inflammatory painPain 133:99–110 [PubMed[]
  • Sugiura T, Kondo S, Sukagawa A, Nakane S, Shinoda A, Itoh K, Yamashita A, Waku K. (1995) 2-Arachidonoylglycerol: a possible endogenous cannabinoid receptor ligand in brainBiochem Biophys Res Commun 215:89–97 [PubMed[]
  • Sugo T, Tachimoto H, Chikatsu T, Murakami Y, Kikukawa Y, Sato S, Kikuchi K, Nagi T, Harada M, Ogi K, et al. (2006) Identification of a lysophosphatidylserine receptor on mast cellsBiochem Biophys Res Commun 341:1078–1087 [PubMed[]
  • Suhara Y, Nakane S, Arai S, Takayama H, Waku K, Ishima Y, Sugiura T. (2001) Synthesis and biological activities of novel structural analogues of 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligandBioorg Med Chem Lett 11:1985–1988 [PubMed[]
  • Suhara Y, Takayama H, Nakane S, Miyashita T, Waku K, Sugiura T. (2000) Synthesis and biological activities of 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligand, and its metabolically stable ether-linked analoguesChem Pharm Bull 48:903–907 [PubMed[]
  • Sun Y, Alexander SP, Garle MJ, Gibson CL, Hewitt K, Murphy SP, Kendall DA, Bennett AJ. (2007a) Cannabinoid activation of PPARα; a novel neuroprotective mechanismBr J Pharmacol 152:734–743 [PMC free article] [PubMed[]
  • Sun Y, O’Sullivan SE, Alexander SPH, Kendall D, Bennett A. (2007b) Both agonists and antagonists of cannabinoid receptors can activate PPARα, in 17th Annual Symposium of the Cannabinoids; 2007 Jun 26–Jul 1; St-Sauveur, Canada p. 16 International Cannabinoid Research Society, Burlington, Vermont []
  • Szabo B, Schlicker E. (2005) Effects of cannabinoids on neurotransmissionHandb Exp Pharmacol 168:327–365 [PubMed[]
  • Tanaka S, Ishii K, Kasai K, Yoon SO, Saeki Y. (2007) Neural expression of G protein-coupled receptors GPR3, GPR6, and GPR12 up-regulates cyclic AMP levels and promotes neurite outgrowthJ Biol Chem 282:10506–10515 [PubMed[]
  • Tanaka S, Shaikh IM, Chiocca EA, Saeki Y. (2009) The Gs-linked receptor GPR3 inhibits the proliferation of cerebellar granule cells during postnatal developmentPLoS ONE 4:e5922. [PMC free article] [PubMed[]
  • Tanaka T, Katsuma S, Adachi T, Koshimizu TA, Hirasawa A, Tsujimoto G. (2008) Free fatty acids induce cholecystokinin secretion through GPR120Naunyn Schmiedebergs Arch Pharmacol 377:523–527 [PubMed[]
  • Thomas A, Baillie GL, Phillips AM, Razdan RK, Ross RA, Pertwee RG. (2007) Cannabidiol displays unexpectedly high potency as an antagonist of CB1 and CB2 receptor agonists in vitroBr J Pharmacol 150:613–623 [PMC free article] [PubMed[]
  • Tominaga M, Caterina MJ, Malmberg AB, Rosen TA, Gilbert H, Skinner K, Raumann BE, Basbaum AI, Julius D. (1998) The cloned capsaicin receptor integrates multiple pain-producing stimuliNeuron 21:531–543 [PubMed[]
  • Tóth A, Blumberg PM, Boczán J. (2009) Anandamide and the vanilloid receptor (TRPV1)Vitam Horm 81:389–419 [PubMed[]
  • Turu G, Várnai P, Gyombolai P, Szidonya L, Offertaler L, Bagdy G, Kunos G, Hunyady L. (2009) Paracrine transactivation of the CB1 cannabinoid receptor by AT1 angiotensin and other Gq/11 protein-coupled receptorsJ Biol Chem 284:16914–16921 [PMC free article] [PubMed[]
  • Uhlenbrock K, Gassenhuber H, Kostenis E. (2002) Sphingosine-1-phosphate is a ligand of the human gpr3, gpr6 and gpr12 family of constitutively active G protein-coupled receptorsCell Signal 14:941–953 [PubMed[]
  • Valk P, Verbakel S, Vankan Y, Hol S, Mancham S, Ploemacher R, Mayen A, Löwenberg B, Delwel R. (1997) Anandamide, a natural ligand for the peripheral cannabinoid receptor is a novel synergistic growth factor for hematopoietic cellsBlood 90:1448–1457 [PubMed[]
  • Valverde O, Célérier E, Baranyi M, Vanderhaeghen P, Maldonado R, Sperlagh B, Vassart G, Ledent C. (2009) GPR3 receptor, a novel actor in the emotional-like responsesPLoS ONE 4:e4704. [PMC free article] [PubMed[]
  • Van den Bossche I, Vanheel B. (2000) Influence of cannabinoids on the delayed rectifier in freshly dissociated smooth muscle cells of the rat aortaBr J Pharmacol 131:85–93 [PMC free article] [PubMed[]
  • van der Stelt M, Trevisani M, Vellani V, De Petrocellis L, Schiano Moriello A, Campi B, McNaughton P, Geppetti P, Di Marzo V. (2005) Anandamide acts as an intracellular messenger amplifying Ca2+ influx via TRPV1 channelsEMBO J 24:3026–3037 [PMC free article] [PubMed[]
  • Van Sickle MD, Duncan M, Kingsley PJ, Mouihate A, Urbani P, Mackie K, Stella N, Makriyannis A, Piomelli D, Davison JS, et al. (2005) Identification and functional characterization of brainstem cannabinoid CB2 receptorsScience 310:329–332 [PubMed[]
  • Vandevoorde S, Jonsson KO, Fowler CJ, Lambert DM. (2003) Modifications of the ethanolamine head in N-palmitoylethanolamine: synthesis and evaluation of new agents interfering with the metabolism of anandamideJ Med Chem 46:1440–1448 [PubMed[]
  • Vásquez C, Lewis DL. (1999) The CB1 cannabinoid receptor can sequester G-proteins, making them unavailable to couple to other receptorsJ Neurosci 19:9271–9280 [PMC free article] [PubMed[]
  • Vaysse PJ, Gardner EL, Zukin RS. (1987) Modulation of rat brain opioid receptors by cannabinoidsJ Pharmacol Exp Ther 241:534–539 [PubMed[]
  • Veale EL, Buswell R, Clarke CE, Mathie A. (2007) Identification of a region in the TASK3 two pore domain potassium channel that is critical for its blockade by methanandamideBr J Pharmacol 152:778–786 [PMC free article] [PubMed[]
  • Venkatachalam K, Montell C. (2007) TRP channelsAnnu Rev Biochem 76:387–417 [PMC free article] [PubMed[]
  • Venkataraman C, Kuo F. (2005) The G-protein coupled receptor, GPR84 regulates IL-4 production by T lymphocytes in response to CD3 crosslinkingImmunol Lett 101:144–153 [PubMed[]
  • Vennekens R, Owsianik G, Nilius B. (2008) Vanilloid transient receptor potential cation channels: an overviewCurr Pharm Des 14:18–31 [PubMed[]
  • Vignali M, Benfenati V, Caprini M, Anderova M, Nobile M, Ferroni S. (2009) The endocannabinoid anandamide inhibits potassium conductance in rat cortical astrocytesGlia 57:791–806 [PubMed[]
  • Wager-Miller J, Westenbroek R, Mackie K. (2002) Dimerization of G protein-coupled receptors: CB1 cannabinoid receptors as an exampleChem Phys Lipids 121:83–89 [PubMed[]
  • Waldeck-Weiermair M, Zoratti C, Osibow K, Balenga N, Goessnitzer E, Waldhoer M, Malli R, Graier WF. (2008) Integrin clustering enables anandamide-induced Ca2+ signaling in endothelial cells via GPR55 by protection against CB1-receptor-triggered repressionJ Cell Sci 121:1704–1717 [PMC free article] [PubMed[]
  • Wallace VC, Segerdahl AR, Lambert DM, Vandevoorde S, Blackbeard J, Pheby T, Hasnie F, Rice AS. (2007) The effect of the palmitoylethanolamide analogue, palmitoylallylamide (L-29) on pain behaviour in rodent models of neuropathyBr J Pharmacol 151:1117–1128 [PMC free article] [PubMed[]
  • Wang J, Simonavicius N, Wu X, Swaminath G, Reagan J, Tian H, Ling L. (2006b) Kynurenic acid as a ligand for orphan G protein-coupled receptor GPR35J Biol Chem 281:22021–22028 [PubMed[]
  • Wang J, Wu X, Simonavicius N, Tian H, Ling L. (2006a) Medium-chain fatty acids as ligands for orphan G protein-coupled receptor GPR84J Biol Chem 281:34457–34464 [PubMed[]
  • Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. (2003) Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channelsNature 424:434–438 [PubMed[]
  • Wetter JA, Revankar C, Hanson BJ. (2009) Utilization of the Tango β-Arrestin Recruitment technology for cell-based EDG receptor assay development and interrogationJ Biomol Screen 14:1134–1141 [PubMed[]
  • White R, Ho WS, Bottrill FE, Ford WR, Hiley CR. (2001) Mechanisms of anandamide-induced vasorelaxation in rat isolated coronary arteriesBr J Pharmacol 134:921–929 [PMC free article] [PubMed[]
  • Whyte LS, Ryberg E, Sims NA, Ridge SA, Mackie K, Greasley PJ, Ross RA, Rogers MJ. (2009) The putative cannabinoid receptor GPR55 affects osteoclast function in vitro and bone mass in vivoProc Natl Acad Sci USA 106:16511–16516 [PMC free article] [PubMed[]
  • Williams JR, Khandoga AL, Goyal P, Fells JI, Perygin DH, Siess W, Parrill AL, Tigyi G, Fujiwara Y. (2009) Unique ligand selectivity of the GPR92/LPA5 lysophosphatidate receptor indicates role in human platelet activationJ Biol Chem 284:17304–17319 [PMC free article] [PubMed[]
  • Wotherspoon G, Fox A, McIntyre P, Colley S, Bevan S, Winter J. (2005) Peripheral nerve injury induces cannabinoid receptor 2 protein expression in rat sensory neuronsNeuroscience 135:235–245 [PubMed[]
  • Xiong W, Hosoi M, Koo BN, Zhang L. (2008) Anandamide inhibition of 5-HT3A receptors varies with receptor density and desensitizationMol Pharmacol 73:314–322 [PubMed[]
  • Yamamoto I, Kimura T, Yoshida H, Watanabe K, Yoshimura H. (1992) Cannabinoid metabolite interacts with benzodiazepine receptorRes Commun Subst Abuse 13:299–313 []
  • Yan ZC, Liu DY, Zhang LL, Shen CY, Ma QL, Cao TB, Wang LJ, Nie H, Zidek W, Tepel M, et al. (2007) Exercise reduces adipose tissue via cannabinoid receptor type 1 which is regulated by peroxisome proliferator-activated receptor-δBiochem Biophys Res Commun 354:427–433 [PubMed[]
  • Yanagida K, Masago K, Nakanishi H, Kihara Y, Hamano F, Tajima Y, Taguchi R, Shimizu T, Ishii S. (2009) Identification and characterization of a novel lysophosphatidic acid receptor, p2y5/LPA6J Biol Chem 284:17731–17741 [PMC free article] [PubMed[]
  • Yang Z, Aubrey KR, Alroy I, Harvey RJ, Vandenberg RJ, Lynch JW. (2008) Subunit-specific modulation of glycine receptors by cannabinoids and N-arachidonyl-glycineBiochem Pharmacol 76:1014–1023 [PubMed[]
  • Yin H, Chu A, Li W, Wang B, Shelton F, Otero F, Nguyen DG, Caldwell JS, Chen YA. (2009) Lipid G protein-coupled receptor ligand identification using β-arrestin PathHunter assayJ Biol Chem 284:12328–12338 [PMC free article] [PubMed[]
  • Zhang X, Maor Y, Wang JF, Kunos G, Groopman JE. (2010) Endocannabinoid-like N-arachidonoyl serine is a novel pro-angiogenic mediatorBr J Pharmacol 160:1583–1594 [PMC free article] [PubMed[]
  • Zhang HT, Parker J, Shepherd N, Creazzo TL. (2009) Developmental expression of a functional TASK-I 2P domain K+ channel in embryonic chick heartJ Biomed Sci 16:104. [PMC free article] [PubMed[]
  • Zygmunt PM, Petersson J, Andersson DA, Chuang H, Sørgård M, Di Marzo V, Julius D, Högestätt ED. (1999) Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamideNature 400:452–457 [PubMed[]

Articles from Pharmacological Reviews are provided here courtesy of American Society for Pharmacology and Experimental Therapeutics

potp font 1