Cannabinoid Receptor Ligands

Professor R. G. Pertwee
Department of Biomedical Sciences, Institute of Medical Sciences, University of Aberdeen, Foresterhill, Aberdeen AB25 2ZD, Scotland

Roger Pertwee is Professor of Neuropharmacology at the University of Aberdeen. His research interests include the pharmacology of cannabinoid receptors, the physiological and pathophysiological roles of endogenous cannabinoid receptor ligands, and the therapeutic potential of cannabinoids.

The endocannabinoid system

Two types of cannabinoid receptor have so far been identified, CB1, cloned in 1990,1 and CB2, cloned in 1993.2 These have now been detected in several species including man, rat and mouse3 and both CB1 and CB2 receptor knockout mice have been developed.4-6 The CB1 receptors of different mammalian species exhibit a high level of similarity. For example, CB1 nucleotide sequences of human and rat are 90% identical, those of human and mouse 91% identical and those of rat and mouse 96% identical.7,8 CB2 receptors show greater interspecies differences, the deduced amino acid sequence of the mouse CB2 receptor differing from that of the human CB2 receptor in 60 residues (82% similarity).9 These differences are apparent mainly in the N-terminal extramembrane region. A spliced variant of CB1 cDNA, CB1a, has also been isolated.10,11 However, CB1a mRNA exists only as a minor transcript and there is no evidence for any notable difference between the pharmacology of CB1 and CB1a receptors.

CB1 mRNA has been detected both in the central nervous system and in certain peripheral tissues including pituitary gland, immune cells, reproductive tissues, gastrointestinal tissues, superior cervical ganglion, heart, lung, urinary bladder and adrenal gland.12 Some CB1 receptors are located at central and peripheral nerve terminals12,13,14 and, when activated, these receptors seem to suppress the neuronal release of one or other of a range of excitatory and inhibitory transmitters that include acetylcholine, noradrenaline, dopamine, 5­hydroxy-tryptamine, g-aminobutyric acid, glutamate and aspartate.14 CB2 mRNA is present mainly in immune cells with particularly high levels in B-cells and natural killer cells.15 Little is yet known in any detail about the physiological or pathophysiological roles of CB2 receptors. Most likely, these include immunomodulation which may depend at least in part on CB2 receptor-mediated suppression of proinflammatory cytokine release and enhancement of antiinflammatory cytokine release from immune cells.16,17 Thus one major role that CB1 and CB2 receptors may have in common is modulation of ongoing release of chemical messengers.

The central distribution pattern of CB1 receptors is heterogeneous and accounts for several prominent pharmacological properties of CB1 receptor agonists, for example their ability to impair cognition and memory and to alter the control of motor function. Thus the cerebral cortex, hippocampus, lateral caudate-putamen, substantia nigra pars reticulata, globus pallidus, entopeduncular nucleus and the molecular layer of the cerebellum are all populated with particularly high concentrations of CB1 receptors.12,18 In line with the analgesic properties of cannabinoid receptor agonists, CB1 receptors are also found on pain pathways in brain and spinal cord and at the peripheral terminals of primary sensory neurones.14 Although the concentration of CB1 receptors is considerably less in peripheral tissues than in the central nervous system, this does not imply that peripheral CB1 receptors are unimportant. This is because some peripheral tissues may contain high concentrations of CB1 receptors, localized in discrete regions such as nerve terminals that form only a small part of the total tissue mass. Peripheral tissues in which CB1 receptors are expressed on neurones include the heart, vas deferens, urinary bladder and small intestine.12

As detailed elsewhere,12,19 both CB1 and CB2 receptors are coupled through Gi/o proteins, negatively to adenylate cyclase and positively to mitogen-activated protein kinase. In addition, CB1 receptors are coupled to ion channels through Gi/o proteins, positively to A-type and inwardly rectifying potassium channels and negatively to N-type and P/Q-type calcium channels and to D-type potassium channels.12,19,20 The coupling to A-type and D-type potassium channels is thought to be through adenylate cyclase.20,21 Inwardly rectifying potassium channels can also serve as a signalling mechanism for the CB2 receptor, at least in Xenopus oocytes that have been transfected with such channels together with this receptor type.22,23 In addition, there is evidence from experiments with rat hippocampal CA1 pyramidal neurones that CB1 receptors are negatively coupled to M-type potassium channels.24 CB1 receptors may also mobilize arachidonic acid and close 5-HT3 receptor ion channels12 and, under certain conditions, couple to Gs proteins to activate adenylate cyclase25,26 and/or to reduce outward potassium K current, possibly through arachidonic acid-mediated stimulation of protein kinase C.27 The questions of whether CB1 receptor coupling to Gs proteins has physiological importance and of whether such coupling increases after Gi/o protein sequestration by co-localized non-cannabinoid Gi/o protein-coupled receptors have yet to be resolved. CB1 receptors have also been reported to be positively coupled to phospholipase C through G proteins in COS7 cells co-transfected with CB1 receptors and Ga subunits22 and negatively coupled to voltage-gated L-type calcium channels in cat cerebral arterial smooth muscle cells.28 One other recent finding is that CB1 receptors on cultured cerebellar granule neurones can operate through a phospholipase C-sensitive mechanism to enhance NMDA-elicited calcium release from inositol 1,4,5-triphosphate-gated intracellular stores.29

The discovery of cannabinoid receptors was followed in 1992 by the demonstration that arachidonoyl ethanolamide (anandamide) is an endogenous ligand for these receptors.30 Other endogenous cannabinoids (endocannabinoids) have since been identified, of which the most important is 2-arachidonoyl glycerol.31-33 Both anandamide and 2-arachidonoyl glycerol undergo depolarization-induced synthesis/release from neurones and are removed from the extracellular space by a carrier-mediated, saturable uptake process that is present in the membranes of neurones and astrocytes.31,34-37 There is also evidence that anandamide release in rat dorsal striatum can be triggered by the activation of dopamine D2, D3 and/or D4 receptors.38 Once within the cell, anandamide is hydrolysed to arachidonic acid and ethanolamine by fatty acid amide hydrolase (FAAH).31,36,39 FAAH can also catalyse the hydrolysis of 2-arachidonoyl glycerol,31,40 an indication that it has esterase as well as amidase activity. The distribution in rat brain of FAAH immunoreactivity is heterogeneous, as is the brain distribution of FAAH mRNA and FAAH enzymatic activity. In line with the putative roles of anandamide and 2-arachidonoyl glycerol as endogenous cannabinoids, these measures of FAAH brain distribution exhibit considerable though not complete overlap with the brain distribution of CB1 receptors.34,39,41-43 Unlike the endocannabinoid membrane transporter, which remains to be fully characterized, FAAH has been cloned44 and FAAH knockout mice developed (personal communication from Dr Benjamin Cravatt). At least one other anandamide-hydrolysing enzyme also exists, probably in lysozomes.43 Cannabinoid receptors and their endogenous ligands together constitute what is now usually referred to as the ‘endocannabinoid system’. These receptors, together with FAAH and the endocannabinoid membrane transporter, constitute important new molecular targets and the remainder of this review focuses on ligands that have been developed to interact selectively with these targets.

CB1 and CB2 receptor agonists

Cannabinoid receptor agonists fall essentially into four chemical groups: classical, nonclassical, aminoalkylindole and eicosanoid,45 (Table 1). Classical cannabinoids consist of dibenzopyran derivatives and are either plant-derived cannabinoids or their synthetic analogues. The most investigated of these include the psychotropic plant cannabinoids, D9-tetrahydrocannabinol (D9-THC) and D8-THC, and the synthetic cannabinoid, 11-hydroxy-D8-THC-dimethylheptyl (HU-210). The nonclassical cannabinoids were developed by a Pfizer research team.48 They are quite similar in structure to classical cannabinoids, consisting as they do of bicyclic and tricyclic analogues of D9-THC that lack a pyran ring. Important examples are CP 55940, CP 55244, CP 50556 (L-nantradol) and desacetyl-L-nantradol. The aminoalkylindole group was developed by a Sterling Winthrop research team,49,50 the prototype of this group being WIN 55,212-2 (R-(+)-WIN55212). This group contains compounds that are structurally quite different from classical or nonclassical cannabinoids and, indeed, there is evidence that WIN 55,212-2 binds differently to the CB1 receptor than classical and nonclassical cannabinoids, albeit it in a manner that still permits mutual displacement between WIN 55,212-2 and non-aminoalkylindole cannabinoids at CB1 and CB2 binding sites.45 The prototypic member of the eicosanoid group of cannabinoid receptor agonists is the endocannabinoid, anandamide (see above). Cannabinoid receptor agonists often contain chiral centres and these generally confer marked stereoselectivity in pharmacological assays. WIN 55,212-2 is more active than WIN 55,212-3 (S-(-)-WIN55212) and classical and nonclassical cannabinoids with the same absolute stereochemistry as (-)-D9-THC at 6a and 10a (6aR, 10aR ) have the greater activity. Anandamide itself does not contain any chiral centres. However, some of its synthetic analogues do, one example being methanandamide, the R-(+)-isomer of which has nine times greater affinity for CB1 receptors than the S-(-)-isomer.51

Table 1. Pharmacological properties of some important cannabinoid receptor agonists

Ligand CB1 Ki (nM) CB2 Ki (nM) Likely Relative Efficacy
CB1 CB2
ACEA (1319) 1.4 > 2000†† +++++ ?
ACPA (1318) 2.2 715†† ++++ ?
Methanandamide (1121) 18, 20 868††, 815†† ++++ ?
Anandamide (1339) 89* 371* ++++ +
2-AG 58.3, 472 145, 1400 ++++ ++
2-AG 13.9** 58** ++++ ++
HU-210 (0966) 0.06, 0.73 0.52, 0.22 +++++ +++++
D9-THC 40.7 36.4 +++ +
CP 55940 (0949) 0.58, 3.72, 5 0.69, 2.55, 1.8 +++++ +++++
WIN 55,212-2 (1038) 1.89, 62.3, 123 0.28, 3.3, 4.1 +++++ +++++
JWH-133 677 3.4 ? +++++

2-AG, 2-arachidonoyl glycerol. See text for full names of other agonists. *PMSF present. ** Inhibitors of the enzymic hydrolysis of 2-AG present.46 Binding assays were performed with [3H]CP 55940 or [3H]HU-243 (2-AG) using rat brain and rat or mouse spleen†† membranes or membranes from cultured cells transfected with CB1 or CB2 receptors. For further information see text and references 45 and 47.

The ability of cannabinoid receptor agonists that are commonly used in the laboratory to interact with cannabinoid receptors has been reviewed elsewhere.45,52,53 Essentially, (-)-D9-THC binds equally well to CB1 and CB2 receptors and is a partial agonist at both these receptor types. It has even less efficacy at CB2 than at CB1 receptors and, indeed, has been reported in one CB2 bioassay system to behave as an antagonist.54 (-)-D8-THC resembles (-)-D9-THC both in its affinities for CB1 and CB2 receptors and in its CB1 receptor efficacy. CP 55940 and WIN 55,212-2 have CB1 and CB2 affinities in the low nanomolar range and exhibit relatively high efficacy at both these receptor types. Results from binding experiments with CB1- and CB2-transfected cells indicate that CP 55940 has essentially the same affinity for CB1 and CB2 receptors. WIN 55,212-2 has slightly greater affinity for CB2 than for CB1 receptors whilst anandamide binds marginally more readily to CB1 than to CB2 receptors and, when protected from enzymic hydrolysis, exhibits a CB1 affinity similar to that of (-)-D9-THC. Anandamide also resembles (-)-D9-THC in behaving as a partial agonist at CB1 and CB2 receptors and in exhibiting lower CB2 than CB1 efficacy. One commercially available classical cannabinoid, HU-210, has efficacies at CB1 and CB2 receptors that match those of CP 55940 and WIN 55,212-2 and affinities for CB1 and CB2 receptors that exceed those of these other cannabinoids. As a result it is a particularly potent cannabinoid receptor agonist. Its pharmacological effects in vivo are also exceptionally long-lasting. The enhanced affinity and efficacy shown by HU-210 at cannabinoid receptors can be largely attributed to the replacement of the pentyl side chain of D8-THC with a dimethylheptyl group.

One other classical cannabinoid receptor agonist that merits special mention is 3-(5´-cyano-1´,1´-dimethyl-pentyl)-1-(4-N-morpholinobutyryloxy)-D8-THC hydrochloride (O-1057).55 This stands out from established cannabinoid receptor agonists in being readily soluble in water. The potency of O-1057 relative to that of CP 55940 is just 2.9 times less at CB1 receptors and 6.4 times less at CB2 receptors. The availability of a water-soluble cannabinoid will facilitate cannabinoid delivery not only in the laboratory but also in the clinic, particularly where administration to patients is to be by injection or aerosol inhalation. Another important advance has been the development of cannabinoid receptor agonists that, unlike the agents already mentioned, show marked differences in their abilities to interact with CB1 and CB2 receptors.45,47 For the development of CB1-selective agonists, the starting point has been the anandamide molecule, the marginal CB1 selectivity of which can be significantly enhanced by inserting a fluorine atom on the terminal 2´ carbon to form O-585 and/or by replacing a hydrogen atom on the 1´ or 2 carbon with a methyl group to form R-(+)-methanandamide or O-689.45 Another important consequence of inserting a methyl group on the 1´ or 2 carbon is greater resistance to the hydrolytic action of FAAH and, indeed, R-(+)-methanandamide was first synthesized in order to meet the need for a metabolically more stable anandamide analogue. The most CB1-selective agonists so far developed are arachidonyl-2´-chloroethylamide (ACEA) and arachidonylcyclo-propylamide (ACPA), both of which behave as potent CB1 receptor agonists with reasonably high efficacy.56 However, unlike methanandamide and O-689,45 neither ACEA nor ACPA show any sign of reduced susceptibility to enzymic hydrolysis. This is presumably because they lack a methyl substituent on the 1´ or 2 carbon and, indeed, it was recently shown that the addition of a methyl group to the 1´ carbon of ACEA does markedly decrease the susceptibility of this molecule to FAAH-mediated hydrolysis.57 This structural change also reduces the affinity of ACEA for CB1 receptors by about 14-fold. As to CB2-selective agonists, the best to have been developed so far are all classical cannabinoids: L-759633, L-759656, JWH-133 and HU-308. Each of these agents not only binds more readily to CB2 than to CB1 receptors but also behaves as a potent CB2-selective agonist in functional assays.47,58,59 Conformational requirements for the interaction of endocannabinoids not only with cannabinoid receptors but also with FAAH and with the anandamide membrane transporter (see below) have been reviewed recently by Reggio and Traore.52

Selective CB1 receptor antagonists/inverse agonists

The first of these, the diarylpyrazole, SR141716A, was developed by Sanofi.45,60 This is a highly potent and selective CB1 receptor ligand (Table 2) that readily prevents or reverses CB1-mediated effects both in vitro and in vivo. There is convincing evidence that SR141716A is not a “silent” antagonist. Thus, as well as attenuating effects of CB1 receptor agonists, SR141716A can by itself also elicit responses in some CB1 receptor-containing tissues that are opposite in direction from those elicited by CB1 receptor agonists. Whilst such “inverse cannabimimetic effects” may in some instances be attributable to a direct antagonism of responses evoked at CB1 receptors by released endocannabinoids, there is evidence that this is not the only possible mechanism and that SR141716A is in fact an inverse agonist.63-65 Thus SR141716A may produce inverse cannabimimetic effects in at least some tissues by somehow reducing the constitutive activity of CB1 receptors (the coupling of CB1 receptors to their effector mechanisms that it is thought can occur in the absence of exogenously added or endogenously produced CB1 agonists). Another notable CB1-selective antagonist that exhibits inverse CB1 receptor properties in some assay systems is LY320135.45 This agent, which was developed by Eli Lilly, shares the ability of SR141716A to bind much more readily to CB1 than CB2 receptors (Table 2). However it has less affinity for CB1 receptors than SR141716A and, at concentrations in the low micromolar range, also binds to muscarinic and 5HT2 receptors. Although neither SR141716A nor LY320135 are commercially available, it is possible to purchase structural analogues of SR141716A. These are AM 251 and AM 281 which have been found to be respectively three and eight times less potent than SR141716A in displacing [3H]SR141716A from binding sites on mouse cerebellar membranes.66 AM 281 has also been reported to bind more readily to CB1 than CB2 receptors (Table 2) and to attenuate the ability of WIN 55,212-2 or CP 55940 to decrease rat locomotor activity, to inhibit single population spikes and evoked acetylcholine release in rat hippocampal slices, to inhibit electrically-evoked contractions of the myenteric plexus-longitudinal muscle preparation of guinea-pig ileum and to suppress guinea-pig intestinal peristalsis.62,67-70 Like SR141716A, AM 281 can behave as an inverse agonist when administered alone.68-70

Table 2. Pharmacological properties of some important receptor partial agonists/antagonists/inverse agonists

Ligand CB1 Ki (nM) CB2 Ki (nM) Pharmacological Classification
CB1 CB2
LY320135 141 14900 A/I ?
SR141716A 5.6, 11.8, 11.8, 12.3 > 1000, 973, 13200, 702 A/I A/(I?)
AM 281 (1115) 12 4200†† A/I ?
O-1238 3.5 7.8 P A/P
O-1184 5.2 7.4 A/P A/I*
AM 630 (1120) 5152 31.2 A/P/I A/I*
SR144528 437, > 10,000 0.6, 5.6 A/(I?) A/(I?)
MAFP 20 (IC50) ? iA ?

A, surmountable antagonist; iA, ireversible antagonist; I, inverse agonist; P, partial agonist. Main pharmacological classification for each ligand is shown in bold. See text for full names of ligands. *Less inverse agonist efficacy at CB2 receptors than SR144528. MAFP is also a potent irreversible inhibitor of FAAH (see text). Binding assays were performed with [3H]CP 55940 using rat forebrain and mouse spleen†† membranes or membranes from cultured cells transfected with CB1 or CB2 receptors. For further information see text and references 45, 47, 61 and 62.

In some in vitro bioassay systems, for example rat and guinea-pig isolated arteries and mouse isolated vas deferens, SR141716A has been reported to be less potent against anandamide than against non-eicosanoid CB1 receptor agonists.71,72 There are also reports that in certain in vivo bioassay systems, SR141716A fails to antagonize effects of anandamide altogether when it is given at doses that markedly antagonize the effects of other CB1 receptor agonists in the same bioassays.73-75 For the isolated tissue preparations at least, this discrepancy often seems to stem from the ability of anandamide to activate vanilloid as well as cannabinoid receptors.71,72,76,77 Analogues of anandamide that share this ability to bind to both cannabinoid and vanilloid receptors include two anandamide membrane transport inhibitors, AM 404 and arvanil.71,78-80 Of these, arvanil is a hybrid of anandamide and the established vanilloid receptor agonist, capsaicin. As is to be expected, the structural requirements for activation of cannabinoid and vanilloid receptors are not the same.71,72,76,77 Whether the ability of anandamide to activate vanilloid receptors has physiological as well as pharmacological significance remains to be established.

CB2 receptor antagonists/inverse agonists

The most notable of these is the Sanofi compound, SR144528, a diarylpyrazole that binds with markedly higher affinity to CB2 than CB1 receptors (Table 2).81 There is good evidence that, like SR141716A, SR144528 is not a “silent” antagonist but rather an inverse agonist that can, by itself, produce inverse cannabimimetic effects at CB2 receptors.59,81 Another CB2-selective antagonist/inverse agonist is 6-iodopravadoline (AM 630; Table 2) which potently reverses CP 55940-induced inhibition of forskolin-stimulated cyclic AMP production by human CB2-transfected CHO cell preparations (EC50 = 129 nM) and enhances forskolin-stimulated cyclic AMP production by the same cell line when administered by itself (EC50 = 230 nM).59 The inverse efficacy of AM 630 at CB2 receptors appears to be less than that of SR144528.82 As to the ability of AM 630 to interact with CB1 receptors, results from several investigations when taken together suggest that this ligand has mixed agonist-antagonist properties and that it is a low-affinity partial CB1 agonist.45,59,83-85 There is also one report that it can behave as a low-potency inverse agonist at CB1 receptors.86

One cannabinoid receptor ligand that is close to being a silent cannabinoid receptor antagonist, albeit at both CB1 and CB2 receptors, is the classical cannabinoid, 6´-azidohex-2´-yne-D8-THC (O-1184; Table 2). In addition to a terminal N3 group, the alkyl side chain of this ligand contains a carbon-carbon triple bond, a structural modification that reduces CB1 and CB2 efficacy but not CB1 or CB2 affinity.82 O-1184 behaves as a high-affinity low-efficacy agonist at CB1 receptors and as a high-affinity low-efficacy inverse agonist at CB2 receptors.82,87 O-1238, in which the carbon-carbon triple bond of O-1184 is replaced by a carbon-carbon double bond, has higher efficacy than O-1184 at CB1 receptors and behaves as a high-affinity low-efficacy partial agonist at CB2 receptors (Table 2).82

Radiolabelled cannabinoid receptor ligands

Tritiated cannabinoid receptor ligands that have been most widely used in binding assays or for autoradiography are the CB1-selective [3H]SR141716A (CB1 Kd = 0.19 to 1.24 nM) and [3H]CP 55940,
[3H]WIN 55,212-2 and [3H]HU-243, all three of which bind more or less equally well to CB1 and CB2 receptors. Typical Kd values for [3H]CP 55940, [3H]WIN 55,212-2 and [3H]HU-243 are 0.07 to 4 nM, 1.9 to 16.2 nM and 0.045 nM respectively at CB1 receptors and 0.2 to 7.4 nM, 2.1 to 3.8 nM and
0.061 nM respectively at CB2 receptors.45 Thus [3H]HU-243 is particularly potent. Three radiolabelled ligands have also been developed as potential probes for human single photon emission computed tomography (SPECT) or positron emission tomography (PET) experiments. These are
123I-labelled analogues of AM 251 (CB1 Kd = 0.23 to 0.62 nM) and AM 28166,88,89 and an 18F-labelled analogue of SR141716A (SR144385).90 Particularly promising results have been obtained from animal experiments with [123I]AM 281.66

Inhibitors of the enzymic hydrolysis of endocannabinoids

The presence of FAAH in many cannabinoid bioassay systems has created the need for FAAH inhibitors that can be used to protect endocannabinoids from enzymic hydrolysis.12 The need for FAAH inhibitors also stems from the possibility that FAAH inhibitors might have therapeutic potential as indirect cannabinoid receptor agonists, the expectation being that drugs that activate the endocannabinoid system by increasing the concentration of endocannabinoids at cannabinoid receptors would be more selective than direct agonists. This is because they are unlikely to affect all parts of the endocannabinoid system at one time, producing instead effects only at sites where on-going production of endocannabinoids is taking place. The drug that has been most widely used to inhibit the enzymic hydrolysis of anandamide is the non-specific serine protease inhibitor, phenylmethyl-sulphonyl fluoride.12

FAAH inhibitors with much greater potency are now available, two notable examples being palmitylsulphonyl fluoride (AM 374) and stearylsulphonyl fluoride (AM 381). Both ligands inhibit FAAH irreversibly and show good separation between potency for FAAH inhibition and ability to bind to CB1 receptors. Thus the EC50 value of AM 374 is 7 nM for inhibition of FAAH and 520 nM for displacement of [3H]CP 55940 from specific binding sites on rat forebrain membranes whilst the corresponding EC50 values of AM 381 are 4 nM and 18.5 mM respectively.91 AM 374 potentiates both anandamide-induced inhibition of evoked [3H]acetylcholine release in rat hippocampal slices92 and anandamide-induced suppression of rat operant lever pressing and open field locomotor activity.93 When administered by themselves, both AM 374 and the anandamide membrane transport inhibitor AM 404 (see below) share the ability of cannabinoid receptor agonists to ameliorate spasticity in mice with chronic relapsing experimental allergic encephalomyelitis (CREAE), an induced syndrome that serves as an animal model of multiple sclerosis.94,95 This they most likely do by augmenting levels of endogenously released endocannabinoid(s) as brain and spinal cord concentrations of both anandamide and 2-AG are greater in spastic CREAE mice than in control animals.95 The ability of AM 381 to potentiate anandamide remains to be reported.

Methyl arachidonyl fluorophosphonate (MAFP) is another potent irreversible inhibitor of FAAH (EC50 = 2.5 nM).61 However, it also potently displaces [3H]CP 55940 from specific binding sites on rat brain membranes in an irreversible manner (EC50 = 20 nM)61 and produces insurmountable antagonism at CB1 receptors (Table 2).96 One recently developed analogue of MAFP, O-1887, shows much greater separation of FAAH inhibitory potency (EC50 = 15 nM) from potency for binding to CB1 receptors (CB1 Ki > 10 mM).97 Interestingly, this compound shares the ability of PMSF to elicit cannabimimetic responses when administered to mice by itself.97,98 Whether this is through inhibition of the enzymic hydrolysis of endogenously released cannabinoids has yet to be investigated. Other reasonably potent FAAH inhibitors are diazomethyl arachidonyl ketone (EC50 = 520 nM)99 and (E)-6-(bromomethylene) tetrahydro-3-(1-naphthalenyl)-2H-pyran-2-one (EC50 = 800 nM),100 which act irreversibly, and arachidonyl trifluoromethyl ketone (EC50 = 900 nM), which is a tight-binding but reversible FAAH inhibitor that also inhibits cytosolic phospholipase A2.91,101 Arachidonyl trifluoromethyl ketone and diazomethyl arachidonyl ketone have been reported to bind to CB1 receptors with Ki values of 0.65 and 1.3 mM respectively99,102 and there is also a report that the EC50 of arachidonyl trifluoromethyl ketone for displacement of [3H]CP 55940 from CB1 binding sites is 2.5 mM.91 Significantly more potent as FAAH inhibitors are a series of a-keto bicyclic heterocycles with alkyl or phenylalkyl side chains that have recently been developed by Boger et al.103 These inhibit the enzyme competitively, some with Ki values in the picomolar or low nanomolar range. The structure of the most potent of this new generation of FAAH inhibitors is shown in Figure 6 (Compound 59; Ki = 140 pM). Other pharmacological properties of these inhibitors, for example their ability to interact with cannabinoid or vanilloid receptors or to potentiate endocannabinoids, have yet to be reported.

It has been proposed that since endogenous substrates for FAAH such as linoleoylethanolamide, oleoylethanolamide and oleamide serve as reversible inhibitors of this enzyme,36,104-106 anandamide may be protected from enzymic hydrolysis by endogenously produced FAAH inhibitors.32 This putative “entourage” effect may well extend to 2-arachidonoyl glycerol which can be protected from the esterase action of FAAH (and so potentiated) by the endogenous fatty acid derivatives, 2-linoleyl glycerol and 2-palmitoyl glycerol.32,46,107 Other inhibitors of FAAH have also been identified and details of these appear elsewhere.18,39,43,57,97,101,102,108-110

Inhibitors of endocannabinoid membrane transport

One notable membrane transport inhibitor to have been developed is N-(4-hydroxyphenyl) arachidonylamide (AM 404). This has been reported to inhibit anandamide uptake by rat cultured cortical neurones (EC50 = 1 mM) and astrocytes (EC50 = 5 mM) and to potentiate anandamide both in vitro and in vivo.35,111 When administered to rats by itself, AM 404 increases plasma levels of anandamide and shares the ability of this endocannabinoid to decrease locomotor activity, depress plasma levels of prolactin and alter tyrosine hydroxylase activity in the hypothalamus (increase) and substantia nigra (decrease).112-114 The inhibitory effect of AM 404 on locomotor activity has been found to be susceptible to antagonism by SR141716A.113,114 AM 404 does not, however, elicit two other typical responses to CB1 receptor agonists in rats: catalepsy and signs of analgesia in the hot plate test.113

Structure-activity experiments with AM 404 analogues have revealed major differences between the structural requirements of the transporter for ligand recognition and those for ligand translocation.35 At concentrations in the low micromolar range, it displaces [3H]CP 55940 from specific binding sites on rat forebrain membranes (Ki = 1.76 mM).115 Even so, there are no reports that AM 404 behaves as a cannabinoid receptor agonist or antagonist. There is also one report that AM 404 inhibits FAAH with an EC50 of 0.5 mM.57 However, this conflicts with other reports that the EC50 of AM 404 for inhibition of FAAH is 22 mM116 or exceeds 30 mM.111 Submicromolar concentrations of AM 404 can activate vanilloid receptors,78,80 prompting the need for a more selective inhibitor of endocannabinoid membrane transport. This need was recently addressed by the development of the AM 404 analogue, VDM11, which exhibits the same potency as AM 404 as an endocannabinoid membrane transport inhibitor in C6 rat glioma cells (EC50 = 10 mM) but markedly less efficacy than AM 404 at vanilloid receptors.116 Although membrane transport-inhibiting concentrations of VDM11 undergo some degree of binding to CB1 receptors, these concentrations (5-10 mM) displace [3H]SR141716A from CB1 binding sites by less than 50%.116 VDM11 also exhibits relatively low potency as an inhibitor of FAAH (EC50 > 50 mM). Another ligand that inhibits the endocannabinoid membrane transporter (EC50 = 3.6 mM) more readily than it inhibits FAAH (EC50 = 32 mM), at least in RBL-2H3 cells, is arvanil.79 However this agent not only has reasonably high affinity for CB1 receptors (Ki = 0.5 mM) but is also a potent vanilloid receptor agonist (human and rat VR1 EC50 = 0.5 nM).71,79,116 Details of other membrane transport inhibitors are to be found elsewhere.35,37,57,79,110,117

One other important recent advance has been the development of a fluorescent substrate for anandamide uptake (SKM 4-45-1) that should serve as a useful experimental tool for transporter studies.117 This agent is more potent as an inhibitor of endocannabinoid membrane transport (EC50 = 7.8 mM) than as an FAAH inhibitor (EC50 > 10 mM). Whilst the ability of SKM 4-45-1 to interact with vanilloid receptors remains to be investigated, it is known that, at 3 mM, it does not displace [3H]CP 55940 from CB1 binding sites on rat brain membranes.117

Other cannabinoid receptors?

Calignano et al118 have postulated the existence of an SR144528-sensitive non-CB2 cannabinoid receptor (‘CB2-like’ receptor). This hypothesis is based on evidence that even though palmitoylethanolamide lacks significant affinity for CB1 or CB2 receptors,30,119-121 its ability to produce signs of anti-hyperalgesia in the mouse formalin paw test is readily attenuated by the CB2-selective antagonist/inverse agonist, SR144528.118 The existence of CB2-like receptors in the mouse vas deferens has also been proposed.122 More recently, evidence was obtained for the presence in vascular endothelium of an SR141716A-sensitive non-CB1, non-CB2, non-vanilloid receptor that is unresponsive to established non-eicosanoid CB1/CB2 receptor agonists but can be activated both by the eicosanoid cannabinoids, anandamide and methanandamide, and by certain classical cannabinoids that do not act through CB1 or vanilloid receptors (“abnormal cannabidiol” and its more potent analogue, (O-1602).123,124 Interestingly, two effects of abnormal cannabidiol, hypotension and mesenteric vasodilation, were found to be antagonized by the non-psychotropic classical cannabinoid, cannabidiol. So too was anandamide-induced mesenteric vasodilation. Finally, the existence in the brain of non-CB1, non-CB2, SR141716A-insensitive G protein-coupled receptors for anandamide has recently been proposed to explain results obtained from experiments with CB1 knockout mice.125

Conclusions

The field of cannabinoid research is now at a particularly exciting stage. Thus, whilst a number of major advances have recently been made, many important questions still remain unanswered or incompletely addressed so prompting the need for more research at both non-clinical and clinical levels. Of particular importance at the pharmacological level is the need:

  • to understand the modes of action of anandamide and 2-arachidonoyl glycerol more fully and to characterize their effects and the processes that terminate these effects more completely,
  • to develop silent cannabinoid receptor antagonists possibly through further elucidation of the structural features that attenuate agonist and inverse agonist efficacy at cannabinoid receptors without reducing cannabinoid receptor affinity,
  • to seek out any additional endocannabinoids or novel types of cannabinoid receptor,
  • to achieve a fuller pharmacological characterization of some existing inhibitors of endocannabinoid membrane transport and enzymic hydrolysis,
  • to develop inhibitors of endocannabinoid membrane transport and enzymic hydrolysis that show greater separation between their sought-after inhibitory actions and other pharmacological actions, particularly the ability to activate or block cannabinoid, vanilloid or other receptors or to inhibit other enzymes or transporters,
  • to obtain more detailed information about endocannabinoid release in both health and disease and hence gain a better appreciation of the therapeutic potential of drugs that inhibit endocannabinoid membrane transport and/or endocannabinoid enzymic hydrolysis,
  • to search for beneficial and unwanted consequences of selective inhibition of both endocannabinoid membrane transport and the enzymic hydrolysis of endocannabinoids.

Note added in proof: the presence in the brain of a novel CB1-selective endocannabinoid, 2-arachidonoyl glyceryl ether (noladin ether), has just been announced.126

References

1. Matsuda et al (1990) Nature 346 561.
2. Munro et al (1993) Nature 365 61.
3. Matsuda (1997) Crit.Revs.Neurobiol. 11 143.
4. Buckley et al (2000) Eur.J.Pharmacol. 396 141.
5. Ledent et al (1999) Science 283 401.
6. Zimmer et al (1999) Proc.Natl.Acad.Sci.USA 96 5780.
7. Chakrabarti et al (1995) DNA Sequence 5 385.
8. Gérard et al (1990) Nucleic Acids Res. 18 7142.
9. Shire et al (1996) Biochim.Biophys.Acta 1307 132.
10. Rinaldi-Carmona et al (1996) J.Pharmacol.Exp.Ther. 278 871.
11. Shire et al (1995) J.Biol.Chem. 270 3726.
12. Pertwee (1997) Pharmacol.Ther. 74 129.
13. Ong and Mackie (1999) Neuroscience 92 1177.
14. Pertwee (2001) Progr.Neurobiol. 63 569.
15. Galiègue et al (1995) Eur.J.Biochem. 232 54.
16. Molina-Holgado et al (1999) Biodrugs 12 317.
17. Berdyshev (2000) Chem.Phys.Lipids 108 169.
18. Pertwee (1998) In Cannabis: The Genus Cannabis. Ed. Brown, D.T., p 125, Harwood Academic Publishers, Amsterdam.
19. Howlett and Mukhopadhyay (2000) Chem.Phys.Lipids 108 53.
20. Mu et al (1999) J.Pharmacol.Exp.Ther. 291 893.
21. Mu et al (2000) Pflügers Arch.Eur.J.Physiol. 439 541.
22. Ho et al (1999) Receptors & Channels 6 363.
23. McAllister et al (1999) J.Pharmacol.Exp.Ther. 291 618.
24. Schweitzer (2000) J.Neurosci. 20 51.
25. Glass and Felder (1997) J.Neurosci. 17 5327. 25.
26. Calandra et al (1999) Eur.J.Pharmacol. 374 445.
27. Hampson et al (2000) J.Neurophysiol. 84 2356.
28. Gebremedhin et al (1999) Am.J.Physiol. 276 H2085.
29. Netzeband et al (1999) J.Neurosci. 19 8765.
30. Devane et al (1992) Science 258 1946.
31. Di Marzo et al (1998) TiNS 21 521.
32. Mechoulam et al (1998) Eur.J.Pharmacol. 359 1.
33. Sugiura and Waku (2000) Chem.Phys.Lipids 108 89.
34. Di Marzo (1999) Life Sci. 65 645.
35. Piomelli et al (1999) Proc.Natl.Acad.Sci.USA 96 5802.
36. Maccarrone et al (1998) J.Biol.Chem. 273 32332.
37. Hillard and Jarrahian (2000) Chem.Phys.Lipids 108 123.
38. Giuffrida et al (1999) Nature Neurosci. 2 358.
39. Ueda et al (2000) Chem.Phys.Lipids 108 107.
40. Khanolkar and Makriyannis (1999) Life Sci. 65 607.
41. Thomas et al (1997) J.Neurosci.Res. 50 1047.
42. Tsou et al (1998) Neurosci.Lett. 254 137.
43. Ueda and Yamamoto (2000) Prostaglandins & Other Lipid Mediators 61 19.
44. Cravatt et al (1996) Nature 384 83.
45. Pertwee (1999) Curr.Med.Chem. 6 635.
46. Ben-Shabat et al (1998) Eur.J.Pharmacol. 353 23.
47. Pertwee (2000) Expert Opin.Invest.Drugs 9 1553.
48. Johnson and Melvin (1986) In Cannabinoids as Therapeutic Agents. Ed. Mechoulam, R., p 121, CRC Press, Boca Raton.
49. Martin et al (1991) Pharmacol.Biochem.Behav. 40 471.
50. Pacheco et al (1991) J.Pharmacol.Exp.Ther. 257 170.
51. Abadji et al (1994) J.Med.Chem. 37 1889.
52. Reggio and Traore (2000) Chem.Phys.Lipids 108 15.
53. Khanolkar et al (2000) Chem.Phys.Lipids 108 37.
54. Bayewitch et al (1996) J.Biol.Chem. 271 9902.
55. Pertwee et al (2000) Br.J.Pharmacol. 129 1577.
56. Hillard et al (1999) J.Pharmacol.Exp.Ther. 289 1427.
57. Jarrahian et al (2000) J.Neurochem. 74 2597.
58. Hanus et al (1999) Proc.Natl.Acad.Sci.USA 96 14228.
59. Ross et al (1999) Br.J.Pharmacol. 126 665.
60. Rinaldi-Carmona et al (1994) FEBS Lett. 350 240.
61. Deutsch et al (1997) Biochem.Pharmacol. 53 255.
62. Lan et al (1999) AAPS Pharmsci. 1 U13, article 4, http://www.pharmsci.org/.
63. Bouaboula et al (1997) J.Biol.Chem. 272 22330.
64. Coutts et al (2000) Br.J.Pharmacol. 129 645.
65. MacLennan et al (1998) Br.J.Pharmacol. 124 619.
66. Gatley et al (1998) J.Neurochem. 70 417.
67. Al-Hayani and Davies (2000) Br.J.Pharmacol. 131 663.
68. Izzo et al (2000) Br.J.Pharmacol. 129 984.
69. Gifford et al (1997) Neurosci.Lett. 238 84.
70. Cosenza et al (2000) Synapse 38 477.
71. Ross et al (2001) Br.J.Pharmacol. 132 631.
72. Zygmunt et al (1999) Nature 400 452.
73. Adams et al (1998) J.Pharmacol.Exp.Ther. 284 1209.
74. Di Marzo et al (2000) Eur.J.Pharmacol. 406 363.
75. Smith et al (1998) Pharmacol.Biochem.Behav. 60 183.
76. Smart et al (2000) Br.J.Pharmacol. 129 227.
77. Smart and Jerman (2000) TiPS 21 134.
78. Jerman et al (2000) Br.J.Pharmacol. (Proc. Suppl.) 129 73P.
79. Melck et al (1999) Biochem.Biophys.Res.Comm. 262 275.
80. Zygmunt et al (2000) Eur.J.Pharmacol. 396 39.
81. Rinaldi-Carmona et al (1998) J.Pharmacol.Exp.Ther. 284 644.
82. Ross et al (1999) Br.J.Pharmacol. 128 735.
83. Hosohata et al (1997) Life Sci. 61 PL115.
84. Hosohata et al (1997) Eur.J.Pharmacol. 321 R1.
85. Pertwee et al (1996) Br.J.Pharmacol. 118 2199.
86. Landsman et al (1998) Life Sci. 62 PL109.
87. Ross et al (1998) Br.J.Pharmacol. 125 1345.
88. Gatley et al (1997) Life Sci. 61 PL191.
89. Lan et al (1996) J.Labelled Comp.Radiopharm. 38 875.
90. Barth (1998) Exp.Opin.Ther.Patents 8 301.
91. Deutsch et al (1997) Biochem.Biophys.Res.Comm. 231 217.
92. Gifford et al (1999) Eur.J.Pharmacol. 383 9.
93. Cervone et al (1999) Proc.Soc.Neurosci. 25 2077.
94. Baker et al (2000) Nature 404 84.
95. Baker et al (2001) FASEB J. 15 300.
96. Fernando and Pertwee (1997) Br.J.Pharmacol. 121 1716.
97. Martin et al (2000) J.Pharmacol.Exp.Ther. 294 1209.
98. Compton and Martin (1997) J.Pharmacol.Exp.Ther. 283 1138.
99. Edgemond et al (1998) J.Pharmacol.Exp.Ther. 286 184.
100. Beltramo et al (1997) FEBS Lett. 403 263.
101. Di Marzo and Deutsch (1998) Neurobiol.Dis. 5 386.
102. Koutek et al (1994) J.Biol.Chem. 269 22937.
103. Boger et al (2000) Proc.Natl.Acad.Sci.USA 97 5044.
104. Desarnaud et al (1995) J.Biol.Chem. 270 6030.
105. Maurelli et al (1995) FEBS Lett. 377 82.
106. Mechoulam et al (1997) Nature 389 25.
107. Gallily et al (2000) Eur.J.Pharmacol. 406 R5.
108. Hillard and Campbell (1997) J.Lipid Res. 38 2383.
109. Pertwee (1998) Biochem.Soc.Transact. 26 267.
110. Bisogno et al (2000) Biochem.J. 351 817.
111. Beltramo et al (1997) Science 277 1094.
112. González et al (1999) Life Sci. 65 327.
113. Beltramo et al (2000) J.Neurosci. 20 3401.
114. Giuffrida et al (2000) Eur.J.Pharmacol. 408 161.
115. Khanolkar et al (1996) J.Med.Chem. 39 4515.
116. De Petrocellis et al (2000) FEBS Lett. 483 52.
117. Muthian et al (2000) J.Pharmacol.Exp.Ther. 293 289.
118. Calignano et al (1998) Nature 394 277.
119. Lambert et al (1999) Biochim.Biophys.Acta 1440 266.
120. Sheskin et al (1997) J.Med.Chem. 40 659.
121. Showalter et al (1996) J.Pharmacol.Exp.Ther. 278 989.
122. Griffin et al (1997) Eur.J.Pharmacol. 339 53.
123. Járai et al (1999) Proc.Natl.Acad.Sci.USA 96 14136.
124. Kunos et al (2000) Chem.Phys.Lipids 108 159.
125. Di Marzo et al (2000) J.Neurochem. 75 2434.
126. Hanus et al (2001) Proc.Natl.Acad.Sci.USA 98 3662.

Published May 2001