Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Jan 10.
Published in final edited form as: Pharmacol Res. 2007 Sep 8;56(5):393–405. doi: 10.1016/j.phrs.2007.09.005

Endocannabinoid system involvement in brain reward processes related to drug abuse

Marcello Solinas 1, Sevil Yasar 2, Steven R Goldberg 3
PMCID: PMC2189556  NIHMSID: NIHMS34758  PMID: 17936009

Abstract

Cannabis is the most commonly abused illegal drug in the world and its main psychoactive ingredient, delta-9-tetrahydrocannabinol (THC), produces rewarding effects in humans and non-human primates. Over the last several decades, an endogenous system comprised of cannabinoid receptors, endogenous ligands for these receptors and enzymes responsible for the synthesis and degradation of these endogenous cannabinoid ligands has been discovered and partly characterized. Experimental findings strongly suggest a major involvement of the endocannabinoid system in general brain reward functions and drug abuse. First, natural and synthetic cannabinoids and endocannabinoids can produce rewarding effects in humans and laboratory animals. Second, activation or blockade of the endogenous cannabinoid system has been shown to modulate the rewarding effects of non-cannabinoid psychoactive drugs. Third, most abused drugs alter brain levels of endocannabinoids in the brain. In addition to reward functions, the endocannabinoid cannabinoid system appears to be involved in the ability of drugs and drug-related cues to reinstate drug-seeking behavior in animal models of relapse. Altogether, evidence points to the endocannadinoid system as a promising target for the development of medications for the treatment of drug abuse.

1. Introduction

Although recreational use of cannabis has been known for centuries, its abuse has dramatically increased in Europe and the United States over the last few decades. Cannabis and its derivates produce feelings of euphoria, “high” and well-being in humans (1) which probably are central in the reinforcement of repeated cannabis use and the development of cannabis dependence. Until recently the brain mechanisms underlying the reinforcing and dependence-producing effects of cannabis remained unclear. However, the characterization of delta-9-tetrahydrocannaibinol (THC) as the main active ingredient in cannabis (2), the identification and cloning of the first cannabinoid receptors, and the designing of potent and stereoselective synthetic cannabinoid agonists and antagonists led to the discovery of an entire endogenous system, which not only mediates the behavioral and neurobiological effects of cannabis but also plays an important role in the regulation of a wide variety of brain functions (37).

The endogenous cannabinoid system consists of cannabinoid (CB) receptors, endogenous ligands for these receptors and enzymes involved in the synthesis and degradation of these endogenous ligands (37). Two types of cannabinoid receptors, CB1 and CB2, have been cloned and well characterized and additional cannabinoid receptors have been proposed but not yet identified (8, 9). Although it has been recently shown that CB2 receptors are present in the brain (1012), it appears that the abuse-related effects of cannabinoids are mostly, if not exclusively, mediated by CB1 receptors. The two identified and best characterized endogenous ligands for cannabinoid receptors are anandamide (13) and 2-AG (14, 15). These lipophylic compounds are derived from arachidonic acid and are not stored in vesicles but, instead, are formed and released “on demand” and are quickly inactivated by a two-step process involving transport into neurons, presumably by a yet-to-be identified active membrane transport mechanism (16, 17), and metabolic inactivation, primarily by fatty acid amide hydrolase (FAAH) in the case of anandamide (1821) and monoacylglycerol lipase (MGL) in the case of 2-AG (22, 23).

Here, we will focus on the involvement of the endogenous cannabinoid system in brain reward processes that are believed to play pivotal roles in the development and maintenance of drug abuse and dependence.

2. Rewarding effects of cannabinoids

Clinical evidence supporting the hypothesized role of the cannabinoid system in brain reward processes (24, 25) are the wide spread abuse of cannabis, the self-administration of marijuana and THC by human subjects in experimental settings, and the subjective reports of “high, well being and euphoria” following administration of cannabis extracts or THC in humans (2628). Despite the clinical evidence for rewarding effects of cannabis and its psychoactive ingredient THC in humans, rewarding effects of THC or other synthetic cannabinoid agonists in animal models of drug abuse have been difficult to demonstrate (24, 2932). It has been long debated whether cannabis is primarily used for its rewarding effects or whether it lacks clear rewarding effects and is, instead, voluntarily administered by humans for its mind-altering effects. However, several recent findings have demonstrated that, cannabinoid agonists can initiate and sustain active self-administration behavior in monkeys and rodents, although the range of conditions under which this occurs is more limited than with drugs such as cocaine and heroin, (3335). A further step in demonstrating that the endogenous cannabinoid system is involved in brain reward processes is the recent demonstration that the endogenous cannabinoid receptor ligand anandamide can initiate and sustain active self-administration behavior in monkeys (32).

2.1. Self-administration of cannabinoids

Until a decade ago, demonstrations of THC self-administration by experimental animals had been limited and unconvincing (32). Several studies (3642) failed to demonstrate that persistent, dose-related, intravenous drug self-administration behavior could be maintained by THC or synthetic cannabinoid agonists, in a manner that was susceptible to vehicle extinction and subsequent reinstatement.

In the 1990s, with the expansion in cannabinoid research that occurred after the cloning of cannabinoid receptors, several groups reassessed the rewarding effects of cannabinoids in experimental animals using intravenous drug self-administration techniques. In 1998, Martellotta and colleagues, were the first to report that mice would show some evidence of self-administration behavior with a synthetic CB1 receptor agonist WIN55,212-2 (43). However, special conditions were required for this demonstration. First, self-administration was inferred from a dose-dependent shift to nose pokes in an active hole associated with a WIN55,212-2 injection from nose pokes in an inactive hole, but total number of nose-pokes did not vary from one condition to the other. In addition, self-administration was studied during a single acute self-administration session and required restraint and acute insertion of a needle into the mouse’s tail, raising questions about the involvement of cannabinoid-induced anxiolytic effects relieving restraint-induced stress (44) or cannabinoid-induced analgesic effects (4547), rather than true brain reward, in the reinforcement of active-hole nose-poke behavior. Finally, the acute nature of the experiment, by itself, made results difficult to interpret and compare with data from self-administration studies in catheterized animals which employed repeated daily sessions.

More recently, self-administration of WIN55,212-2 was demonstrated in catheterized mice during repeated daily testing when they were food-deprived (34). WIN55,212-2 self-administration behavior was rapidly acquired in mice lacking kappa-opioid receptors; however acquisition was markedly delayed in wild-type mice unless they received an intraperitoneal injection of WIN55,212-2 the day before the start of self-administration sessions. This strategy of pre-exposure to cannabinoid was previously used by the same research group to obtain significant conditioned place preference with THC in wild-type mice (48). The authors speculated that rewarding effects of THC and synthetic agonists such as WIN55,212-2 are normally masked in mice by strong aversive effects mediated by kappa-opioid receptors (49), which are preponderant during the first exposure to THC or WIN55,212-2 and can be unmasked by a priming injection of THC or WIN (34, 48).

Significant intravenous self-administration of WIN55,212-2 was again reported in food-restricted Long Evans rats by Fattore and colleagues (33). Self-administration of WIN55, 212-2 was dose-dependent (with peak responding at 12.5 μg/kg/injection) and behavior extinguished (although very slowly) when vehicle was substituted for WIN55,212-2. Vehicle-like responding was also found when the cannabinoid CB1 receptor antagonist rimonabant (SR-141716, Acomplia) was administered before the session. Self-administration of WIN55,212 in rats has been confirmed several times by the same group as well as other by groups (5054). Interestingly, whereas Deiana et al., (51) reported that Sprague-Dawley rats (in contrast to Long Evans and Lister Hooded rats) do not acquire WIN55,212-2 self-administration behavior, Lecca et al., (53) reported significant acquisition of self-administration behavior in Sprague-Dawley rats under conditions similar to those used by Deiana et al. (51).

Further investigations are needed to resolve these discrepancies, but, it should be noted, that we and others have been unable to demonstrate reliable intravenous self-administration behavior with THC under similar conditions (unpublished findings). However, we recently demonstrated that THC maintains self-administration behavior in Sprague-Dawley rats when it is directly injected into the ventral tegmental area (VTA) or the nucleus accumbens (55). This intracranial self-administration behavior is acquired quickly, extinguishes within a single session when vehicle is substituted for THC or when rats are injected with the CB1 antagonist rimonabant before the session, and self-administration behavior is immediately re-acquired when THC is again available. Thus, it is possible that THC has negative effects in rats that mask its rewarding effects and that brain mechanisms mediating these negative effects are bypassed when THC is directly injected into the VTA or nucleus accumbens.

In 2000, we were the first to report robust and persistent self-administration of THC in non-human primates (35). Self-administration of THC was maintained at rates as high as those of cocaine in the same experimental conditions, was dose dependent, extinguished rapidly when vehicle was substituted for THC or when rimonabant was administered before the session, and immediately returned to normal levels when THC was made available again (35). Whereas the monkeys in this study had a history of cocaine self-administration behavior that had been extinguished for several weeks before the study began, (30), we subsequently found that naïve squirrel monkeys, with no drug or experimental history, also learned to self-administer THC and, in fact, the final levels of drug-taking behavior that developed were about two-fold higher than in our earlier study with cocaine-experienced monkeys (56). Interestingly, we have also found that THC can sustain long sequences of drug-seeking behaviour under a second-order schedule of intravenous THC self-administration (Justinova et al., unpublished findings) where responding during experimental sessions is maintained by cues associated with THC injection only at the end of the session (5759).

Finally, we have recently demonstrated that, like THC, the endogenous cannabinoid anandamide and its synthetic analogue methanandamide maintain high rates of self-administration behavior when they are intravenously administered (32). Self-administration of anandamide and methanandamide was dose dependent (with both compounds being about 10 times less potent than THC), and was sensitive to vehicle extinction and to pharmacological blockade of CB1 receptors by rimonabant (32).

2.2. Conditioned place preference and aversion with cannabinoids

Conditioned place preference (CPP) is an alternative measure of reward in animals. Although CPP provides a less direct estimation of rewarding effects of drugs than intravenous self-administration, it has several advantages. Animals do not need to be catheterized, allowing the use of subcutaneous and intraperitoneal routes of administration, and experiments can usually be completed within one or two weeks. Also, both positive (rewarding) and negative (aversive) effects of drugs can be evaluated within the same experiment, which can be particularly interesting with cannabinoid agonists, where quite contrasting findings ranging from positive place preferences to no effect to place aversions can be found.

Gardner and colleagues were among the first to evaluate the effects of THC on place conditioning (60). In their experiments with Long-Evans rats, THC-induced conditioned place preferences were observed but they depended not only on the dose of THC but also on the regimen of THC administration. With place-conditioning procedures, conditioning sessions with drug and vehicle alternate. Thus, when conditioning sessions were performed every day and THC was given every second day, conditioned place preferences developed at THC doses of 2 and 4 mg/kg with no effect at 1mg/kg, whereas, when conditioning sessions were performed every second day and THC was given every fourth day, preferences only developed at the 1 mg/kg dose of THC and higher THC doses produced conditioned place aversions (60). Subsequently, a few studies have shown cannabinoid-induced conditioned place preferences in rats and all of them used a standard every day conditioning procedure. One study using the potent synthetic CB1 receptor agonist CP 55, 940 and Wistar rats found conditioned place preferences at a dose of 20 μg/kg but there were no effects at lower or higher doses (61); another study using Sprague-Dawley rats and THC found conditioned place preferences at a dose of 0.1 mg/kg but no effect at lower or higher doses (62); a third study using Wistar rats and WIN55,212-2 found conditioned place preferences at a dose of 1 mg/kg in rats housed in enriched conditions but not in rats housed in standard conditions (63). A number of studies have found that cannabinoid agonists produce conditioned place aversions and not place preferences (61, 6468). These studies used the same compounds, the same rat strains, similar conditioning procedures, and similar dose range as the studies that demonstrated conditioned place preferences. Thus, drawing general conclusions on whether cannabinoids have reinforcing or aversive effects in place conditioning procedures is difficult. If we consider the reinforcing effects of cannabinoids in self-administration procedures, it appears that the rewarding effects of THC in place preference procedures may be masked and, in fact, often reversed by some aversive effects. It is also possible that differences exist in the rewarding and aversive effects of cannabinoids in mice and rats, since a survey of results with cannabinoid agonists on another emotional measure, anxiety, found that mice show predominantly anxiolytic effects while rats show predominantly anxiogenic effects (69).

Important advances in the characterization and understanding of the rewarding and aversive effects of cannabinoids in mice have been provided by the work of Maldonado and colleagues. These studies suggest that the first administration of THC, even at a low dose (1mg/kg), produces aversive effects that prevent the development of subsequent conditioned place preferences (48). However, these aversive effects appear to undergo fast tolerance, since mice that are given a first injection of THC 24 hours before the start of conditioning go on to develop positive conditioned place preferences (48). In addition, whereas 5 mg/kg of THC induced conditioned place aversions in their experiments under normal conditions, this aversive effect was lost when a priming injection of THC was given 24h before the start of conditioning (48). In subsequent studies, it was found that the THC-induced conditioned place aversions in mice depended on kappa-opioid receptors (49) and endogenous dynorphin (70), whereas THC-induced conditioned place preferences depended on mu-opioid receptors (49).

The activation of these systems by cannabinoid agonists can be region-specific, since we have recently shown that THC induces conditioned place preferences when injected directly into the posterior part of the nucleus accumbens shell or the posterior ventral tegmental area (55). Only one study has addressed the question on whether endogenous cannabinoids have rewarding or aversive effects in place conditioning procedures. Mallet and Beninger (67) compared THC and the endocannabinoid anandamide in Wistar rats and found that THC at doses of 1 and 1.5 mg/kg (but not lower or higher doses), but not anandamide (0–16 mg/kg i.p.) induced significant conditioned place aversions. In this study, anandamide injections (as well as vehicle injections) were preceded by injection of the non-specific protease inhibitor phenylmethylsulfonyl fluoride in order to slow down anandamide’s rapid metabolic inactivation. We have recently found that anandamide does not have effects on place conditioning when injected intravenously by itself but it induces dose-related conditioned place aversions when its metabolic inactivation is inhibited with the selective fatty acid amide hydrolase (FAAH) inhibitor, URB-597 (Scherma et al., unpublished findings). URB-597 alone does not produce either conditioned place preferences or aversions (19, 71), even at doses 10 to 30-fold higher than the 0.1 to 0.3 mg/kg doses needed to produce significant elevations in brain levels of anandamide (19, 71) and potentiate the in vivo effects of systemically administered anandamide (72, 73). In contrast, the endocannabinoid transport inhibitor, AM-404 induces a small but significant conditioned place preference in rats housed in enriched conditions (63). Interestingly, the dose of 2.5 mg/kg AM-404 that induces conditioned place preferences does not increase either anandamide or 2-AG in the brain areas investigated (63). Thus, the involvement of the endogenous cannabinoid system in AM-404-induced place preferences remains doubtful.

Finally, although most studies have found no effects of the cannabinoid CB1 antagonist rimonabant on place conditioning (7476), a few studies have found that rimonabant induces conditioned place preferences in rats (66, 68). Again, as with AM-404, it is not possible to determine whether rimonabant induces conditioned place preferences in these studies by blocking an endogenous cannabinoid tone or by acting as an inverse agonist (77). These results, together with those showing conditioned place aversions induced by cannabinoid CB1 agonists, have led some authors to hypothesize that the endogenous cannabinoid system is an “aversive or counter-rewarding system” (66).

2.3. Discriminative stimulus effects of cannabinoids

A protocol that we and others have used to characterize the abuse-related effects of cannabinoids is the two-lever choice drug discrimination procedure with THC or other cannabinoid CB1 agonists as the training drug (7888). Rats or monkeys readily learn to discriminate even relatively low doses of THC from vehicle, although the development of stable discrimination performance usually requires 30 sessions or more. While discriminative stimulus effects of drugs are not a direct measure of reward, it is clear that discriminative effects of drugs play an important role in the initiation and maintenance of drug-taking behavior (89, 90). Importantly, the range of effects measured by drug discrimination are wider than those of direct measures of reward and reinforcement and can include aversive, anxiogenic or anxiolytic effects of cannabinoids (89, 90). From a practical point of view, discriminative stimulus effects of cannabinoid CB1 receptor agonists are very strong and they provide a behavioral baseline that remains stable over long periods of time (84). The discriminative effects of cannabinoids are also pharmacologically selective so that, generally, only cannabinoid CB1 receptor agonists produce discriminative effects similar to those of THC and only CB1 receptor antagonists block them (72, 8486, 9193). There are examples of non-CB1 agonists or antagonists having THC-like discriminative effects, but when this occurs results often suggest increased or decreased release of endogenous cannabinoid CB1 agonists as a mechanism. For example, we have recently demonstrated that nicotine produces dose related THC-like discriminative effects after FAAH inhibition with URB-597 (93).

Drug discrimination procedures allow for extensive collection of data, for the qualitative and quantitative evaluation of the effects of new cannabinoid compounds and for the evaluation of the influences of an infinite number of compounds on the behavioral effects of cannabinoids. Drug discrimination procedures also appear to have good predictive validity for self-administration of drugs of abuse (89). For example, we have recently demonstrated that the alpha-7 nicotinic receptor antagonist MLA reduce both the discriminative effects of THC and the reinforcing effects of WIN55,212-2 (54). Thus, given the difficulties of having stable and replicable models of cannabinoid self-administration and conditioned place preference in rodents, drug discrimination is an excellent alternative for studying the pharmacology of cannabinoids (2931).

Several studies have investigated whether endogenous cannabinoid ligands produce THC-like discriminative effects when systemically administered and found that anandamide either does not produce THC-like discriminative effects or it does so only at very high doses that dramatically depress rates of responding (80, 91, 94). Since, under the same or similar conditions, metabolically stable synthetic analogues of anandamide such as methanandamide, O-1812 and AM-1346, produce complete generalization to a THC training stimulus (80, 87, 91, 95, 96), it is likely that anandamide’s fast metabolic inactivation is responsible for the weak effects observed. In fact, in a recent study, we found that anandamide produces dose-related THC-like discriminative effects when its metabolic inactivation by FAAH was blocked by the FAAH inhibitor URB-597 (72). Importantly, when URB-597 was given alone it did not produce any THC-like effects, even at doses 10 times higher than those that potentiated the effects of anandamide (71), supporting the conclusion that this compound has little or no potential for abuse. Interestingly, AM-404, an inhibitor of endocannabinoid transport, produced no THC-like effects itself and, surprisingly, did not potentiate the THC-like effects of anandamide (72), suggesting that, in those brain regions mediating the discriminative effects of THC, membrane transport is not a main mechanism for anandamide inactivation.

2.4. Activation of endogenous dopamine systems involved in brain reward functions

Most psychoactive drugs that have reinforcing or rewarding effects in experimental animals and are abused by humans increase dopamine levels in the nucleus accumbens shell (97100). Although dopamine elevations in the nucleus accumbens are not a direct measure of reward, they are often considered an important neurochemical correlate of rewarding and reinforcing effects of abused drugs and a potential mechanism for these effects. Importantly, THC and other cannabinoid CB1 agonists activate dopamine neurotransmission, like other abused drugs. In microdialysis studies, cannabinoid CB1 agonists increase dopamine levels in the nucleus accumbens (61, 101103) and in particular in its shell sub-region (104). Also, in studies using fast-scan cyclic voltammetry, the CB1 agonist WIN55, 212-2 increases the frequency of dopamine concentration transients (105). Finally, burst firing of ventral tegmental area dopaminergic neurons is increased by cannabinoid CB1 agonists (106109). In contrast, cannabinoid CB1 antagonists are ineffective by themselves on all these measures but they block the effects of cannabinoid CB1 agonists, which is considered a demonstration that the effects of cannabinoid agonists on dopaminergic neurotransmission are mediated through a CB1 receptor mechanism (106, 107, 109).

We have recently shown that intravenously injected anandamide and methanandamide also increase dopamine levels in the shell of the nucleus accumbens by a CB1 receptor mediated mechanism (73). The FAAH inhibitor URB-597 dramatically enhanced and prolonged the effects of anandamide at doses of URB-597 that had no effect by themselves (73). However, when anandamide was injected intraperitoneally rather than intravenously, it only increased dopamine levels in the nucleus accumbens after FAAH inhibition by URB-597. Surprisingly, AM-404, an inhibitor of endogenous cannabinoid transport, which also did not increase dopamine levels in the nucleus accumbens shell by itself, did not enhance or prolong the effects of intravenously injected anandamide (72). Although dopamine elevations in the nucleus accumbens shell are believed to play an important role in the reinforcing effects of THC (24, 29, 31), the fact that a number of differences between the dopamine-increasing profile and the discriminative stimulus profiles of endogenous cannabinoids were found (72) and that dopaminergic compounds do not produce THC-like discriminative effects (88) indicates that the behavioral effects of cannabinoids are only partially mediated by the dopaminergic system.

2.5. Activation of endogenous opioid systems involved in brain reward functions

The endogenous opioid system, like the endogenous dopamine system, appears to play a pivotal role in brain reward and reinforcement processes (110, 111). This system is comprised of three main types of receptors, mu-, delta- and kappa-opioid receptors, and of several endogenous ligands. These ligands are peptides that have different affinities for different types of opioid receptors: endomorphins are selective for mu receptors, beta-endorphin preferentially activates mu receptors, but also activates delta receptors, enkephalin preferentially activates delta receptors but also activates mu receptors, and dynorphin selectively activates kappa opioid receptors. Importantly, whereas mu and delta receptors are thought to mediate positive reinforcement and reward processes, kappa opioid receptors appear to mediate negative reinforcement and aversion (110, 111). Cannabinoids appear to have very strong interactions with the opioid system and, interestingly, these interactions appear to be important both for the reinforcing and rewarding and for the aversive effects of cannabinoids.

Administration of THC increases levels of enkephalins in the nucleus accumbens (112) and levels of beta-endorphin both in the nucleus accumbens and the VTA, with the VTA playing the more relevant role in the mediation or modulation of the behavioral effects of THC by mu-opioid compounds (86). Pharmacological blockade or genetic ablation of mu-opioid receptors reduces the reinforcing (49, 61, 113), discriminative (86, 92) and dopamine releasing (103, 104) of THC, but not the effects of THC on firing of dopaminergic mesolimbic neurons (107, 109). Although these results suggest that many THC effects are mediated by released endopioids, other mechanisms may participate in these interactions. Cannabinoid and opioid receptors, especially mu-opioid receptors, show similar brain distributions, show at least a partial degree of co-localization in brain areas involved in motivated behaviors (61, 114, 115), and share similar second-messenger cascades (116, 117), which indicates that cannabinoid and opioid receptors may interact at the level of the cell membrane (direct protein-protein hetero-dimerization) or at the level of signalling pathways. This is supported by recent demonstrations of allosteric modulation of mu- and delta-opioid receptors (118) and reductions in the signalling strength of cannabinoid CB1 receptor agonists in mu-opioid receptor-deficient “knock-out” mice (119).

As previously discussed, kappa-opioid receptors may mediate the aversive effects of THC and other cannabinoid CB1 agonists (34, 49). Although increases in dynorphin levels following injections of cannabinoid CB1 receptor agonists have not been reported, it is likely that dynorphin plays a role in these aversive effects since dynorphin deficient mice do not develop conditioned place aversions with high doses of THC (70) and show a leftward shift in the dose-response curve for WIN55, 212-2 self-administration (34).

2.6. Summary

The existence of reliable models of cannabinoid and endocannabinoid self-administration in squirrel monkeys and of WIN 55,212-2 self-administration in rats and mice, the limited evidence that THC and WIN 55,212-2 can induce the development of conditioned place preferences in rodents under certain conditions, together with of the consistent findings that THC and WIN 55,212-2 can activate dopaminergic brain reward circuits, demonstrates that cannabinoid CB1 agonists have rewarding and reinforcing effects under certain experimental conditions. On the other hand, the clear difficulties encountered in obtaining reliable and persistent self-administration of THC and other CB1 agonists and the frequent reports that they induce dose-related conditioned place aversions under many conditions, demonstrates that cannabinoids can also have aversive effects, particularly in rodents. The balance between these opposing effects (that at the moment appears to be influenced by several factors in unpredictable ways) determines the behavioral output measured in experimental settings. However, in humans, and in the non-human primate (squirrel monkey) model, this balance appears to be shifted to the side of reinforcing effects (1, 2628). Thus, it remains to be established how the large amounts of pharmacological and neurobiological data collected in preclinical rodent models reflects and predicts effects in humans.

3. The endocannabinoid system and non-cannabinoid drugs of abuse

The recent availability of selective CB1 receptor antagonists such as rimonabant (120) and AM-251 (121) and of mice genetically lacking CB1 receptors (122, 123) has greatly facilitated the study of cannabinoid system involvement in the rewarding and reinforcing effects of abused and addictive drugs. The involvement of the endogenous cannabinoid system in the modulation or mediation of the rewarding and reinforcing effects of opioids, psychostimulants alcohol and nicotine has now been clearly demonstrated in a series of animal studies.

3.1. Opioids

Opioids and cannabinoids have many similar effects and mu-opioid and cannabinoid-CB1 receptors are similarly expressed in brain areas involved in reward processes (115, 124127) where they share common signalling cascades (116, 117). Not surprisingly, an impressive number of interactions have been described to date (128132). Concerning the reinforcing effects of opioids, it has been demonstrated that pharmacological blockade or genetic deletion of CB1 receptors reduces self-administration of heroin and morphine in rats and mice (85, 122, 133135). Consistent with these findings, the CB1 receptor antagonist rimonabant reduces the reinforcing effects of self-administered heroin in rats (85, 133, 134). The effects of cannabinoid CB1 receptor antagonists on opioid reinforcement appear to be mediated primarily by the nucleus accumbens and its projections to the ventral pallidum, since injections of rimonabant directly into the nucleus accumbens decrease heroin self-administration and morphine-induced release of GABA in the ventral pallidum (136). Importantly, the effects of cannabinoid CB1 antagonists appear to be relatively weak when the effort animals have to make to self-administer an opioid is low (i.e. under continuous reinforcement, fixed-ratio 1 schedules) but become more pronounced when the effort is high (i.e. progressive ratio schedules) (85, 133, 134).

The involvement of the endogenous cannabinoid system in the reinforcing effects of opioids is further supported by the findings that administration of CB1 agonists dramatically increases the motivation for self-administering heroin under a progressive-ratio schedule (137). In that study, however, administration of compounds that should increase endogenous cannabinoid system activity by interfering with the inactivation of endocannabinoids, such as the FAAH inhibitor URB-597 or the endocannabinoid transport inhibitor AM-404, did not increase self-administration of heroin and, instead, at high doses decreased it (137). These results do not provide support for the hypothesis that opioid-induced release of endocannabinoids is a major mechanism underlying the influences of cannabinoids on opioid reinforcement and suggest, instead, that interactions at other levels may be responsible for these effects.

In one study, mice lacking CB1 receptors were shown to develop morphine-induced conditioned place preferences (138). However, another study found that mice lacking CB1 receptors did not develop morphine-induced conditioned place preferences (138). Furthermore, administration of the CB1 receptor antagonist rimonabant blocks the development of morphine-induced conditioned place preferences in mice (135) and rats (74, 76). Heroin or morphine-induced elevations in dopamine levels in the nucleus accumbens are not mediated by CB1 receptors, since rimonabant does not block them (104, 133) and mice lacking CB1 receptors show normal dopamine accumbal elevations following administration of morphine (139).

3.2. Psychostimulants

The involvement of the endogenous cannabinoid system in psychostimulant reinforcement has also been investigated, but results have not been consistent. For example, self-administration of cocaine under fixed-ratio schedules is not altered by administration of the CB1 antagonist rimonabant in monkeys (35), rats (140, 141) and in restrained CB1 knock out mice (142). However, a recent study has shown that non-restrained CB1 knock out mice actually show reduced acquisition of cocaine self-administration and that these mice, or wild type mice treated with rimonabant, show a reduced motivation to self-administer cocaine injections under progressive-ratio schedules (143). In addition, the cannabinoid system appears to be involved in reinstatement of extinguished cocaine self-administration behavior (see next section). In contrast, treatment with CB1 antagonists or genetic ablation of CB1 receptors does not alter cocaine-induced conditioned place preferences (74, 142, 144) or cocaine-induced elevations in dopamine levels in the nucleus accumbens (133, 143). However, dopamine concentration transients in the nucleus accumbens are significantly reduced by the CB1 antagonist rimonabant (145).

The involvement of CB1 receptors in the reinforcing effects of amphetamines has been also studied and, again, results have not been consistent. On one hand, amphetamine is self-administered in CB1 knock out mice (142). On the other hand, the CB1 antagonist AM-251 reduces self-administration of methamphetamine and the endogenous cannabinoid ligand anandamide, and its synthetic analogue methanandamide, increase self-administration of methamphetamine in rats (146).

3.3. Alcohol

The endogenous cannabinoid system also appears to play an important role in the rewarding and reinforcing effects of alcohol (114, 147). Pharmacological blockade or genetic ablation of CB1 receptors decrease operant self-administration of ethanol (148, 149) and decrease voluntary consumption of ethanol in rats (150154) and in mice (155159). Conversely, cannabinoid CB1 receptor agonists increase voluntary consumption of ethanol in rats (160, 161). A caveat for the results obtained with oral self-administration or voluntary consumption of ethanol is that cannabinoid CB1 agonists have facilitating effects and cannabinoid CB1 antagonists have inhibitory effects on food and fluid intake, on the motivation to respond to obtain food, and on ingestive behaviour and palatability (162165), and these actions may contribute to the behavioral effects of CB1 receptor blockade or deletion on ethanol self-administration and voluntary ethanol consumption. Place conditioning procedures, where ethanol is normally given by the intraperitoneal route and the oral route of administration is bypassed, may provide important information about the role of the cannabinoid system in ethanol reward. Such studies have confirmed that the rewarding effects of ethanol require CB1 receptor activation since CB1 knock out mice do not develop preferences for ethanol-paired compartments (144, 159). In addition, pharmacological blockade or genetic ablation of FAAH enzymes in mice increase ethanol preference and intake in a two-bottle free-choice procedure (166, 167). The prefrontal cortex and the nucleus accumbens appear two brain regions implicated in the modulation of the reinforcing effects of ethanol by cannabinoids (168, 169).

Finally, pharmacological blockade or genetic deletion of CB1 receptors results in the inhibition of ethanol-induced elevations in dopamine extracellular levels as well as dopamine transients in the nucleus accumbens (145, 156, 170) and inhibition of ethanol-induced increases in firing of dopaminergic neurons in the VTA (171).

3.4. Nicotine

Characterization of the interactions between nicotine and the endogenous cannabinoid system may have very important implications for our understanding of the biological basis for the co-administration of cannabis and tobacco in humans. In addition, preclinical results obtained over the last ten years have led to clinical trials that indicate that cannabinoid CB1 receptor antagonists could have important therapeutic utility as smoking cessation agents (172) (http://en.sanofi-aventis.com/press/ppc_1960.asp).

In preclinical studies, the CB1 antagonist rimonabant reduces or blocks both the intravenous self-administration of nicotine and the development and expression of nicotine-induced conditioned place preferences in rats (170, 173). Rimonabant also reduces the motivational impact of drug-related cues or environments on nicotine-seeking behavior (174176). Although CB1 knock out mice do develop nicotine-induced conditioned place preferences (177), they do not acquire intravenous nicotine self-administration (142). In addition, in place conditioning procedures, doses of THC and nicotine that are ineffective by themselves, induce significant conditioned place preferences in mice when they are given in combination (44) and these effects are paralleled at the molecular level by synergistic effects on the expression of immediate early genes such as c-FOS in several brain areas, including the shell of the nucleus accumbens (44). Thus, the rewarding effects of low doses of THC and nicotine appear to be synergistic. Consistent with these findings, we have recently found that nicotine potentiates the discriminative effects of low doses of THC and that these effects are at least in part mediated by release of the endogenous cannabinoid anandamide (93).

3.6. Summary

Experimental data supports a role for the endocannabinoid system in the rewarding and reinforcing effects of non-cannabinoid drugs of abuse. The facilitating effects of cannabinoids are clearest for opioids and alcohol, where pharmacological blockade or genetic ablation of cannabinoid receptors have been repeatedly shown to reduce opioid and alcohol self-administration and opioid- and alcohol-induced conditioned place preferences. In the case of nicotine, although the preclinical data is still incomplete, clinical studies strongly support the involvement of endocannabinoids in nicotine reward and cannabinoid antagonists are promising medications for smoking cessation. On the other hand, the endocannabinoid system does not appear to play an important role in the primary reinforcing effects of psychostimulants but they may be involved in incentive motivational effects of these drugs.

4. Release of endogenous cannabinoids by abused drugs

A mechanism that could explain the modulation by the endocannabinoid system of the rewarding effects of non-cannabinoid drugs and that would support the role of the endogenous cannabinoid system in the signalling of rewarding events is the release of endogenous cannabinoids by non-cannabinoid drugs. Unfortunately, because of the high lipophilic nature and instability of endocannabinoids, techniques such as in vivo microdialysis that allow monitoring the dynamics of neurotransmitter release have proven very difficult to perform routinely for anandamide and 2-AG. Thus, until very recently, only one microdialysis study existed showing that intrastriatal administrations of the dopamine D2 agonist quinpirole locally elevate levels of anandamide but not those of 2-AG (178). All other studies had investigated tissue levels at a single time point (most often 2 hours) after administration of drugs. Since drugs of abuse produce their reinforcing effects rapidly, it may be difficult to correlate endocannabinoid levels and behavioral consequences in order to establish whether measured levels represent a direct effect of the drug or a reaction to the drug effects. Moreover, most studies have investigated brain levels of endocannabinoids after chronic drug treatment. In these studies one group received drug at all times (chronic) and one group received saline, vehicle or water at all times (control) but no group received saline or vehicle chronically and then drug acutely and vice versa. Thus, it is difficult to establish with certainty whether measured levels of endocannabinoids reflect the consequences of chronic drug administration or of drug withdrawal. For example, it has been shown that chronic administration of THC itself decreases the levels of anandamide and 2-AG in the striatum which contains the nucleus accumbens (179); ethanol decreases the levels of anandamide and 2-AG in the midbrain which contains the VTA (180); nicotine decreases the levels of anandamide in the striatum (180); cocaine does not alter anandamide or 2-AG levels either in the striatum or the midbrain (180); morphine decreases 2-AG levels in the striatum without altering anandamide levels (181). Interestingly, Vigano and colleagues in a more recent paper used a protocol of morphine administration that induced behavioral sensitization and investigated the levels of anandamide and 2-AG at different time points: at the end of induction phase, after fifteen days of withdrawal and after the expression of morphine sensitization by a morphine challenge (182). In addition they had a saline group and an acute morphine group control. They found that chronic morphine administration did not change anandamide levels but decreased 2-AG levels. More importantly, they found that acute administration as well as a challenge administration of morphine increased levels of anandamide and decreased levels of 2-AG in the striatum (182). These findings underscore the critical role of the protocol of administration and the time point at which endocannabinoids are measured in determining the quality and quantity of alterations in endocannabinoid levels observed.

A promising approach to the study of the involvement of released endocannabinoids in the rewarding and reinforcing effects of abused drugs reward has recently been employed by Caille et al. (169), involving the combination of drug self-administration and in vivo microdialysis techniques in rats. They found that, in the nucleus accumbens, self-administration of heroin increased anandamide and decreased 2-AG levels, self-administration of alcohol increased 2-AG and did not alter anandamide levels and self-administration of cocaine did not alter either anandamide or 2-AG levels (169). Hopefully, such a protocol will be soon used to study the effects of self-administration of other drugs and the effects of self-administration of different types of drugs on changes in the levels of endocannabinoids in other brain areas, such as the VTA, that play an important role in the reinforcing effects of cannabinoids (55).

Although our knowledge is still limited and there is a need for further research to investigate conditions more relevant to drug abuse and dependence, it appears likely that the alterations found in the referenced papers reflect the involvement of released endocannabinoids in the effects of non-cannabinoid drugs of abuse. On the other hand, as discussed in the sections describing modulation of the behavioral and neurochemical effects of cannabinoids and opioids (sections 2.4. and 2.5.), other mechanisms, such as receptor-receptor interactions or interactions at the signalling cascade level already described for mu-opioid and dopamine D2 receptors (118, 119, 183186), could play a role in the regulation of the rewarding effects of abused drugs by the endogenous cannabinoid system.

5. The endocannabinoid system and reinstatement of drug-seeking

A key feature of drug abuse and dependence is relapse to drug use even after long period of forced or voluntary withdrawal and detoxification (187). A commonly used preclinical model of relapse is reinstatement of extinguished intravenous drug self-administration by a priming injection of drug or by drug-associated stimuli (188190). In a typical version of this procedure (137), animals learn to intravenously self-administer a drug over repeated daily sessions and then extinction sessions are conducted when responding has no programmed consequences (injections and, in some instances, drug-associated stimuli are not presented) until the animal’s responding drops to a low, arbitrary level; reinstatement of operant drug-seeking behavior can then be tested by injecting a drug (drug-induced), by presenting visual or auditory stimuli previously associated with the drug (cue-induced) or by exposing animals to acute stress (stress induced).

The endogenous cannabinoid system appears to be involved in reinstatement of extinguished self-administration of several drugs of abuse and, in particular, appears to be involved in drug- and cue-induced reinstatement (191, 192). The first evidence for this role was provided by De Vries et al., (140) in a study that found that the CB1 agonist HU210 could reinstate extinguished cocaine-seeking behaviour. The cannabinoid CB1 antagonist rimonabant not only blocked this effect but also reduced cocaine- and cue-induced reinstatement; however, rimonabant had no effect on stress-induced reinstatement. Cocaine-induced reinstatement of drug-seeking behavior is also blocked by another cannabinoid CB1 antagonist, AM-251, and appears to be mediated by inhibition of glutamate release in the nucleus accumbens (193). Rimonabant also reduces heroin-induced reinstatement (134, 194); methamphetamine-induced reinstatement (195), nicotine-induced reinstatement (196) and ethanol-induced reinstatement (148, 149). Interestingly, although the endocannabinoid system appears to play an important role in the regulation of stress reactions (197), stress-induced reinstatement does not seem involve endocannabinoid signalling (140, 148, 149). These preclinical data have recently received strong support by preliminary data from clinical trials (STRATUS, Studies with Rimonabant And Tobacco Use) showing that rimonabant significantly increases the odds of quitting smoking (172) (http://en.sanofi-aventis.com/press/ppc_1960.asp).

6. Conclusions

The endogenous cannabinoid is a recently discovered system that appears to play an important and pervasive role in many types of drug abuse and dependence. Endogenous cannabinoids are neuromodulators that are involved in the signalling of rewarding events and can produce reinforcing and rewarding effects in experimental animals, as they do in humans. Endogenous cannabinoids can also activate other brain systems involved in reward signalling, can modulate the reinforcing and rewarding effects of other non-cannabinoid abused drugs, and are released by drugs of abuse in brain areas involved in reward and reinforcement processes. Accumulating evidence points to the endocannabinoid system as a major target for the development of new pharmacological agents for the treatment of many different types of drug abuse and dependence.

Acknowledgments

This work was supported by the Centre National de la Recherche Scientifique, by Johns Hopkins University, School of Medicine and by the Intramural Research Program of the National Institute on Drug Abuse, National Institutes of Health, Department of Health and Human Services.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Goodman J, Gilman L. The Pharmacological Basis of Therapeutics. 9. New York: McGraw-Hill; 2006. [Google Scholar]
  • 2.Gaoni Y, Mechoulam R. Isolation, structure and partial sythesis of an active constituent of hashish. J Am Chem Soc. 1964;86:1646–1647. [Google Scholar]
  • 3.Di Marzo V, Bifulco M, De Petrocellis L. The endocannabinoid system and its therapeutic exploitation. Nat Rev Drug Discov. 2004;3:771–84. doi: 10.1038/nrd1495. [DOI] [PubMed] [Google Scholar]
  • 4.Freund TF, Katona I, Piomelli D. Role of endogenous cannabinoids in synaptic signaling. Physiol Rev. 2003;83:1017–66. doi: 10.1152/physrev.00004.2003. [DOI] [PubMed] [Google Scholar]
  • 5.Fride E. Endocannabinoids in the central nervous system: from neuronal networks to behavior. Curr Drug Targets CNS Neurol Disord. 2005;4:633–42. doi: 10.2174/156800705774933069. [DOI] [PubMed] [Google Scholar]
  • 6.Piomelli D. The molecular logic of endocannabinoid signalling. Nat Rev Neurosci. 2003;4:873–84. doi: 10.1038/nrn1247. [DOI] [PubMed] [Google Scholar]
  • 7.Wilson RI, Nicoll RA. Endocannabinoid signaling in the brain. Science. 2002;296:678–82. doi: 10.1126/science.1063545. [DOI] [PubMed] [Google Scholar]
  • 8.Howlett AC, Barth F, Bonner TI, Cabral G, Casellas P, Devane WA, Felder CC, Herkenham M, Mackie K, Martin BR, Mechoulam R, Pertwee RG. International Union of Pharmacology. XXVII. Classification of cannabinoid receptors. Pharmacol Rev. 2002;54:161–202. doi: 10.1124/pr.54.2.161. [DOI] [PubMed] [Google Scholar]
  • 9.Pertwee RG, Ross RA. Cannabinoid receptors and their ligands. Prostaglandins Leukot Essent Fatty Acids. 2002;66:101–21. doi: 10.1054/plef.2001.0341. [DOI] [PubMed] [Google Scholar]
  • 10.Van Sickle MD, Duncan M, Kingsley PJ, Mouihate A, Urbani P, Mackie K, Stella N, Makriyannis A, Piomelli D, Davison JS, Marnett LJ, Di Marzo V, Pittman QJ, Patel KD, Sharkey KA. Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science. 2005;310:329–32. doi: 10.1126/science.1115740. [DOI] [PubMed] [Google Scholar]
  • 11.Onaivi ES, Ishiguro H, Gong JP, Patel S, Perchuk A, Meozzi PA, Myers L, Mora Z, Tagliaferro P, Gardner E, Brusco A, Akinshola BE, Liu QR, Hope B, Iwasaki S, Arinami T, Teasenfitz L, Uhl GR. Discovery of the presence and functional expression of cannabinoid CB2 receptors in brain. Ann N Y Acad Sci. 2006;1074:514–36. doi: 10.1196/annals.1369.052. [DOI] [PubMed] [Google Scholar]
  • 12.Gong JP, Onaivi ES, Ishiguro H, Liu QR, Tagliaferro PA, Brusco A, Uhl GR. Cannabinoid CB2 receptors: immunohistochemical localization in rat brain. Brain Res. 2006;1071:10–23. doi: 10.1016/j.brainres.2005.11.035. [DOI] [PubMed] [Google Scholar]
  • 13.Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D, Mandelbaum A, Etinger A, Mechoulam R. Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science. 1992;258:1946–9. doi: 10.1126/science.1470919. [DOI] [PubMed] [Google Scholar]
  • 14.Sugiura T, Kondo S, Sukagawa A, Nakane S, Shinoda A, Itoh K, Yamashita A, Waku K. 2-Arachidonoylglycerol: a possible endogenous cannabinoid receptor ligand in brain. Biochem Biophys Res Commun. 1995;215:89–97. doi: 10.1006/bbrc.1995.2437. [DOI] [PubMed] [Google Scholar]
  • 15.Mechoulam R, Ben-Shabat S, Hanus L, Ligumsky M, Kaminski NE, Schatz AR, Gopher A, Almog S, Martin BR, Compton DR, et al. Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors. Biochem Pharmacol. 1995;50:83–90. doi: 10.1016/0006-2952(95)00109-d. [DOI] [PubMed] [Google Scholar]
  • 16.Glaser ST, Kaczocha M, Deutsch DG. Anandamide transport: a critical review. Life Sci. 2005;77:1584–604. doi: 10.1016/j.lfs.2005.05.007. [DOI] [PubMed] [Google Scholar]
  • 17.McFarland MJ, Barker EL. Anandamide transport. Pharmacol Ther. 2004;104:117–35. doi: 10.1016/j.pharmthera.2004.07.008. [DOI] [PubMed] [Google Scholar]
  • 18.Piomelli D, Tarzia G, Duranti A, Tontini A, Mor M, Compton TR, Dasse O, Monaghan EP, Parrott JA, Putman D. Pharmacological profile of the selective FAAH inhibitor KDS-4103 (URB597) CNS Drug Rev. 2006;12:21–38. doi: 10.1111/j.1527-3458.2006.00021.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kathuria S, Gaetani S, Fegley D, Valino F, Duranti A, Tontini A, Mor M, Tarzia G, La Rana G, Calignano A, Giustino A, Tattoli M, Palmery M, Cuomo V, Piomelli D. Modulation of anxiety through blockade of anandamide hydrolysis. Nat Med. 2003;9:76–81. doi: 10.1038/nm803. [DOI] [PubMed] [Google Scholar]
  • 20.Cravatt BF, Demarest K, Patricelli MP, Bracey MH, Giang DK, Martin BR, Lichtman AH. Supersensitivity to anandamide and enhanced endogenous cannabinoid signaling in mice lacking fatty acid amide hydrolase. Proc Natl Acad Sci U S A. 2001;98:9371–6. doi: 10.1073/pnas.161191698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Cravatt BF, Lichtman AH. Fatty acid amide hydrolase: an emerging therapeutic target in the endocannabinoid system. Curr Opin Chem Biol. 2003;7:469–75. doi: 10.1016/s1367-5931(03)00079-6. [DOI] [PubMed] [Google Scholar]
  • 22.Dinh TP, Freund TF, Piomelli D. A role for monoglyceride lipase in 2-arachidonoylglycerol inactivation. Chem Phys Lipids. 2002;121:149–58. doi: 10.1016/s0009-3084(02)00150-0. [DOI] [PubMed] [Google Scholar]
  • 23.Makara JK, Mor M, Fegley D, Szabo SI, Kathuria S, Astarita G, Duranti A, Tontini A, Tarzia G, Rivara S, Freund TF, Piomelli D. Selective inhibition of 2-AG hydrolysis enhances endocannabinoid signaling in hippocampus. Nat Neurosci. 2005;8:1139–41. doi: 10.1038/nn1521. [DOI] [PubMed] [Google Scholar]
  • 24.Gardner EL. Endocannabinoid signaling system and brain reward: emphasis on dopamine. Pharmacol Biochem Behav. 2005;81:263–84. doi: 10.1016/j.pbb.2005.01.032. [DOI] [PubMed] [Google Scholar]
  • 25.Maldonado R, Valverde O, Berrendero F. Involvement of the endocannabinoid system in drug addiction. Trends Neurosci. 2006;29:225–32. doi: 10.1016/j.tins.2006.01.008. [DOI] [PubMed] [Google Scholar]
  • 26.Ward AS, Comer SD, Haney M, Foltin RW, Fischman MW. The effects of a monetary alternative on marijuana self-administration. Behav Pharmacol. 1997;8:275–86. doi: 10.1097/00008877-199708000-00001. [DOI] [PubMed] [Google Scholar]
  • 27.Haney M, Comer SD, Ward AS, Foltin RW, Fischman MW. Factors influencing marijuana self-administration by humans. Behav Pharmacol. 1997;8:101–12. [PubMed] [Google Scholar]
  • 28.Hart CL, Haney M, Vosburg SK, Comer SD, Foltin RW. Reinforcing effects of oral Delta9-THC in male marijuana smokers in a laboratory choice procedure. Psychopharmacology (Berl) 2005;181:237–43. doi: 10.1007/s00213-005-2234-2. [DOI] [PubMed] [Google Scholar]
  • 29.Maldonado R, Rodriguez de Fonseca F. Cannabinoid addiction: behavioral models and neural correlates. J Neurosci. 2002;22:3326–31. doi: 10.1523/JNEUROSCI.22-09-03326.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Maldonado R. Study of cannabinoid dependence in animals. Pharmacol Ther. 2002;95:153–64. doi: 10.1016/s0163-7258(02)00254-1. [DOI] [PubMed] [Google Scholar]
  • 31.Tanda G, Goldberg SR. Cannabinoids: reward, dependence, and underlying neurochemical mechanisms--a review of recent preclinical data. Psychopharmacology (Berl) 2003;169:115–34. doi: 10.1007/s00213-003-1485-z. [DOI] [PubMed] [Google Scholar]
  • 32.Justinova Z, Solinas M, Tanda G, Redhi GH, Goldberg SR. The endogenous cannabinoid anandamide and its synthetic analog R(+)-methanandamide are intravenously self-administered by squirrel monkeys. J Neurosci. 2005;25:5645–50. doi: 10.1523/JNEUROSCI.0951-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Fattore L, Cossu G, Martellotta CM, Fratta W. Intravenous self-administration of the cannabinoid CB1 receptor agonist WIN 55,212-2 in rats. Psychopharmacology (Berl) 2001;156:410–6. doi: 10.1007/s002130100734. [DOI] [PubMed] [Google Scholar]
  • 34.Mendizabal V, Zimmer A, Maldonado R. Involvement of kappa/dynorphin system in WIN 55,212-2 self-administration in mice. Neuropsychopharmacology. 2006;31:1957–66. doi: 10.1038/sj.npp.1300957. [DOI] [PubMed] [Google Scholar]
  • 35.Tanda G, Munzar P, Goldberg SR. Self-administration behavior is maintained by the psychoactive ingredient of marijuana in squirrel monkeys. Nat Neurosci. 2000;3:1073–4. doi: 10.1038/80577. [DOI] [PubMed] [Google Scholar]
  • 36.Harris RT, Waters W, McLendon D. Evaluation of reinforcing capability of delta-9-tetrahydrocannabinol in rhesus monkeys. Psychopharmacologia. 1974;37:23–9. doi: 10.1007/BF00426679. [DOI] [PubMed] [Google Scholar]
  • 37.Carney JM, Uwaydah IM, Balster RL. Evaluation of a suspension system for intravenous self-administration studies of water-insoluble compounds in the rhesus monkey. Pharmacol Biochem Behav. 1977;7:357–64. doi: 10.1016/0091-3057(77)90232-5. [DOI] [PubMed] [Google Scholar]
  • 38.van Ree JM, Slangen JL, de Wied D. Intravenous self-administration of drugs in rats. J Pharmacol Exp Ther. 1978;204:547–57. [PubMed] [Google Scholar]
  • 39.Takahashi RN, Singer G. Self-administration of delta 9-tetrahydrocannabinol by rats. Pharmacol Biochem Behav. 1979;11:737–40. doi: 10.1016/0091-3057(79)90274-0. [DOI] [PubMed] [Google Scholar]
  • 40.Takahashi RN, Singer G. Effects of body weight levels on cannabis self-injection. Pharmacol Biochem Behav. 1980;13:877–81. doi: 10.1016/0091-3057(80)90222-1. [DOI] [PubMed] [Google Scholar]
  • 41.Mansbach RS, Nicholson KL, Martin BR, Balster RL. Failure of Delta(9)-tetrahydrocannabinol and CP 55,940 to maintain intravenous self-administration under a fixed-interval schedule in rhesus monkeys. Behav Pharmacol. 1994;5:219–225. doi: 10.1097/00008877-199404000-00014. [DOI] [PubMed] [Google Scholar]
  • 42.Pickens R, Thompson T, Muchow DC. Cannabis and phencyclidine self-administered by animals. In: Goldberg FHaL., editor. Psychic dependence: definition, assessment in animals and man. New York: Springer-Verlag; 1973. pp. 78–86. [Google Scholar]
  • 43.Martellotta MC, Cossu G, Fattore L, Gessa GL, Fratta W. Self-administration of the cannabinoid receptor agonist WIN 55,212-2 in drug-naive mice. Neuroscience. 1998;85:327–30. doi: 10.1016/s0306-4522(98)00052-9. [DOI] [PubMed] [Google Scholar]
  • 44.Valjent E, Mitchell JM, Besson MJ, Caboche J, Maldonado R. Behavioural and biochemical evidence for interactions between Delta 9-tetrahydrocannabinol and nicotine. Br J Pharmacol. 2002;135:564–78. doi: 10.1038/sj.bjp.0704479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hohmann A, Suplita R. Endocannabinoid mechanisms of pain modulation. Aaps J. 2006;8:E693–708. doi: 10.1208/aapsj080479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chaperon F, Thiebot MH. Behavioral effects of cannabinoid agents in animals. Crit Rev Neurobiol. 1999;13:243–81. doi: 10.1615/critrevneurobiol.v13.i3.20. [DOI] [PubMed] [Google Scholar]
  • 47.Ameri A. The effects of cannabinoids on the brain. Prog Neurobiol. 1999;58:315–48. doi: 10.1016/s0301-0082(98)00087-2. [DOI] [PubMed] [Google Scholar]
  • 48.Valjent E, Maldonado R. A behavioural model to reveal place preference to delta 9-tetrahydrocannabinol in mice. Psychopharmacology (Berl) 2000;147:436–8. doi: 10.1007/s002130050013. [DOI] [PubMed] [Google Scholar]
  • 49.Ghozland S, Matthes HW, Simonin F, Filliol D, Kieffer BL, Maldonado R. Motivational effects of cannabinoids are mediated by mu-opioid and kappa-opioid receptors. J Neurosci. 2002;22:1146–54. doi: 10.1523/JNEUROSCI.22-03-01146.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Spano MS, Fattore L, Cossu G, Deiana S, Fadda P, Fratta W. CB1 receptor agonist and heroin, but not cocaine, reinstate cannabinoid-seeking behaviour in the rat. Br J Pharmacol. 2004;143:343–50. doi: 10.1038/sj.bjp.0705932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Deiana S, Fattore L, Sabrina Spano M, Cossu G, Porcu E, Fadda P, Fratta W. Strain and schedule-dependent differences in the acquisition, maintenance and extinction of intravenous cannabinoid self-administration in rats. Neuropharmacology. 2007;52:646–654. doi: 10.1016/j.neuropharm.2006.09.007. [DOI] [PubMed] [Google Scholar]
  • 52.Fadda P, Scherma M, Spano MS, Salis P, Melis V, Fattore L, Fratta W. Cannabinoid self-administration increases dopamine release in the nucleus accumbens. Neuroreport. 2006;17:1629–32. doi: 10.1097/01.wnr.0000236853.40221.8e. [DOI] [PubMed] [Google Scholar]
  • 53.Lecca D, Cacciapaglia F, Valentini V, Di Chiara G. Monitoring extracellular dopamine in the rat nucleus accumbens shell and core during acquisition and maintenance of intravenous WIN 55,212-2 self-administration. Psychopharmacology (Berl) 2006;188:63–74. doi: 10.1007/s00213-006-0475-3. [DOI] [PubMed] [Google Scholar]
  • 54.Solinas M, Scherma M, Fattore L, Stroik J, Wertheim C, Tanda G, Fratta W, Goldberg SR. Nicotinic alpha 7 receptors as a new target for treatment of cannabis abuse. J Neurosci. 2007;27:5615–20. doi: 10.1523/JNEUROSCI.0027-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Zangen A, Solinas M, Ikemoto S, Goldberg SR, Wise RA. Two brain sites for cannabinoid reward. J Neurosci. 2006;26:4901–7. doi: 10.1523/JNEUROSCI.3554-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Justinova Z, Tanda G, Redhi GH, Goldberg SR. Self-administration of delta9-tetrahydrocannabinol (THC) by drug naive squirrel monkeys. Psychopharmacology (Berl) 2003;169:135–40. doi: 10.1007/s00213-003-1484-0. [DOI] [PubMed] [Google Scholar]
  • 57.Everitt BJ, Robbins TW. Second-order schedules of drug reinforcement in rats and monkeys: measurement of reinforcing efficacy and drug-seeking behaviour. Psychopharmacology (Berl) 2000;153:17–30. doi: 10.1007/s002130000566. [DOI] [PubMed] [Google Scholar]
  • 58.Schindler CW, Panlilio LV, Goldberg SR. Second-order schedules of drug self-administration in animals. Psychopharmacology (Berl) 2002;163:327–44. doi: 10.1007/s00213-002-1157-4. [DOI] [PubMed] [Google Scholar]
  • 59.Goldberg SR, Kelleher RT, Morse WH. Second-order schedules of drug injection. Fed Proc. 1975;34:1771–6. [PubMed] [Google Scholar]
  • 60.Lepore M, Vorel SR, Lowinson J, Gardner EL. Conditioned place preference induced by delta 9-tetrahydrocannabinol: comparison with cocaine, morphine, and food reward. Life Sci. 1995;56:2073–80. doi: 10.1016/0024-3205(95)00191-8. [DOI] [PubMed] [Google Scholar]
  • 61.Braida D, Pozzi M, Cavallini R, Sala M. Conditioned place preference induced by the cannabinoid agonist CP 55,940: interaction with the opioid system. Neuroscience. 2001;104:923–6. doi: 10.1016/s0306-4522(01)00210-x. [DOI] [PubMed] [Google Scholar]
  • 62.Le Foll B, Wiggins M, Goldberg SR. Nicotine pre-exposure does not potentiate the locomotor or rewarding effects of Delta-9-tetrahydrocannabinol in rats. Behav Pharmacol. 2006;17:195–9. doi: 10.1097/01.fbp.0000197460.16516.81. [DOI] [PubMed] [Google Scholar]
  • 63.Bortolato M, Campolongo P, Mangieri RA, Scattoni ML, Frau R, Trezza V, La Rana G, Russo R, Calignano A, Gessa GL, Cuomo V, Piomelli D. Anxiolytic-like properties of the anandamide transport inhibitor AM404. Neuropsychopharmacology. 2006;31:2652–9. doi: 10.1038/sj.npp.1301061. [DOI] [PubMed] [Google Scholar]
  • 64.Parker LA, Gillies T. THC-induced place and taste aversions in Lewis and Sprague-Dawley rats. Behav Neurosci. 1995;109:71–8. doi: 10.1037//0735-7044.109.1.71. [DOI] [PubMed] [Google Scholar]
  • 65.McGregor IS, Issakidis CN, Prior G. Aversive effects of the synthetic cannabinoid CP 55,940 in rats. Pharmacol Biochem Behav. 1996;53:657–64. doi: 10.1016/0091-3057(95)02066-7. [DOI] [PubMed] [Google Scholar]
  • 66.Sanudo-Pena MC, Tsou K, Delay ER, Hohman AG, Force M, Walker JM. Endogenous cannabinoids as an aversive or counter-rewarding system in the rat. Neurosci Lett. 1997;223:125–8. doi: 10.1016/s0304-3940(97)13424-3. [DOI] [PubMed] [Google Scholar]
  • 67.Mallet PE, Beninger RJ. Delta9-tetrahydrocannabinol, but not the endogenous cannabinoid receptor ligand anandamide, produces conditioned place avoidance. Life Sci. 1998;62:2431–9. doi: 10.1016/s0024-3205(98)00226-4. [DOI] [PubMed] [Google Scholar]
  • 68.Cheer JF, Kendall DA, Marsden CA. Cannabinoid receptors and reward in the rat: a conditioned place preference study. Psychopharmacology (Berl) 2000;151:25–30. doi: 10.1007/s002130000481. [DOI] [PubMed] [Google Scholar]
  • 69.Haller J, Matyas F, Soproni K, Varga B, Barsy B, Nemeth B, Mikics E, Freund TF, Hajos N. Correlated species differences in the effects of cannabinoid ligands on anxiety and on GABAergic and glutamatergic synaptic transmission. Eur J Neurosci. 2007;25:2445–56. doi: 10.1111/j.1460-9568.2007.05476.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Zimmer A, Valjent E, Konig M, Zimmer AM, Robledo P, Hahn H, Valverde O, Maldonado R. Absence of delta -9-tetrahydrocannabinol dysphoric effects in dynorphin-deficient mice. J Neurosci. 2001;21:9499–505. doi: 10.1523/JNEUROSCI.21-23-09499.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Gobbi G, Bambico FR, Mangieri R, Bortolato M, Campolongo P, Solinas M, Cassano T, Morgese MG, Debonnel G, Duranti A, Tontini A, Tarzia G, Mor M, Trezza V, Goldberg SR, Cuomo V, Piomelli D. Antidepressant-like activity and modulation of brain monoaminergic transmission by blockade of anandamide hydrolysis. Proc Natl Acad Sci U S A. 2005;102:18620–5. doi: 10.1073/pnas.0509591102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Solinas M, Tanda G, Justinova Z, Wertheim CE, Yasar S, Piomelli D, Vadivel SK, Makriyannis A, Goldberg SR. The endogenous cannabinoid anandamide produces delta-9-tetrahydrocannabinol-like discriminative and neurochemical effects that are enhanced by inhibition of fatty acid amide hydrolase but not by inhibition of anandamide transport. J Pharmacol Exp Ther. 2007;321:370–80. doi: 10.1124/jpet.106.114124. [DOI] [PubMed] [Google Scholar]
  • 73.Solinas M, Justinova Z, Goldberg SR, Tanda G. Anandamide administration alone and after inhibition of fatty acid amide hydrolase (FAAH) increases dopamine levels in the nucleus accumbens shell in rats. J Neurochem. 2006;98:408–19. doi: 10.1111/j.1471-4159.2006.03880.x. [DOI] [PubMed] [Google Scholar]
  • 74.Chaperon F, Soubrie P, Puech AJ, Thiebot MH. Involvement of central cannabinoid (CB1) receptors in the establishment of place conditioning in rats. Psychopharmacology (Berl) 1998;135:324–32. doi: 10.1007/s002130050518. [DOI] [PubMed] [Google Scholar]
  • 75.Mas-Nieto M, Pommier B, Tzavara ET, Caneparo A, Da Nascimento S, Le Fur G, Roques BP, Noble F. Reduction of opioid dependence by the CB(1) antagonist SR141716A in mice: evaluation of the interest in pharmacotherapy of opioid addiction. Br J Pharmacol. 2001;132:1809–16. doi: 10.1038/sj.bjp.0703990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Singh ME, Verty AN, McGregor IS, Mallet PE. A cannabinoid receptor antagonist attenuates conditioned place preference but not behavioural sensitization to morphine. Brain Res. 2004;1026:244–53. doi: 10.1016/j.brainres.2004.08.027. [DOI] [PubMed] [Google Scholar]
  • 77.Landsman RS, Burkey TH, Consroe P, Roeske WR, Yamamura HI. SR141716A is an inverse agonist at the human cannabinoid CB1 receptor. Eur J Pharmacol. 1997;334:R1–2. doi: 10.1016/s0014-2999(97)01160-6. [DOI] [PubMed] [Google Scholar]
  • 78.Jarbe TU, Henriksson BG. Discriminative response control produced with hashish, tetrahydrocannabinols (delta 8-THC and delta 9-THC), and other drugs. Psychopharmacologia. 1974;40:1–16. doi: 10.1007/BF00429443. [DOI] [PubMed] [Google Scholar]
  • 79.Jarbe TU, Lamb RJ, Makriyannis A, Lin S, Goutopoulos A. Delta9-THC training dose as a determinant for (R)-methanandamide generalization in rats. Psychopharmacology (Berl) 1998;140:519–22. doi: 10.1007/s002130050797. [DOI] [PubMed] [Google Scholar]
  • 80.Burkey RT, Nation JR. (R)-methanandamide, but not anandamide, substitutes for delta 9-THC in a drug-discrimination procedure. Exp Clin Psychopharmacol. 1997;5:195–202. doi: 10.1037//1064-1297.5.3.195. [DOI] [PubMed] [Google Scholar]
  • 81.De Vry J, Jentzsch KR. Intrinsic activity estimation of cannabinoid CB1 receptor ligands in a drug discrimination paradigm. Behav Pharmacol. 2003;14:471–6. doi: 10.1097/01.fbp.0000087739.21047.d8. [DOI] [PubMed] [Google Scholar]
  • 82.McMahon LR, Amin MR, France CP. SR 141716A differentially attenuates the behavioral effects of delta9-THC in rhesus monkeys. Behav Pharmacol. 2005;16:363–72. doi: 10.1097/00008877-200509000-00008. [DOI] [PubMed] [Google Scholar]
  • 83.Wiley JL, Barrett RL, Lowe J, Balster RL, Martin BR. Discriminative stimulus effects of CP 55,940 and structurally dissimilar cannabinoids in rats. Neuropharmacology. 1995;34:669–76. doi: 10.1016/0028-3908(95)00027-4. [DOI] [PubMed] [Google Scholar]
  • 84.Wiley JL. Cannabis: discrimination of “internal bliss”? Pharmacol Biochem Behav. 1999;64:257–60. doi: 10.1016/s0091-3057(99)00059-3. [DOI] [PubMed] [Google Scholar]
  • 85.Solinas M, Panlilio LV, Antoniou K, Pappas LA, Goldberg SR. The cannabinoid CB1 antagonist N-piperidinyl-5-(4-chlorophenyl)-1-(2,4-dichlorophenyl) -4-methylpyrazole-3-carboxamide (SR-141716A) differentially alters the reinforcing effects of heroin under continuous reinforcement, fixed ratio, and progressive ratio schedules of drug self-administration in rats. J Pharmacol Exp Ther. 2003;306:93–102. doi: 10.1124/jpet.102.047928. [DOI] [PubMed] [Google Scholar]
  • 86.Solinas M, Zangen A, Thiriet N, Goldberg SR. Beta-endorphin elevations in the ventral tegmental area regulate the discriminative effects of Delta-9-tetrahydrocannabinol. Eur J Neurosci. 2004;19:3183–92. doi: 10.1111/j.0953-816X.2004.03420.x. [DOI] [PubMed] [Google Scholar]
  • 87.Alici T, Appel JB. Increasing the selectivity of the discriminative stimulus effects of delta 9-tetrahydrocannabinol: complete substitution with methanandamide. Pharmacol Biochem Behav. 2004;79:431–7. doi: 10.1016/j.pbb.2004.08.020. [DOI] [PubMed] [Google Scholar]
  • 88.Browne RG, Weissman A. Discriminative stimulus properties of delta 9-tetrahydrocannabinol: mechanistic studies. J Clin Pharmacol. 1981;21:227S–234S. doi: 10.1002/j.1552-4604.1981.tb02599.x. [DOI] [PubMed] [Google Scholar]
  • 89.Solinas M, Panlilio LV, Justinova Z, Yasar S, Goldberg SR. Using drug-discrimination techniques to study the abuse-related effects of psychoactive drugs in rats. Nature Protocols. 2006;1:1194–1206. doi: 10.1038/nprot.2006.167. [DOI] [PubMed] [Google Scholar]
  • 90.Colpaert FC. Drug discrimination in neurobiology. Pharmacol Biochem Behav. 1999;64:337–45. doi: 10.1016/s0091-3057(99)00047-7. [DOI] [PubMed] [Google Scholar]
  • 91.Jarbe TU, Lamb RJ, Lin S, Makriyannis A. (R)-methanandamide and Delta 9-THC as discriminative stimuli in rats: tests with the cannabinoid antagonist SR-141716 and the endogenous ligand anandamide. Psychopharmacology (Berl) 2001;156:369–80. doi: 10.1007/s002130100730. [DOI] [PubMed] [Google Scholar]
  • 92.Solinas M, Goldberg SR. Involvement of mu-, delta- and kappa-opioid receptor subtypes in the discriminative-stimulus effects of delta-9-tetrahydrocannabinol (THC) in rats. Psychopharmacology (Berl) 2005;179:804–12. doi: 10.1007/s00213-004-2118-x. [DOI] [PubMed] [Google Scholar]
  • 93.Solinas M, Scherma M, Tanda G, Wertheim CE, Fratta W, Goldberg SR. Nicotinic facilitation of delta9-tetrahydrocannabinol discrimination involves endogenous anandamide. J Pharmacol Exp Ther. 2007;321:1127–34. doi: 10.1124/jpet.106.116830. [DOI] [PubMed] [Google Scholar]
  • 94.Wiley JL, Golden KM, Ryan WJ, Balster RL, Razdan RK, Martin BR. Evaluation of cannabimimetic discriminative stimulus effects of anandamide and methylated fluoroanandamide in rhesus monkeys. Pharmacol Biochem Behav. 1997;58:1139–43. doi: 10.1016/s0091-3057(97)00327-4. [DOI] [PubMed] [Google Scholar]
  • 95.Jarbe TU, Lamb RJ, Liu Q, Makriyannis A. Discriminative stimulus functions of AM-1346, a CB1R selective anandamide analog in rats trained with Delta9-THC or (R)-methanandamide (AM-356) Psychopharmacology (Berl) 2006;188:315–23. doi: 10.1007/s00213-006-0517-x. [DOI] [PubMed] [Google Scholar]
  • 96.Wiley JL, LaVecchia KL, Karp NE, Kulasegram S, Mahadevan A, Razdan RK, Martin BR. A comparison of the discriminative stimulus effects of delta(9)-tetrahydrocannabinol and O-1812, a potent and metabolically stable anandamide analog, in rats. Exp Clin Psychopharmacol. 2004;12:173–9. doi: 10.1037/1064-1297.12.3.173. [DOI] [PubMed] [Google Scholar]
  • 97.Di Chiara G, Bassareo V, Fenu S, De Luca MA, Spina L, Cadoni C, Acquas E, Carboni E, Valentini V, Lecca D. Dopamine and drug addiction: the nucleus accumbens shell connection. Neuropharmacology. 2004;47(Suppl 1):227–41. doi: 10.1016/j.neuropharm.2004.06.032. [DOI] [PubMed] [Google Scholar]
  • 98.Everitt BJ, Robbins TW. Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nat Neurosci. 2005;8:1481–9. doi: 10.1038/nn1579. [DOI] [PubMed] [Google Scholar]
  • 99.Koob GF. Neurobiology of addiction. Toward the development of new therapies. Ann N Y Acad Sci. 2000;909:170–85. doi: 10.1111/j.1749-6632.2000.tb06682.x. [DOI] [PubMed] [Google Scholar]
  • 100.Wise RA. Dopamine, learning and motivation. Nat Rev Neurosci. 2004;5:483–94. doi: 10.1038/nrn1406. [DOI] [PubMed] [Google Scholar]
  • 101.Chen JP, Paredes W, Lowinson JH, Gardner EL. Strain-specific facilitation of dopamine efflux by delta 9-tetrahydrocannabinol in the nucleus accumbens of rat: an in vivo microdialysis study. Neurosci Lett. 1991;129:136–80. doi: 10.1016/0304-3940(91)90739-g. [DOI] [PubMed] [Google Scholar]
  • 102.Chen J, Marmur R, Pulles A, Paredes W, Gardner EL. Ventral tegmental microinjection of delta 9-tetrahydrocannabinol enhances ventral tegmental somatodendritic dopamine levels but not forebrain dopamine levels: evidence for local neural action by marijuana’s psychoactive ingredient. Brain Res. 1993;621:65–70. doi: 10.1016/0006-8993(93)90298-2. [DOI] [PubMed] [Google Scholar]
  • 103.Chen JP, Paredes W, Li J, Smith D, Lowinson J, Gardner EL. Delta 9-tetrahydrocannabinol produces naloxone-blockable enhancement of presynaptic basal dopamine efflux in nucleus accumbens of conscious, freely-moving rats as measured by intracerebral microdialysis. Psychopharmacology (Berl) 1990;102:156–62. doi: 10.1007/BF02245916. [DOI] [PubMed] [Google Scholar]
  • 104.Tanda G, Pontieri FE, Di Chiara G. Cannabinoid and heroin activation of mesolimbic dopamine transmission by a common mu1 opioid receptor mechanism. Science. 1997;276:2048–50. doi: 10.1126/science.276.5321.2048. [DOI] [PubMed] [Google Scholar]
  • 105.Cheer JF, Wassum KM, Heien ML, Phillips PE, Wightman RM. Cannabinoids enhance subsecond dopamine release in the nucleus accumbens of awake rats. J Neurosci. 2004;24:4393–400. doi: 10.1523/JNEUROSCI.0529-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gessa GL, Melis M, Muntoni AL, Diana M. Cannabinoids activate mesolimbic dopamine neurons by an action on cannabinoid CB1 receptors. Eur J Pharmacol. 1998;341:39–44. doi: 10.1016/s0014-2999(97)01442-8. [DOI] [PubMed] [Google Scholar]
  • 107.Melis M, Gessa GL, Diana M. Different mechanisms for dopaminergic excitation induced by opiates and cannabinoids in the rat midbrain. Prog Neuropsychopharmacol Biol Psychiatry. 2000;24:993–1006. doi: 10.1016/s0278-5846(00)00119-6. [DOI] [PubMed] [Google Scholar]
  • 108.French ED, Dillon K, Wu X. Cannabinoids excite dopamine neurons in the ventral tegmentum and substantia nigra. Neuroreport. 1997;8:649–52. doi: 10.1097/00001756-199702100-00014. [DOI] [PubMed] [Google Scholar]
  • 109.French ED. delta9-Tetrahydrocannabinol excites rat VTA dopamine neurons through activation of cannabinoid CB1 but not opioid receptors. Neurosci Lett. 1997;226:159–62. doi: 10.1016/s0304-3940(97)00278-4. [DOI] [PubMed] [Google Scholar]
  • 110.van Ree JM, Gerrits MA, Vanderschuren LJ. Opioids, reward and addiction: An encounter of biology, psychology, and medicine. Pharmacol Rev. 1999;51:341–96. [PubMed] [Google Scholar]
  • 111.Van Ree JM, Niesink RJ, Van Wolfswinkel L, Ramsey NF, Kornet MM, Van Furth WR, Vanderschuren LJ, Gerrits MA, Van den Berg CL. Endogenous opioids and reward. Eur J Pharmacol. 2000;405:89–101. doi: 10.1016/s0014-2999(00)00544-6. [DOI] [PubMed] [Google Scholar]
  • 112.Valverde O, Noble F, Beslot F, Dauge V, Fournie-Zaluski MC, Roques BP. Delta9-tetrahydrocannabinol releases and facilitates the effects of endogenous enkephalins: reduction in morphine withdrawal syndrome without change in rewarding effect. Eur J Neurosci. 2001;13:1816–24. doi: 10.1046/j.0953-816x.2001.01558.x. [DOI] [PubMed] [Google Scholar]
  • 113.Justinova Z, Tanda G, Munzar P, Goldberg SR. The opioid antagonist naltrexone reduces the reinforcing effects of Delta 9 tetrahydrocannabinol (THC) in squirrel monkeys. Psychopharmacology (Berl) 2004;173:186–94. doi: 10.1007/s00213-003-1693-6. [DOI] [PubMed] [Google Scholar]
  • 114.Hungund BL, Basavarajappa BS, Vadasz C, Kunos G, Rodriguez de Fonseca F, Colombo G, Serra S, Parsons L, Koob GF. Ethanol, endocannabinoids, and the cannabinoidergic signaling system. Alcohol Clin Exp Res. 2002;26:565–74. [PubMed] [Google Scholar]
  • 115.Mansour A, Khachaturian H, Lewis ME, Akil H, Watson SJ. Autoradiographic differentiation of mu, delta, and kappa opioid receptors in the rat forebrain and midbrain. J Neurosci. 1987;7:2445–64. [PMC free article] [PubMed] [Google Scholar]
  • 116.Reisine T, Law SF, Blake A, Tallent M. Molecular mechanisms of opiate receptor coupling to G proteins and effector systems. Ann N Y Acad Sci. 1996;780:168–75. doi: 10.1111/j.1749-6632.1996.tb15121.x. [DOI] [PubMed] [Google Scholar]
  • 117.Howlett AC. The cannabinoid receptors. Prostaglandins Other Lipid Mediat. 2002;68–69:619–31. doi: 10.1016/s0090-6980(02)00060-6. [DOI] [PubMed] [Google Scholar]
  • 118.Kathmann M, Flau K, Redmer A, Trankle C, Schlicker E. Cannabidiol is an allosteric modulator at mu- and delta-opioid receptors. Naunyn Schmiedebergs Arch Pharmacol. 2006;372:354–61. doi: 10.1007/s00210-006-0033-x. [DOI] [PubMed] [Google Scholar]
  • 119.Berrendero F, Mendizabal V, Murtra P, Kieffer BL, Maldonado R. Cannabinoid receptor and WIN 55 212-2-stimulated [35S]-GTPgammaS binding in the brain of mu-, delta- and kappa-opioid receptor knockout mice. Eur J Neurosci. 2003;18:2197–202. doi: 10.1046/j.1460-9568.2003.02951.x. [DOI] [PubMed] [Google Scholar]
  • 120.Rinaldi-Carmona M, Barth F, Heaulme M, Shire D, Calandra B, Congy C, Martinez S, Maruani J, Neliat G, Caput D, et al. SR141716A, a potent and selective antagonist of the brain cannabinoid receptor. FEBS Lett. 1994;350:240–4. doi: 10.1016/0014-5793(94)00773-x. [DOI] [PubMed] [Google Scholar]
  • 121.Gatley SJ, Gifford AN, Volkow ND, Lan R, Makriyannis A. 123I-labeled AM251: a radioiodinated ligand which binds in vivo to mouse brain cannabinoid CB1 receptors. Eur J Pharmacol. 1996;307:331–8. doi: 10.1016/0014-2999(96)00279-8. [DOI] [PubMed] [Google Scholar]
  • 122.Ledent C, Valverde O, Cossu G, Petitet F, Aubert JF, Beslot F, Bohme GA, Imperato A, Pedrazzini T, Roques BP, Vassart G, Fratta W, Parmentier M. Unresponsiveness to cannabinoids and reduced addictive effects of opiates in CB1 receptor knockout mice. Science. 1999;283:401–4. doi: 10.1126/science.283.5400.401. [DOI] [PubMed] [Google Scholar]
  • 123.Zimmer A, Zimmer AM, Hohmann AG, Herkenham M, Bonner TI. Increased mortality, hypoactivity, and hypoalgesia in cannabinoid CB1 receptor knockout mice. Proc Natl Acad Sci U S A. 1999;96:5780–5. doi: 10.1073/pnas.96.10.5780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Rodriguez JJ, Mackie K, Pickel VM. Ultrastructural localization of the CB1 cannabinoid receptor in mu-opioid receptor patches of the rat Caudate putamen nucleus. J Neurosci. 2001;21:823–33. doi: 10.1523/JNEUROSCI.21-03-00823.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Pickel VM, Chan J, Kash TL, Rodriguez JJ, MacKie K. Compartment-specific localization of cannabinoid 1 (CB1) and mu-opioid receptors in rat nucleus accumbens. Neuroscience. 2004;127:101–12. doi: 10.1016/j.neuroscience.2004.05.015. [DOI] [PubMed] [Google Scholar]
  • 126.Herkenham M, Lynn AB, Johnson MR, Melvin LS, de Costa BR, Rice KC. Characterization and localization of cannabinoid receptors in rat brain: a quantitative in vitro autoradiographic study. J Neurosci. 1991;11:563–83. doi: 10.1523/JNEUROSCI.11-02-00563.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Herkenham M, Lynn AB, Little MD, Johnson MR, Melvin LS, de Costa BR, Rice KC. Cannabinoid receptor localization in brain. Proc Natl Acad Sci U S A. 1990;87:1932–6. doi: 10.1073/pnas.87.5.1932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Fattore L, Deiana S, Spano SM, Cossu G, Fadda P, Scherma M, Fratta W. Endocannabinoid system and opioid addiction: behavioural aspects. Pharmacol Biochem Behav. 2005;81:343–59. doi: 10.1016/j.pbb.2005.01.031. [DOI] [PubMed] [Google Scholar]
  • 129.Manzanares J, Corchero J, Romero J, Fernandez-Ruiz JJ, Ramos JA, Fuentes JA. Pharmacological and biochemical interactions between opioids and cannabinoids. Trends Pharmacol Sci. 1999;20:287–94. doi: 10.1016/s0165-6147(99)01339-5. [DOI] [PubMed] [Google Scholar]
  • 130.Maldonado R, Valverde O. Participation of the opioid system in cannabinoid-induced antinociception and emotional-like responses. Eur Neuropsychopharmacol. 2003;13:401–10. doi: 10.1016/j.euroneuro.2003.08.001. [DOI] [PubMed] [Google Scholar]
  • 131.Corchero J, Manzanares J, Fuentes JA. Cannabinoid/opioid crosstalk in the central nervous system. Crit Rev Neurobiol. 2004;16:159–72. doi: 10.1615/critrevneurobiol.v16.i12.170. [DOI] [PubMed] [Google Scholar]
  • 132.Vigano D, Rubino T, Parolaro D. Molecular and cellular basis of cannabinoid and opioid interactions. Pharmacol Biochem Behav. 2005;81:360–8. doi: 10.1016/j.pbb.2005.01.021. [DOI] [PubMed] [Google Scholar]
  • 133.Caille S, Parsons LH. SR141716A reduces the reinforcing properties of heroin but not heroin-induced increases in nucleus accumbens dopamine in rats. Eur J Neurosci. 2003;18:3145–9. doi: 10.1111/j.1460-9568.2003.02961.x. [DOI] [PubMed] [Google Scholar]
  • 134.De Vries TJ, Homberg JR, Binnekade R, Raaso H, Schoffelmeer AN. Cannabinoid modulation of the reinforcing and motivational properties of heroin and heroin-associated cues in rats. Psychopharmacology (Berl) 2003;168:164–9. doi: 10.1007/s00213-003-1422-1. [DOI] [PubMed] [Google Scholar]
  • 135.Navarro M, Carrera MR, Fratta W, Valverde O, Cossu G, Fattore L, Chowen JA, Gomez R, del Arco I, Villanua MA, Maldonado R, Koob GF, Rodriguez de Fonseca F. Functional interaction between opioid and cannabinoid receptors in drug self-administration. J Neurosci. 2001;21:5344–50. doi: 10.1523/JNEUROSCI.21-14-05344.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Caille S, Parsons LH. Cannabinoid modulation of opiate reinforcement through the ventral striatopallidal pathway. Neuropsychopharmacology. 2006;31:804–13. doi: 10.1038/sj.npp.1300848. [DOI] [PubMed] [Google Scholar]
  • 137.Solinas M, Panlilio LV, Tanda G, Makriyannis A, Matthews SA, Goldberg SR. Cannabinoid agonists but not inhibitors of endogenous cannabinoid transport or metabolism enhance the reinforcing efficacy of heroin in rats. Neuropsychopharmacology. 2005;30:2046–57. doi: 10.1038/sj.npp.1300754. [DOI] [PubMed] [Google Scholar]
  • 138.Rice OV, Gordon N, Gifford AN. Conditioned place preference to morphine in cannabinoid CB1 receptor knockout mice. Brain Res. 2002;945:135–8. doi: 10.1016/s0006-8993(02)02890-1. [DOI] [PubMed] [Google Scholar]
  • 139.Mascia MS, Obinu MC, Ledent C, Parmentier M, Bohme GA, Imperato A, Fratta W. Lack of morphine-induced dopamine release in the nucleus accumbens of cannabinoid CB(1) receptor knockout mice. Eur J Pharmacol. 1999;383:R1–2. doi: 10.1016/s0014-2999(99)00656-1. [DOI] [PubMed] [Google Scholar]
  • 140.De Vries TJ, Shaham Y, Homberg JR, Crombag H, Schuurman K, Dieben J, Vanderschuren LJ, Schoffelmeer AN. A cannabinoid mechanism in relapse to cocaine seeking. Nat Med. 2001;7:1151–4. doi: 10.1038/nm1001-1151. [DOI] [PubMed] [Google Scholar]
  • 141.Filip M, Golda A, Zaniewska M, McCreary AC, Nowak E, Kolasiewicz W, Przegalinski E. Involvement of cannabinoid CB(1) receptors in drug addiction: effects of rimonabant on behavioral responses induced by cocaine. Pharmacol Rep. 2006;58:806–19. [PubMed] [Google Scholar]
  • 142.Cossu G, Ledent C, Fattore L, Imperato A, Bohme GA, Parmentier M, Fratta W. Cannabinoid CB1 receptor knockout mice fail to self-administer morphine but not other drugs of abuse. Behav Brain Res. 2001;118:61–5. doi: 10.1016/s0166-4328(00)00311-9. [DOI] [PubMed] [Google Scholar]
  • 143.Soria G, Mendizabal V, Tourino C, Robledo P, Ledent C, Parmentier M, Maldonado R, Valverde O. Lack of CB1 cannabinoid receptor impairs cocaine self-administration. Neuropsychopharmacology. 2005;30:1670–80. doi: 10.1038/sj.npp.1300707. [DOI] [PubMed] [Google Scholar]
  • 144.Houchi H, Babovic D, Pierrefiche O, Ledent C, Daoust M, Naassila M. CB1 receptor knockout mice display reduced ethanol-induced conditioned place preference and increased striatal dopamine D2 receptors. Neuropsychopharmacology. 2005;30:339–49. doi: 10.1038/sj.npp.1300568. [DOI] [PubMed] [Google Scholar]
  • 145.Cheer JF, Wassum KM, Sombers LA, Heien ML, Ariansen JL, Aragona BJ, Phillips PE, Wightman RM. Phasic dopamine release evoked by abused substances requires cannabinoid receptor activation. J Neurosci. 2007;27:791–5. doi: 10.1523/JNEUROSCI.4152-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Vinklerova J, Novakova J, Sulcova A. Inhibition of methamphetamine self-administration in rats by cannabinoid receptor antagonist AM 251. J Psychopharmacol. 2002;16:139–43. doi: 10.1177/026988110201600204. [DOI] [PubMed] [Google Scholar]
  • 147.Colombo G, Serra S, Vacca G, Carai MA, Gessa GL. Endocannabinoid system and alcohol addiction: pharmacological studies. Pharmacol Biochem Behav. 2005;81:369–80. doi: 10.1016/j.pbb.2005.01.022. [DOI] [PubMed] [Google Scholar]
  • 148.Cippitelli A, Bilbao A, Hansson AC, del Arco I, Sommer W, Heilig M, Massi M, Bermudez-Silva FJ, Navarro M, Ciccocioppo R, de Fonseca FR. Cannabinoid CB1 receptor antagonism reduces conditioned reinstatement of ethanol-seeking behavior in rats. Eur J Neurosci. 2005;21:2243–51. doi: 10.1111/j.1460-9568.2005.04056.x. [DOI] [PubMed] [Google Scholar]
  • 149.Economidou D, Mattioli L, Cifani C, Perfumi M, Massi M, Cuomo V, Trabace L, Ciccocioppo R. Effect of the cannabinoid CB1 receptor antagonist SR-141716A on ethanol self-administration and ethanol-seeking behaviour in rats. Psychopharmacology (Berl) 2006;183:394–403. doi: 10.1007/s00213-005-0199-9. [DOI] [PubMed] [Google Scholar]
  • 150.Colombo G, Agabio R, Fa M, Guano L, Lobina C, Loche A, Reali R, Gessa GL. Reduction of voluntary ethanol intake in ethanol-preferring sP rats by the cannabinoid antagonist SR-141716. Alcohol Alcohol. 1998;33:126–30. doi: 10.1093/oxfordjournals.alcalc.a008368. [DOI] [PubMed] [Google Scholar]
  • 151.Freedland CS, Sharpe AL, Samson HH, Porrino LJ. Effects of SR141716A on ethanol and sucrose self-administration. Alcohol Clin Exp Res. 2001;25:277–82. [PubMed] [Google Scholar]
  • 152.Gallate JE, McGregor IS. The motivation for beer in rats: effects of ritanserin, naloxone and SR 141716. Psychopharmacology (Berl) 1999;142:302–8. doi: 10.1007/s002130050893. [DOI] [PubMed] [Google Scholar]
  • 153.Gessa GL, Serra S, Vacca G, Carai MA, Colombo G. Suppressing effect of the cannabinoid CB1 receptor antagonist, SR147778, on alcohol intake and motivational properties of alcohol in alcohol-preferring sP rats. Alcohol Alcohol. 2005;40:46–53. doi: 10.1093/alcalc/agh114. [DOI] [PubMed] [Google Scholar]
  • 154.Lallemand F, De Witte P. SR147778, a CB1 cannabinoid receptor antagonist, suppresses ethanol preference in chronically alcoholized Wistar rats. Alcohol. 2006;39:125–34. doi: 10.1016/j.alcohol.2006.08.001. [DOI] [PubMed] [Google Scholar]
  • 155.Arnone M, Maruani J, Chaperon F, Thiebot MH, Poncelet M, Soubrie P, Le Fur G. Selective inhibition of sucrose and ethanol intake by SR 141716, an antagonist of central cannabinoid (CB1) receptors. Psychopharmacology (Berl) 1997;132:104–6. doi: 10.1007/s002130050326. [DOI] [PubMed] [Google Scholar]
  • 156.Hungund BL, Szakall I, Adam A, Basavarajappa BS, Vadasz C. Cannabinoid CB1 receptor knockout mice exhibit markedly reduced voluntary alcohol consumption and lack alcohol-induced dopamine release in the nucleus accumbens. J Neurochem. 2003;84:698–704. doi: 10.1046/j.1471-4159.2003.01576.x. [DOI] [PubMed] [Google Scholar]
  • 157.Naassila M, Pierrefiche O, Ledent C, Daoust M. Decreased alcohol self-administration and increased alcohol sensitivity and withdrawal in CB1 receptor knockout mice. Neuropharmacology. 2004;46:243–53. doi: 10.1016/j.neuropharm.2003.09.002. [DOI] [PubMed] [Google Scholar]
  • 158.Wang L, Liu J, Harvey-White J, Zimmer A, Kunos G. Endocannabinoid signaling via cannabinoid receptor 1 is involved in ethanol preference and its age-dependent decline in mice. Proc Natl Acad Sci U S A. 2003;100:1393–8. doi: 10.1073/pnas.0336351100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Thanos PK, Dimitrakakis ES, Rice O, Gifford A, Volkow ND. Ethanol self-administration and ethanol conditioned place preference are reduced in mice lacking cannabinoid CB1 receptors. Behav Brain Res. 2005;164:206–13. doi: 10.1016/j.bbr.2005.06.021. [DOI] [PubMed] [Google Scholar]
  • 160.Colombo G, Serra S, Brunetti G, Gomez R, Melis S, Vacca G, Carai MM, Gessa L. Stimulation of voluntary ethanol intake by cannabinoid receptor agonists in ethanol-preferring sP rats. Psychopharmacology (Berl) 2002;159:181–7. doi: 10.1007/s002130100887. [DOI] [PubMed] [Google Scholar]
  • 161.Gallate JE, Saharov T, Mallet PE, McGregor IS. Increased motivation for beer in rats following administration of a cannabinoid CB1 receptor agonist. Eur J Pharmacol. 1999;370:233–40. doi: 10.1016/s0014-2999(99)00170-3. [DOI] [PubMed] [Google Scholar]
  • 162.Horvath TL. The unfolding cannabinoid story on energy homeostasis: central or peripheral site of action? Int J Obes (Lond) 2006;30(Suppl 1):S30–2. doi: 10.1038/sj.ijo.0803275. [DOI] [PubMed] [Google Scholar]
  • 163.Cota D, Tschop MH, Horvath TL, Levine AS. Cannabinoids, opioids and eating behavior: the molecular face of hedonism? Brain Res Brain Res Rev. 2006;51:85–107. doi: 10.1016/j.brainresrev.2005.10.004. [DOI] [PubMed] [Google Scholar]
  • 164.Matias I, Bisogno T, Di Marzo V. Endogenous cannabinoids in the brain and peripheral tissues: regulation of their levels and control of food intake. Int J Obes (Lond) 2006;30(Suppl 1):S7–S12. doi: 10.1038/sj.ijo.0803271. [DOI] [PubMed] [Google Scholar]
  • 165.Solinas M, Goldberg SR. Motivational effects of cannabinoids and opioids on food reinforcement depend on simultaneous activation of cannabinoid and opioid systems. Neuropsychopharmacology. 2005;30:2035–45. doi: 10.1038/sj.npp.1300720. [DOI] [PubMed] [Google Scholar]
  • 166.Basavarajappa BS, Hungund BL. Role of the endocannabinoid system in the development of tolerance to alcohol. Alcohol Alcohol. 2005;40:15–24. doi: 10.1093/alcalc/agh111. [DOI] [PubMed] [Google Scholar]
  • 167.Blednov YA, Cravatt BF, Boehm SL, 2nd, Walker D, Harris RA. Role of endocannabinoids in alcohol consumption and intoxication: studies of mice lacking Fatty Acid amide hydrolase. Neuropsychopharmacology. 2007;32:1570–82. doi: 10.1038/sj.npp.1301274. [DOI] [PubMed] [Google Scholar]
  • 168.Hansson AC, Bermudez-Silva FJ, Malinen H, Hyytia P, Sanchez-Vera I, Rimondini R, Rodriguez de Fonseca F, Kunos G, Sommer WH, Heilig M. Genetic impairment of frontocortical endocannabinoid degradation and high alcohol preference. Neuropsychopharmacology. 2007;32:117–26. doi: 10.1038/sj.npp.1301034. [DOI] [PubMed] [Google Scholar]
  • 169.Caille S, Alvarez-Jaimes L, Polis I, Stouffer DG, Parsons LH. Specific alterations of extracellular endocannabinoid levels in the nucleus accumbens by ethanol, heroin, and cocaine self-administration. J Neurosci. 2007;27:3695–702. doi: 10.1523/JNEUROSCI.4403-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Cohen C, Perrault G, Voltz C, Steinberg R, Soubrie P. SR141716, a central cannabinoid (CB(1)) receptor antagonist, blocks the motivational and dopamine-releasing effects of nicotine in rats. Behav Pharmacol. 2002;13:451–63. doi: 10.1097/00008877-200209000-00018. [DOI] [PubMed] [Google Scholar]
  • 171.Perra S, Pillolla G, Melis M, Muntoni AL, Gessa GL, Pistis M. Involvement of the endogenous cannabinoid system in the effects of alcohol in the mesolimbic reward circuit: electrophysiological evidence in vivo. Psychopharmacology (Berl) 2005;183:368–77. doi: 10.1007/s00213-005-0195-0. [DOI] [PubMed] [Google Scholar]
  • 172.Fernandez JR, Allison DB. Rimonabant Sanofi-Synthelabo. Curr Opin Investig Drugs. 2004;5:430–5. [PubMed] [Google Scholar]
  • 173.Forget B, Hamon M, Thiebot MH. Cannabinoid CB1 receptors are involved in motivational effects of nicotine in rats. Psychopharmacology (Berl) 2005;181:722–34. doi: 10.1007/s00213-005-0015-6. [DOI] [PubMed] [Google Scholar]
  • 174.Forget B, Barthelemy S, Saurini F, Hamon M, Thiebot MH. Differential involvement of the endocannabinoid system in short- and long-term expression of incentive learning supported by nicotine in rats. Psychopharmacology (Berl) 2006;189:59–69. doi: 10.1007/s00213-006-0525-x. [DOI] [PubMed] [Google Scholar]
  • 175.Le Foll B, Goldberg SR. Rimonabant, a CB1 antagonist, blocks nicotine-conditioned place preferences. Neuroreport. 2004;15:2139–43. doi: 10.1097/00001756-200409150-00028. [DOI] [PubMed] [Google Scholar]
  • 176.Cohen C, Perrault G, Griebel G, Soubrie P. Nicotine-associated cues maintain nicotine-seeking behavior in rats several weeks after nicotine withdrawal: reversal by the cannabinoid (CB1) receptor antagonist, rimonabant (SR141716) Neuropsychopharmacology. 2005;30:145–55. doi: 10.1038/sj.npp.1300541. [DOI] [PubMed] [Google Scholar]
  • 177.Castane A, Valjent E, Ledent C, Parmentier M, Maldonado R, Valverde O. Lack of CB1 cannabinoid receptors modifies nicotine behavioural responses, but not nicotine abstinence. Neuropharmacology. 2002;43:857–67. doi: 10.1016/s0028-3908(02)00118-1. [DOI] [PubMed] [Google Scholar]
  • 178.Giuffrida A, Parsons LH, Kerr TM, Rodriguez de Fonseca F, Navarro M, Piomelli D. Dopamine activation of endogenous cannabinoid signaling in dorsal striatum. Nat Neurosci. 1999;2:358–63. doi: 10.1038/7268. [DOI] [PubMed] [Google Scholar]
  • 179.Di Marzo V, Berrendero F, Bisogno T, Gonzalez S, Cavaliere P, Romero J, Cebeira M, Ramos JA, Fernandez-Ruiz JJ. Enhancement of anandamide formation in the limbic forebrain and reduction of endocannabinoid contents in the striatum of delta9-tetrahydrocannabinol-tolerant rats. J Neurochem. 2000;74:1627–35. doi: 10.1046/j.1471-4159.2000.0741627.x. [DOI] [PubMed] [Google Scholar]
  • 180.Gonzalez S, Cascio MG, Fernandez-Ruiz J, Fezza F, Di Marzo V, Ramos JA. Changes in endocannabinoid contents in the brain of rats chronically exposed to nicotine, ethanol or cocaine. Brain Res. 2002;954:73–81. doi: 10.1016/s0006-8993(02)03344-9. [DOI] [PubMed] [Google Scholar]
  • 181.Vigano D, Grazia Cascio M, Rubino T, Fezza F, Vaccani A, Di Marzo V, Parolaro D. Chronic morphine modulates the contents of the endocannabinoid, 2-arachidonoyl glycerol, in rat brain. Neuropsychopharmacology. 2003;28:1160–7. doi: 10.1038/sj.npp.1300117. [DOI] [PubMed] [Google Scholar]
  • 182.Vigano D, Valenti M, Cascio MG, Di Marzo V, Parolaro D, Rubino T. Changes in endocannabinoid levels in a rat model of behavioural sensitization to morphine. Eur J Neurosci. 2004;20:1849–57. doi: 10.1111/j.1460-9568.2004.03645.x. [DOI] [PubMed] [Google Scholar]
  • 183.Pickel VM, Chan J, Kearn CS, Mackie K. Targeting dopamine D2 and cannabinoid-1 (CB1) receptors in rat nucleus accumbens. J Comp Neurol. 2006;495:299–313. doi: 10.1002/cne.20881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Glass M, Felder CC. Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors augments cAMP accumulation in striatal neurons: evidence for a Gs linkage to the CB1 receptor. J Neurosci. 1997;17:5327–33. doi: 10.1523/JNEUROSCI.17-14-05327.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Jarrahian A, Watts VJ, Barker EL. D2 dopamine receptors modulate Galpha-subunit coupling of the CB1 cannabinoid receptor. J Pharmacol Exp Ther. 2004;308:880–6. doi: 10.1124/jpet.103.057620. [DOI] [PubMed] [Google Scholar]
  • 186.Kearn CS, Blake-Palmer K, Daniel E, Mackie K, Glass M. Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors enhances heterodimer formation: a mechanism for receptor cross-talk? Mol Pharmacol. 2005;67:1697–704. doi: 10.1124/mol.104.006882. [DOI] [PubMed] [Google Scholar]
  • 187.DSM-IV. Diagnostic and Statistical Manual of Mental Disorders. 4. Washington, DC: American Psychiatry Association Press; 1994. [Google Scholar]
  • 188.Shaham Y, Shalev U, Lu L, De Wit H, Stewart J. The reinstatement model of drug relapse: history, methodology and major findings. Psychopharmacology (Berl) 2003;168:3–20. doi: 10.1007/s00213-002-1224-x. [DOI] [PubMed] [Google Scholar]
  • 189.Katz JL, Higgins ST. The validity of the reinstatement model of craving and relapse to drug use. Psychopharmacology (Berl) 2003;168:21–30. doi: 10.1007/s00213-003-1441-y. [DOI] [PubMed] [Google Scholar]
  • 190.Epstein DH, Preston KL. The reinstatement model and relapse prevention: a clinical perspective. Psychopharmacology (Berl) 2003;168:31–41. doi: 10.1007/s00213-003-1470-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.De Vries TJ, Schoffelmeer AN. Cannabinoid CB1 receptors control conditioned drug seeking. Trends Pharmacol Sci. 2005;26:420–6. doi: 10.1016/j.tips.2005.06.002. [DOI] [PubMed] [Google Scholar]
  • 192.Fattore L, Spano MS, Deiana S, Melis V, Cossu G, Fadda P, Fratta W. An endocannabinoid mechanism in relapse to drug seeking: a review of animal studies and clinical perspectives. Brain Res Brain Res Rev. 2007;53:1–16. doi: 10.1016/j.brainresrev.2006.05.003. [DOI] [PubMed] [Google Scholar]
  • 193.Xi ZX, Gilbert JG, Peng XQ, Pak AC, Li X, Gardner EL. Cannabinoid CB1 receptor antagonist AM251 inhibits cocaine-primed relapse in rats: role of glutamate in the nucleus accumbens. J Neurosci. 2006;26:8531–6. doi: 10.1523/JNEUROSCI.0726-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Fattore L, Spano MS, Cossu G, Deiana S, Fratta W. Cannabinoid mechanism in reinstatement of heroin-seeking after a long period of abstinence in rats. Eur J Neurosci. 2003;17:1723–6. doi: 10.1046/j.1460-9568.2003.02607.x. [DOI] [PubMed] [Google Scholar]
  • 195.Anggadiredja K, Nakamichi M, Hiranita T, Tanaka H, Shoyama Y, Watanabe S, Yamamoto T. Endocannabinoid system modulates relapse to methamphetamine seeking: possible mediation by the arachidonic acid cascade. Neuropsychopharmacology. 2004;29:1470–8. doi: 10.1038/sj.npp.1300454. [DOI] [PubMed] [Google Scholar]
  • 196.De Vries TJ, de Vries W, Janssen MC, Schoffelmeer AN. Suppression of conditioned nicotine and sucrose seeking by the cannabinoid-1 receptor antagonist SR141716A. Behav Brain Res. 2005;161:164–8. doi: 10.1016/j.bbr.2005.02.021. [DOI] [PubMed] [Google Scholar]
  • 197.Viveros MP, Marco EM, File SE. Endocannabinoid system and stress and anxiety responses. Pharmacol Biochem Behav. 2005;81:331–42. doi: 10.1016/j.pbb.2005.01.029. [DOI] [PubMed] [Google Scholar]

RESOURCES