Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Dec 1.
Published in final edited form as: Pharmacol Ther. 2011 Jul 18;132(3):215–241. doi: 10.1016/j.pharmthera.2011.06.005

ENDOCANNABINOID INFLUENCE IN DRUG REINFORCEMENT, DEPENDENCE AND ADDICTION-RELATED BEHAVIORS

Antonia Serrano 1,2, Loren H Parsons 1
PMCID: PMC3209522  NIHMSID: NIHMS323099  PMID: 21798285

Abstract

The endogenous cannabinoid system is an important regulatory system involved in physiological homeostasis. Endocannabinoid signaling is known to modulate neural development, immune function, metabolism, synaptic plasticity and emotional state. Accumulating evidence also implicates brain endocannabinoid signaling in the etiology of drug addiction which is characterized by compulsive drug seeking, loss of control in limiting drug intake, emergence of a negative emotional state in the absence of drug use and a persistent vulnerability toward relapse to drug use during protracted abstinence. In this review we discuss the effects of drug intake on brain endocannabinoid signaling, evidence implicating the endocannabinoid system in the motivation for drug consumption, and drug-induced alterations in endocannabinoid function that may contribute to various aspects of addiction including dysregulated synaptic plasticity, increased stress responsivity, negative affective states, drug craving and relapse to drug taking. Current knowledge of genetic variants in endocannabinoid signaling associated with addiction is also discussed.

Keywords: endocannabinoid, CB1 receptor, addiction, craving, relapse, synaptic plasticity

1. Introduction

The recreational use of psychoactive substances by humans has been documented for centuries, and it is well known that the repeated use of many abused drugs can lead to dependence and a progression toward addiction. Addiction is a persistent state characterized by compulsion to seek and take a drug, a loss of control in limiting drug intake even when serious negative consequences ensue and the emergence of a negative emotional state (e.g. anxiety, depression, irritability) when access to the drug is prevented (Hyman, 2005; Hyman et al., 2006; Jaffe, 1990; Koob et al., 2010; O’Brien et al., 1996; O’Brien et al., 2005). Addicts often have a persistent vulnerability to relapse to drug use after days or even years of abstinence, and this aspect of addiction presents the greatest difficulty in treating this psychiatric disease (Nestler, 2002; O’Brien, 1997; O’Brien, 2008; Shaham et al., 2005; Yahyavi-Firouz-Abadi et al., 2009).

An important goal of addiction research is to understand the neuropharmacological and neuroadaptive mechanisms that mediate the transition from occasional, controlled drug use to the loss of behavioral control over drug-seeking and drug-taking that defines addiction. Investigations of the neural substrates mediating the positive reinforcing properties of various drugs have dominated the field of addiction research for several decades. This work has established an important role for increased mesolimbic dopamine signaling in the mediation of acute drug reward. Indeed, all drugs of abuse have been shown to increase extracellular dopamine levels in the nucleus accumbens (NAc) (Di Chiara et al., 2004) though this effect is relatively less pronounced for ethanol, nicotine and opioids as compared with psychostimulants. Further studies have demonstrated significant involvement of serotonin, glutamate, GABA, acetylcholine, and various neuroactive peptides in the positive reinforcing effects of various abused drugs (Bardo, 1998; Hyman et al., 2006; Koob et al., 2004; Ross et al., 2009) and the relative importance of a large number of brain structures in addition to the mesolimbic system has been characterized (Bardo, 1998; McBride et al., 1999; Paulus, 2007).

In recent years there has been substantial effort to characterize the neural mechanisms underlying the shift from controlled to compulsive drug use and the associated transfer in motivational drive from positive to negative reinforcement that results from long-term drug exposure. The data suggest that the transition to drug dependence and addiction involves adaptations in many of the neurochemical systems implicated in the positive reinforcing effects of acute drug use, a dysregulation of synaptic plasticity and the development of maladaptive stress responses (Chen et al., 2010; Kalivas, 2009; Kalivas et al., 2008; Koob, 2008; Koob, 2009; Martin-Fardon et al., 2010; Wise et al., 2010).

A growing body of evidence points to an involvement of the endogenous cannabinoid system (ECS) in the acquisition and maintenance of drug taking behavior and in various physiological and behavioral processes associated with addiction. An assessment of this evidence is presented in the sections that follow, with a focus on the potential endocannabinoid (eCB) influence on drug reinforcement, drug-related synaptic plasticity, drug-seeking behavior (relapse), stress responsivity and affective state. We will first provide a brief overview of the ECS and a consideration of the behavioral effects produced by enhanced eCB signaling.

2. The endogenous cannabinoid system

In the early 1990s two arachidonic acid derivatives were identified as endogenous cannabinoid receptor ligands. The first eCB to be discovered was anandamide (N-arachidonoyl-ethanolamine; AEA) (Devane et al., 1992) and within two years a second endogenous lipid, 2-arachidonoyl-glycerol (2-AG), was shown to function as signaling molecule at CB1 and CB2 receptors (Mechoulam et al., 1995; Sugiura et al., 1995). Other endogenous molecules have been identified that exert cannabinoid-like effects including 2-arachidonyl-glycerol ether (noladin ether) (Hanus et al., 2001), N-arachidonoyl-dopamine (NADA) (Bisogno et al., 2000; Huang et al., 2002), virodhamine (Porter et al., 2002), N-homo-γ-linolenoylethanolamine (HEA) and N-docosatetraenoylethanolamine (DEA) (Hanus et al., 1993; Pertwee et al., 1994). However, the presence of many of these lipids in intact tissue has been a matter of debate and their pharmacological activity and metabolism have not yet been thoroughly characterized. Endocannabinoid congeners including palmitoylethanolamide (PEA) (Re et al., 2007), and oleoylethanolamide (OEA) (Rodriguez de Fonseca et al., 2001) have also been identified, though these moieties do not interact with cannabinoid receptors. Accordingly, AEA and 2-AG are still considered the primary endogenous mediators of cannabinoid signaling.

Unlike classical neurotransmitters, eCBs are not stored in intracellular compartments but are synthesized on an “as needed” basis by cleavage from membrane lipid precursors and immediate extrusion from neurons through distinct calcium-dependent mechanisms. AEA derives from the phospholipid precursor N-arachidonoyl-phosphatidylethanolamine (NAPE) (Cadas et al., 1996; Piomelli, 2003). The precise pathways through which NAPE is converted to AEA remain controversial, and at least four routes have been proposed including a direct transacylation-phosphodiesterase pathway catalyzed by a N-acyl-phosphatidylethanolamine-selective phosphodiesterase (NAPE-PLD) (Di Marzo et al., 2007; Liu et al., 2008; Okamoto et al., 2004) whose activity is regulated by depolarization and/or activation of ionotropic (Piomelli, 2003; Stella et al., 2001) or metabotropic receptors (Giuffrida et al., 1999; Kim et al., 2002; Varma et al., 2001). Although this has been considered to be a major route for AEA production, NAPE-PLD knockout mice were found to have unaltered AEA levels in brain (Leung et al., 2006) clearly pointing to the presence of additional metabolic pathways for AEA formation. At least two pathways distinct from NAPE-PLD have been proposed. One pathway involves the double-O-deacylation of NAPEs by α,β-hydrolase (ABHD4) to form glycerophospho-N-acylethanolamines (GP-NAEs), followed by conversion of these intermediates to NAEs by glycerophosphodiesterase-1 (GDE1) (Simon et al., 2006; Simon et al., 2008). Another pathway utilizes a phospholipase C (PLC) to produce phopho-N-arachidonoylethanolamine (pAEA) from NAPE, followed by conversion of pAEA into AEA by phosphatases such as PTPN22 and SHIP1 (Liu et al., 2006; Liu et al., 2008). 2-AG derives primarily from the hydrolytic metabolism of 1,2-diacylglycerol (DAG) mediated by two sn-1-selective DAG lipases, DAGL-α and DAGL-β (Bisogno et al., 2003; Piomelli, 2003; Stella et al., 1997) though alternate biosynthetic routes have been described (Sugiura et al., 1995).

Inactivation of eCB signaling is mediated by cellular reuptake into both neurons and glial cells (Beltramo et al., 1997; Hillard et al., 2000) and subsequent intracellular hydrolysis. Fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) have been identified as enzymes primarily responsible for the degradation of AEA and 2-AG, respectively. FAAH can also hydrolyze 2-AG (Goparaju et al., 1998) and other fatty acid primary amides, N-acyltaurines and N-acyl amino acids. The cloning, structural and kinetic properties, distribution in the body and crystal structure of FAAH have been described (Ahn et al., 2008; McKinney et al., 2005). FAAH is abundantly expressed in the CNS and FAAH-positive neurons are found in proximity to CB1 receptor-containing terminals, supporting a role for this enzyme in eCB inactivation (McKinney et al., 2005). 2-AG hydrolysis is performed by multiple enzymes, though the primary mechanism of clearance appears to be mediated by MAGL (Blankman et al., 2007; Dinh et al., 2002). MAGL cloning, structural and catalytic properties have recently been reviewed (Saario et al., 2007). In addition to MAGL, the enzymes ABHD6 and ABHD12 appear to play an important role in 2-AG metabolism (Blankman et al., 2007; Marrs et al., 2010).

Two major types of cannabinoid receptors have been characterized and cloned: CB1 and CB2. CB1 receptors are highly expressed in brain and are also found in peripheral tissues (Batkai et al., 2001; Herkenham et al., 1991a; Herkenham et al., 1991b; Herkenham et al., 1990; Howlett et al., 1990). CB2 receptors are mainly located in immune cells (Munro et al., 1993), although there is evidence for CB2 receptor expression on neurons, glia and endothelial cells in brain (Ashton et al., 2007; Atwood et al., 2010; Brusco et al., 2008a; Brusco et al., 2008b; Jhaveri et al., 2008; Onaivi et al., 2006; Onaivi et al., 2008; Van Sickle et al., 2005; Viscomi et al., 2009). CB1 and CB2 receptors are both coupled to similar transduction systems through Gi or Go proteins. Their activation reduces adenylate cyclase activity and Ca2+ influx through N-, P/Q- and L-type Ca2+ channels (Gebremedhin et al., 1999; Mackie et al., 1992; Twitchell et al., 1997). CB receptor activation also stimulates inwardly rectifying potassium channels and the mitogen-activated protein kinase pathway (Howlett, 2005; Mackie et al., 1995).

AEA and 2-AG each exert agonist activity at both CB1 and CB2 receptors. AEA binds with slightly higher affinity to CB1 than CB2, and similar to Δ9-THC it functions as a partial agonist at both CB receptors, though it is less efficacious at CB2 as compared with CB1 receptors (Pertwee, 2010). 2-AG binds with more or less equal affinity to CB1 and CB2 receptors, and has greater potency and efficacy than AEA at these receptors (Pertwee, 2010). AEA and 2-AG also function as agonists at orphan G-protein coupled receptors that may be members of the cannabinoid receptor family such as GPR55 and GPR119 (Godlewski et al., 2009; Lauckner et al., 2008; Overton et al., 2006; Pertwee, 2010; Ryberg et al., 2007; Sharir et al., 2010), and AEA potently activates non-cannabinoid receptors such as transient receptor potential vanilloid type-1 (TRPV1) receptors (Di Marzo et al., 2010; Zygmunt et al., 1999). Thus, the physiological and behavioral effects produced by eCBs result from agonist effects at both cannabinoid (CB1 and CB2) and non-cannabinoid receptors. For purposes of clarity, however, this review will focus primarily on the influence of eCB signaling through CB1 receptors.

Since its discovery the ECS has been implicated in the regulation of a range of physiological processes including neural development (Fride, 2008), immune function (Pandey et al., 2009; Stella, 2009), metabolism and energy homeostasis (Viveros et al., 2008), synaptic plasticity and learning (Heifets et al., 2009), pain (Guindon et al., 2009) and emotional state (Lutz, 2009). There has also been substantial interest in the possible involvement of the ECS in the etiology of drug abuse and addiction, largely due to the dense CB1 receptor expression in brain regions involved the motivational and addictive properties of abused drugs, and the similar pharmacological effects produced by eCBs and exogenous cannabinoids such as Δ9-tetrahydrocannabinol (Δ9-THC).

3. Behavioral effects produced by endocannabinoids

In the early 1990s Billy Martin and colleagues developed a battery of behavioral tests as an in vivo assay for evaluating cannabimimetic effects (Martin et al., 1991). Collectively these procedures were shown to provide a sensitive index of a drug affinity and efficacy at centrally located cannabinoid receptors. Although the full battery includes various tests in multiple species, the primary collection of tests is the rodent “tetrad”. Cannabinoid agonists such as Δ9-THC, WIN 55,212-2, and CP 55,940 produce a characteristic combination of four symptoms; hypoactivity, catalepsy, hypothermia and analgesia. Each of these effects are reversed by the CB1 receptor antagonist SR 141716, providing good evidence for the involvement of CB1-mediated mechanisms. Exogenous AEA administration produces behavioral effects similar to the CB1 agonists mentioned above, though with much lower potency and of shorter duration (Crawley et al., 1993; Fride et al., 1993; Smith et al., 1994), most likely due to the rapid metabolism of this eCB. Although initial evaluations demonstrated that the hypothermia, analgesia and catalepsy induced by AEA are not reversed by the CB1 receptor antagonist SR141716A (Adams, Compton, & Martin, 1998) subsequent tests revealed that the hypothermic effects of AEA and two AEA analogs (ACEA and ACPA) are attenuated by SR141716A pretreatment (Hillard et al., 1999). Further studies have revealed that selective FAAH and MAGL inhibitors produce an intriguing subset of the behavioral effects observed with direct CB1 agonists (Kinsey et al., 2009; Lichtman et al., 2004; Long et al., 2009a; Long et al., 2009c; Russo et al., 2007). For example, inhibition of MAGL, but not FAAH, causes CB1-dependent hypomotility (Long et al., 2009c) and neither FAAH nor MAGL inhibition induces catalepsy as typically observed with direct CB1 agonists. Thus, while AEA and 2-AG each produce some behavioral responses characteristic of Δ9-THC there are notable distinctions between the behavioral effects of eCBs and CB1 agonists.

Several preclinical studies have investigated whether enhanced eCB signaling produces rewarding effects, and much of this work has compared eCB-induced behaviors with those produced by CB1 agonists. With one published exception (Wiley et al., 1995), exogenous AEA has not substituted for Δ9-THC in drug discrimination tests except at quite high doses that also produce substantial motor disturbances (Burkey & Nation, 1997; Wiley, et al., 1997; Wiley, Ryan, Razdan, & Martin, 1998; Solinas, et al., 2007). Very rapid AEA metabolism likely confers a very short duration of action, and this may be responsible for the lack of generalization observed in most studies. Consistent with this hypothesis is evidence that the metabolically stable AEA analogs (R)-methanandamide and O-1812 as well as the CB1-selective AEA analog AM1346 each fully substitute for Δ9-THC in discrimination tasks performed in rats (Alici et al., 2004; Burkey et al., 1997; Jarbe et al., 2000; Jarbe et al., 2001; Jarbe et al., 2006; Jarbe et al., 1998; Solinas et al., 2007; Wiley et al., 2004) though (R)-methanandamide does not generalize to the Δ9-THC stimulus in C58BL/6J mice or Rhesus monkeys (McMahon, 2006; McMahon et al., 2008). In rats and mice, treatment with the selective FAAH inhibitor URB597 or the non-selective serine hydrolase inhibitor PMSF potentiates the discriminative stimulus properties of AEA sufficiently to fully substitute for Δ9-THC (Solinas et al., 2007; Vann et al., 2009), though URB597 alone does not produce Δ9-THC appropriate responding (Gobbi et al., 2005; Solinas et al., 2007). Recent studies have also shown that Δ9-THC and (R)-methanandamide each fully generalize to the discriminative stimulus produced by the CB1-selective AEA analog AM1346 (Jarbe et al., 2009) and Δ9-THC, WIN 55,212-2 and the selective CB1 agonist AM678 fully generalize to the discriminative stimulus of (R)-methanandamide through a CB1 receptor-sensitive mechanism (Jarbe et al., 2010). The discriminative properties of (R)-methanandamide, AEA (following URB597 administration), WIN 55,212-2, AM678, AM1346 and Δ9-THC are each blocked by SR 141716A and are therefore thought to be CB1 receptor dependent (Jarbe et al., 2000; Jarbe et al., 2001; Jarbe et al., 2009; Jarbe et al., 2010; Jarbe et al., 1998; Solinas et al., 2007). These studies provide strong evidence of similar pharmacological effects produced by phytocannabinoids (e.g. Δ9-THC), synthetic CB1 agonists and AEA-like moieties. However, reported differences in discriminative stimulus potency and CB1 antagonist surmountability suggest that CB1 receptor activation by these ligands may engage distinct signaling pathways (Jarbe et al., 2009; Jarbe et al., 2010). Recent work has shown that the MAGL inhibitor JZL184 produces only partial generalization to the Δ9-THC discriminative stimulus, though dual inhibition of MAGL and FAAH by JZL195 produces much more reliable Δ9-THC appropriate responding (Long et al., 2009b). Collectively these findings indicate that systemic administration of eCBs produces CB1-dependent Δ9-THC-like discriminative stimulus effects, particularly following inhibition of eCB clearance mechanisms. In addition, recent evidence from the Goldberg lab suggests that enhanced AEA production induced by overt dopamine D2 receptor activation can augment the discriminative stimulus properties of Δ9-THC (Solinas et al., 2010).

Neither exogenous AEA administration (Mallet et al., 1998) nor the FAAH inhibitor URB597 (Gobbi et al., 2005) produce place-conditioning effects and moderate doses of the FAAH inhibitors URB597 and PMSF or the putative eCB clearance inhibitor OMDM-2 do not alter brain stimulation reward thresholds (Vlachou et al., 2006), suggesting that enhancement of AEA tone does not produce rewarding effects in these paradigms. In contrast, the eCB transport and FAAH inhibitor AM404 produces rewarding effects in rats in the CPP paradigm (Bortolato, et al., 2006). AM404 enhances brain levels of both AEA and 2-AG (Beltramo & Piomelli, 2000; Beltramo, et al., 1997; Hajos, Kathuria, Dinh, Piomelli, & Freund, 2004) and this more broad increase in eCB tone may contribute to the reinforcing properties of this compound. Similar to Δ9-THC (Justinova et al., 2003; Tanda et al., 2000) intravenous AEA infusion supports operant self-administration behavior in squirrel monkeys in a CB1 receptor-dependent manner (Justinova et al., 2005). Intravenous AEA administration significantly increases extracellular dopamine (DA) levels in the NAc of rats in a CB1 receptor-dependent manner (Solinas et al., 2006) and it’s likely this effect contributes to the reinforcing effects of exogenous AEA. However, the FAAH inhibitor URB597 does not support operant self-administration in squirrel monkeys despite the fact that its delivery produces a substantial elevation in AEA throughout the monkey brain (Justinova et al., 2008b).

Two studies have evaluated whether repeated AEA administration produces physical dependence, and the sum of these reports is inconclusive. Aceto and colleagues characterized the effects of continuous intraperitoneal AEA infusion (10 – 100 mg/kg/24h) in rats over the course of 4 days. While significant behavioral effects were observed during the AEA infusion (immobility, eyelid ptosis), no somatic symptoms of withdrawal were observed when AEA administration was abruptly terminated (Aceto et al., 1998). Treatment with SR141716A at the end of the AEA administration period also did not precipitate a withdrawal syndrome. In contrast, Costa and colleagues observed that daily i.p. administration of 20 mg/kg AEA to rats resulted in significant withdrawal symptoms under conditions of both spontaneous (e.g. cessation of AEA treatment) and precipitated (e.g. SR141716A challenge) withdrawal conditions (Costa et al., 2000). Significant tolerance to some of the behavioral effects of AEA was observed over the course of treatment (including hypothermia, catalepsy, and splayed hind limbs), which could reflect a downregulation of systems activated by AEA, or an increase in AEA clearance. Withdrawal symptoms were similar to those observed during Δ9-THC withdrawal including ptosis, wet dog shakes, teeth chattering, forepaw fluttering and piloerection and these were substantially more pronounced during precipitated vs. spontaneous withdrawal. In one experiment in this report the spontaneous withdrawal symptoms were remarkably short-lived (<30 min), though SR141716A challenge produced clear withdrawal symptoms that persisted for >80 min. Recent studies have also evaluated subchronic administration of selective FAAH and MAGL inhibitors in tests of dependence (Schlosburg et al., 2009; Schlosburg et al., 2010). Mice given daily injections of URB597 or the more potent FAAH inhibitor PF-3845 for 6 days and subsequently challenged with SR141716A displayed no significant withdrawal symptoms (paw tremors, head twitching), suggesting prolonged FAAH inhibition does not lead to physical dependence. In contrast, significant SR141716A-induced somatic withdrawal was observed following 6 days of treatment with the selective MAGL inhibitor JZL184. This treatment regimen also resulted in significant reductions in CB1 receptor expression and function, suggesting that long-term reductions in 2-AG clearance produces a dependent state similar to that produced by repeated Δ9-THC administration (Schlosburg et al., 2010). It is worth noting that the JZL184 dose employed in this study is more than 3-fold higher than required to elicit significant behavioral effects in animal models of pain, anxiety and drug withdrawal (Kinsey et al., 2011; Kinsey et al., 2009; Schlosburg et al., 2009). It is presently unclear whether repeated treatment with low, behaviorally efficacious JZL184 doses also dysregulates CB1 receptor signaling.

Collectively these findings suggest that administration of exogenous eCBs can induce behavioral effects similar to those produced by CB1 agonists, including reinforcing effects in a variety of paradigms. However, most evidence indicates that pharmacological compounds that enhance eCB tone (e.g. eCB clearance inhibitors) do not produce robust CB1 agonist-like effects and likely do not possess abuse liability. It is notable, however, that much of this work has focused on AEA and FAAH due to the limited availability of tools for manipulating 2-AG. As described above, the recent development of selective MAGL inhibitors and dual FAAH and MAGL inhibitors (Long et al., 2009b; Long et al., 2009c) will allow for important additional analyses regarding the behavioral processes modulated by eCB signaling.

4. Endocannabinoid modulation of non-cannabinoid drug reinforcement

The motivational and addictive properties of abused drugs are mediated through drug-induced changes in neurotransmitter and neuromodulatory signaling in the CNS. Although there are differences in effects caused by various classes of abused drugs, substantial evidence points to an involvement of monoamines (dopamine, serotonin, norepinephrine), excitatory and inhibitory amino acids (primarily glutamate and GABA, respectively), acetylcholine, opioid peptides and stress-related peptides in the etiology of addiction. As reviewed below, there is now substantial evidence implicating the ECS in the motivational effects produced by several classes of abused drugs such as ethanol, nicotine, opiates and psychostimulants. Much of this is based on the influence of CB1 receptor signaling on drug-induced behaviors, though some studies have more directly evaluated manipulations of eCB tone on drug reinforcement.

4.1 Alcohol

A large amount of evidence points to an important influence of CB1 receptor signaling on ethanol-related behaviors. CB1 receptor knockout mice exhibit less ethanol-induced conditioned place preference than do wildtype controls (Houchi et al., 2005; Thanos et al., 2005) suggesting a role for CB1 receptors in the mediation of ethanol reward and/or interoreceptive-related conditioned associations. CB1 receptors also exert substantial regulatory influence on ethanol consumption. The selective CB1 receptor antagonist SR 141716A decreases ethanol consumption in alcohol-preferring C57BL/6 mice (Arnone et al., 1997; Wang et al., 2003), Sardinian alcohol-preferring sP rats (Colombo et al., 1998), Long Evans rats (Freedland et al., 2001), and both non-dependent (Freedland et al., 2001; Gallate et al., 1999) and alcohol-dependent (Lallemand et al., 2001) Wistar rats. Repeated administration of SR 141716A or the more selective CB1 antagonist SR147778 also prevents the acquisition of ethanol drinking by Sardinian alcohol-preferring rats given free choice between 10% ethanol (v/v) and water (Gessa et al., 2005; Serra et al., 2001). CB1 receptor knockout mice display low ethanol preference and intake compared with wild-type mice, and the suppressive effects of SR141716A on ethanol preference and consumption are absent in these mice (Hungund et al., 2003; Naassila et al., 2004; Thanos et al., 2005; Wang et al., 2003). Moreover, ethanol-induced increases in NAc dopamine levels are absent in CB1 knockout mice and are attenuated in wildtype animals by SR141716A pretreatment (Cohen et al., 2002; Hungund et al., 2003). Based on these data most investigators have concluded that ethanol reward depends in part on intact CB1 receptor signaling, though some evidence suggests that reduced ethanol preference and consumption in CB1 knockout mice results from an increased sensitivity to ethanol intoxication and severity of withdrawal (Naassila et al., 2004). It is also possible that CB1 antagonist-induced suppression of ethanol consumption results in part from non-specific reductions in consummatory behavior. For example, although some studies report that SR 141716A reduces ethanol intake at doses that do not attenuate food consumption (Arnone et al., 1997; Colombo et al., 1998), this antagonist can produce comparable dose-dependent reductions in the consumption of both ethanol and sucrose (Arnone et al., 1997; Freedland et al., 2001). Although not fully explored, it is conceivable such an influence results from interactions between the cannabinoid, opioid and dopamine systems in the regulation of palatable food consumption and reward (Cooper, 2004; Cota et al., 2006; Gardner, 2005). It is notable, however, that CB1 receptor knockout mice do not display phenotypic differences in sucrose, quinine or food consumption (Hungund et al., 2003; Naassila et al., 2004). Finally, some evidence suggests that the CB1 receptor influence on ethanol intake may also be dependent on the animals’ stress state (Racz et al., 2003).

Despite these caveats, the findings described above suggest that CB1 receptors exert a facilitory influence on ethanol preference and consumption. Consistently CB1 receptor activation has been shown to increase ethanol consumption. The synthetic CB1 agonists WIN 55,212-2 and CP 55,940 dose-dependently increase spontaneous drinking in Sardinian alcohol preferring sP rats (Colombo et al., 2002) and C57BL/6J mice (Wang et al., 2003) and CP 55,940 dose-dependently increases the breaking point of responding for beer by Wistar rats on a lick-based progressive ratio schedule of reinforcement (Gallate et al., 1999). In each case the CB1 agonist-induced increase in ethanol intake was reversed by SR141716A doses that by themselves did not alter ethanol consumption. CB1 agonist pretreatment non-selectively increased the consumption of beer, non-alcoholic beer (“near-beer”) and sucrose by Wistar rats (Gallate, et al., 1999) suggesting a non-specific stimulatory effect of CB1 receptor stimulation on appetite for palatable beverages. In contrast, both WIN 55,212-2 and CP 55,940 selectively increased ethanol, but not sucrose or food intake in alcohol-preferring sP rats (Colombo, et al., 2002). These differential effects may result from innate differences in CB1 receptor expression and function in this line of alcohol-preferring rats (Vinod et al., 2011). It is worth noting, however, that not all studies have observed CB1 agonist-induced facilitation of ethanol intake. Cippitelli and colleagues did not observe any alteration in ethanol self-administration by Wistar rats following pretreatment with the selective CB1 agonists ACEA or HU-210, and observed a dose-dependent reduction in ethanol self-administration by WIN 55,212-2 (Cippitelli et al., 2007).

Studies in FAAH knockout mice provide more direct evidence for an eCB influence on EtOH consumption. These animals display an increased preference for ethanol and consume significantly more ethanol than do wildtype mice, though there is no genotypic difference in consumption of non-alcoholic tastants such as saccharine or quinine (Basavarajappa et al., 2006; Blednov et al., 2007; Vinod et al., 2008). Similarly, acute administration of the FAAH inhibitor URB597 increases ethanol preference and consumption in wildtype mice. This effect is not observed in FAAH or CB1 receptor knockout mice, indicating that the effects of URB597 are mediated through its actions on FAAH, and that the resultant effect of FAAH inhibition on ethanol consumption relies on CB1 receptor signaling. There is some evidence that increased ethanol consumption by FAAH deficient mice results from a diminished sensitivity to ethanol-induced motor incoordination and intoxication (Blednov et al., 2007), and this is analogous to the association between decreased ethanol consumption and increased sensitivity to ethanol intoxication in CB1 deficient mice (Naassila et al., 2004). In addition to these findings in mice, Hansson and colleagues reported that rats selectively bred for high ethanol preference and consumption are characterized by reduced FAAH expression and activity in the prefrontal cortex (PFC) (Hansson et al., 2007). Moreover, these authors found that localized URB597 administration into the PFC increases ethanol consumption by outbred Wistar rats. In contrast, Cippitelli and colleagues reported that systemically administered URB597 does not alter voluntary homecage ethanol consumption by Marchigian Sardinian alcohol-preferring (msP) rats, and does not facilitate operant ethanol self-administration by Wistar rats under either continuous or progressive ratio schedules of reinforcement (Cippitelli et al., 2008). In a separate study these authors observed that the putative eCB reuptake inhibitor AM404 significantly reduces operant ethanol self-administration by Wistar rats at doses that do not alter saccharin self-administration (Cippitelli et al., 2007). Thus, while FAAH inhibition appears to consistently increase ethanol preference and consumption in mice, less consistent effects are observed in rats.

Several studies have sought to characterize the neurochemical mechanisms and neuronal substrates through which CB1 receptors modulate ethanol intake. Pharmacological blockade or genetic deletion of CB1 receptors diminishes ethanol-induced increases in dopamine levels and dopamine transients in the NAc (Cheer et al., 2007; Hungund et al., 2003; Perra et al., 2005) and ethanol-induced increases in the firing of ventral tegmental area (VTA) dopamine cells that project to the NAc appear to involve a CB1 receptor mechanism (Perra et al., 2005). Localized administration of SR141716A into VTA reduces ethanol self-administration by Wistar rats (Alvarez-Jaimes et al., 2009a; Malinen et al., 2008) and this effect appears to be restricted to the posterior aspect of the VTA (Alvarez-Jaimes et al., 2009a) consistent with evidence of a specific involvement of the posterior VTA in the regulation of ethanol intake (Gatto et al., 1994; Rodd et al., 2004; Rodd-Henricks et al., 2000). Ethanol self-administration is also dose-dependently reduced by SR141716A infusions into the NAc shell, but is unaffected by infusions of this antagonist into the medial PFC (Alvarez-Jaimes et al., 2009a). This regional pattern of CB1 receptor influence on ethanol consumption is consistent with the regional specificity of ethanol-induced increases in interstitial eCB levels determined by microdialysis (Alvarez-Jaimes et al., 2009a; Caille et al., 2007) (see discussion below).

4.2 Nicotine

Nicotinic acetylcholine (nACh) receptors are the primary pharmacological target of nicotine. The overlapping distribution of CB1 and nACh receptors in several brain structures such as the hippocampus and the amygdala (Picciotto et al., 2000) allows the possibility of functional interactions between these two systems. Indeed, nicotine facilitates several physiological and behavioral effects induced by acute Δ9-THC administration, including hypothermia, antinociception, hypolocomotion and anxiolytic-like responses (Valjent et al., 2002). Co-administration of sub-threshold doses of nicotine and Δ9-THC induces significant anxiolytic-like effects in the light-dark transfer test and open field, and a significant conditioned place preference (Valjent et al., 2002). Moreover, co-administration of Δ9-THC and nicotine produces a strong potentiation of c-Fos immunoreactivity in various limbic and cortical structures, including the shell of the NAc, central and basolateral nucleus of amygdala (CeA & BLA, respectively), dorsolateral bed nucleus stria terminalis (BNST), cingulate and piriform cortices and PVN (Valjent et al., 2002).

Several studies implicate a CB1 receptor involvement in the rewarding effects of nicotine. For example, nicotine doses that produce significant conditioned place preference in wildtype mice do not produce place preference in CB1 knockout mice (Castane et al., 2002). Similarly, SR 141716A blocks the acquisition of nicotine-induced conditioned place-preference in rats (Forget et al., 2005; Le Foll et al., 2004) and wildtype mice (Merritt et al., 2008). SR141716A also blocks short-term expression of previously established nicotine place preference (24h after the final nicotine pairing), but not the long-term expression of nicotine-conditioned preference (3 or 12 weeks after nicotine pairing) (Forget et al., 2005; Forget et al., 2006). SR 141716A and the more selective CB1 antagonist AM251 (Krishnamurthy et al., 2004; Lan et al., 1999) each dose-dependently reduce nicotine self-administration by rats (Cohen et al., 2002; Shoaib, 2008) and CB1 receptor knockout mice self-administer less nicotine than wildtypes (Cossu et al., 2001). Further, in vivo microdialysis evaluations reveal that SR141716A blocks the dopamine-releasing effects of nicotine in limbic regions of the rat (Cohen et al., 2002) though subsequent studies failed to observe a significant effect of SR 141716A on nicotine-induced increases in mesolimbic dopamine cell firing (Melis et al., 2008). Finally, SR141716A does not alter the discriminative stimulus properties of nicotine (Cohen et al., 2002; Le Foll et al., 2004).

Two recent studies have characterized the effect of FAAH inhibition on nicotine-induced conditioned place preference and self-administration. In the first study, Merritt and colleagues observed that FAAH knockout mice and URB597-treated wildtype mice display enhanced sensitivity to low nicotine doses relative to vehicle treated wildtype mice in the conditioned place-preference assay (Merritt et al., 2008). The genotypic and URB597-related enhanced effect on nicotine conditioning was lost with higher nicotine doses. The enhanced sensitivity to the rewarding effects of low dose nicotine was blocked by the SR141716A in FAAH knockout mice, confirming an involvement of CB1 receptors in this behavioral effect. The URB597-induced enhancement of nicotine conditioned place preference occurred without a significant effect on locomotor activity, and URB597 did not alter the potency of nicotine in acute antinociception or hypothermia tests, suggesting that the FAAH-related enhancement of nicotine reward was not due to a general increase in sensitivity to nicotine. Collectively these results suggest that FAAH inhibition enhances nicotine reward in mice through a CB1 receptor mechanism, most likely due to elevated levels of AEA.

In contrast, Scherma and colleagues reported that URB597 blocks the development of nicotine-induced conditioned place preference in rats (Scherma et al., 2008a). This effect was observed across a wide range of nicotine doses that included a subthreshold dose for inducing place preference. URB597 pretreatments also attenuated acquisition of intravenous nicotine self-administration, and reduced the effect of acute nicotine on NAc DA levels. These findings suggest that while FAAH inhibition enhances the rewarding effects of low nicotine doses in mice (Merritt et al., 2008) the reinforcing and neurochemical effects of nicotine in rats are reduced by FAAH inhibition. Further in vivo studies using anesthetized rats revealed that FAAH inhibition attenuates nicotine-induced excitation of VTA dopamine cells in a CB1-independent manner through activation of nuclear peroxisome proliferator-activated receptor-α (PPAR-α) receptors by oleoylethanolamide (OEA) and/or palmitoylethanolamide (PEA) (Luchicchi et al., 2010; Melis et al., 2008). Similar to AEA, both OEA and PEA are metabolized by FAAH and thus FAAH inhibition results in enhanced levels of these and other ethanolamines in brain (Ahn et al., 2008). However, unlike AEA neither OEA nor PEA possess significant CB1 receptor affinity. The effects of PPAR-α activation on drug-induced dopamine cell activation are selective for nicotine (Luchicchi et al., 2010) and administration of the metabolically stable OEA analog methOEA and the selective PPAR-α agonist WY14643 selectively decrease nicotine, but not cocaine or food self-administration in both rats and monkeys (Mascia et al., 2011).

In summary, the behavioral, physiological and neurochemical effects of nicotine appear to be facilitated by CB1 receptor activation. While some evidence suggests that enhanced eCB tone resulting from FAAH inhibition increases the rewarding effects of low nicotine doses in mice, similar effects of FAAH inhibition have not been observed in rats or monkeys due to a dominant opposing influence of non-CB1 mechanisms engaged by increased OEA and PEA levels.

4.3 Opioids

Opioids, such as heroin and morphine exert their physiological and behavioral effects through specific interactions with mu (μ)-, delta (δ)- and, kappa (κ)-opioid receptors, though opiate dependence appears to primarily involve μ receptors (Kieffer, 1999; Matthes et al., 1996). CB1 and μ-opioid receptors are similarly expressed in many brain areas involved in reward processes (Delfs et al., 1994; Herkenham et al., 1991b; Mansour et al., 1995; Matsuda et al., 1993; Navarro et al., 1998; Rodriguez et al., 2001), and these receptors share common signaling cascades (Childers et al., 1992; Howlett, 2002; Reisine et al., 1996). Growing evidence suggests a functional interaction between the endogenous cannabinoid and opioid systems (for review see (Corchero et al., 2004; Maldonado et al., 2003; Manzanares et al., 1999b; Vigano et al., 2005)). Activation of each system produces similar behavioral responses including motor suppression, analgesia and reward (Manzanares et al., 1999b; Massi et al., 2001) and the ECS appears to exert a modulatory influence on many opiate-induced processes (Corchero et al., 2004; Hine et al., 1975; Manzanares et al., 1999b; Valverde et al., 2000; Vela et al., 1995). CB1 receptor efficacy is diminished in the striatum of μ-opioid receptor knockout mice (Berrendero et al., 2003) while heroin self-administration or repeated morphine administration leads to increased CB1 receptor function in reward-related brain regions (Fattore et al., 2007; Gonzalez et al., 2002a). The rewarding effects of Δ9-THC are suppressed in opioid receptor knockout mice (Castane et al., 2003; Ghozland et al., 2002) and are attenuated by opioid receptor antagonists (Braida et al., 2001; Justinova et al., 2004). In contrast, CB1 receptor deletion increases δ- and κ-opioid receptor efficacy in the caudate putamen without altering the density of opioid receptor expression (Uriguen et al., 2005). CB1 receptor activation increases the release of endogenous opioid peptides in various brain regions (Corchero et al., 1997; Manzanares et al., 2005; Pugh et al., 1997; Valverde et al., 2001) and opiate-induced increases in brain eCB content have also been reported (Caille et al., 2007; Vigano et al., 2004; Vigano et al., 2003) (see section 5).

Several studies have demonstrated an important CB1 receptor influence in the rewarding effects of opiates. The CB1 receptor antagonist SR141716A blocks the development of morphine-induced conditioned place preference in rats and mice (Chaperon et al., 1998; Navarro et al., 2001; Singh et al., 2004), and mice lacking CB1 receptors display reduced morphine-induced CPP (Martin et al., 2000) though this effect was not observed in another test of CB1 knockouts (Rice et al., 2002). CB1 receptor knockout mice do not acquire heroin self-administration (Cossu et al., 2001; Ledent et al., 1999) and SR141716A dose-dependently reduces heroin self-administration in rats (Caille et al., 2003; De Vries et al., 2003; Navarro et al., 2001; Solinas et al., 2003). The suppressant effect of SR141716A on heroin self-administration is more pronounced under schedules that require high levels of operant behavior (Caille et al., 2003; De Vries et al., 2003; Solinas et al., 2003).

Increased CB1 receptor signaling appears to facilitate opiate reward. Pre-exposure to the CB1 receptor agonist WIN 55,212-2 potentiates morphine-induced conditioned place preference (Manzanedo et al., 2004), and both WIN 55,212-2 and Δ9-THC pretreatment increase operant responding for heroin self-administration under a progressive ratio schedule in rats (Solinas et al., 2005). However, neither the FAAH inhibitor URB597 nor the putative eCB uptake inhibitor AM404 increase progressive ratio responding for heroin (Solinas et al., 2005). These findings suggest that while direct CB1 receptor agonists facilitate opioid reward, opioid-induced increases in eCB formation (Caille et al., 2007) likely do not influence the motivation for opiate self-administration.

CB1 and opioid agonists appear to increase NAc dopamine levels through a common mechanism (Tanda et al., 1997), and thus it is possible that CB1 receptors modulate opiate reward through a mesolimbic dopamine mechanism. However, while morphine-induced increases in NAc dopamine are attenuated in CB1 knockout mice (Mascia et al., 1999), acute administration of the CB1 receptor antagonist SR141716A does not alter morphine- or heroin-induced increases in NAc dopamine (Caille et al., 2003; Caille et al., 2006; Tanda et al., 1997). Moreover, recent evidence demonstrates that FAAH inhibition by URB597 does not alter morphine-induced alterations in VTA dopamine cell firing, though as previously reviewed this manipulation does block nicotine-induced increases in VTA dopamine cell firing in a CB1-dependent manner (Luchicchi et al., 2010). In addition to the mesolimbic dopamine system, opiate reward is mediated in part through reduced GABA signaling in the ventral pallidum (Bardo, 1998; Caille et al., 2004; Xi et al., 2000; Xi et al., 2002) and SR 141716A dose-dependently prevents morphine-induced reductions in ventral pallidal GABA release through a dopamine-independent mechanism (Caille et al., 2006). Similarly, WIN 55,212-2 induced reductions in ventral pallidal GABA release are reversed by the opiate receptor antagonist naloxone, suggesting a cross talk between opioid and cannabinoid signaling mechanisms in this neurochemical response. However, this effect appears to be selective for opiates as CB1 antagonism does not alter cocaine-induced reductions in pallidal GABA (Caille et al., 2006). Finally, heroin self-administration is dose-dependently reduced by SR 141716A infusions into the NAc, but not ventral pallidum (Caille et al., 2006). Collectively these findings suggest that NAc CB1 receptors may modulate opiate reward by regulating opiate-induced inhibition of GABAergic medium spiny neurons projecting to the ventral pallidum.

4.4 Psychostimulants

The behavioral and addictive effects of psychostimulants such as cocaine, 3,4-methylenedioxymethamphetamine (MDMA, “Ecstasy”) and methamphetamine are produced by direct interactions with the brain monoamine neurotransmitters (dopamine, serotonin and norepinephrine). Substantial evidence indicates that dopamine is the primary monoamine involved in psychostimulant reward (Howell et al., 2008; Rothman et al., 2003) and drugs in this class increase extracellular dopamine concentrations through inhibition of dopamine reuptake (e.g. cocaine) or the promotion of impulse-independent dopamine release (e.g. amphetamine and MDMA) (Riddle et al., 2005; Ritz et al., 1987; Sulzer et al., 1990). This direct influence on dopamine release and clearance differentiates psychostimulants from other classes of abused drugs (i.e. nicotine, alcohol, opiates) that primarily increase dopamine cell firing through indirect mechanisms (Brodie et al., 1999; Gysling et al., 1983; Matthews et al., 1984; Pierce et al., 2006).

The distinct mechanisms through which these drugs increase mesolimbic dopamine levels may underlie the apparent differential influence of CB1/eCB signaling on the reinforcing effects of psychostimulants vs. other classes of abused drugs. In contrast to reports that CB1 receptor deletion attenuates the reinforcing effects of alcohol, nicotine and opiates, CB1 receptor deletion does not affect psychostimulant-induced place conditioning (Houchi et al., 2005; Martin et al., 2000) or self-administration (Cossu et al., 2001). In addition, doses of the CB1 antagonist SR141716A that significantly reduce alcohol, nicotine and opiate self-administration do not alter cocaine self-administration by squirrel monkeys (Tanda et al., 2000), rats (Caille et al., 2006; Caille et al., 2007; De Vries et al., 2001; Fattore et al., 1999; Filip et al., 2006) or mice (Lesscher et al., 2005). Cocaine-induced increases in NAc dopamine levels are unaltered by SR141716A pretreatment or CB1 receptor deletion (Caille et al., 2003; Soria et al., 2005) and cocaine-induced enhancement in the sensitivity to brain stimulation reward is unaffected by SR141716A (Vlachou et al., 2003; Xi et al., 2008).

These findings suggest that CB1 receptors do not play a critical role in the rewarding effects of cocaine. However, these reports are countered by several studies that provide evidence for a CB1 receptor modulation of cocaine-induced behavioral and neurochemical events. For example, WIN 55,212-2 dose-dependently reduces both cocaine-induced facilitation of brain stimulation reward in rats (Vlachou et al., 2003) and cocaine self-administration by rats through a CB1 receptor-dependent mechanism (Fattore et al., 1999). Soria and colleagues observed that CB1 receptor deletion impairs the acquisition of cocaine self-administration by mice, and both genetic and pharmacological CB1 receptor blockade reduces the motivation for cocaine under a progressive ratio schedule of reinforcement (Soria et al., 2005). Moreover, in contrast to SR141716A the structurally similar CB1 receptor antagonist AM251 significantly attenuates the motivation for cocaine self-administration under a progressive ratio schedule of reinforcement (Xi et al., 2008), reduces methamphetamine self-administration (Vinklerova et al., 2002) and attenuates cocaine-induced enhancement in the sensitivity to brain stimulation reward (Xi et al., 2008). Additionally, Li and colleagues recently reported that CB1 receptor deletion results in diminished basal dopamine release and attenuated cocaine-induced increases in NAc dopamine (Li et al., 2009).

A recent report from Orio and colleagues suggests that the CB1 receptor influence on cocaine reward is enhanced by long periods of cocaine self-administration that result in progressive increases in cocaine intake (Orio et al., 2009). For example, SR 141716A dose-dependently reduces the breakpoint for cocaine self-administration under a progressive ratio schedule selectively in rats previously given daily 6h access to cocaine with much lesser efficacy in rats given only 1h daily access to cocaine. Similar group-related differences in SR141716A efficacy were observed following direct SR141716A infusions into the NAc. Consistent with these findings levels of both phosphorylated and total CB1 receptor protein were increased in the NAc and amygdala of rats given extended daily access to cocaine. These observations suggest that neuroadaptations induced by extended cocaine exposure may recruit a CB1 receptor involvement in a progressive escalation of drug intake that results from extended periods of cocaine use.

Few studies have directly evaluated an eCB influence in psychostimulant reinforcement through manipulation of eCB clearance mechanisms. Two recent reports demonstrate that FAAH inhibition does not alter fixed ratio cocaine self-administration by either squirrel monkeys (Justinova et al., 2008b) or rats (Adamczyk et al., 2009), though both URB597 and PMSF attenuate cocaine- and cue-induced reinstatement of cocaine-seeking behavior in rats (Adamczyk et al., 2009) (see section 7). Luchicchi and colleagues have observed that URB597 does not affect cocaine-induced alterations in dopamine cell activity in the rat VTA, though this FAAH inhibitor does attenuate cocaine-induced alterations in the firing of medium spiny neurons in the shell of the NAc (Luchicchi et al., 2010). In contrast, the eCB uptake/FAAH inhibitor AM404 attenuates cocaine-induced facilitation of brain stimulation reward through a CB1 receptor-dependent mechanism (Vlachou et al., 2008) in a manner similar to the effects of the synthetic CB1 agonist WIN 55,212-2 (Vlachou et al., 2003). Collectively these findings suggest that increased levels of AEA (and other FAAH substrates) do not substantially alter the reinforcing effects of cocaine or cocaine-induced alterations in mesolimbic dopamine transmission, though eCB reuptake blockade by AM404 appears to attenuate cocaine-induced enhancement of reward function. Because AM404 increases brain levels of both AEA and 2-AG (Beltramo et al., 2000; Beltramo et al., 1997; Hajos et al., 2004) these latter findings may suggest a more potent influence of 2-AG than AEA on cocaine-induced behavioral effects, though this possibility clearly requires further investigation of the effects of MAGL inhibition on cocaine-related behaviors.

Overall, while there is some evidence for a CB1 receptor and/or eCB involvement in the reinforcing and neurochemical effects of psychostimulants the comparable number of papers reporting negative findings suggests a relatively modest eCB influence in the behavioral effects of this class of abused drugs.

5. Drug-induced alterations in brain eCB content

As reviewed above, a substantial literature indicates that the reinforcing effects of various non-cannabinoid drugs are modulated by CB1 receptor manipulations and in some cases by manipulations of eCB signaling (through administration of clearance inhibitors). From these observations it has been hypothesized that brain eCB levels are altered by exposure to drugs of abuse.

The most common approach for quantifying brain eCB content has been lipid extraction and purification from bulk postmortal brain tissue, with subsequent analysis by liquid or gas chromatography coupled with mass spectrometry (for review see (Buczynski et al., 2010)). Several groups have employed this approach to evaluate the effects of acute or repeated drug exposure on eCB levels in reward-relevant brain regions. In general the findings demonstrate that brain AEA and 2-AG content is altered following administration of abused drugs, though dissimilar effects on these two eCBs are often observed, with distinct profiles produced by different classes of drugs.

For example, while repeated Δ9-THC administration (10 mg/kg 1x/day for 8 days) significantly reduces both AEA and 2-AG content in the caudate putamen, only AEA content was decreased in the midbrain and diencephalon (Di Marzo et al., 2000; Gonzalez et al., 2004b). In contrast, this Δ9-THC exposure regimen robustly increased AEA (but not 2-AG) content in the limbic forebrain, while 2-AG (but not AEA) content is significantly increased in the hippocampus, brainstem and cerebellum. Interestingly, one clinical study in humans revealed that antipsychotic-naïve schizophrenics who used cannabis at least 20 time in their life had lower CSF AEA levels compared with subjects who had 5 or fewer lifetime exposures, though no group differences in serum AEA levels were evident (Leweke et al., 2007). Although preliminary and perhaps selective to a specific patient population, these findings provide some indication that cannabis use can alter eCB function in the CNS. Results of one preclinical study also suggest that prenatal CB1 agonist exposure can result in long-term alterations in brain eCB levels. Castelli and colleagues observed that adult rats exposed to moderate WIN 55,212-2 doses (0.5 mg/kg, 1x/day) on gestation days 5 – 20 are characterized by significantly increased striatal AEA content and significantly reduced AEA levels in the limbic forebrain (Castelli et al., 2007). Interestingly, these effects may have resulted from altered AEA clearance mechanisms as FAAH activity was significantly reduced in striatum and increased in the limbic forebrain of animals give prenatal WIN 55,212-2 exposure.

Several lines of evidence indicate that ethanol exposure stimulates central eCB production and leads to alterations in eCB signaling mechanisms. Pioneering work in this area was performed in the laboratory of Basalingappa Hungund demonstrating that chronic ethanol exposure increases both 2-AG and AEA formation in human neuroblastoma cells and primary cultures of rodent cerebellar granule neurons (Basavarajappa et al., 1999; Basavarajappa et al., 2000; Basavarajappa et al., 2003). Consistently it was shown that chronic ethanol vapor inhalation downregulates CB1 receptor expression an desensitizes cannabinoid-activated signal transduction in Swiss-Webster mice (Basavarajappa et al., 1999; Basavarajappa et al., 1998; Vinod et al., 2006) similar to the effects produced by chronic exposure to both synthetic and endogenous CB1 agonists (Oviedo et al., 1993; Romero et al., 1995; Romero et al., 1998). The effects of voluntary ethanol consumption on brain eCB content has been evaluated by several groups, and while the results clearly demonstrate alcohol-induced alterations in brain eCBs it is difficult to define a clear picture of regional effects from the published literature. For example, while some studies report increased AEA content in the limbic forebrain and NAc of alcohol-exposed rats (Gonzalez et al., 2002b; Gonzalez et al., 2004a; Malinen et al., 2009), other studies report significant AEA decreases in components of the limbic forebrain such as the amygdala, hippocampus, PFC (Malinen et al., 2009; Rubio et al., 2007). Moreover, ethanol consumption is reported to both increase and decrease AEA content in the caudate putamen, and to increase and decrease 2-AG content in the PFC (Malinen et al., 2009; Rubio et al., 2007). These differential findings may result from differences in rat strain, gender, amount of daily ethanol consumption duration of ethanol exposure, and post-exposure time that was evaluated. Moreover, quantification of tissue eCB content is highly sensitive to the procedures employed for tissue harvesting, eCB extraction and analysis and methodological differences likely also underlie between-study differences in observations (Buczynski et al., 2010).

Nicotine-induced alterations in brain eCB content have also been reported. For example, repeated nicotine administration (1 mg/kg/day for 7 days) significantly increases both AEA and 2-AG levels in the brainstem and significantly decreases levels of these eCBs in the cortex, while some regions such as limbic forebrain and caudate putamen display significant elevations in AEA but not 2-AG content. (Gonzalez et al., 2002b). Marco and colleagues observed a significant increase in hippocampal CB1 receptor expression and decrease in striatal CB1 receptor expression 1 month after a 10-day period of nicotine administration in Wistar rats, though no changes in CB1 expression were observed in tissue harvested 2h after the final nicotine treatment (Marco et al., 2007). Werling and colleagues compared the effects of subchronic nicotine exposure on CB1 receptor expression in periadolescent and adult Sprague Dawley rats (Werling et al., 2009). Twenty-four hours after the final nicotine exposure CB1 receptor binding was significantly increased in the prelimbic PFC, VTA and hippocampus of adolescent, but not adult rats. Collectively these findings suggest that repeated nicotine exposure alters CB1 receptor function presumably through nicotine-induced alterations in brain eCB levels.

Opiate administration has also been shown to differentially alter both AEA and 2-AG content in rat brain. Vigano and colleagues report that repeated morphine injections significantly reduce 2-AG content in the striatum, cortex, hippocampus, hypothalamus and limbic forebrain, while AEA content in these same samples was unaltered (Vigano et al., 2003). Higher morphine doses also significantly reduce 2-AG content in the ventral striatum and hippocampus, but increase AEA content in the dorsal and ventral striatum, hippocampus and PFC (Vigano et al., 2004). Chronic morphine exposure significantly alters CB1 receptor expression in several brain regions including the NAc, caudate, hippocampus and amygdala (Gonzalez et al., 2002a) and this may result from morphine-induced alterations in brain eCB production.

Several studies have evaluated the effect of psychostimulant administration on brain eCB levels. Patel and colleagues observed that AEA or 2-AG content in mouse limbic forebrain tissue was unaltered by acute cocaine doses up to 10 mg/kg (Patel et al., 2003). However, 20 mg/kg cocaine produced a subtle, but significant increase in limbic forebrain 2-AG content with no concurrent alteration in tissue AEA content. Gonzalez and colleagues reported that repeated cocaine exposure (15 mg/kg, 2x/day for 10 days) does not alter AEA or 2-AG content in hippocampus, cerebral cortex, caudate putamen, midbrain, mesencephalon, brainstem or cerebellum (Gonzalez et al., 2002b). Unlike the stimulatory effects of acute cocaine on 2-AG content in mouse limbic forebrain (Patel et al., 2003), 10 days’ exposure to cocaine induced subtle but significant reductions in 2-AG levels in rat limbic forebrain tissue (AEA content was unaltered (Gonzalez et al., 2002b)). In contrast to the subtle effects of cocaine observed in whole animal studies, in vitro studies in rat brain striatal slices have shown that acute superfusion with 10 μM cocaine robustly increases AEA levels in the superfusion media, most likely through a combined increase in NAPE-PLD activity and decrease in FAAH activity (e.g. both increased AEA formation and decreased AEA clearance (Centonze et al., 2004)). Interestingly, these cocaine-induced effects were reliant on dopamine D2 receptor signaling, and were mimicked by superfusion with the D2-selective agonist quinpirole (Centonze et al., 2004), consistent with other evidence of a facilitory D2 receptor influence on brain AEA levels (Giuffrida et al., 1999; Giuffrida et al., 2004; Patel et al., 2003). Moreover, cocaine-induced reductions in GABA mediated spontaneous IPSCs were significantly attenuated by SR141716A (Centonze et al., 2004), suggesting an AEA involvement in cocaine-induced inhibition of presynaptic GABA transmission in the striatum. The effects of cocaine on striatal 2-AG levels were not evaluated in this study. The more robust effects of cocaine on eCB levels observed in in vitro vs. in vivo preparations may result from the simplified neuronal environment provided by the isolated brain slice preparation, or possibly from differences in sustained brain cocaine concentrations in these paradigms. In addition to cocaine, Patel and colleagues evaluated the effects produced by acute administration of the dopamine/norepinephrine releasing drug methylphenidate (Ritalin) and the selective dopamine transport inhibitor GBR12909 (Patel et al., 2003). Both of these stimulant drugs dose-dependently reduced AEA levels in limbic forebrain tissue, and methylphenidate also significantly reduced 2-AG content in these samples (a non-significant trend toward GBR12909-induced reductions in 2-AG was also evident). The differential eCB effects produced by cocaine, GBR12909 and methylphenidate may result from differences in monoaminergic selectivity or perhaps differences in pharmacodynamic mechanism.

Collectively these studies indicate that Δ9-THC, ethanol, nicotine and opiates can alter rat brain eCB content. Although drug-induced increases and decreases have been reported, these findings are generally consistent with the substantial literature demonstrating a CB1 receptor influence in the behavioral effects of each of these drugs. Although there is relatively less information available on cocaine-induced alterations in brain eCB content, the subtle changes in eCB levels observed following relatively high dose cocaine administration correlates with behavioral evidence suggesting only a subtle CB1 influence on psychostimulant-induced behaviors.

As a complement to eCB measures from bulk brain tissue in vivo microdialysis methods have recently been developed for quantifying eCB levels in the brain interstitial space (Buczynski et al., 2010) and a few studies have utilized this approach to evaluate drug-induced alterations in brain eCB levels. Bequet and colleagues observed that acute systemic WIN 55,212-2 administration significantly increases 2-AG and decreases AEA levels in rat hypothalamic microdialysates (Bequet et al., 2007). This effect of WIN 55,212-2 was blocked by moderate SR141716A doses, and high dose SR141716A administration itself was found to produce an opposite effect (e.g. decrease in 2-AG, increase in AEA). These data suggest that CB1 receptor activation activates 2-AG production and inhibits AEA production in the rat hypothalamus. Caillé and colleagues observed that ethanol, heroin and cocaine self-administration results in dose-dependent and drug-specific alterations in eCB levels in microdialysates collected from the NAc (Caille et al., 2007). Ethanol self-administration increased 2-AG levels without altering dialysate AEA levels, and the rise and fall of dialysate 2-AG content followed the pattern of blood alcohol concentrations typically observed following oral consumption. In contrast, heroin self-administration led to dose-dependent increases in dialysate AEA content with delayed decreases in dialysate 2-AG levels, while cocaine self-administration did not alter dialysate levels of either AEA or 2-AG. Consistent with the drug-specific alterations in NAc eCB levels ethanol and heroin self-administration were dose-dependently reduced by either systemic or intra-accumbens CB1 antagonist administration, while these same antagonist treatments did not alter cocaine self-administration (Caille et al., 2006; Caille et al., 2007). These findings suggest that drug-induced increases in NAc eCB formation modulate the reinforcing properties of certain abused substances. Regionally specific effects of drug intake on interstitial eCB levels have also been observed. For example, ethanol self-administration increases 2-AG content in dialysates collected from the NAc (Caille et al., 2007) and ventral tegmental area (Parsons, unpublished observations), but does not alter either 2-AG or AEA levels in dialysates collected from the medial PFC (Alvarez-Jaimes et al., 2009a). Additional studies have revealed that several factors can affect drug-induced alterations in interstitial eCB levels, including the route, dose and contingency of drug administration as well as the prior history of drug exposure (Alvarez-Jaimes et al., 2009b; Ferrer et al., 2007; Orio et al., 2009). These factors have also been shown to affect measures of drug-induced alterations in bulk tissue eCB content.

Collectively these observations provide initial evidence that brain eCB levels are altered following exposure to various drugs of abuse. However, due to the variability associated with between-study quantitative measures of brain eCB content (Buczynski et al., 2010) and differences in drug dosing regimen, a clear picture of the effects of distinct classes of abused drugs on brain eCB formation has yet to emerge, and more work is needed in this regard.

6. Endocannabinoids, synaptic plasticity and addiction

Synaptic plasticity is considered one of the primary mechanisms underlying learning and memory processes, and is thought to be vital for experience-dependent modifications in neural function that underlies behavioral flexibility. Several conceptualizations of addiction posit that drugs of abuse “hijack” plasticity mechanisms in circuits responsible for reinforcement and reward, and in this way addiction represents a powerful pathological form of learning and memory (Berke et al., 2000; Everitt et al., 2001; Everitt et al., 1999; Hyman, 2005; Hyman et al., 2006; Jentsch et al., 1999; Kalivas et al., 2008; Kauer, 2004; Kauer et al., 2007; Kelley, 2004; Thomas et al., 2008; Torregrossa et al., 2011; White, 1996). Substantial evidence supports a fundamental role of synaptic plasticity in the VTA and regions innervated by VTA dopamine projections in the behavioral response to initial drug exposure. Moreover, plastic events in regions such as the NAc, PFC and other limbic structures likely participate in the formation of conditioned associations between drug reward and a variety of internal and external cues associated with drug intake. These conditioned cues subsequently exert powerful motivational effects on drug intake and relapse to drug consumption following periods of abstinence (Kalivas et al., 2008). Drug-induced alterations in long-term plastic events in limbic circuitry involved in stress, reward and emotional memory have also been described (Childress et al., 1999; Dumont et al., 2005; Fu et al., 2007; Grueter et al., 2006; Weitlauf et al., 2004), and these processes likely contribute to dysregulated stress responses and affective state associated with drug dependence and protracted withdrawal (Koob, 2008).

Endocannabinoids are well known to suppress neurotransmitter release at both excitatory and inhibitory synapses with both short- and long-term effects (Alger, 2002; Chevaleyre et al., 2006; Diana et al., 2004; Kano et al., 2009; Lovinger, 2008; Marsicano et al., 2006; Wilson et al., 2002). Endocannabinoid-mediated short-term synaptic plasticity includes depolarization-induced suppression of excitation (DSE) that results from suppression of presynaptic glutamate release, and depolarization-induced suppression of inhibition (DSI) resulting from inhibition of presynaptic GABA release. These effects typically persist for a minute or less. Endocannabinoids are also involved in long-term forms of synaptic plasticity including long-term potentiation (LTP) that is a long-lasting increase in the strength of synaptic signaling, and long-term depression (LTD) that is a prolonged weakening of synaptic strength. LTP and LTD can persist for hours to weeks and are particularly important for various forms of learning and memory.

In 2003 the Lupica group reported that eCB-mediated LTD of excitatory signaling in the rat NAc is abolished following chronic treatment with either Δ9-THC or WIN 55,212-2, and this effect was correlated with decreased CB1 receptor influence at both excitatory and inhibitory synapses (Hoffman et al., 2003). Soon thereafter the Manzoni lab reported that a single Δ9-THC exposure abolishes eCB-mediated LTD of both excitatory and inhibitory transmission in the mouse NAc and hippocampus (Mato et al., 2004). This effect was transient (< 24h) and associated with tolerance to CB1 receptor agonists, consistent with the findings by Hoffman and colleagues. Subsequently the Manzoni group reported that LTD of NAc excitatory transmission is not altered following chronic Δ9-THC treatment despite the presence of diminished CB1 receptor G-protein coupling and tolerance to CB1 receptor agonist effects on excitatory transmission (Mato et al., 2005). This surprising preservation of LTD following repeated Δ9-THC exposure was found to result from recruitment of mGluR2/3 mechanisms that were posited to compensate for reduced eCB influence resultant from desensitized CB1 receptor function. The reasons for the differential reported effects of repeated Δ9-THC exposure on NAc LTD are not clear, though several experimental factors distinguish these studies including nearly 7-fold differences in Δ9-THC dose and the rodent species under study (Sprague Dawley rats vs. C57Bl/6 mice).

Acute and repeated alcohol exposure also disrupts eCB-mediated synaptic plasticity. Short periods of low-frequency stimulation produce a CB1-dependent long-lasting disinhibition (DLL) of striatal output neurons as a result of reduced synaptic strength at inhibitory synapses (Adermark et al., 2009). Acute exposure of striatal slices to moderate ethanol doses substantially reduces eCB-mediated DLL in the dorsolateral striatum (Clarke et al., 2010). DLL is also significantly reduced in the dorsolateral striatum of rats following long-term voluntary alcohol consumption (Adermark et al., 2011). Endocannabinoid-mediated LTD at inhibitory striatal synapses is also reduced by acute ethanol, though no significant ethanol effects are evident on LTD of excitatory synapses (Clarke et al., 2010). Because the dorsal striatum is involved in reward-guided learning and habitual behavior (Volkow et al., 2007; Yin et al., 2008) it is possible that ethanol-induced interference in eCB-LTD contributes to maladaptive habitual behavior associated with addiction.

Similar to Δ9-THC, a single injection of a moderately high cocaine dose (20 mg/kg) abolishes eCB-mediated LTD of excitatory transmission in the mouse NAc (Fourgeaud et al., 2004), thereby resulting in heightened excitatory signaling in this region. This was not associated with reduced CB1 receptor function but was correlated with reduced mGluR5 expression and influence suggesting that cocaine-induced abolition of eCB-LTD results from diminished mGluR5-mediated eCB production. In concert with this, repeated cocaine exposure reduces the inhibition of rat VTA dopamine cells through facilitated induction of eCB-mediated LTD of GABAergic cells (Pan et al., 2008). This effect is reliant on dopamine D2, mGluR5 and CB1 receptors (Liu et al., 2005), and is attenuated following inhibition of PLC or DAGL function (Pan et al., 2008), implicating a role for 2-AG in this process. Collectively these findings suggest that cocaine-induced disruption of eCB-LTD results in imbalanced mesolimbic dopamine function characterized by diminished inhibitory control over VTA dopamine cells and heightened excitatory signaling in the NAc. Cocaine-induced disruption of eCB-mediated LTD of excitatory transmission has also been observed in the BNST (Grueter et al., 2006). This effect was induced by either repeated administration of non-contingently administered cocaine injections or response-contingent cocaine self-administration. However, in contrast to the effects reported in NAc (Fourgeaud et al., 2004; Swanson et al., 2001), cocaine-induced disruption of eCB-LTD in the BNST was evident only after repeated cocaine exposure and was not present following a single cocaine injection. The BNST is a stress-responsive structure and excitatory transmission in this region is critical in mediating stress – reward interactions and anxiety-like behavior (Delfs et al., 2000; McElligott et al., 2009; Stewart, 2000). The BNST extends a substantial projection to the VTA (Dong et al., 2006) and accordingly drug-related disruption of BNST plasticity is likely to influence motivational responses to stress. As reviewed below, dysregulation of stress reactivity is a major adverse consequence of long-term drug exposure that plays a primary role in relapse to drug use.

7. Endocannabinoids, drug-seeking and relapse

Drug addiction is characterized by compulsive drug-seeking and use that is associated with a high incidence of relapse to drug use following periods of abstinence (Leshner, 1997; McLellan et al., 2000; O’Brien et al., 1996; O’Brien et al., 1998). Substantial difficulties with relapse have been described for a range of abused substances including opiates, ethanol, nicotine and cocaine (Chutuape et al., 2001; Cooney et al., 1997; Covey et al., 2000; Gossop et al., 1989; Kaplan et al., 1985; Laberg, 1986; Mendelson et al., 1996; O’Brien, 1997; Silagy et al., 2004). Several factors are believed to be causal in drug relapse including craving induced by exposure to environments or situations previously associated with drug use (e.g. conditioning factors) (Carter et al., 1999; O’Brien et al., 1992), acute exposure to the drug itself or a pharmacologically related agent during abstinence (e.g. drug priming) (de Wit, 2000; Jaffe et al., 1989) and stressful events (Kosten et al., 1986; Kreek et al., 1998; Sinha et al., 1999).

Recently developed animal models of drug seeking have contributed substantially to our understanding of the neurobiological mechanisms underlying drug relapse (See, 2005; Shaham et al., 2000; Shalev et al., 2002; Stewart, 2000; Weiss, 2005). These models have identified important roles for glutamate, dopamine, GABA, serotonin, opioid peptides and stress-related peptides (e.g. CRF) in relapse-related behavior (Crombag et al., 2008; Weiss, 2005; Yahyavi-Firouz-Abadi et al., 2009). Recent studies have also demonstrated an important influence of cannabinoid signaling in the re-initiation of drug-seeking and taking behaviors. For example, direct acting CB1 agonists such as Δ9-THC, HU-210, WIN 55,212-2, and CP 55,940 reinstate extinguished drug-seeking behavior for cannabinoids (Justinova et al., 2008a; Spano et al., 2004), opioids (De Vries et al., 2003; Fattore et al., 2003; Fattore et al., 2005b), ethanol (Lopez-Moreno et al., 2004; McGregor et al., 2005), nicotine (Biala et al., 2008) and cocaine (De Vries et al., 2001).

Drug-seeking behavior associated with a variety of abused substance is attenuated following CB1 receptor antagonism. For example, SR141716A significantly reduces operant responding by squirrel monkeys reinforced by a conditioned cue that was previously paired with Δ9-THC self-administration (Justinova et al., 2008a). In a similar manner, SR1417161A dose-dependently attenuates cue-induced and heroin-primed reinstatement of heroin-seeking behavior in rats (De Vries et al., 2003; Fattore et al., 2003; Fattore et al., 2005a). Bilateral SR141716A infusions into the PFC and NAc shell dose-dependently attenuate cue-induced reinstatement of heroin-seeking behavior (Alvarez-Jaimes et al., 2008), consistent with evidence that expression of goal-directed behaviors in response to drug-paired stimuli relies on these structures (Kalivas et al., 2005). However, despite substantial evidence of an important involvement of the BLA in cue-induced drug-seeking behavior (See, 2005; Weiss, 2005) and the dense expression of CB1 receptors in this region (Herkenham et al., 1991b), cue-induced heroin-seeking behavior was unaltered by intra-BLA SR141716A administration (Alvarez-Jaimes et al., 2008).

CB1 receptor antagonism also attenuates ethanol-seeking behavior. SR 141716A reduces cue-induced reinstatement of ethanol-seeking behavior (Cippitelli et al., 2005; Economidou et al., 2006) and reduces excessive ethanol intake present following periods of abstinence (an animal model of relapse-like alcohol consumption (Spanagel et al., 2000)). Both SR141716A and the similar CB1 antagonist SR147778 reduce this “alcohol deprivation effect” when the antagonists are systemically administered prior to the post-deprivation access to alcohol (Gessa et al., 2005; Serra et al., 2002). In contrast, subchronic WIN 55,212-2 administration during alcohol abstinence increases the magnitude and duration of post-deprivation elevations in alcohol consumption (Alen et al., 2008; Lopez-Moreno et al., 2004). These observations suggest that CB1 receptors exert a facilitory influence on post-abstinence increases in alcohol consumption.

Reinstatement of nicotine-seeking behavior evoked by non-contingent presentation of a drug-paired stimulus is dose-dependently reduced by SR 141716A (Cohen et al., 2005; De Vries et al., 2005) as is nicotine-seeking behavior precipitated by exposure to environments previously paired with nicotine self-administration (Diergaarde et al., 2008). These findings suggest that CB1 receptor blockade reduces the motivational saliency of both discrete and contextual cues associated with drug intake. However, as demonstrated by De Vries and colleagues (De Vries et al., 2005) SR 141716A also suppresses cue-induced sucrose-seeking behavior, suggesting a CB1 receptor influence on both drug-associated and generalized conditioned reinforcement. Similar to the work on heroin-seeking behavior described above, bilateral SR141716A infusions into the NACsh dose-dependently reduce cue-induced nicotine-seeking behavior (Kodas et al., 2007). Significant reductions were also observed following CB1 antagonist infusion into the PFC and BLA.

Despite inconsistent and subtle CB1 receptor effects on psychostimulant self-administration, growing evidence demonstrates a substantial CB1 receptor influence on psychostimulant-seeking behavior. In fact, the first evidence of CB1 antagonist effects on drug-seeking behavior was gathered by De Vries and colleagues who observed that SR 141716A dose-dependently attenuates both cocaine-primed and cue-induced reinstatement of cocaine-seeking behavior in rats (De Vries et al., 2001). These findings were subsequently replicated by others using higher SR 141716A doses (Filip et al., 2006) and the more selective CB1 receptor antagonist AM251 (Xi et al., 2006). Interestingly, Xi and colleagues observed that AM251 selectively attenuates reinstatement of cocaine- but not sucrose-seeking behavior which stands in contrast to earlier work demonstrating that SR141716A significantly reduces both drug- and sucrose-seeking (De Vries et al., 2005). Cue-induced reinstatement of cocaine self-administration by C57/BL6 mice is also significantly reduced by SR 141716A, as is “priming”-induced reinstatement of oral corn oil self-administration (Ward et al., 2009). This latter finding again suggests that CB1 receptors modulate the motivational saliency of cues conditioned to both drug and non-drug rewards. A CB1 receptor influence on psychostimulant-seeking may not be restricted to cocaine, as SR 141716A also significantly reduces both priming- and cue-induced reinstatement of methamphetamine-seeking behavior (Anggadiredja et al., 2004) though comparable effects were not observed using AM251 (Boctor et al., 2007). Finally, a recent study by Gerdeman and colleagues demonstrates that expression of behavioral sensitization to cocaine is reduced by SR 141716A in a context-dependent manner, such that attenuation is observed only when the animals are tested in a cocaine-paired environment and not when tested with identical drug treatments in a non-paired environment (Gerdeman et al., 2008). This observation further supports a CB1 receptor involvement in modulation of the perception and/or motivational saliency of drug-paired cues.

Recent studies have begun evaluating the effect of eCB clearance inhibitors on drug-seeking behavior as a more direct means of investigating the effect of eCB tone on relapse-related behaviors. In squirrel monkeys administration of the FAAH inhibitor URB597 does not promote reinstatement of drug-seeking behavior previously maintained by Δ9-THC, cocaine or AEA (Justinova et al., 2008b). Similarly, neither URB597 nor the eCB reuptake inhibitor AM404 alter cue- or stress-induced reinstatement of ethanol-seeking behavior in rats (Cippitelli et al., 2007; Cippitelli et al., 2008). However, two studies have determined that URB597 significantly reduces both drug priming- and cue-induced nicotine-seeking behavior (Forget et al., 2009; Scherma et al., 2008b), and both URB597 and PMSF significantly attenuate both priming- and cue-induced cocaine-seeking behavior (Adamczyk et al., 2009). These observations are somewhat surprising in light of evidence that CB1 receptor stimulation enhances drug-seeking and FAAH inhibition is expected to produce substantial increases in brain AEA levels. However, FAAH hydrolyzes a large number of fatty acid moieties in addition to AEA (Ahn et al., 2008) and it is conceivable that the effects of FAAH inhibition on drug-seeking are mediated through a non-AEA lipid signaling molecule. Moreover, it should be borne in mind that AEA potently activates non-CB1 signaling mechanisms such as TRPV1 receptors that are known to mediate behavioral effects that are sometimes opposite to those produced by CB1 activation (Micale et al., 2009; Santos et al., 2008; Terzian et al., 2009).

8. Endocannabinoids, stress responsivity and addiction

A substantial clinical and preclinical literature indicates that various forms of stress are involved in the etiology and maintenance of drug dependence. Stress often precedes the development of substance use disorders (Jose et al., 2000; Richman et al., 1996; Rospenda et al., 2000), and preclinical studies have shown that stress and glucocorticoids increase the acquisition of drug self-administration (Goeders, 2002; Goeders et al., 1996; Koob et al., 2007; Lynch et al., 1999; Vengeliene et al., 2003). There is some evidence to suggest that both stress- and drug-induced activation of the hypothalamic-pituitary-adrenal (HPA) axis sensitizes the function of reward pathways through glucocorticoid mechanisms (Piazza et al., 1998; Rouge-Pont et al., 1998; Rouge-Pont et al., 1995; Tidey et al., 1997). High glucocorticoid levels and increased levels of stress-related peptides have been proposed to contribute to an allostatic state that results from long-term drug exposure. In the context of drug use, initial exposure produces a positive hedonic response followed by a minor negative response (e.g. withdrawal, or “hangover”) that is hypothesized to involve recruitment of brain stress systems. Continued drug use prior to the resolution of the negative response prevents the reestablishment of homeostasis, and is hypothesized to lead to the transition to an allostatic reward state that is driven in part by an overactive HPA axis and recruitment of stress peptide systems ultimately resulting in an escalation of drug consumption (Koob et al., 2007; Koob et al., 1997; Le Moal, 2009).

In this context, it is now well established that eCB signaling serves a homeostatic role in the constraint of HPA axis activation (Cota, 2008; Gorzalka et al., 2008; Hill et al., 2010). There is substantial CB1 receptor expression in brain regions involved in stress responsivity, including the hippocampus, PFC, amygdala, hypothalamus and midbrain monoaminergic nuclei such as the locus coeruleus and dorsal raphe. Tasker and colleagues have demonstrated that CB1 receptor activation inhibits glutamatergic activation of CRF neurosecretory neurons, and that glucocorticoid inhibition of CRF release is mediated by eCB/CB1 signaling (Di et al., 2003). CB1 receptor knockout mice exhibit enhanced HPA axis activity with heightened basal plasma ACTH and corticosterone levels, enhanced pituitary ACTH secretion in response to CRF and increased CRF mRNA levels in the PVN (Barna et al., 2004; Cota et al., 2007; Haller et al., 2004). CB1 receptor deficient mice also exhibit exaggerated stress-induced secretion of both ACTH and corticosterone as compared with wildtypes (Barna et al., 2004; Haller et al., 2004; Uriguen et al., 2004; Valverde, 2006). Consistently, CB1 antagonist administration increases basal and stress-induced corticosterone secretion and enhances stress-induced neuronal activation in the PVN (Manzanares et al., 1999a; Patel et al., 2004; Wade et al., 2006) while CB1 agonists and eCB clearance inhibitors significantly attenuate stress-induced corticosterone secretion (Patel et al., 2004). Collectively these findings indicate that eCB signaling is capable of dampening HPA axis activity.

Several studies indicate that eCB production is increased in response to stress. In humans the Trier social stress test rapidly increases circulating levels of 2-AG without affecting circulating levels of AEA (Hill et al., 2009a). Both in vitro and in vivo studies in rodents have demonstrated that glucocorticoids such as corticosterone increase AEA and 2-AG production in the PVN, amygdala, hippocampus and hypothalamus (Di et al., 2003; Di et al., 2009; Di et al., 2005a; Di et al., 2005b; Hill et al., 2010; Malcher-Lopes et al., 2006). Moreover, brief footshock stress increases eCB levels in the periaqueductal grey (Hohmann et al., 2005) and acute restraint stress produces 2-AG elevations in PFC, hippocampus and hypothalamus. Sensitization of stress-induced increases in forebrain and amygdalar 2-AG levels following repeated episodes of restraint has been linked to the behavioral and physiological habituation to repeated homotypic stress (Patel et al., 2005b; Patel et al., 2004). These findings demonstrate that stress can increase eCB levels in the CNS, and that this effect provides a homeostatic feedback mechanism dampening stress responses.

An alternate hypothesis posits that eCBs exert a tonic steady-state regulation of HPA activity, and that stress-induced reductions in AEA content in the amygdala, PFC and hippocampus (Hill et al., 2008b; Hill et al., 2009b; Patel et al., 2005b; Rademacher et al., 2008) allow for enhanced HPA axis activation and the subsequent expression of stress responses (Patel et al., 2004). Restoration of AEA tone or the prevention of AEA decrements (through eCB clearance inhibition) would therefore maintain constraint of the HPA axis and attenuate stress responses. Regardless of the specific mechanism that prevails, it is evident that eCB signaling serves to limit HPA axis activation and may contribute to the termination of stress responses.

In light of the evidence for an eCB influence in the regulation of stress responses, it is conceivable that an alteration in eCB function following long-term drug exposure may contribute to the allostatic shift toward persistent dysregulation of HPA axis function and increased sensitivity to stress that contributes to the maintenance of high levels of drug intake and vulnerability toward relapse associated with drug dependence (Koob et al., 2007; Koob et al., 1997; Le Moal, 2009). However, direct tests of this hypothesis have not been published and this remains an open question for investigation.

9. Endocannabinoids and affective state related to drug dependence

Withdrawal is a critical component of most definitions of drug dependence and is characterized by impaired physiological function and increased negative affect. Physical symptoms of withdrawal can include fatigue, tremor, perspiration, convulsions and nausea. These symptoms may vary as a function of drug class and typically dissipate after the first few days of abstinence. While unpleasant, physical withdrawal symptoms are not thought to play a significant role in relapse to drug taking (with the exception of opiates). Withdrawal from most classes of abused drugs is also associated with increased negative affective symptoms such as anxiety and depression (Alling et al., 1982; Carmody, 1992; Coffey et al., 2000; Janiri et al., 2005; Nunes et al., 2004a; Nunes et al., 2004b; Voltaire-Carlsson et al., 1996). These negative affective states can persist for many months during protracted abstinence and their severity is hypothesized to be closely associated with susceptibility to relapse (Annis et al., 1998; Miller et al., 2000). Accordingly there has been substantial interest in elucidating the neural mechanisms contributing to withdrawal-related affective dysregulation with the aim of identifying viable treatment targets for the prolongation of abstinence.

As discussed below, substantial evidence implicates the ECS in the regulation of affective state, and dysfunctional eCB signaling has been associated with increased anxiety and depression. In light of the evidence for drug-induced changes in brain eCB formation and subsequent alterations of eCB signaling following long-term drug exposure (see sections 4 – 6) it is conceivable that dysregulated eCB function contributes to affective disturbances associated with drug dependence and protracted withdrawal.

9.1 Anxiety

CB1 receptors are expressed in high density in key regions implicated in the regulation of anxiety such as the amygdala, hippocampus, anterior cingulate cortex and PFC (Glass et al., 1997; Herkenham et al., 1991a; Herkenham et al., 1991b; Herkenham et al., 1990; Katona et al., 2001). It has long been recognized that cannabis produces a bimodal effect on anxiety in humans, with low doses having anxiolytic effects and high doses having anxiogenic effects (Ashton, 2001; Dannon et al., 2004; Hall et al., 1998; Tournier et al., 2003). Similar biphasic effects of cannabinoid agonists on anxiety-like behaviors have been observed in rats and mice (Viveros et al., 2005). The mechanisms contributing to these dose-related differences in cannabinoid-induced anxiety-like behavior have not been fully characterized. It is possible that different behavioral effects are produced by activation of CB1 receptors versus novel non-CB1 cannabinoid sensitive receptors (Hajos et al., 2002; Haller et al., 2002). For example, Rubino and colleagues observed that activation of CB1 receptors by low doses of exogenous AEA produces an anxiolytic-like effect, while higher AEA doses produce an anxiogenic-like effect through the activation of transient receptor potential vanilloid (TRPV1) receptors (Rubino et al., 2008). Opposing influences of these receptors in the control of anxiety-like behavior has been well documented in rodent studies (Aguiar et al., 2009; Micale et al., 2009; Terzian et al., 2009). It is also possible that the CB1 receptor-mediated effects on anxiety-like behavior vary in a region-dependent manner, with anxiolytic-like effects produced in the PFC, ventral hippocampus and periaqueductal grey (Lisboa et al., 2008; Moreira et al., 2007; Rubino et al., 2008) and anxiogenic-like effects produced in dorsal hippocampus and amygdala (Onaivi et al., 1995; Roohbakhsh et al., 2007; Rubino et al., 2008).

In addition, there is considerable evidence that manipulation of the ECS can substantially modulate anxiety-related behavioral responses. Elevation of interstitial eCB levels through inhibition of eCB clearance mechanisms produces anxiolytic-like effects in the elevated zero-maze, elevated plus-maze, light-dark transfer test, defensive withdrawal tests and the isolation-induced ultrasonic emission test in pups (Bortolato et al., 2006; Braida et al., 2007; Haller et al., 2009; Kathuria et al., 2003; Micale et al., 2009; Moise et al., 2008; Moreira et al., 2008; Naderi et al., 2008; Naidu et al., 2007; Patel et al., 2006; Rutkowska et al., 2006; Scherma et al., 2008a). Unlike traditional anxiolytics such as benzodiazepines, enhancement of eCB tone appears to produce effects on anxiety-like behavior only under aversive or stressful conditions (Haller et al., 2009; Naidu et al., 2007). Moreover, unlike exogenous CB1 receptor agonists, eCB clearance inhibitors do not produce dramatic bi-phasic effects on anxiety-like behavior but appear to only produce anxiolytic-like behavioral effects. Thus, increased eCB tone produces anxiolytic-like effects, particularly under stressful or aversive conditions.

In contrast, reduced CB1 receptor signaling produces anxiogenic-like behavioral effects. CB1 knockout mice display increased aggression in the resident-intruder test, and anxiogenic-like responses in the light/dark box, elevated plus-maze and social interaction tests and hypersensitivity to the effects of restraint stress on HPA activation (Haller et al., 2002; Martin et al., 2002; Uriguen et al., 2004). There is some indication that the anxiogenic profile observed in CB1 knockout mice is most apparent under conditions of heightened stress such as bright light on the EPM and an unfamiliar environment in the social interaction test (Haller et al., 2002), and this may explain why some investigators have not observed anxiogenic-like responses on the EPM (Marsicano et al., 2002). In wildtype mice SR141716A produces anxiogenic-like effects in the defensive withdrawal and elevated plus maze tests, and the isolation-induced ultrasonic emission test in pups (Arevalo et al., 2001; McGregor et al., 1996; Navarro et al., 1997). SR141716A also increases Fos expression in anxiety-related structures (Patel et al., 2005a). Based on these observations, deficient eCB tone appears to increase anxiety-like behavior with effects more consistently observed under stressful or aversive conditions.

In addition to the expression of generalized symptoms of anxiety, the ECS may play a particular role in phobias and post-traumatic stress disorder (PTSD) that are characterized by a reduced capacity to suppress emotionally aversive memories. Disruption of CB1 receptor signaling impairs the extinction of aversive memories (Cannich et al., 2004; Chhatwal et al., 2005; Chhatwal et al., 2006; Chhatwal et al., 2009; de Oliveira Alvares et al., 2008; Harloe et al., 2008; Kamprath et al., 2006; Marsicano et al., 2002; Niyuhire et al., 2007; Pamplona et al., 2008; Pamplona et al., 2006; Reich et al., 2008; Suzuki et al., 2004) without altering the capacity to extinguish appetitive memories (Harloe et al., 2008; Holter et al., 2005; Niyuhire et al., 2007). Conversely, CB1 receptor activation either through direct pharmacologic stimulation or facilitation of eCB signaling accelerates the extinction of aversive memory (Bitencourt et al., 2008; Chhatwal et al., 2005; de Oliveira Alvares et al., 2008; Pamplona et al., 2008; Pamplona et al., 2006). This appears to be mediated primarily in the BLA where eCB levels are increased during extinction trials that serve to suppress aversive memories (Marsicano et al., 2002).

Collectively the evidence published to date indicates the eCB system participates in a negative feedback system that limits the expression of anxiety under stressful circumstances and that contributes to the suppression of aversive or fearful memories. Accordingly, dysregulation of eCB function may contribute to abnormal anxiety-like responses to stress and reduced capacity to recover from emotionally traumatic and stressful experiences. To the extent that aversive memory is involved in relapse to drug taking behavior (Kaplan et al., 2011; Quirk et al., 2003; Stewart et al., 2001) these findings suggest the ECS as a potential pharmacological target for enhancing the efficacy of extinction therapy for addiction.

9.2 Depression

Deficiencies in CB1 receptor signaling appear to produce depressive-like symptoms in rodents and humans. CB1 knockout mice more rapidly become anhedonic during chronic mild stress exposure than do wildtype mice (Martin et al., 2002) and exhibit increased passive stress coping behaviors in paradigms such as the forced swim test and tail suspension test (Aso et al., 2008; Steiner et al., 2008). As previously noted, CB1 knockout mice display a heightened HPA axis response to stress (Uriguen et al., 2004), in line with one of the most consistent findings in major depressive disorder. Although acute CB1 receptor antagonist administration produces antidepressive-like responses in rodents (Griebel et al., 2005; Shearman et al., 2003; Takahashi et al., 2008; Tzavara et al., 2003), recent evidence suggests that a depression-like profile is evident following chronic antagonist administration (Beyer et al., 2010). Long-lasting CB1 receptor blockade induces a depressive behavioral phenotype in rats that is associated with well-known biochemical markers of depression along with diminished neuroplasticity and synaptic efficiency in the PFC (Rubino et al., 2009; Rubino et al., 2008). Perhaps the most important data in this regard are clinical reports from the use of Rimonabant (SR141716A) for the treatment of obesity in which significant signs of anxiety and depression were presented even in individuals with no prior history of these affective disorders (Christensen et al., 2007; Nissen et al., 2008). Similar effects were observed in trials with the structurally distinct CB1 receptor antagonist Taranabant (Aronne et al., 2010). Recent evidence also suggests that a single nucleotide polymorphism (SNP) in the CB1 receptor gene (rs7766029) is highly associated with increased symptoms of depression following recent negative life events (Juhasz et al., 2009), and a different CB1 receptor gene SNP (rs1049353) is associated with a diminished neuronal response to social reward stimuli and increased resistance to antidepressant treatment (Domschke et al., 2008). Reductions in circulating eCBs have also been documented in women with major depression (Hill et al., 2008a). All of these data support the hypothesis that impairment of CB1 receptor signaling leads to depression-like sympomatology.

In general, enhancement of CB1 receptor signaling produces antidepressant-like behavioral effects. Systemic administration of direct CB1 receptor agonists reverses behavioral despair in the forced swim test under both acute (Adamczyk et al., 2008; Bambico et al., 2007; Hill et al., 2005; Rutkowska et al., 2004) and chronic conditions (Jiang et al., 2005; Morrish et al., 2009). Enhancement of eCB tone through inhibition of FAAH activity also produces antidepressant-like effects in the forced swim and tail suspension tests, as well as in the chronic stress-induced anhedonia model (Adamczyk et al., 2008; Bortolato et al., 2006; Gobbi et al., 2005; Hill et al., 2005; McLaughlin et al., 2007; Naidu et al., 2007; Rademacher et al., 2007). FAAH knockout mice also display an antidepressant profile in the forced swim and tail suspension tests (Bambico et al., 2007; Naidu et al., 2007). Finally, administration of direct CB1 receptor agonists or enhancement of eCB tone through eCB clearance inhibition increases hippocampal neurogenesis (Goncalves et al., 2008; Hill et al., 2006; Jiang et al., 2005; Marchalant et al., 2009), which may play a role in the antidepressant profile produced by these manipulations (Drew et al., 2007; Perera et al., 2008).

10. Genetic variants in endocannabinoid signaling and addiction

10.1 CNR1 polymorphisms

Several genetic polymorphisms of eCB-related genes have been linked to drug and alcohol abuse. Most studies have focused the CNR1 gene that encodes the cannabinoid receptor CB1. Among the first polymorphisms reported in this system was a triplet repeat of varying number (AAT)n in the 3′-flanking region of the CNR1 (CB1) genetic locus (Dawson et al., 1995). In a 1997 report Comings and colleagues reported a significant linkage to psychostimulant and cannabis dependence in non-Hispanic Caucasians who are homozygous for more than 5 repeats of (AAT)n (Comings et al., 1997), and a linkage between this polymorphism was subsequently confirmed for cocaine dependence in an African-Caribbean population (Ballon et al., 2006). In addition, a significant relationship between this triplet repeat marker and decreased P300 event-related potential amplitudes in subjects with alcohol and substance abuse disorders has been reported (Johnson et al., 1997). A 6-repeat allele of the triplet repeat polymorphism (AATn/A6) was recently found to be significantly associated with impulsivity in a Native American population with a high lifetime prevalence of substance dependence (Ehlers et al., 2007) and a similar polymorphism is reportedly associated with ADHD in Spanish alcoholics (Ponce et al., 2003). However, other studies have failed to find a significant linkage between the (AAT)n polymorphism and drug dependence (Covault et al., 2001; Heller et al., 2001; Li et al., 2000; Zhang et al., 2004).

Another CNR1 gene polymorphism is a silent mutation that results in the substitution of G to A at nucleotide position 1359 in codon 435 (Thr) (Gadzicki et al., 1999; Hoehe et al., 1991). A modest association between the A/A genotype and alcohol withdrawal delirium has been reported in German patients who are A/A homozygous (Schmidt et al., 2002), though further studies in a similar population did not replicate this finding (Preuss et al., 2003).

Several single nucleotide polymorphisms (SNPs) in CNR1 appear to be associated with problem drug and alcohol use. A ‘TAG’ haplotype resulting from a combination of three SNPs in intron 2 (rs806379, rs1535255, rs2023239) has been associated with substance abuse in European-Americans, African-Americans and Japanese populations (Zhang et al., 2004). Consistent with these observations, two recent reports indicate that SNPs proximal to the TAG haplotypes (rs806380, rs806368) are associated with susceptibility to cannabis dependence while other haplotypes (rs6454674, rs806380, rs806377, rs1049353) are associated with fewer cannabis dependence problems in adolescent and adult European- and African-Americans (Agrawal et al., 2009; Hopfer et al., 2006). However, these findings were not replicated in a more recent study (Hartman et al., 2009). Another CNR1 SNP (rs2023239) is proposed to be a significant predictor of a cannabis withdrawal syndrome and overall levels of cannabis craving during abstinence (Haughey et al., 2008). Carriers the rs2023239 G allele also display significantly greater activity in reward-related brain areas during exposure to marijuana-associated cues as compared with carriers of the A/A genotype for this SNP (Filbey et al., 2010).

CNR1-associated SNPs, particularly rs6454674 and rs806368, are also associated with cocaine, opioid and alcohol dependence in European Americans and African-Americans (Zuo et al., 2007; Zuo et al., 2009). However, other studies have not found linkages between CNR1 SNP haplotypes and cocaine, opioid, cannabis or polysubstance dependence (Herman et al., 2006), though one SNP (rs1535255 T/G) has been linked to alcohol dependence (Herman et al., 2006). The C allele of rs2023239 appears to be associated with greater CB1 binding in the PFC, greater alcohol cue–elicited brain activation in the midbrain and PFC and greater subjective reward when consuming alcohol (Filbey et al., 2008; Hutchison et al., 2008). Interestingly, this allele is also associated with more positive outcomes in individuals receiving Olanzapine, a medication that targets mesocorticolimbic circuitry (Hutchison et al., 2008). Four SNPs in or near the CNR1 gene (rs1535225, rs2023239, rs1049353, rs806368) are associated with increased impulsivity in a Native American population characterized by a high prevalence of alcoholism and substance abuse (Ehlers et al., 2007). Finally, three SNPs (rs2023239, rs12720071, rs806368) are associated with nicotine dependence in Caucasian American females (Chen et al., 2008).

Recently Benyamina and colleagues systematically reviewed all association studies of CNR1 polymorphisms with drug dependence, and identified 11 studies that were used for meta-analysis of rs1049353, rs806379 and the AAT repeat (Benyamina et al., 2010). Of these polymorphisms only the AAT repeat showed a significant association with illicit substance dependence, and only in Caucasian populations with a risk allele definition of >16 repeats. However, a wide heterogeneity between studies was acknowledged with particular regard to between study differences in ethnic population, lack of gender delineation, differences in phenotype definition and ambiguity regarding illicit substances. Thus, while this meta-analysis indicates a significant influence of the CNR1 AAT repeat polymorphism in drug dependence, it also highlights the need for a greater body of well-described studies to determine the relative influence of CNR1 polymorphisms in addiction.

10.2 FAAH polymorphisms

Several recent reports suggest that genetic alterations in FAAH expression or function confer susceptibility to problem drug and alcohol use. For example, FAAH expression and function are reduced in the PFC of rats selectively bred for high ethanol preference and consumption (Hansson et al., 2007). Consistently, genetic deletion of FAAH results in increased ethanol preference and consumption in mice (Basavarajappa et al., 2006; Blednov et al., 2007; Vinod et al., 2008). Animals lacking FAAH are also less sensitive to ethanol-induced motor incoordination and exhibit reduced signs of ethanol withdrawal, leading to the theory that reductions in the aversive effects of ethanol intake confer increased motivation for ethanol consumption. FAAH knockout mice also display enhanced sensitivity to both the rewarding effects of nicotine and the aversive effects of nicotine withdrawal (Merritt et al., 2008), though Δ9-THC dependence and precipitated withdrawal appear to be unaltered by FAAH deletion (Schlosburg et al., 2009). Collectively these findings in rodents suggest that genetic disruption of FAAH function results in altered sensitivity to some abused drugs such as alcohol and nicotine.

In humans, a missense SNP in the FAAH DNA sequence has been identified (C385A, rs324420) that leads to a conserved proline to threonine (P129T) conversion in the FAAH amino acid sequence (Chiang et al., 2004). The resultant enzyme variant is more susceptible to proteolytic degradation, thereby resulting in substantially reduced FAAH activity and a presumed increase in levels of AEA and other FAAH substrates. Initial reports described a significant association between the C385A polymorphism and problem drug use (Flanagan et al., 2006; Sipe et al., 2002). Further characterization has shown this FAAH mutation to be strongly associated with risk for frequent sedative use, though individuals with this SNP do not appear to be at greater risk for alcohol, nicotine or methamphetamine use or dependence (Iwasaki et al., 2007; Tyndale et al., 2007). Among human marijuana users the C allele of this SNP has been associated with increased risk of cannabis dependence including increased severity of withdrawal symptoms, increased craving during abstinence from marijuana use and heightened happiness after renewed marijuana smoking (Haughey et al., 2008; Schacht et al., 2009; Tyndale et al., 2007). Moreover, a recent study by Filbey and colleagues determined that individuals homozygous for the C allele of this FAAH SNP have a heightened neural response to marijuana-associated cues in reward-related structures (orbitofrontal cortex, anterior cingulated gyrus and NAc) (Filbey et al., 2010). A similar increase in cue-induced neural activity was observed in carriers of the G allele of rs2023239 in CNR1, though the genetic effects were greater for the FAAH gene than for the CNR1 gene in terms of activation cluster size. This pattern of heightened response to marijuana cues was found to increase as the number of risk alleles increased. Another recent study found that carriers of the FAAH C385A SNP display increased ventral striatal reactivity associated with delay discounting, a behavioral index of impulsivity and reward sensitivity (Hariri et al., 2009). Moreover, C385A carriers exhibited a markedly decreased relationship between threat-related amygdala reactivity and trait anxiety, similar to patterns observed in individuals with high familial risk for alcoholism (Glahn et al., 2007). These findings suggest dysregulation of FAAH function through the C385A polymorphism confers increased impulsivity and decreased threat perception that result in increased risk-taking behavior associated with addiction.

11. Concluding remarks

While enhancement of eCB signaling per se does not produce robust addiction-related behaviors, growing evidence implicates eCB signaling in the modulation of the motivational effects produced by ethanol, nicotine, opiates and to a lesser degree psychostimulant drugs. Endocannabinoids have also recently been implicated in mediating the effects of abused inhalants (such as toluene) on prefrontal cortical neural activity (Beckley et al., 2011). Further, preclinical studies point to an important involvement of CB1 receptors in drug-seeking behaviors induced by exposure to both drug-conditioned cues and drug administration itself.

Most preclinical investigations of an eCB influence in drug abuse-related behaviors have focused on pharmacological or genetic manipulations of CB1 receptor function. In this regard it should be noted that the inverse agonist properties of SR141716A and AM251 (Bergman et al., 2008; Janero et al., 2009; Pertwee, 2005) might produce effects that are distinct from the simple blockade of eCB-mediated CB1 receptor activation. In addition, while both of these compounds are typically used to characterize CB1 receptor-mediated processes it is worth bearing in mind that AM251 also activates GPR55 receptors (Henstridge et al., 2009; Kapur et al., 2009; Whyte et al., 2009; Yin et al., 2009) and SR141716A binds to non-CB1, non-CB2 G protein-coupled receptors (including μ and κ opioid receptors and tachykinin NK2 receptors) and ion channels (including TRPV1 receptors) in the low micromolar range (vs. a low nanomolar interaction with CB1) (Cinar et al., 2009; De Petrocellis et al., 2001; Kathmann et al., 2006; Price et al., 2004). Despite these caveats, the aforementioned preclinical studies along with growing clinical evidence for an association between CNR1 polymorphisms and problem drug use/dependence strongly support a cannabinoid receptor influence in the etiology of addiction.

More direct evaluations of an eCB influence on drug-related behaviors have focused on the effects of increased eCB tone induced by inhibition of eCB clearance mechanisms. Inhibition of eCB clearance significantly alters the behavioral effects of alcohol and nicotine, suppresses drug-seeking behavior and ameliorates affective disruptions commonly associated with drug dependence and withdrawal. While many of these tests provide evidence of CB1-dependent eCB influences, not all results are consistent with this pharmacological mechanism. eCBs interact with several non-CB1 putative cannabinoid receptors such as GPR55 and GPR119 (Godlewski et al., 2009; Lauckner et al., 2008; Overton et al., 2006; Pertwee, 2010; Ryberg et al., 2007; Sharir et al., 2010) and the influence of these receptors on addiction-related behaviors has not been characterized. The agonist effects of eCBs at CB2 receptors should also not be discounted as these receptors are present on brain neuronal cells and emerging evidence suggests a possible CB2 receptor influence on drug-related behaviors (Garcia-Gutierrez et al., 2011; Garcia-Gutierrez et al., 2010; Ishiguro et al., 2007; Onaivi et al., 2008). Moreover, AEA and other eCB-like moieties potently activate non-cannabinoid receptors such as TRPV1 receptors (Di Marzo et al., 2010; Zygmunt et al., 1999) and peroxisome proliferators-ativated receptors (PPAR)-α and -γ receptors (O’Sullivan, 2007). As such, behavioral effects produced by manipulations of eCB clearance may result from activation of these non-cannabinoid targets that may produce distinct effects from CB1 receptors. An example of this circumstance is the anxiogenic-like behavioral effects produced by AEA activation of TRPV1 receptors following FAAH inhibition that counters the anxiolytic-like effects of CB1 receptor activation (Micale et al., 2009). Inhibition of eCB clearance mechanisms can also produce effects mediated by non-cannabinoid lipids. A good example of this is the inhibitory effects of PPAR activation on nicotine-induced behaviors and neurochemical effects resulting from increased OEA and/or PEA levels following FAAH inhibition (Luchicchi et al., 2010; Melis et al., 2008; Melis et al., 2010; Scherma et al., 2008b). Thus the behavioral effects produced by eCB clearance inhibition likely result from activation of both CB1 and non-CB1 receptor mechanisms, and this may in part explain some incongruities in the literature such as the attenuation of drug-seeking behavior by either CB1 receptor antagonism (Alvarez-Jaimes et al., 2008; Cippitelli et al., 2005; Cohen et al., 2005; De Vries et al., 2005; De Vries et al., 2003; De Vries et al., 2001; Economidou et al., 2006; Fattore et al., 2003; Justinova et al., 2008a; Xi et al., 2006) or eCB clearance inhibition (Adamczyk et al., 2009; Forget et al., 2009; Scherma et al., 2008b) and the attenuation of nicotine-induced activation of mesolimbic dopamine cells by either CB1 receptor antagonism (Cohen et al., 2002) or eCB clearance inhibition (Luchicchi et al., 2010; Melis et al., 2008; Scherma et al., 2008b).

Because several classes of abused drugs appear to increase eCB formation in brain it is possible that long-term drug exposure results in neuroadaptations that serve to diminish the influence of eCB signaling. Indeed, drug exposure generally leads to a transient loss of eCB-mediated synaptic plasticity at both inhibitory and excitatory synapses and this may be related to a downregulation of CB1 receptor expression and/or function (for review see (Sidhpura et al., 2011)). Drug dependence and withdrawal may also be associated with deficient eCB levels resulting from altered eCB biosynthetic and/or clearance mechanisms, or disruptions in non-cannabinoid mechanisms that can stimulate eCB production (e.g. mGluR5, dopamine D2, etc.). Deficits in eCB signaling may contribute to decreased extinction of drug-related memories, drug craving and relapse, increased stress responsivity and affective disruptions (including anxiety and depression) that are associated with drug dependence and protracted withdrawal. Accordingly, therapeutic approaches aimed at restoring or bolstering cannabinoid signaling may have clinical benefit for the treatment of addiction. The use of exogenous CB1 receptor agonists for this purpose would likely be problematic as these compounds induce drug craving, increase relapse and exacerbate abstinence-related affective disorders at higher doses. However, enhancement of eCB tone through inhibition of eCB clearance may be a viable therapeutic approach in light of evidence that these manipulations facilitate extinction of aversive memory, decrease drug-seeking behavior, constrain excessive stress responses and produce anxiolytic- and antidepressive-like behaviors. Importantly, because eCBs are generally produced in response to specific stimuli, the behavioral effects of moderate eCB clearance inhibition may be preferentially evident in limited circumstances such as exposure to stress or drug-associated cues. Supporting this hypothesis is in vivo evidence that eCB clearance inhibition does not alter extracellular eCB levels under quiescent conditions but enhances and prolongs elevations in extracellular eCB levels induced by neural activation (Alvarez-Jaimes et al., 2009b; Buczynski et al., 2010; Long et al., 2009c). Endocannabinoid clearance inhibition may also facilitate region-specific increases in brain eCB signaling as a result of regionally distinct stimulus-induced eCB production. For this reason moderate doses of eCB clearance inhibitors may produce fewer unwanted behavioral effects than exogenous CB1 agonists that induce widespread CB1 receptor activation. These and related considerations have supported the proposed therapeutic use of eCB clearance inhibitors for a variety of pathologies (Gaetani et al., 2009; Hwang et al., 2010; Lutz, 2004; Petrosino et al., 2010).

Nearly all evaluations of the behavioral effects of eCB clearance inhibition have employed acute dosing procedures and further studies characterizing the effects of long-term administration of eCB clearance inhibitors are needed. The importance of this is underscored by clinical evidence that polymorphisms in the FAAH gene (C385A) conferring reduced FAAH activity are associated with problem sedative use and cannabis dependence. Moreover, although several preclinical studies have characterized the behavioral effects of FAAH inhibition, knowledge of the effects of MAGL inhibition on addiction-related behaviors is sparse. Work in this area will be greatly facilitated by the recent development of selective and efficacious MAGL inhibitors (e.g. (Long et al., 2009c)).

Table 1.

Pharmacological agents frequently cited in this review

CB1 Agonist CB1 Antagonist eCB clearance
inhibitors
Δ9-THC (CB1 = CB2) SR141716A (CB1>CB2) URB597 FAAH
SR147778 (CB1>CB2) PF-
3845
FAAH
WIN 55,212-2 (CB1 ≤ CB2) AM251 (CB1>CB2)
CP 55,940 (CB1 = CB2) JZL184 MAGL
HU-210 (CB1 = CB2)
AM678 (JWH-
018)
(CB1 ≥ CB2) JZL195 Dual
FAAH/MAGL
R-
methanandamide
(CB1 ≥ CB2) PMSF Non-selective
serine
hydrolase
ACEA (CB1 > CB2)
O-1812 (CB1 ≥ CB2) AM404 Reuptake/FAAH
AM1346 (CB1 ≥ CB2) OMDM-
2
Reuptake

Table 2.

Discriminative Stimulus effects of eCBs

Training Drug Test Drug Effect
Δ9-THC exogenous AEA no generalization
FAAH inhibition no generalization
AEA + FAAH inhibition full generalization
R-methanandamide full generalization
O-1812 full generalization
AM1346 full generalization
JZL184 partial generalization
JZL195 full generalization
R-methanandamide Δ9-THC full generalization
WIN 55,212-2 full generalization
AM678 full generalization
AM1346 Δ9-THC full generalization
R-methanandamide full generalization

Table 3.

Rewarding effects of increased eCB signaling

Paradigm Test Drug Effect
Place Conditioning AEA no conditioning effect
URB597 no conditioning effect
AM404 significant place preference
Brain stimulation reward URB597 no effect
PMSF no effect
OMDM-2 no effect
Intravenous self-administration AEA supports operant behavior (squirrel monkey)
URB597 does not support operant behavior (squirrel monkey)
AM404 supports operant behavior (squirrel monkey)

Table 4.

Endocannabinoid influence on alcohol-related behaviors

Manipulation Effect
CB1 receptor CB1 knockout Reduced alcohol conditioned place preference
Reduced alcohol preference and consumption
Absence of alcohol-induced increase in NAc dopamine
Increased sensitivity to intoxication
Increased withdrawal
CB1 antagonists Attenuate acquisition of alcohol consumption
Reduce alcohol preference and consumption
Attenuate alcohol-induced increases in NAc dopamine
CB1 agonists WIN 55,212-2 and CP 55,940 increase the motivation for
and quantity of alcohol consumption
eCB clearance inhibition FAAH knockout Increased alcohol preference and consumption
Rats selectively
bred for high
alcohol
preference
Decreased FAAH expression and activity in PFC
Pharmacological
Manipulations
URB597 increases alcohol preference and consumption in
mice, but not rats
AM404 decreases alcohol consumption in rats
Alcohol-induced alterations
in brain eCB function
Increased 2-AG and AEA formation in culture
Regionally distinct alterations in brain tissue eCB content
Alcohol self-administration increases interstitial levels of 2-
AG but not AEA in the NAc
Decreased CB1 expression and function by chronic alcohol

Table 5.

Endocannabinoid influence on nicotine-related behaviors

Manipulation Effect
CB1 receptor CB1 knockout Reduced nicotine conditioned place preference
Reduced intravenous nicotine self-administration
CB1 antagonists Reduce nicotine conditioned place preference
Reduce intravenous nicotine self-administration
Attenuate nicotine-induced increases in NAc dopamine
No effect on nicotine-induced discriminative stimulus
eCB clearance inhibition FAAH knockout Increased sensitivity to low-dose nicotine place preference
Pharmacological
Manipulations
URB597 blocks nicotine place preference in rats
(CB1 receptor independent)
URB597 attenuates nicotine self-administration in rats
(CB1 receptor independent)
URB597 attenuates nicotine-induced increases in NAc
dopamine in rats (CB1 receptor independent)
Nicotine-induced alterations
in brain eCB function
Regionally distinct alterations in brain tissue eCB content
Increased CB1 expression during protracted nicotine
withdrawal

Table 6.

Endocannabinoid influence on opioid-related behaviors

Manipulation Effect
CB1 receptor CB1 knockout Reduced morphine conditioned place preference
Reduced intravenous heroin self-administration
CB1 antagonists Reduce morphine conditioned place preference
Reduce intravenous heroin self-administration
CB1 agonists WIN 55,212-2 enhaces morphine conditioned place
preference
WIN 55,212-2 and Δ9-THC increase the motivation for
heroin self-administration
eCB clearance inhibition Pharmacological
Manipulations
Neither URB597 nor AM404 alter heroin self-administration
Opiate-induced alterations in
brain eCB function
Regionally distinct alterations in brain tissue eCB content
Heroin self-administration increases interstitial levels of AEA
but not 2-AG in the NAc
CB1 expression significantly altered by chronic morphine

Table 7.

Endocannabinoid influence on psychostimulant-related behaviors

Manipulation Effect
CB1 receptor CB1 knockout Unaltered cocaine place preference
Modest or no change in cocaine self-administration
CB1 antagonists AM251, but not SR141716A, decreases cocaine self-
administration
AM251 decreases methamphetamine self-administration
AM251, but not SR141716A, decreases cocaine-induced
facilitation of brain stimulation reward
SR141716A does not alter cocaine-induced increases in
NAc dopamine
CB1 agonists WIN 55,212-2 reduces cocaine self-administration and
blocks cocaine-induced facilitation of brain stimulation
reward
eCB clearance inhibition Pharmacological
Manipulations
URB597 does not alter cocaine self-administration
URB597 does not modulate cocaine-induced alterations in
VTA dopamine cell activity
AM404 attenuates cocaine-induced facilitation of brain
stimulation reward
Cocaine-induced alterations in
brain eCB function
Tissue eCB content unaltered by chronic cocaine exposure
Cocaine self-administration does not alter interstitial eCB
levels in the NAc
Extended periods of cocaine self-administration result in
increased levels of phosphorylated and total CB1 protein in
the NAc and amygdala

Table 8.

Endocannabinoid influence on drug-seeking behaviors (relapse)

Manipulation Effect
CB1 receptor CB1 agonists reinstate extinguished drug-seeking for:
     cannabinoids
     alcohol
     opioids
     nicotine
     cocaine
CB1 antagonists block:
  Cue-induced reinstatement of drug-seeking for:
     Δ9-THC
     Alcohol
     Heroin
     Nicotine
     cocaine
     sucrose
  Drug-primed reinstatement of drug-seeking for:
     Alcohol (deprivation effect)
     Heroin
     Cocaine
eCB clearance inhibition URB597 does not reinstate extinguished drug-seeking for:
     Δ9-THC
     Cocaine
     AEA
URB597 blocks cue-induced reinstatement of drug-seeking for:
     nicotine
     cocaine
URB597 blocks drug-primed reinstatement of drug-seeking for:
     nicotine
     cocaine
Neither URB597 nor AM404 block cue- or stress-induced
 reinstatement of alcohol-seeking

Acknowledgments

This work was supported by NIH grants PO1-DA017259, P20-DA024194 and P60-AA006420 and grants from the Spanish National Institute of Health ‘Carlos III’, Red de Trastornos Adictivos EU-ERDF (RD06/0001/0000) and National Plan on Drugs PND 2009/049. The authors would like to thank Dr. Matthew Buczynski for assistance in the preparation of Figure 1. This is publication number 21298 from The Scripps Research Institute.

Figure 1. Synaptic organization of the endocannabinoid system.

Figure 1

AEA and 2-AG are synthesized from phospholipids in the postsynaptic cell. The most commonly accepted route for AEA synthesis involves its direct formation from N-arachidonoyl-phosphatidylethanolamine (NAPE) via catalysis by a specific phospholipase D (NAPE-PLD). Alternatively, recent evidence points to AEA formation via the ABHD4/GDE1 pathway and also through the PLC pathway via either PTPN22 or SHIP1 activity. Because the specific enzymes for AEA generation remain unclear no specific pathway is represented in this figure. 2-AG derives primarily from the hydrolytic metabolism of 1,2-diacylglycerol (DAG) via the sn-1-selective DAG lipases DAGLα and DAGLβ. Once released into the synaptic cleft AEA and 2-AG exert signaling effects through CB1 and other receptors including CB2 and GPR55 (Pertwee, 2010). CB1 receptors are expressed on presynaptic terminals, though the synaptic positions of CB2 and GPR55 in the CNS have not been confirmed. AEA also functions as a potent agonist at transient receptor potential vanilloid receptor 1 (TRPV1), though unlike GPCRs these receptors contain an intracellular lipid binding site. Termination of extracellular eCB signaling is initiated by cellular reuptake, and once inside the cell eCBs are shuttled via chaperone mechanisms (including fatty acid binding proteins (FABPs; (Maccarrone et al., 2010)) to specific enzymes for hydrolytic cleavage. Degradation of 2-AG is attributed primarily to presynaptic monoacylglycerol lipase (MAGL) that cleaves 2-AG into arachidonic acid (AA) and glycerol. Degradation of AEA occurs in postsynaptic cells through fatty acid amide hydrolase (FAAH) that cleaves AEA into AA and ethanolamine. Evidence suggests that FAAH can also conjugate AA and various neurotransmitters into n-arachidonoyl moieties such as n-arachidonoyl dopamine (NADA) (Hu et al., 2009) that have potent bioactivity at TRPV1 and CB1 receptors (Pertwee, 2010).

Abbreviations

AEA

anandamide (N-arachidonoyl-ethanolamine)

2-AG

2-arachidonoylglycerol

BLA

basolateral amygdala

BNST

bed nucleus of the stria terminalis

CB1

cannabinoid receptor type 1

CeA

central nucleus of the amygdala

CNR1

cannabinoid receptor type 1 gene

CPA

conditioned place aversion

CPP

conditioned place preference

ECS

endogenous cannabinoid system

FAAH

fatty acid amide hydrolase

MAGL

monoacylglycerol lipase

NAc

nucleus accumbens

nACh

nicotinic acetylcholine ion channel receptors

OEA

oleoylethanolamide

PEA

palmitoylethanolamide

PVN

paraventricular nucleus of the hypothalamus

sP

Sardinian alcohol-preferring rats

THC

Δ9-tetrahydrocannabinol

VTA

ventral tegmental area

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. Aceto MD, Scates SM, Razdan RK, Martin BR. Anandamide, an endogenous cannabinoid, has a very low physical dependence potential. J Pharmacol Exp Ther. 1998;287(2):598–605. [PubMed] [Google Scholar]
  2. Adamczyk P, Golda A, McCreary AC, Filip M, Przegalinski E. Activation of endocannabinoid transmission induces antidepressant-like effects in rats. J Physiol Pharmacol. 2008;59(2):217–228. [PubMed] [Google Scholar]
  3. Adamczyk P, McCreary AC, Przegalinski E, Mierzejewski P, Bienkowski P, Filip M. The effects of fatty acid amide hydrolase inhibitors on maintenance of cocaine and food self-administration and on reinstatement of cocaine-seeking and food-taking behavior in rats. J Physiol Pharmacol. 2009;60(3):119–125. [PubMed] [Google Scholar]
  4. Adermark L, Lovinger DM. Frequency-dependent inversion of net striatal output by endocannabinoid-dependent plasticity at different synaptic inputs. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2009;29(5):1375–1380. doi: 10.1523/JNEUROSCI.3842-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Adermark L, Jonsson S, Ericson M, Soderpalm B. Intermittent ethanol consumption depresses endocannabinoid-signaling in the dorsolateral striatum of rat. Neuropharmacology. 2011 doi: 10.1016/j.neuropharm.2011.01.014. [DOI] [PubMed] [Google Scholar]
  6. Agrawal A, Wetherill L, Dick DM, Xuei X, Hinrichs A, Hesselbrock V, Kramer J, Nurnberger JI, Jr., Schuckit M, Bierut LJ, Edenberg HJ, Foroud T. Evidence for association between polymorphisms in the cannabinoid receptor 1 (CNR1) gene and cannabis dependence. Am J Med Genet B Neuropsychiatr Genet. 2009;150B(5):736–740. doi: 10.1002/ajmg.b.30881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Aguiar DC, Terzian AL, Guimaraes FS, Moreira FA. Anxiolytic-like effects induced by blockade of transient receptor potential vanilloid type 1 (TRPV1) channels in the medial prefrontal cortex of rats. Psychopharmacology (Berl) 2009;205(2):217–225. doi: 10.1007/s00213-009-1532-5. [DOI] [PubMed] [Google Scholar]
  8. Ahn K, McKinney MK, Cravatt BF. Enzymatic pathways that regulate endocannabinoid signaling in the nervous system. Chem Rev. 2008;108(5):1687–1707. doi: 10.1021/cr0782067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Alen F, Moreno-Sanz G, de Tena A. Isabel, Brooks RD, Lopez-Jimenez A, Navarro M, Lopez-Moreno JA. Pharmacological activation of CB1 and D2 receptors in rats: predominant role of CB1 in the increase of alcohol relapse. Eur J Neurosci. 2008;27(12):3292–3298. doi: 10.1111/j.1460-9568.2008.06302.x. [DOI] [PubMed] [Google Scholar]
  10. Alger BE. Retrograde signaling in the regulation of synaptic transmission: focus on endocannabinoids. Prog Neurobiol. 2002;68(4):247–286. doi: 10.1016/s0301-0082(02)00080-1. [DOI] [PubMed] [Google Scholar]
  11. 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(3):431–437. doi: 10.1016/j.pbb.2004.08.020. [DOI] [PubMed] [Google Scholar]
  12. Alling C, Balldin J, Bokstrom K, Gottfries CG, Karlsson I, Langstrom G. Studies on duration of a late recovery period after chronic abuse of ethanol. A cross-sectional study of biochemical and psychiatric indicators. Acta Psychiatr Scand. 1982;66(5):384–397. doi: 10.1111/j.1600-0447.1982.tb06720.x. [DOI] [PubMed] [Google Scholar]
  13. Alvarez-Jaimes L, Parsons LH. Regional influence of CB1 receptor signaling on ethanol self-administration by rats. The Open Neuropsychopharmacology. 2009a;2:77–85. doi: 10.2174/1876523800902020077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Alvarez-Jaimes L, Polis I, Parsons LH. Attenuation of cue-induced heroin-seeking behavior by cannabinoid CB1 antagonist infusions into the nucleus accumbens core and prefrontal cortex, but not basolateral amygdala. Neuropsychopharmacology. 2008;33(10):2483–2493. doi: 10.1038/sj.npp.1301630. [DOI] [PubMed] [Google Scholar]
  15. Alvarez-Jaimes L, Stouffer DG, Parsons LH. Chronic ethanol treatment potentiates ethanol-induced increases in interstitial nucleus accumbens endocannabinoid levels in rats. J Neurochem. 2009b;111(1):37–48. doi: 10.1111/j.1471-4159.2009.06301.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. 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(8):1470–1478. doi: 10.1038/sj.npp.1300454. [DOI] [PubMed] [Google Scholar]
  17. Annis HM, Sklar SM, Moser AE. Gender in relation to relapse crisis situations, coping, and outcome among treated alcoholics. Addict Behav. 1998;23(1):127–131. doi: 10.1016/s0306-4603(97)00024-5. [DOI] [PubMed] [Google Scholar]
  18. Arevalo C, de Miguel R, Hernandez-Tristan R. Cannabinoid effects on anxiety-related behaviours and hypothalamic neurotransmitters. Pharmacol Biochem Behav. 2001;70(1):123–131. doi: 10.1016/s0091-3057(01)00578-0. [DOI] [PubMed] [Google Scholar]
  19. 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(1):104–106. doi: 10.1007/s002130050326. [DOI] [PubMed] [Google Scholar]
  20. Aronne LJ, Tonstad S, Moreno M, Gantz I, Erondu N, Suryawanshi S, Molony C, Sieberts S, Nayee J, Meehan AG, Shapiro D, Heymsfield SB, Kaufman KD, Amatruda JM. A clinical trial assessing the safety and efficacy of taranabant, a CB1R inverse agonist, in obese and overweight patients: a high-dose study. Int J Obes (Lond) 2010 doi: 10.1038/ijo.2010.21. [DOI] [PubMed] [Google Scholar]
  21. Ashton CH. Pharmacology and effects of cannabis: a brief review. Br J Psychiatry. 2001;178:101–106. doi: 10.1192/bjp.178.2.101. [DOI] [PubMed] [Google Scholar]
  22. Ashton JC, Glass M. The cannabinoid CB2 receptor as a target for inflammation-dependent neurodegeneration. Curr Neuropharmacol. 2007;5(2):73–80. doi: 10.2174/157015907780866884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Aso E, Ozaita A, Valdizan EM, Ledent C, Pazos A, Maldonado R, Valverde O. BDNF impairment in the hippocampus is related to enhanced despair behavior in CB1 knockout mice. J Neurochem. 2008;105(2):565–572. doi: 10.1111/j.1471-4159.2007.05149.x. [DOI] [PubMed] [Google Scholar]
  24. Atwood BK, Mackie K. CB2: a cannabinoid receptor with an identity crisis. Br J Pharmacol. 2010;160(3):467–479. doi: 10.1111/j.1476-5381.2010.00729.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Ballon N, Leroy S, Roy C, Bourdel MC, Charles-Nicolas A, Krebs MO, Poirier MF. (AAT)n repeat in the cannabinoid receptor gene (CNR1): association with cocaine addiction in an African-Caribbean population. Pharmacogenomics J. 2006;6(2):126–130. doi: 10.1038/sj.tpj.6500352. [DOI] [PubMed] [Google Scholar]
  26. Bambico FR, Katz N, Debonnel G, Gobbi G. Cannabinoids elicit antidepressant-like behavior and activate serotonergic neurons through the medial prefrontal cortex. J Neurosci. 2007;27(43):11700–11711. doi: 10.1523/JNEUROSCI.1636-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Bardo MT. Neuropharmacological mechanisms of drug reward: beyond dopamine in the nucleus accumbens. Crit Rev Neurobiol. 1998;12(1-2):37–67. doi: 10.1615/critrevneurobiol.v12.i1-2.30. [DOI] [PubMed] [Google Scholar]
  28. Barna I, Zelena D, Arszovszki AC, Ledent C. The role of endogenous cannabinoids in the hypothalamo-pituitary-adrenal axis regulation: in vivo and in vitro studies in CB1 receptor knockout mice. Life Sci. 2004;75(24):2959–2970. doi: 10.1016/j.lfs.2004.06.006. [DOI] [PubMed] [Google Scholar]
  29. Basavarajappa BS, Hungund BL. Chronic ethanol increases the cannabinoid receptor agonist anandamide and its precursor N-arachidonoylphosphatidylethanolamine in SK-N-SH cells. J Neurochem. 1999;72(2):522–528. doi: 10.1046/j.1471-4159.1999.0720522.x. [DOI] [PubMed] [Google Scholar]
  30. Basavarajappa BS, Cooper TB, Hungund BL. Chronic ethanol administration down-regulates cannabinoid receptors in mouse brain synaptic plasma membrane. Brain Res. 1998;793(1-2):212–218. doi: 10.1016/s0006-8993(98)00175-9. [DOI] [PubMed] [Google Scholar]
  31. Basavarajappa BS, Saito M, Cooper TB, Hungund BL. Stimulation of cannabinoid receptor agonist 2-arachidonylglycerol by chronic ethanol and its modulation by specific neuromodulators in cerebellar granule neurons. Biochim Biophys Acta. 2000;1535(1):78–86. doi: 10.1016/s0925-4439(00)00085-5. [DOI] [PubMed] [Google Scholar]
  32. Basavarajappa BS, Saito M, Cooper TB, Hungund BL. Chronic ethanol inhibits the anandamide transport and increases extracellular anandamide levels in cerebellar granule neurons. Eur J Pharmacol. 2003;466(1-2):73–83. doi: 10.1016/s0014-2999(03)01557-7. [DOI] [PubMed] [Google Scholar]
  33. Basavarajappa BS, Yalamanchili R, Cravatt BF, Cooper TB, Hungund BL. Increased ethanol consumption and preference and decreased ethanol sensitivity in female FAAH knockout mice. Neuropharmacology. 2006;50(7):834–844. doi: 10.1016/j.neuropharm.2005.12.005. [DOI] [PubMed] [Google Scholar]
  34. Batkai S, Jarai Z, Wagner JA, Goparaju SK, Varga K, Liu J, Wang L, Mirshahi F, Khanolkar AD, Makriyannis A, Urbaschek R, Garcia N, Jr., Sanyal AJ, Kunos G. Endocannabinoids acting at vascular CB1 receptors mediate the vasodilated state in advanced liver cirrhosis. Nat Med. 2001;7(7):827–832. doi: 10.1038/89953. [DOI] [PubMed] [Google Scholar]
  35. Beckley JT, Woodward JJ. The abused inhalant toluene differentially modulates excitatory and inhibitory synaptic transmission in deep-layer neurons of the medial prefrontal cortex. Neuropsychopharmacology. 2011;36(7):1531–1542. doi: 10.1038/npp.2011.38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Beltramo M, Piomelli D. Carrier-mediated transport and enzymatic hydrolysis of the endogenous cannabinoid 2-arachidonylglycerol. Neuroreport. 2000;11(6):1231–1235. doi: 10.1097/00001756-200004270-00018. [DOI] [PubMed] [Google Scholar]
  37. Beltramo M, Stella N, Calignano A, Lin SY, Makriyannis A, Piomelli D. Functional role of high-affinity anandamide transport, as revealed by selective inhibition. Science. 1997;277(5329):1094–1097. doi: 10.1126/science.277.5329.1094. [DOI] [PubMed] [Google Scholar]
  38. Benyamina A, Kebir O, Blecha L, Reynaud M, Krebs MO. CNR1 gene polymorphisms in addictive disorders: a systematic review and a meta-analysis. Addict Biol. 2010 doi: 10.1111/j.1369-1600.2009.00198.x. [DOI] [PubMed] [Google Scholar]
  39. Bequet F, Uzabiaga F, Desbazeille M, Ludwiczak P, Maftouh M, Picard C, Scatton B, Le Fur G. CB1 receptor-mediated control of the release of endocannabinoids (as assessed by microdialysis coupled with LC/MS) in the rat hypothalamus. Eur J Neurosci. 2007;26(12):3458–3464. doi: 10.1111/j.1460-9568.2007.05900.x. [DOI] [PubMed] [Google Scholar]
  40. Bergman J, Delatte MS, Paronis CA, Vemuri K, Thakur GA, Makriyannis A. Some effects of CB1 antagonists with inverse agonist and neutral biochemical properties. Physiol Behav. 2008;93(4-5):666–670. doi: 10.1016/j.physbeh.2007.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Berke JD, Hyman SE. Addiction, dopamine, and the molecular mechanisms of memory. Neuron. 2000;25(3):515–532. doi: 10.1016/s0896-6273(00)81056-9. [DOI] [PubMed] [Google Scholar]
  42. 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(8):2197–2202. doi: 10.1046/j.1460-9568.2003.02951.x. [DOI] [PubMed] [Google Scholar]
  43. Beyer CE, Dwyer JM, Piesla MJ, Platt BJ, Shen R, Rahman Z, Chan K, Manners MT, Samad TA, Kennedy JD, Bingham B, Whiteside GT. Depression-like phenotype following chronic CB(1) receptor antagonism. Neurobiol Dis. 2010 doi: 10.1016/j.nbd.2010.03.020. [DOI] [PubMed] [Google Scholar]
  44. Biala G, Budzynska B. Calcium-dependent mechanisms of the reinstatement of nicotine-conditioned place preference by drug priming in rats. Pharmacol Biochem Behav. 2008;89(1):116–125. doi: 10.1016/j.pbb.2007.12.005. [DOI] [PubMed] [Google Scholar]
  45. Bisogno T, Melck D, Bobrov M, Gretskaya NM, Bezuglov VV, De Petrocellis L, Di Marzo V. N-acyl-dopamines: novel synthetic CB(1) cannabinoid-receptor ligands and inhibitors of anandamide inactivation with cannabimimetic activity in vitro and in vivo. Biochem J. 2000;351(Pt 3):817–824. [PMC free article] [PubMed] [Google Scholar]
  46. Bisogno T, Howell F, Williams G, Minassi A, Cascio MG, Ligresti A, Matias I, Schiano-Moriello A, Paul P, Williams EJ, Gangadharan U, Hobbs C, Di Marzo V, Doherty P. Cloning of the first sn1-DAG lipases points to the spatial and temporal regulation of endocannabinoid signaling in the brain. J Cell Biol. 2003;163(3):463–468. doi: 10.1083/jcb.200305129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Bitencourt RM, Pamplona FA, Takahashi RN. Facilitation of contextual fear memory extinction and anti-anxiogenic effects of AM404 and cannabidiol in conditioned rats. Eur Neuropsychopharmacol. 2008;18(12):849–859. doi: 10.1016/j.euroneuro.2008.07.001. [DOI] [PubMed] [Google Scholar]
  48. Blankman JL, Simon GM, Cravatt BF. A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem Biol. 2007;14(12):1347–1356. doi: 10.1016/j.chembiol.2007.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. 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(7):1570–1582. doi: 10.1038/sj.npp.1301274. [DOI] [PubMed] [Google Scholar]
  50. Boctor SY, Martinez JL, Jr., Koek W, France CP. The cannabinoid CB1 receptor antagonist AM251 does not modify methamphetamine reinstatement of responding. Eur J Pharmacol. 2007;571(1):39–43. doi: 10.1016/j.ejphar.2007.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. 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(12):2652–2659. doi: 10.1038/sj.npp.1301061. [DOI] [PubMed] [Google Scholar]
  52. 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(4):923–926. doi: 10.1016/s0306-4522(01)00210-x. [DOI] [PubMed] [Google Scholar]
  53. Braida D, Limonta V, Malabarba L, Zani A, Sala M. 5-HT1A receptors are involved in the anxiolytic effect of Delta9-tetrahydrocannabinol and AM 404, the anandamide transport inhibitor, in Sprague-Dawley rats. Eur J Pharmacol. 2007;555(2-3):156–163. doi: 10.1016/j.ejphar.2006.10.038. [DOI] [PubMed] [Google Scholar]
  54. Brodie MS, Pesold C, Appel SB. Ethanol directly excites dopaminergic ventral tegmental area reward neurons. Alcohol Clin Exp Res. 1999;23(11):1848–1852. [PubMed] [Google Scholar]
  55. Brusco A, Tagliaferro PA, Saez T, Onaivi ES. Ultrastructural localization of neuronal brain CB2 cannabinoid receptors. Ann N Y Acad Sci. 2008a;1139:450–457. doi: 10.1196/annals.1432.037. [DOI] [PubMed] [Google Scholar]
  56. Brusco A, Tagliaferro P, Saez T, Onaivi ES. Postsynaptic localization of CB2 cannabinoid receptors in the rat hippocampus. Synapse. 2008b;62(12):944–949. doi: 10.1002/syn.20569. [DOI] [PubMed] [Google Scholar]
  57. Buczynski MW, Parsons LH. Quantification of brain endocannabinoid levels: methods, interpretations and pitfalls. Br J Pharmacol. 2010;160:423–442. doi: 10.1111/j.1476-5381.2010.00787.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Burkey RT, Nation JR. (R)-methanandamide, but not anandamide, substitutes for delta 9-THC in a drug-discrimination procedure. Exp Clin Psychopharmacol. 1997;5(3):195–202. doi: 10.1037//1064-1297.5.3.195. [DOI] [PubMed] [Google Scholar]
  59. Cadas H, Gaillet S, Beltramo M, Venance L, Piomelli D. Biosynthesis of an endogenous cannabinoid precursor in neurons and its control by calcium and cAMP. J Neurosci. 1996;16(12):3934–3942. doi: 10.1523/JNEUROSCI.16-12-03934.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. 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(11):3145–3149. doi: 10.1111/j.1460-9568.2003.02961.x. [DOI] [PubMed] [Google Scholar]
  61. Caille S, Parsons LH. Intravenous heroin self-administration decreases GABA efflux in the ventral pallidum: an in vivo microdialysis study in rats. Eur J Neurosci. 2004;20(2):593–596. doi: 10.1111/j.1460-9568.2004.03497.x. [DOI] [PubMed] [Google Scholar]
  62. Caille S, Parsons LH. Cannabinoid modulation of opiate reinforcement through the ventral striatopallidal pathway. Neuropsychopharmacology. 2006;31(4):804–813. doi: 10.1038/sj.npp.1300848. [DOI] [PubMed] [Google Scholar]
  63. 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(14):3695–3702. doi: 10.1523/JNEUROSCI.4403-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Cannich A, Wotjak CT, Kamprath K, Hermann H, Lutz B, Marsicano G. CB1 cannabinoid receptors modulate kinase and phosphatase activity during extinction of conditioned fear in mice. Learn Mem. 2004;11(5):625–632. doi: 10.1101/lm.77904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Carmody TP. Affect regulation, nicotine addiction, and smoking cessation. J Psychoactive Drugs. 1992;24(2):111–122. doi: 10.1080/02791072.1992.10471632. [DOI] [PubMed] [Google Scholar]
  66. Carter BL, Tiffany ST. Cue-reactivity and the future of addiction research. Addiction. 1999;94(3):349–351. [PubMed] [Google Scholar]
  67. Castane A, Robledo P, Matifas A, Kieffer BL, Maldonado R. Cannabinoid withdrawal syndrome is reduced in double mu and delta opioid receptor knockout mice. Eur J Neurosci. 2003;17(1):155–159. doi: 10.1046/j.1460-9568.2003.02409.x. [DOI] [PubMed] [Google Scholar]
  68. 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(5):857–867. doi: 10.1016/s0028-3908(02)00118-1. [DOI] [PubMed] [Google Scholar]
  69. Castelli MP, Piras A. Paola, D’Agostino A, Pibiri F, Perra S, Gessa GL, Maccarrone M, Pistis M. Dysregulation of the endogenous cannabinoid system in adult rats prenatally treated with the cannabinoid agonist WIN 55,212-2. Eur J Pharmacol. 2007;573(1-3):11–19. doi: 10.1016/j.ejphar.2007.06.047. [DOI] [PubMed] [Google Scholar]
  70. Centonze D, Battista N, Rossi S, Mercuri NB, Finazzi-Agro A, Bernardi G, Calabresi P, Maccarrone M. A critical interaction between dopamine D2 receptors and endocannabinoids mediates the effects of cocaine on striatal gabaergic Transmission. Neuropsychopharmacology. 2004;29(8):1488–1497. doi: 10.1038/sj.npp.1300458. [DOI] [PubMed] [Google Scholar]
  71. 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(4):324–332. doi: 10.1007/s002130050518. [DOI] [PubMed] [Google Scholar]
  72. 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(4):791–795. doi: 10.1523/JNEUROSCI.4152-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Chen BT, Hopf FW, Bonci A. Synaptic plasticity in the mesolimbic system: therapeutic implications for substance abuse. Ann N Y Acad Sci. 2010;1187:129–139. doi: 10.1111/j.1749-6632.2009.05154.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Chen X, Williamson VS, An SS, Hettema JM, Aggen SH, Neale MC, Kendler KS. Cannabinoid receptor 1 gene association with nicotine dependence. Arch Gen Psychiatry. 2008;65(7):816–824. doi: 10.1001/archpsyc.65.7.816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Chevaleyre V, Takahashi KA, Castillo PE. Endocannabinoid-mediated synaptic plasticity in the CNS. Annu Rev Neurosci. 2006;29:37–76. doi: 10.1146/annurev.neuro.29.051605.112834. [DOI] [PubMed] [Google Scholar]
  76. Chhatwal JP, Davis M, Maguschak KA, Ressler KJ. Enhancing cannabinoid neurotransmission augments the extinction of conditioned fear. Neuropsychopharmacology. 2005;30(3):516–524. doi: 10.1038/sj.npp.1300655. [DOI] [PubMed] [Google Scholar]
  77. Chhatwal JP, Stanek-Rattiner L, Davis M, Ressler KJ. Amygdala BDNF signaling is required for consolidation but not encoding of extinction. Nat Neurosci. 2006;9(7):870–872. doi: 10.1038/nn1718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Chhatwal JP, Gutman AR, Maguschak KA, Bowser ME, Yang Y, Davis M, Ressler KJ. Functional interactions between endocannabinoid and CCK neurotransmitter systems may be critical for extinction learning. Neuropsychopharmacology. 2009;34(2):509–521. doi: 10.1038/npp.2008.97. [DOI] [PubMed] [Google Scholar]
  79. Chiang KP, Gerber AL, Sipe JC, Cravatt BF. Reduced cellular expression and activity of the P129T mutant of human fatty acid amide hydrolase: evidence for a link between defects in the endocannabinoid system and problem drug use. Hum Mol Genet. 2004;13(18):2113–2119. doi: 10.1093/hmg/ddh216. [DOI] [PubMed] [Google Scholar]
  80. Childers SR, Fleming L, Konkoy C, Marckel D, Pacheco M, Sexton T, Ward S. Opioid and cannabinoid receptor inhibition of adenylyl cyclase in brain. Ann N Y Acad Sci. 1992;654:33–51. doi: 10.1111/j.1749-6632.1992.tb25954.x. [DOI] [PubMed] [Google Scholar]
  81. Childress AR, Mozley PD, McElgin W, Fitzgerald J, Reivich M, O’Brien CP. Limbic activation during cue-induced cocaine craving. Am J Psychiatry. 1999;156(1):11–18. doi: 10.1176/ajp.156.1.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Christensen R, Kristensen PK, Bartels EM, Bliddal H, Astrup A. Efficacy and safety of the weight-loss drug rimonabant: a meta-analysis of randomised trials. Lancet. 2007;370(9600):1706–1713. doi: 10.1016/S0140-6736(07)61721-8. [DOI] [PubMed] [Google Scholar]
  83. Chutuape MA, Jasinski DR, Fingerhood MI, Stitzer ML. One-, three-, and six-month outcomes after brief inpatient opioid detoxification. Am J Drug Alcohol Abuse. 2001;27(1):19–44. doi: 10.1081/ada-100103117. [DOI] [PubMed] [Google Scholar]
  84. Cinar R, Szucs M. CB1 receptor-independent actions of SR141716 on G-protein signaling: coapplication with the mu-opioid agonist Tyr-D-Ala-Gly-(NMe)Phe-Gly-ol unmasks novel, pertussis toxin-insensitive opioid signaling in mu-opioid receptor-Chinese hamster ovary cells. J Pharmacol Exp Ther. 2009;330(2):567–574. doi: 10.1124/jpet.109.152710. [DOI] [PubMed] [Google Scholar]
  85. Cippitelli A, Bilbao A, Gorriti MA, Navarro M, Massi M, Piomelli D, Ciccocioppo R, Rodriguez de Fonseca F. The anandamide transport inhibitor AM404 reduces ethanol self-administration. Eur J Neurosci. 2007;26(2):476–486. doi: 10.1111/j.1460-9568.2007.05665.x. [DOI] [PubMed] [Google Scholar]
  86. Cippitelli A, Cannella N, Braconi S, Duranti A, Tontini A, Bilbao A, Defonseca FR, Piomelli D, Ciccocioppo R. Increase of brain endocannabinoid anandamide levels by FAAH inhibition and alcohol abuse behaviours in the rat. Psychopharmacology (Berl) 2008;198(4):449–460. doi: 10.1007/s00213-008-1104-0. [DOI] [PubMed] [Google Scholar]
  87. 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(8):2243–2251. doi: 10.1111/j.1460-9568.2005.04056.x. [DOI] [PubMed] [Google Scholar]
  88. Clarke RB, Adermark L. Acute ethanol treatment prevents endocannabinoid-mediated long-lasting disinhibition of striatal output. Neuropharmacology. 2010;58(4-5):799–805. doi: 10.1016/j.neuropharm.2009.12.006. [DOI] [PubMed] [Google Scholar]
  89. Coffey SF, Dansky BS, Carrigan MH, Brady KT. Acute and protracted cocaine abstinence in an outpatient population: a prospective study of mood, sleep and withdrawal symptoms. Drug Alcohol Depend. 2000;59(3):277–286. doi: 10.1016/s0376-8716(99)00126-x. [DOI] [PubMed] [Google Scholar]
  90. 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(1):145–155. doi: 10.1038/sj.npp.1300541. [DOI] [PubMed] [Google Scholar]
  91. 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(5-6):451–463. doi: 10.1097/00008877-200209000-00018. [DOI] [PubMed] [Google Scholar]
  92. 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(2):126–130. doi: 10.1093/oxfordjournals.alcalc.a008368. [DOI] [PubMed] [Google Scholar]
  93. 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(2):181–187. doi: 10.1007/s002130100887. [DOI] [PubMed] [Google Scholar]
  94. Comings DE, Muhleman D, Gade R, Johnson P, Verde R, Saucier G, MacMurray J. Cannabinoid receptor gene (CNR1): association with i.v. drug use. Mol Psychiatry. 1997;2(2):161–168. doi: 10.1038/sj.mp.4000247. [DOI] [PubMed] [Google Scholar]
  95. Cooney NL, Litt MD, Morse PA, Bauer LO, Gaupp L. Alcohol cue reactivity, negative-mood reactivity, and relapse in treated alcoholic men. J Abnorm Psychol. 1997;106(2):243–250. doi: 10.1037//0021-843x.106.2.243. [DOI] [PubMed] [Google Scholar]
  96. Cooper SJ. Endocannabinoids and food consumption: comparisons with benzodiazepine and opioid palatability-dependent appetite. Eur J Pharmacol. 2004;500(1-3):37–49. doi: 10.1016/j.ejphar.2004.07.009. [DOI] [PubMed] [Google Scholar]
  97. Corchero J, Fuentes JA, Manzanares J. delta 9-Tetrahydrocannabinol increases proopiomelanocortin gene expression in the arcuate nucleus of the rat hypothalamus. Eur J Pharmacol. 1997;323(2-3):193–195. doi: 10.1016/s0014-2999(97)00144-1. [DOI] [PubMed] [Google Scholar]
  98. Corchero J, Manzanares J, Fuentes JA. Cannabinoid/opioid crosstalk in the central nervous system. Crit Rev Neurobiol. 2004;16(1-2):159–172. doi: 10.1615/critrevneurobiol.v16.i12.170. [DOI] [PubMed] [Google Scholar]
  99. 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(1):61–65. doi: 10.1016/s0166-4328(00)00311-9. [DOI] [PubMed] [Google Scholar]
  100. Costa B, Giagnoni G, Colleoni M. Precipitated and spontaneous withdrawal in rats tolerant to anandamide. Psychopharmacology (Berl) 2000;149(2):121–128. doi: 10.1007/s002139900360. [DOI] [PubMed] [Google Scholar]
  101. Cota D. The role of the endocannabinoid system in the regulation of hypothalamic-pituitary-adrenal axis activity. J Neuroendocrinol. 2008;20(Suppl 1):35–38. doi: 10.1111/j.1365-2826.2008.01673.x. [DOI] [PubMed] [Google Scholar]
  102. Cota D, Tschop MH, Horvath TL, Levine AS. Cannabinoids, opioids and eating behavior: the molecular face of hedonism? Brain Res Rev. 2006;51(1):85–107. doi: 10.1016/j.brainresrev.2005.10.004. [DOI] [PubMed] [Google Scholar]
  103. Cota D, Steiner MA, Marsicano G, Cervino C, Herman JP, Grubler Y, Stalla J, Pasquali R, Lutz B, Stalla GK, Pagotto U. Requirement of cannabinoid receptor type 1 for the basal modulation of hypothalamic-pituitary-adrenal axis function. Endocrinology. 2007;148(4):1574–1581. doi: 10.1210/en.2005-1649. [DOI] [PubMed] [Google Scholar]
  104. Covault J, Gelernter J, Kranzler H. Association study of cannabinoid receptor gene (CNR1) alleles and drug dependence. Mol Psychiatry. 2001;6(5):501–502. doi: 10.1038/sj.mp.4000925. [DOI] [PubMed] [Google Scholar]
  105. Covey LS, Sullivan MA, Johnston JA, Glassman AH, Robinson MD, Adams DP. Advances in non-nicotine pharmacotherapy for smoking cessation. Drugs. 2000;59(1):17–31. doi: 10.2165/00003495-200059010-00003. [DOI] [PubMed] [Google Scholar]
  106. Crawley JN, Corwin RL, Robinson JK, Felder CC, Devane WA, Axelrod J. Anandamide, an endogenous ligand of the cannabinoid receptor, induces hypomotility and hypothermia in vivo in rodents. Pharmacol Biochem Behav. 1993;46(4):967–972. doi: 10.1016/0091-3057(93)90230-q. [DOI] [PubMed] [Google Scholar]
  107. Crombag HS, Bossert JM, Koya E, Shaham Y. Review. Context-induced relapse to drug seeking: a review. Philos Trans R Soc Lond B Biol Sci. 2008;363(1507):3233–3243. doi: 10.1098/rstb.2008.0090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Dannon PN, Lowengrub K, Amiaz R, Grunhaus L, Kotler M. Comorbid cannabis use and panic disorder: short term and long term follow-up study. Hum Psychopharmacol. 2004;19(2):97–101. doi: 10.1002/hup.560. [DOI] [PubMed] [Google Scholar]
  109. Dawson E, Gill M, Curtis D, Castle D, Hunt N, Murray R, Powell J. Genetic association between alleles of pancreatic phospholipase A2 gene and bipolar affective disorder. Psychiatr Genet. 1995;5(4):177–180. doi: 10.1097/00041444-199524000-00005. [DOI] [PubMed] [Google Scholar]
  110. de Oliveira Alvares L, Genro B. Pasqualini, Diehl F, Molina VA, Quillfeldt JA. Opposite action of hippocampal CB1 receptors in memory reconsolidation and extinction. Neuroscience. 2008;154(4):1648–1655. doi: 10.1016/j.neuroscience.2008.05.005. [DOI] [PubMed] [Google Scholar]
  111. De Petrocellis L, Bisogno T, Maccarrone M, Davis JB, Finazzi-Agro A, Di Marzo V. The activity of anandamide at vanilloid VR1 receptors requires facilitated transport across the cell membrane and is limited by intracellular metabolism. J Biol Chem. 2001;276(16):12856–12863. doi: 10.1074/jbc.M008555200. [DOI] [PubMed] [Google Scholar]
  112. 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(1):164–168. doi: 10.1016/j.bbr.2005.02.021. [DOI] [PubMed] [Google Scholar]
  113. 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(1-2):164–169. doi: 10.1007/s00213-003-1422-1. [DOI] [PubMed] [Google Scholar]
  114. 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(10):1151–1154. doi: 10.1038/nm1001-1151. [DOI] [PubMed] [Google Scholar]
  115. de Wit H. Laboratory-based assessment of alcohol craving in social drinkers. Addiction. 2000;95(Suppl 2):S165–169. doi: 10.1080/09652140050111735. [DOI] [PubMed] [Google Scholar]
  116. Delfs JM, Zhu Y, Druhan JP, Aston-Jones G. Noradrenaline in the ventral forebrain is critical for opiate withdrawal-induced aversion. Nature. 2000;403(6768):430–434. doi: 10.1038/35000212. [DOI] [PubMed] [Google Scholar]
  117. Delfs JM, Kong H, Mestek A, Chen Y, Yu L, Reisine T, Chesselet MF. Expression of mu opioid receptor mRNA in rat brain: an in situ hybridization study at the single cell level. J Comp Neurol. 1994;345(1):46–68. doi: 10.1002/cne.903450104. [DOI] [PubMed] [Google Scholar]
  118. 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(5090):1946–1949. doi: 10.1126/science.1470919. [DOI] [PubMed] [Google Scholar]
  119. 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):227–241. doi: 10.1016/j.neuropharm.2004.06.032. [DOI] [PubMed] [Google Scholar]
  120. Di Marzo V, Petrosino S. Endocannabinoids and the regulation of their levels in health and disease. Curr Opin Lipidol. 2007;18(2):129–140. doi: 10.1097/MOL.0b013e32803dbdec. [DOI] [PubMed] [Google Scholar]
  121. Di Marzo V, De Petrocellis L. Endocannabinoids as regulators of transient receptor potential (TRP) channels: A further opportunity to develop new endocannabinoid-based therapeutic drugs. Curr Med Chem. 2010;17(14):1430–1449. doi: 10.2174/092986710790980078. [DOI] [PubMed] [Google Scholar]
  122. 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(4):1627–1635. doi: 10.1046/j.1471-4159.2000.0741627.x. [DOI] [PubMed] [Google Scholar]
  123. Di S, Malcher-Lopes R, Halmos KC, Tasker JG. Nongenomic glucocorticoid inhibition via endocannabinoid release in the hypothalamus: a fast feedback mechanism. J Neurosci. 2003;23(12):4850–4857. doi: 10.1523/JNEUROSCI.23-12-04850.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Di S, Maxson MM, Franco A, Tasker JG. Glucocorticoids regulate glutamate and GABA synapse-specific retrograde transmission via divergent nongenomic signaling pathways. J Neurosci. 2009;29(2):393–401. doi: 10.1523/JNEUROSCI.4546-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Di S, Malcher-Lopes R, Marcheselli VL, Bazan NG, Tasker JG. Rapid glucocorticoid-mediated endocannabinoid release and opposing regulation of glutamate and gamma-aminobutyric acid inputs to hypothalamic magnocellular neurons. Endocrinology. 2005a;146(10):4292–4301. doi: 10.1210/en.2005-0610. [DOI] [PubMed] [Google Scholar]
  126. Di S, Boudaba C, Popescu IR, Weng FJ, Harris C, Marcheselli VL, Bazan NG, Tasker JG. Activity-dependent release and actions of endocannabinoids in the rat hypothalamic supraoptic nucleus. J Physiol. 2005b;569(Pt 3):751–760. doi: 10.1113/jphysiol.2005.097477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Diana MA, Marty A. Endocannabinoid-mediated short-term synaptic plasticity: depolarization-induced suppression of inhibition (DSI) and depolarization-induced suppression of excitation (DSE) Br J Pharmacol. 2004;142(1):9–19. doi: 10.1038/sj.bjp.0705726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Diergaarde L, de Vries W, Raaso H, Schoffelmeer AN, De Vries TJ. Contextual renewal of nicotine seeking in rats and its suppression by the cannabinoid-1 receptor antagonist Rimonabant (SR141716A) Neuropharmacology. 2008;55(5):712–716. doi: 10.1016/j.neuropharm.2008.06.003. [DOI] [PubMed] [Google Scholar]
  129. Dinh TP, Carpenter D, Leslie FM, Freund TF, Katona I, Sensi SL, Kathuria S, Piomelli D. Brain monoglyceride lipase participating in endocannabinoid inactivation. Proc Natl Acad Sci U S A. 2002;99(16):10819–10824. doi: 10.1073/pnas.152334899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Domschke K, Dannlowski U, Ohrmann P, Lawford B, Bauer J, Kugel H, Heindel W, Young R, Morris P, Arolt V, Deckert J, Suslow T, Baune BT. Cannabinoid receptor 1 (CNR1) gene: impact on antidepressant treatment response and emotion processing in major depression. Eur Neuropsychopharmacol. 2008;18(10):751–759. doi: 10.1016/j.euroneuro.2008.05.003. [DOI] [PubMed] [Google Scholar]
  131. Dong HW, Swanson LW. Projections from bed nuclei of the stria terminalis, dorsomedial nucleus: implications for cerebral hemisphere integration of neuroendocrine, autonomic, and drinking responses. J Comp Neurol. 2006;494(1):75–107. doi: 10.1002/cne.20790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Drew MR, Hen R. Adult hippocampal neurogenesis as target for the treatment of depression. CNS Neurol Disord Drug Targets. 2007;6(3):205–218. doi: 10.2174/187152707780619353. [DOI] [PubMed] [Google Scholar]
  133. Dumont EC, Mark GP, Mader S, Williams JT. Self-administration enhances excitatory synaptic transmission in the bed nucleus of the stria terminalis. Nat Neurosci. 2005;8(4):413–414. doi: 10.1038/nn1414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. 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(4):394–403. doi: 10.1007/s00213-005-0199-9. [DOI] [PubMed] [Google Scholar]
  135. Ehlers CL, Slutske WS, Lind PA, Wilhelmsen KC. Association between single nucleotide polymorphisms in the cannabinoid receptor gene (CNR1) and impulsivity in southwest California Indians. Twin Res Hum Genet. 2007;10(6):805–811. doi: 10.1375/twin.10.6.805. [DOI] [PubMed] [Google Scholar]
  136. Everitt BJ, Dickinson A, Robbins TW. The neuropsychological basis of addictive behaviour. Brain Res Brain Res Rev. 2001;36(2-3):129–138. doi: 10.1016/s0165-0173(01)00088-1. [DOI] [PubMed] [Google Scholar]
  137. Everitt BJ, Parkinson JA, Olmstead MC, Arroyo M, Robledo P, Robbins TW. Associative processes in addiction and reward. The role of amygdala-ventral striatal subsystems. Ann N Y Acad Sci. 1999;877:412–438. doi: 10.1111/j.1749-6632.1999.tb09280.x. [DOI] [PubMed] [Google Scholar]
  138. Fattore L, Martellotta MC, Cossu G, Mascia MS, Fratta W. CB1 cannabinoid receptor agonist WIN 55,212-2 decreases intravenous cocaine self-administration in rats. Behav Brain Res. 1999;104(1-2):141–146. doi: 10.1016/s0166-4328(99)00059-5. [DOI] [PubMed] [Google Scholar]
  139. 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(8):1723–1726. doi: 10.1046/j.1460-9568.2003.02607.x. [DOI] [PubMed] [Google Scholar]
  140. Fattore L, Spano S, Cossu G, Deiana S, Fadda P, Fratta W. Cannabinoid CB(1) antagonist SR 141716A attenuates reinstatement of heroin self-administration in heroin-abstinent rats. Neuropharmacology. 2005a;48(8):1097–1104. doi: 10.1016/j.neuropharm.2005.01.022. [DOI] [PubMed] [Google Scholar]
  141. Fattore L, Vigano D, Fadda P, Rubino T, Fratta W, Parolaro D. Bidirectional regulation of mu-opioid and CB1-cannabinoid receptor in rats self-administering heroin or WIN 55,212-2. Eur J Neurosci. 2007;25(7):2191–2200. doi: 10.1111/j.1460-9568.2007.05470.x. [DOI] [PubMed] [Google Scholar]
  142. Fattore L, Deiana S, Spano SM, Cossu G, Fadda P, Scherma M, Fratta W. Endocannabinoid system and opioid addiction: behavioural aspects. Pharmacol Biochem Behav. 2005b;81(2):343–359. doi: 10.1016/j.pbb.2005.01.031. [DOI] [PubMed] [Google Scholar]
  143. Ferrer B, Bermudez-Silva FJ, Bilbao A, Alvarez-Jaimes L, Sanchez-Vera I, Giuffrida A, Serrano A, Baixeras E, Khaturia S, Navarro M, Parsons LH, Piomelli D, Rodriguez de Fonseca F. Regulation of brain anandamide by acute administration of ethanol. Biochem J. 2007;404(1):97–104. doi: 10.1042/BJ20061898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Filbey FM, Schacht JP, Myers US, Chavez RS, Hutchison KE. Individual and additive effects of the CNR1 and FAAH genes on brain response to marijuana cues. Neuropsychopharmacology. 2010;35(4):967–975. doi: 10.1038/npp.2009.200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Filbey FM, Ray L, Smolen A, Claus ED, Audette A, Hutchison KE. Differential neural response to alcohol priming and alcohol taste cues is associated with DRD4 VNTR and OPRM1 genotypes. Alcohol Clin Exp Res. 2008;32(7):1113–1123. doi: 10.1111/j.1530-0277.2008.00692.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Filip M, Golda A, Zaniewska M, McCreary AC, Nowak E, Kolasiewicz W, Przegalinski E. Involvement of cannabinoid CB1 receptors in drug addiction: effects of rimonabant on behavioral responses induced by cocaine. Pharmacol Rep. 2006;58(6):806–819. [PubMed] [Google Scholar]
  147. Flanagan JM, Gerber AL, Cadet JL, Beutler E, Sipe JC. The fatty acid amide hydrolase 385 A/A (P129T) variant: haplotype analysis of an ancient missense mutation and validation of risk for drug addiction. Hum Genet. 2006;120(4):581–588. doi: 10.1007/s00439-006-0250-x. [DOI] [PubMed] [Google Scholar]
  148. Forget B, Hamon M, Thiebot MH. Cannabinoid CB1 receptors are involved in motivational effects of nicotine in rats. Psychopharmacology (Berl) 2005;181(4):722–734. doi: 10.1007/s00213-005-0015-6. [DOI] [PubMed] [Google Scholar]
  149. Forget B, Coen KM, Le Foll B. Inhibition of fatty acid amide hydrolase reduces reinstatement of nicotine seeking but not break point for nicotine self-administration--comparison with CB(1) receptor blockade. Psychopharmacology (Berl) 2009;205(4):613–624. doi: 10.1007/s00213-009-1569-5. [DOI] [PubMed] [Google Scholar]
  150. 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(1):59–69. doi: 10.1007/s00213-006-0525-x. [DOI] [PubMed] [Google Scholar]
  151. Fourgeaud L, Mato S, Bouchet D, Hemar A, Worley PF, Manzoni OJ. A single in vivo exposure to cocaine abolishes endocannabinoid-mediated long-term depression in the nucleus accumbens. J Neurosci. 2004;24(31):6939–6945. doi: 10.1523/JNEUROSCI.0671-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Freedland CS, Sharpe AL, Samson HH, Porrino LJ. Effects of SR141716A on ethanol and sucrose self-administration. Alcohol Clin Exp Res. 2001;25(2):277–282. [PubMed] [Google Scholar]
  153. Fride E. Multiple roles for the endocannabinoid system during the earliest stages of life: pre- and postnatal development. J Neuroendocrinol. 2008;20(Suppl 1):75–81. doi: 10.1111/j.1365-2826.2008.01670.x. [DOI] [PubMed] [Google Scholar]
  154. Fride E, Mechoulam R. Pharmacological activity of the cannabinoid receptor agonist, anandamide, a brain constituent. Eur J Pharmacol. 1993;231(2):313–314. doi: 10.1016/0014-2999(93)90468-w. [DOI] [PubMed] [Google Scholar]
  155. Fu Y, Pollandt S, Liu J, Krishnan B, Genzer K, Orozco-Cabal L, Gallagher JP, Shinnick-Gallagher P. Long-term potentiation (LTP) in the central amygdala (CeA) is enhanced after prolonged withdrawal from chronic cocaine and requires CRF1 receptors. J Neurophysiol. 2007;97(1):937–941. doi: 10.1152/jn.00349.2006. [DOI] [PubMed] [Google Scholar]
  156. Gadzicki D, Muller-Vahl K, Stuhrmann M. A frequent polymorphism in the coding exon of the human cannabinoid receptor (CNR1) gene. Mol Cell Probes. 1999;13(4):321–323. doi: 10.1006/mcpr.1999.0249. [DOI] [PubMed] [Google Scholar]
  157. Gaetani S, Dipasquale P, Romano A, Righetti L, Cassano T, Piomelli D, Cuomo V. The endocannabinoid system as a target for novel anxiolytic and antidepressant drugs. Int Rev Neurobiol. 2009;85:57–72. doi: 10.1016/S0074-7742(09)85005-8. [DOI] [PubMed] [Google Scholar]
  158. 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(3):233–240. doi: 10.1016/s0014-2999(99)00170-3. [DOI] [PubMed] [Google Scholar]
  159. Garcia-Gutierrez MS, Manzanares J. Overexpression of CB2 cannabinoid receptors decreased vulnerability to anxiety and impaired anxiolytic action of alprazolam in mice. J Psychopharmacol. 2011;25(1):111–120. doi: 10.1177/0269881110379507. [DOI] [PubMed] [Google Scholar]
  160. Garcia-Gutierrez MS, Perez-Ortiz JM, Gutierrez-Adan A, Manzanares J. Depression-resistant endophenotype in mice overexpressing cannabinoid CB(2) receptors. Br J Pharmacol. 2010;160(7):1773–1784. doi: 10.1111/j.1476-5381.2010.00819.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Gardner EL. Endocannabinoid signaling system and brain reward: emphasis on dopamine. Pharmacol Biochem Behav. 2005;81(2):263–284. doi: 10.1016/j.pbb.2005.01.032. [DOI] [PubMed] [Google Scholar]
  162. Gatto GJ, McBride WJ, Murphy JM, Lumeng L, Li TK. Ethanol self-infusion into the ventral tegmental area by alcohol-preferring rats. Alcohol. 1994;11(6):557–564. doi: 10.1016/0741-8329(94)90083-3. [DOI] [PubMed] [Google Scholar]
  163. Gebremedhin D, Lange AR, Campbell WB, Hillard CJ, Harder DR. Cannabinoid CB1 receptor of cat cerebral arterial muscle functions to inhibit L-type Ca2+ channel current. Am J Physiol. 1999;276(6 Pt 2):H2085–2093. doi: 10.1152/ajpheart.1999.276.6.H2085. [DOI] [PubMed] [Google Scholar]
  164. Gerdeman GL, Schechter JB, French ED. Context-specific reversal of cocaine sensitization by the CB1 cannabinoid receptor antagonist rimonabant. Neuropsychopharmacology. 2008;33(11):2747–2759. doi: 10.1038/sj.npp.1301648. [DOI] [PubMed] [Google Scholar]
  165. 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(1):46–53. doi: 10.1093/alcalc/agh114. [DOI] [PubMed] [Google Scholar]
  166. 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(3):1146–1154. doi: 10.1523/JNEUROSCI.22-03-01146.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. 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(4):358–363. doi: 10.1038/7268. [DOI] [PubMed] [Google Scholar]
  168. Giuffrida A, Leweke FM, Gerth CW, Schreiber D, Koethe D, Faulhaber J, Klosterkotter J, Piomelli D. Cerebrospinal anandamide levels are elevated in acute schizophrenia and are inversely correlated with psychotic symptoms. Neuropsychopharmacology. 2004;29(11):2108–2114. doi: 10.1038/sj.npp.1300558. [DOI] [PubMed] [Google Scholar]
  169. Glahn DC, Lovallo WR, Fox PT. Reduced amygdala activation in young adults at high risk of alcoholism: studies from the Oklahoma family health patterns project. Biol Psychiatry. 2007;61(11):1306–1309. doi: 10.1016/j.biopsych.2006.09.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Glass M, Dragunow M, Faull RL. Cannabinoid receptors in the human brain: a detailed anatomical and quantitative autoradiographic study in the fetal, neonatal and adult human brain. Neuroscience. 1997;77(2):299–318. doi: 10.1016/s0306-4522(96)00428-9. [DOI] [PubMed] [Google Scholar]
  171. 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(51):18620–18625. doi: 10.1073/pnas.0509591102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Godlewski G, Offertaler L, Wagner JA, Kunos G. Receptors for acylethanolamides-GPR55 and GPR119. Prostaglandins Other Lipid Mediat. 2009;89(3-4):105–111. doi: 10.1016/j.prostaglandins.2009.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Goeders NE. Stress and cocaine addiction. J Pharmacol Exp Ther. 2002;301(3):785–789. doi: 10.1124/jpet.301.3.785. [DOI] [PubMed] [Google Scholar]
  174. Goeders NE, Guerin GF. Role of corticosterone in intravenous cocaine self-administration in rats. Neuroendocrinology. 1996;64(5):337–348. doi: 10.1159/000127137. [DOI] [PubMed] [Google Scholar]
  175. Goncalves MB, Suetterlin P, Yip P, Molina-Holgado F, Walker DJ, Oudin MJ, Zentar MP, Pollard S, Yanez-Munoz RJ, Williams G, Walsh FS, Pangalos MN, Doherty P. A diacylglycerol lipase-CB2 cannabinoid pathway regulates adult subventricular zone neurogenesis in an age-dependent manner. Mol Cell Neurosci. 2008;38(4):526–536. doi: 10.1016/j.mcn.2008.05.001. [DOI] [PubMed] [Google Scholar]
  176. Gonzalez S, Fernandez-Ruiz J, Sparpaglione V, Parolaro D, Ramos JA. Chronic exposure to morphine, cocaine or ethanol in rats produced different effects in brain cannabinoid CB(1) receptor binding and mRNA levels. Drug Alcohol Depend. 2002a;66(1):77–84. doi: 10.1016/s0376-8716(01)00186-7. [DOI] [PubMed] [Google Scholar]
  177. 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. 2002b;954(1):73–81. doi: 10.1016/s0006-8993(02)03344-9. [DOI] [PubMed] [Google Scholar]
  178. Gonzalez S, Valenti M, de Miguel R, Fezza F, Fernandez-Ruiz J, Di Marzo V, Ramos JA. Changes in endocannabinoid contents in reward-related brain regions of alcohol-exposed rats, and their possible relevance to alcohol relapse. Br J Pharmacol. 2004a;143(4):455–464. doi: 10.1038/sj.bjp.0705963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Gonzalez S, Fernandez-Ruiz J, Di Marzo V, Hernandez M, Arevalo C, Nicanor C, Cascio MG, Ambrosio E, Ramos JA. Behavioral and molecular changes elicited by acute administration of SR141716 to Delta9-tetrahydrocannabinol-tolerant rats: an experimental model of cannabinoid abstinence. Drug Alcohol Depend. 2004b;74(2):159–170. doi: 10.1016/j.drugalcdep.2003.12.011. [DOI] [PubMed] [Google Scholar]
  180. Goparaju SK, Ueda N, Yamaguchi H, Yamamoto S. Anandamide amidohydrolase reacting with 2-arachidonoylglycerol, another cannabinoid receptor ligand. FEBS Lett. 1998;422(1):69–73. doi: 10.1016/s0014-5793(97)01603-7. [DOI] [PubMed] [Google Scholar]
  181. Gorzalka BB, Hill MN, Hillard CJ. Regulation of endocannabinoid signaling by stress: implications for stress-related affective disorders. Neurosci Biobehav Rev. 2008;32(6):1152–1160. doi: 10.1016/j.neubiorev.2008.03.004. [DOI] [PubMed] [Google Scholar]
  182. Gossop M, Green L, Phillips G, Bradley B. Lapse, relapse and survival among opiate addicts after treatment. A prospective follow-up study. Br J Psychiatry. 1989;154:348–353. doi: 10.1192/bjp.154.3.348. [DOI] [PubMed] [Google Scholar]
  183. Griebel G, Stemmelin J, Scatton B. Effects of the cannabinoid CB1 receptor antagonist rimonabant in models of emotional reactivity in rodents. Biol Psychiatry. 2005;57(3):261–267. doi: 10.1016/j.biopsych.2004.10.032. [DOI] [PubMed] [Google Scholar]
  184. Grueter BA, Gosnell HB, Olsen CM, Schramm-Sapyta NL, Nekrasova T, Landreth GE, Winder DG. Extracellular-signal regulated kinase 1-dependent metabotropic glutamate receptor 5-induced long-term depression in the bed nucleus of the stria terminalis is disrupted by cocaine administration. J Neurosci. 2006;26(12):3210–3219. doi: 10.1523/JNEUROSCI.0170-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Guindon J, Hohmann AG. The endocannabinoid system and pain. CNS Neurol Disord Drug Targets. 2009;8(6):403–421. doi: 10.2174/187152709789824660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Gysling K, Wang RY. Morphine-induced activation of A10 dopamine neurons in the rat. Brain Res. 1983;277(1):119–127. doi: 10.1016/0006-8993(83)90913-7. [DOI] [PubMed] [Google Scholar]
  187. Hajos N, Freund TF. Distinct cannabinoid sensitive receptors regulate hippocampal excitation and inhibition. Chem Phys Lipids. 2002;121(1-2):73–82. doi: 10.1016/s0009-3084(02)00149-4. [DOI] [PubMed] [Google Scholar]
  188. Hajos N, Kathuria S, Dinh T, Piomelli D, Freund TF. Endocannabinoid transport tightly controls 2-arachidonoyl glycerol actions in the hippocampus: effects of low temperature and the transport inhibitor AM404. Eur J Neurosci. 2004;19(11):2991–2996. doi: 10.1111/j.0953-816X.2004.03433.x. [DOI] [PubMed] [Google Scholar]
  189. Hall W, Solowij N. Adverse effects of cannabis. Lancet. 1998;352(9140):1611–1616. doi: 10.1016/S0140-6736(98)05021-1. [DOI] [PubMed] [Google Scholar]
  190. Haller J, Bakos N, Szirmay M, Ledent C, Freund TF. The effects of genetic and pharmacological blockade of the CB1 cannabinoid receptor on anxiety. Eur J Neurosci. 2002;16(7):1395–1398. doi: 10.1046/j.1460-9568.2002.02192.x. [DOI] [PubMed] [Google Scholar]
  191. Haller J, Varga B, Ledent C, Barna I, Freund TF. Context-dependent effects of CB1 cannabinoid gene disruption on anxiety-like and social behaviour in mice. Eur J Neurosci. 2004;19(7):1906–1912. doi: 10.1111/j.1460-9568.2004.03293.x. [DOI] [PubMed] [Google Scholar]
  192. Haller J, Barna I, Barsvari B, Pelczer K. Gyimesi, Yasar S, Panlilio LV, Goldberg S. Interactions between environmental aversiveness and the anxiolytic effects of enhanced cannabinoid signaling by FAAH inhibition in rats. Psychopharmacology (Berl) 2009;204(4):607–616. doi: 10.1007/s00213-009-1494-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. 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(1):117–126. doi: 10.1038/sj.npp.1301034. [DOI] [PubMed] [Google Scholar]
  194. Hanus L, Gopher A, Almog S, Mechoulam R. Two new unsaturated fatty acid ethanolamides in brain that bind to the cannabinoid receptor. J Med Chem. 1993;36(20):3032–3034. doi: 10.1021/jm00072a026. [DOI] [PubMed] [Google Scholar]
  195. Hanus L, Abu-Lafi S, Fride E, Breuer A, Vogel Z, Shalev DE, Kustanovich I, Mechoulam R. 2-arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1 receptor. Proc Natl Acad Sci U S A. 2001;98(7):3662–3665. doi: 10.1073/pnas.061029898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Hariri AR, Gorka A, Hyde LW, Kimak M, Halder I, Ducci F, Ferrell RE, Goldman D, Manuck SB. Divergent effects of genetic variation in endocannabinoid signaling on human threat- and reward-related brain function. Biol Psychiatry. 2009;66(1):9–16. doi: 10.1016/j.biopsych.2008.10.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Harloe JP, Thorpe AJ, Lichtman AH. Differential endocannabinoid regulation of extinction in appetitive and aversive Barnes maze tasks. Learn Mem. 2008;15(11):806–809. doi: 10.1101/lm.1113008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Hartman CA, Hopfer CJ, Haberstick B, Rhee SH, Crowley TJ, Corley RP, Hewitt JK, Ehringer MA. The association between cannabinoid receptor 1 gene (CNR1) and cannabis dependence symptoms in adolescents and young adults. Drug Alcohol Depend. 2009;104(1-2):11–16. doi: 10.1016/j.drugalcdep.2009.01.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Haughey HM, Marshall E, Schacht JP, Louis A, Hutchison KE. Marijuana withdrawal and craving: influence of the cannabinoid receptor 1 (CNR1) and fatty acid amide hydrolase (FAAH) genes. Addiction. 2008;103(10):1678–1686. doi: 10.1111/j.1360-0443.2008.02292.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Heifets BD, Castillo PE. Endocannabinoid signaling and long-term synaptic plasticity. Annu Rev Physiol. 2009;71:283–306. doi: 10.1146/annurev.physiol.010908.163149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Heller D, Schneider U, Seifert J, Cimander KF, Stuhrmann M. The cannabinoid receptor gene (CNR1) is not affected in German i.v. drug users. Addict Biol. 2001;6(2):183–187. doi: 10.1080/13556210020040271. [DOI] [PubMed] [Google Scholar]
  202. Henstridge CM, Balenga NA, Ford LA, Ross RA, Waldhoer M, Irving AJ. The GPR55 ligand L-alpha-lysophosphatidylinositol promotes RhoA-dependent Ca2+ signaling and NFAT activation. Faseb J. 2009;23(1):183–193. doi: 10.1096/fj.08-108670. [DOI] [PubMed] [Google Scholar]
  203. Herkenham M, Lynn AB, de Costa BR, Richfield EK. Neuronal localization of cannabinoid receptors in the basal ganglia of the rat. Brain Res. 1991a;547(2):267–274. doi: 10.1016/0006-8993(91)90970-7. [DOI] [PubMed] [Google Scholar]
  204. 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. 1991b;11(2):563–583. doi: 10.1523/JNEUROSCI.11-02-00563.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. 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(5):1932–1936. doi: 10.1073/pnas.87.5.1932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Herman AI, Kranzler HR, Cubells JF, Gelernter J, Covault J. Association study of the CNR1 gene exon 3 alternative promoter region polymorphisms and substance dependence. Am J Med Genet B Neuropsychiatr Genet. 2006;141B(5):499–503. doi: 10.1002/ajmg.b.30325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Hill MN, Gorzalka BB. Pharmacological enhancement of cannabinoid CB1 receptor activity elicits an antidepressant-like response in the rat forced swim test. Eur Neuropsychopharmacol. 2005;15(6):593–599. doi: 10.1016/j.euroneuro.2005.03.003. [DOI] [PubMed] [Google Scholar]
  208. Hill MN, McEwen BS. Involvement of the endocannabinoid system in the neurobehavioural effects of stress and glucocorticoids. Prog Neuropsychopharmacol Biol Psychiatry. 2009a doi: 10.1016/j.pnpbp.2009.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Hill MN, Karatsoreos IN, Hillard CJ, McEwen BS. Rapid elevations in limbic endocannabinoid content by glucocorticoid hormones in vivo. Psychoneuroendocrinology. 2010 doi: 10.1016/j.psyneuen.2010.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Hill MN, Kambo JS, Sun JC, Gorzalka BB, Galea LA. Endocannabinoids modulate stress-induced suppression of hippocampal cell proliferation and activation of defensive behaviours. Eur J Neurosci. 2006;24(7):1845–1849. doi: 10.1111/j.1460-9568.2006.05061.x. [DOI] [PubMed] [Google Scholar]
  211. Hill MN, Miller GE, Ho WS, Gorzalka BB, Hillard CJ. Serum endocannabinoid content is altered in females with depressive disorders: a preliminary report. Pharmacopsychiatry. 2008a;41(2):48–53. doi: 10.1055/s-2007-993211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Hill MN, Carrier EJ, McLaughlin RJ, Morrish AC, Meier SE, Hillard CJ, Gorzalka BB. Regional alterations in the endocannabinoid system in an animal model of depression: effects of concurrent antidepressant treatment. J Neurochem. 2008b;106(6):2322–2336. doi: 10.1111/j.1471-4159.2008.05567.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Hill MN, McLaughlin RJ, Morrish AC, Viau V, Floresco SB, Hillard CJ, Gorzalka BB. Suppression of amygdalar endocannabinoid signaling by stress contributes to activation of the hypothalamic-pituitary-adrenal axis. Neuropsychopharmacology. 2009b;34(13):2733–2745. doi: 10.1038/npp.2009.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Hillard CJ, Jarrahian A. The movement of N-arachidonoylethanolamine (anandamide) across cellular membranes. Chem Phys Lipids. 2000;108(1-2):123–134. doi: 10.1016/s0009-3084(00)00191-2. [DOI] [PubMed] [Google Scholar]
  215. Hillard CJ, Manna S, Greenberg MJ, DiCamelli R, Ross RA, Stevenson LA, Murphy V, Pertwee RG, Campbell WB. Synthesis and characterization of potent and selective agonists of the neuronal cannabinoid receptor (CB1) J Pharmacol Exp Ther. 1999;289(3):1427–1433. [PubMed] [Google Scholar]
  216. Hine B, Torrelio M, Gershon S. Interactions between cannabidiol and delta9-THC during abstinence in morphine-dependent rats. Life Sci. 1975;17(6):851–857. doi: 10.1016/0024-3205(75)90435-x. [DOI] [PubMed] [Google Scholar]
  217. Hoehe MR, Caenazzo L, Martinez MM, Hsieh WT, Modi WS, Gershon ES, Bonner TI. Genetic and physical mapping of the human cannabinoid receptor gene to chromosome 6q14-q15. New Biol. 1991;3(9):880–885. [PubMed] [Google Scholar]
  218. Hoffman AF, Oz M, Caulder T, Lupica CR. Functional tolerance and blockade of long-term depression at synapses in the nucleus accumbens after chronic cannabinoid exposure. J Neurosci. 2003;23(12):4815–4820. doi: 10.1523/JNEUROSCI.23-12-04815.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Hohmann AG, Suplita RL, Bolton NM, Neely MH, Fegley D, Mangieri R, Krey JF, Walker JM, Holmes PV, Crystal JD, Duranti A, Tontini A, Mor M, Tarzia G, Piomelli D. An endocannabinoid mechanism for stress-induced analgesia. Nature. 2005;435(7045):1108–1112. doi: 10.1038/nature03658. [DOI] [PubMed] [Google Scholar]
  220. Holter SM, Kallnik M, Wurst W, Marsicano G, Lutz B, Wotjak CT. Cannabinoid CB1 receptor is dispensable for memory extinction in an appetitively-motivated learning task. Eur J Pharmacol. 2005;510(1-2):69–74. doi: 10.1016/j.ejphar.2005.01.008. [DOI] [PubMed] [Google Scholar]
  221. Hopfer CJ, Young SE, Purcell S, Crowley TJ, Stallings MC, Corley RP, Rhee SH, Smolen A, Krauter K, Hewitt JK, Ehringer MA. Cannabis receptor haplotype associated with fewer cannabis dependence symptoms in adolescents. Am J Med Genet B Neuropsychiatr Genet. 2006;141B(8):895–901. doi: 10.1002/ajmg.b.30378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. 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(2):339–349. doi: 10.1038/sj.npp.1300568. [DOI] [PubMed] [Google Scholar]
  223. Howell LL, Kimmel HL. Monoamine transporters and psychostimulant addiction. Biochem Pharmacol. 2008;75(1):196–217. doi: 10.1016/j.bcp.2007.08.003. [DOI] [PubMed] [Google Scholar]
  224. Howlett AC. The cannabinoid receptors. Prostaglandins Other Lipid Mediat. 2002;68-69:619–631. doi: 10.1016/s0090-6980(02)00060-6. [DOI] [PubMed] [Google Scholar]
  225. Howlett AC. Cannabinoid receptor signaling. Handb Exp Pharmaco. 2005;168:53–79. doi: 10.1007/3-540-26573-2_2. [DOI] [PubMed] [Google Scholar]
  226. Howlett AC, Bidaut-Russell M, Devane WA, Melvin LS, Johnson MR, Herkenham M. The cannabinoid receptor: biochemical, anatomical and behavioral characterization. Trends Neurosci. 1990;13(10):420–423. doi: 10.1016/0166-2236(90)90124-s. [DOI] [PubMed] [Google Scholar]
  227. Hu SS, Bradshaw HB, Benton VM, Chen JS, Huang SM, Minassi A, Bisogno T, Masuda K, Tan B, Roskoski R, Jr., Cravatt BF, Di Marzo V, Walker JM. The biosynthesis of N-arachidonoyl dopamine (NADA), a putative endocannabinoid and endovanilloid, via conjugation of arachidonic acid with dopamine. Prostaglandins Leukot Essent Fatty Acids. 2009;81(4):291–301. doi: 10.1016/j.plefa.2009.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Huang SM, Bisogno T, Trevisani M, Al-Hayani A, De Petrocellis L, Fezza F, Tognetto M, Petros TJ, Krey JF, Chu CJ, Miller JD, Davies SN, Geppetti P, Walker JM, Di Marzo V. An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors. Proc Natl Acad Sci U S A. 2002;99(12):8400–8405. doi: 10.1073/pnas.122196999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. 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(4):698–704. doi: 10.1046/j.1471-4159.2003.01576.x. [DOI] [PubMed] [Google Scholar]
  230. Hutchison KE, Haughey H, Niculescu M, Schacht J, Kaiser A, Stitzel J, Horton WJ, Filbey F. The incentive salience of alcohol: translating the effects of genetic variant in CNR1. Arch Gen Psychiatry. 2008;65(7):841–850. doi: 10.1001/archpsyc.65.7.841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Hwang J, Adamson C, Butler D, Janero DR, Makriyannis A, Bahr BA. Enhancement of endocannabinoid signaling by fatty acid amide hydrolase inhibition: a neuroprotective therapeutic modality. Life Sci. 2010;86(15-16):615–623. doi: 10.1016/j.lfs.2009.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Hyman SE. Addiction: a disease of learning and memory. Am J Psychiatry. 2005;162(8):1414–1422. doi: 10.1176/appi.ajp.162.8.1414. [DOI] [PubMed] [Google Scholar]
  233. Hyman SE, Malenka RC, Nestler EJ. Neural mechanisms of addiction: the role of reward-related learning and memory. Annu Rev Neurosci. 2006;29:565–598. doi: 10.1146/annurev.neuro.29.051605.113009. [DOI] [PubMed] [Google Scholar]
  234. Ishiguro H, Iwasaki S, Teasenfitz L, Higuchi S, Horiuchi Y, Saito T, Arinami T, Onaivi ES. Involvement of cannabinoid CB2 receptor in alcohol preference in mice and alcoholism in humans. Pharmacogenomics J. 2007;7(6):380–385. doi: 10.1038/sj.tpj.6500431. [DOI] [PubMed] [Google Scholar]
  235. Iwasaki S, Ishiguro H, Higuchi S, Onaivi ES, Arinami T. Association study between alcoholism and endocannabinoid metabolic enzyme genes encoding fatty acid amide hydrolase and monoglyceride lipase in a Japanese population. Psychiatr Genet. 2007;17(4):215–220. doi: 10.1097/YPG.0b013e32809913d8. [DOI] [PubMed] [Google Scholar]
  236. Jaffe JH. Trivializing dependence. Br J Addict. 1990;85(11):1425–1427. doi: 10.1111/j.1360-0443.1990.tb01624.x. discussion 1429-1431. [DOI] [PubMed] [Google Scholar]
  237. Jaffe JH, Cascella NG, Kumor KM, Sherer MA. Cocaine-induced cocaine craving. Psychopharmacology (Berl) 1989;97(1):59–64. doi: 10.1007/BF00443414. [DOI] [PubMed] [Google Scholar]
  238. Janero DR, Makriyannis A. Cannabinoid receptor antagonists: pharmacological opportunities, clinical experience, and translational prognosis. Expert Opin Emerg Drugs. 2009;14(1):43–65. doi: 10.1517/14728210902736568. [DOI] [PubMed] [Google Scholar]
  239. Janiri L, Martinotti G, Dario T, Reina D, Paparello F, Pozzi G, Addolorato G, Di Giannantonio M, De Risio S. Anhedonia and substance-related symptoms in detoxified substance-dependent subjects: a correlation study. Neuropsychobiology. 2005;52(1):37–44. doi: 10.1159/000086176. [DOI] [PubMed] [Google Scholar]
  240. Jarbe TU, Lamb RJ, Lin S, Makriyannis A. Delta9-THC training dose as a determinant for (R)-methanandamide generalization in rats: a systematic replication. Behav Pharmacol. 2000;11(1):81–86. doi: 10.1097/00008877-200002000-00009. [DOI] [PubMed] [Google Scholar]
  241. 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(4):369–380. doi: 10.1007/s002130100730. [DOI] [PubMed] [Google Scholar]
  242. 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(3):315–323. doi: 10.1007/s00213-006-0517-x. [DOI] [PubMed] [Google Scholar]
  243. Jarbe TU, Li C, Liu Q, Makriyannis A. Discriminative stimulus functions in rats of AM1346, a high-affinity CB1R selective anandamide analog. Psychopharmacology (Berl) 2009;203(2):229–239. doi: 10.1007/s00213-008-1199-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Jarbe TU, Li C, Vadivel SK, Makriyannis A. Discriminative stimulus functions of methanandamide and delta(9)-THC in rats: tests with aminoalkylindoles (WIN55,212-2 and AM678) and ethanol. Psychopharmacology (Berl) 2010;208(1):87–98. doi: 10.1007/s00213-009-1708-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. 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(4):519–522. doi: 10.1007/s002130050797. [DOI] [PubMed] [Google Scholar]
  246. Jentsch JD, Taylor JR. Impulsivity resulting from frontostriatal dysfunction in drug abuse: implications for the control of behavior by reward-related stimuli. Psychopharmacology (Berl) 1999;146(4):373–390. doi: 10.1007/pl00005483. [DOI] [PubMed] [Google Scholar]
  247. Jhaveri MD, Elmes SJ, Richardson D, Barrett DA, Kendall DA, Mason R, Chapman V. Evidence for a novel functional role of cannabinoid CB(2) receptors in the thalamus of neuropathic rats. Eur J Neurosci. 2008;27(7):1722–1730. doi: 10.1111/j.1460-9568.2008.06162.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Jiang W, Zhang Y, Xiao L, Van Cleemput J, Ji SP, Bai G, Zhang X. Cannabinoids promote embryonic and adult hippocampus neurogenesis and produce anxiolytic- and antidepressant-like effects. J Clin Invest. 2005;115(11):3104–3116. doi: 10.1172/JCI25509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Johnson JP, Muhleman D, MacMurray J, Gade R, Verde R, Ask M, Kelley J, Comings DE. Association between the cannabinoid receptor gene (CNR1) and the P300 event-related potential. Mol Psychiatry. 1997;2(2):169–171. doi: 10.1038/sj.mp.4000246. [DOI] [PubMed] [Google Scholar]
  250. Jose BS, van Oers HA, van de Mheen HD, Garretsen HF, Mackenbach JP. Stressors and alcohol consumption. Alcohol Alcohol. 2000;35(3):307–312. doi: 10.1093/alcalc/35.3.307. [DOI] [PubMed] [Google Scholar]
  251. Juhasz G, Chase D, Pegg E, Downey D, Toth ZG, Stones K, Platt H, Mekli K, Payton A, Elliott R, Anderson IM, Deakin JF. CNR1 gene is associated with high neuroticism and low agreeableness and interacts with recent negative life events to predict current depressive symptoms. Neuropsychopharmacology. 2009;34(8):2019–2027. doi: 10.1038/npp.2009.19. [DOI] [PubMed] [Google Scholar]
  252. Justinova Z, Tanda G, Redhi GH, Goldberg SR. Self-administration of delta9-tetrahydrocannabinol (THC) by drug naive squirrel monkeys. Psychopharmacology (Berl) 2003;169(2):135–140. doi: 10.1007/s00213-003-1484-0. [DOI] [PubMed] [Google Scholar]
  253. 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(1-2):186–194. doi: 10.1007/s00213-003-1693-6. [DOI] [PubMed] [Google Scholar]
  254. 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(23):5645–5650. doi: 10.1523/JNEUROSCI.0951-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Justinova Z, Munzar P, Panlilio LV, Yasar S, Redhi GH, Tanda G, Goldberg SR. Blockade of THC-seeking behavior and relapse in monkeys by the cannabinoid CB(1)-receptor antagonist rimonabant. Neuropsychopharmacology. 2008a;33(12):2870–2877. doi: 10.1038/npp.2008.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Justinova Z, Mangieri RA, Bortolato M, Chefer SI, Mukhin AG, Clapper JR, King AR, Redhi GH, Yasar S, Piomelli D, Goldberg SR. Fatty acid amide hydrolase inhibition heightens anandamide signaling without producing reinforcing effects in primates. Biol Psychiatry. 2008b;64(11):930–937. doi: 10.1016/j.biopsych.2008.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Kalivas PW. The glutamate homeostasis hypothesis of addiction. Nat Rev Neurosci. 2009;10(8):561–572. doi: 10.1038/nrn2515. [DOI] [PubMed] [Google Scholar]
  258. Kalivas PW, Volkow ND. The neural basis of addiction: a pathology of motivation and choice. Am J Psychiatry. 2005;162(8):1403–1413. doi: 10.1176/appi.ajp.162.8.1403. [DOI] [PubMed] [Google Scholar]
  259. Kalivas PW, O’Brien C. Drug addiction as a pathology of staged neuroplasticity. Neuropsychopharmacology. 2008;33(1):166–180. doi: 10.1038/sj.npp.1301564. [DOI] [PubMed] [Google Scholar]
  260. Kamprath K, Marsicano G, Tang J, Monory K, Bisogno T, Di Marzo V, Lutz B, Wotjak CT. Cannabinoid CB1 receptor mediates fear extinction via habituation-like processes. J Neurosci. 2006;26(25):6677–6686. doi: 10.1523/JNEUROSCI.0153-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Kano M, Ohno-Shosaku T, Hashimotodani Y, Uchigashima M, Watanabe M. Endocannabinoid-mediated control of synaptic transmission. Physiol Rev. 2009;89(1):309–380. doi: 10.1152/physrev.00019.2008. [DOI] [PubMed] [Google Scholar]
  262. Kaplan GB, Heinrichs SC, Carey RJ. Treatment of addiction and anxiety using extinction approaches: neural mechanisms and their treatment implications. Pharmacol Biochem Behav. 2011;97(3):619–625. doi: 10.1016/j.pbb.2010.08.004. [DOI] [PubMed] [Google Scholar]
  263. Kaplan RF, Cooney NL, Baker LH, Gillespie RA, Meyer RE, Pomerleau OF. Reactivity to alcohol-related cues: physiological and subjective responses in alcoholics and nonproblem drinkers. J Stud Alcohol. 1985;46(4):267–272. doi: 10.15288/jsa.1985.46.267. [DOI] [PubMed] [Google Scholar]
  264. Kapur A, Zhao P, Sharir H, Bai Y, Caron MG, Barak LS, Abood ME. Atypical responsiveness of the orphan receptor GPR55 to cannabinoid ligands. J Biol Chem. 2009;284(43):29817–29827. doi: 10.1074/jbc.M109.050187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. 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(5):354–361. doi: 10.1007/s00210-006-0033-x. [DOI] [PubMed] [Google Scholar]
  266. 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(1):76–81. doi: 10.1038/nm803. [DOI] [PubMed] [Google Scholar]
  267. Katona I, Rancz EA, Acsady L, Ledent C, Mackie K, Hajos N, Freund TF. Distribution of CB1 cannabinoid receptors in the amygdala and their role in the control of GABAergic transmission. J Neurosci. 2001;21(23):9506–9518. doi: 10.1523/JNEUROSCI.21-23-09506.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Kauer JA. Learning mechanisms in addiction: synaptic plasticity in the ventral tegmental area as a result of exposure to drugs of abuse. Annu Rev Physiol. 2004;66:447–475. doi: 10.1146/annurev.physiol.66.032102.112534. [DOI] [PubMed] [Google Scholar]
  269. Kauer JA, Malenka RC. Synaptic plasticity and addiction. Nat Rev Neurosci. 2007;8(11):844–858. doi: 10.1038/nrn2234. [DOI] [PubMed] [Google Scholar]
  270. Kelley AE. Memory and addiction: shared neural circuitry and molecular mechanisms. Neuron. 2004;44(1):161–179. doi: 10.1016/j.neuron.2004.09.016. [DOI] [PubMed] [Google Scholar]
  271. Kieffer BL. Opioids: first lessons from knockout mice. Trends Pharmacol Sci. 1999;20(1):19–26. doi: 10.1016/s0165-6147(98)01279-6. [DOI] [PubMed] [Google Scholar]
  272. Kim J, Isokawa M, Ledent C, Alger BE. Activation of muscarinic acetylcholine receptors enhances the release of endogenous cannabinoids in the hippocampus. J Neurosci. 2002;22(23):10182–10191. doi: 10.1523/JNEUROSCI.22-23-10182.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Kinsey SG, O’Neal ST, Long JZ, Cravatt BF, Lichtman AH. Inhibition of endocannabinoid catabolic enzymes elicits anxiolytic-like effects in the marble burying assay. Pharmacol Biochem Behav. 2011;98(1):21–27. doi: 10.1016/j.pbb.2010.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Kinsey SG, Long JZ, O’Neal ST, Abdullah RA, Poklis JL, Boger DL, Cravatt BF, Lichtman AH. Blockade of endocannabinoid-degrading enzymes attenuates neuropathic pain. J Pharmacol Exp Ther. 2009;330(3):902–910. doi: 10.1124/jpet.109.155465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Kodas E, Cohen C, Louis C, Griebel G. Cortico-limbic circuitry for conditioned nicotine-seeking behavior in rats involves endocannabinoid signaling. Psychopharmacology (Berl) 2007;194(2):161–171. doi: 10.1007/s00213-007-0813-0. [DOI] [PubMed] [Google Scholar]
  276. Koob G, Kreek MJ. Stress, dysregulation of drug reward pathways, and the transition to drug dependence. Am J Psychiatry. 2007;164(8):1149–1159. doi: 10.1176/appi.ajp.2007.05030503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Koob GF. A role for brain stress systems in addiction. Neuron. 2008;59(1):11–34. doi: 10.1016/j.neuron.2008.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Koob GF. Neurobiological substrates for the dark side of compulsivity in addiction. Neuropharmacology. 2009;56(Suppl 1):18–31. doi: 10.1016/j.neuropharm.2008.07.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. Koob GF, Le Moal M. Drug abuse: hedonic homeostatic dysregulation. Science. 1997;278(5335):52–58. doi: 10.1126/science.278.5335.52. [DOI] [PubMed] [Google Scholar]
  280. Koob GF, Volkow ND. Neurocircuitry of addiction. Neuropsychopharmacology. 2010;35(1):217–238. doi: 10.1038/npp.2009.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Koob GF, Ahmed SH, Boutrel B, Chen SA, Kenny PJ, Markou A, O’Dell LE, Parsons LH, Sanna PP. Neurobiological mechanisms in the transition from drug use to drug dependence. Neurosci Biobehav Rev. 2004;27(8):739–749. doi: 10.1016/j.neubiorev.2003.11.007. [DOI] [PubMed] [Google Scholar]
  282. Kosten TR, Rounsaville BJ, Kleber HD. A 2.5-year follow-up of depression, life crises, and treatment effects on abstinence among opioid addicts. Arch Gen Psychiatry. 1986;43(8):733–738. doi: 10.1001/archpsyc.1986.01800080019003. [DOI] [PubMed] [Google Scholar]
  283. Kreek MJ, Koob GF. Drug dependence: stress and dysregulation of brain reward pathways. Drug Alcohol Depend. 1998;51(1-2):23–47. doi: 10.1016/s0376-8716(98)00064-7. [DOI] [PubMed] [Google Scholar]
  284. Krishnamurthy M, Li W, Moore BM., 2nd. Synthesis, biological evaluation, and structural studies on N1 and C5 substituted cycloalkyl analogues of the pyrazole class of CB1 and CB2 ligands. Bioorg Med Chem. 2004;12(2):393–404. doi: 10.1016/j.bmc.2003.10.045. [DOI] [PubMed] [Google Scholar]
  285. Laberg JC. Alcohol and expectancy: subjective, psychophysiological and behavioral responses to alcohol stimuli in severely, moderately and non-dependent drinkers. Br J Addict. 1986;81(6):797–808. doi: 10.1111/j.1360-0443.1986.tb00407.x. [DOI] [PubMed] [Google Scholar]
  286. Lallemand F, Soubrie PH, De Witte PH. Effects of CB1 cannabinoid receptor blockade on ethanol preference after chronic ethanol administration. Alcohol Clin Exp Res. 2001;25(9):1317–1323. [PubMed] [Google Scholar]
  287. Lan R, Liu Q, Fan P, Lin S, Fernando SR, McCallion D, Pertwee R, Makriyannis A. Structure-activity relationships of pyrazole derivatives as cannabinoid receptor antagonists. J Med Chem. 1999;42(4):769–776. doi: 10.1021/jm980363y. [DOI] [PubMed] [Google Scholar]
  288. Lauckner JE, Jensen JB, Chen HY, Lu HC, Hille B, Mackie K. GPR55 is a cannabinoid receptor that increases intracellular calcium and inhibits M current. Proc Natl Acad Sci U S A. 2008;105(7):2699–2704. doi: 10.1073/pnas.0711278105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Le Foll B, Goldberg SR. Rimonabant, a CB1 antagonist, blocks nicotine-conditioned place preferences. Neuroreport. 2004;15(13):2139–2143. doi: 10.1097/00001756-200409150-00028. [DOI] [PubMed] [Google Scholar]
  290. Le Moal M. Drug abuse: vulnerability and transition to addiction. Pharmacopsychiatry. 2009;42(Suppl 1):S42–55. doi: 10.1055/s-0029-1216355. [DOI] [PubMed] [Google Scholar]
  291. 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(5400):401–404. doi: 10.1126/science.283.5400.401. [DOI] [PubMed] [Google Scholar]
  292. Leshner AI. Addiction is a brain disease, and it matters. Science. 1997;278(5335):45–47. doi: 10.1126/science.278.5335.45. [DOI] [PubMed] [Google Scholar]
  293. Lesscher HM, Hoogveld E, Burbach JP, van Ree JM, Gerrits MA. Endogenous cannabinoids are not involved in cocaine reinforcement and development of cocaine-induced behavioural sensitization. Eur Neuropsychopharmacol. 2005;15(1):31–37. doi: 10.1016/j.euroneuro.2004.04.003. [DOI] [PubMed] [Google Scholar]
  294. Leung D, Saghatelian A, Simon GM, Cravatt BF. Inactivation of N-acyl phosphatidylethanolamine phospholipase D reveals multiple mechanisms for the biosynthesis of endocannabinoids. Biochemistry. 2006;45(15):4720–4726. doi: 10.1021/bi060163l. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Leweke FM, Giuffrida A, Koethe D, Schreiber D, Nolden BM, Kranaster L, Neatby MA, Schneider M, Gerth CW, Hellmich M, Klosterkotter J, Piomelli D. Anandamide levels in cerebrospinal fluid of first-episode schizophrenic patients: impact of cannabis use. Schizophr Res. 2007;94(1-3):29–36. doi: 10.1016/j.schres.2007.04.025. [DOI] [PubMed] [Google Scholar]
  296. Li T, Liu X, Zhu ZH, Zhao J, Hu X, Ball DM, Sham PC, Collier DA. No association between (AAT)n repeats in the cannabinoid receptor gene (CNR1) and heroin abuse in a Chinese population. Mol Psychiatry. 2000;5(2):128–130. doi: 10.1038/sj.mp.4000670. [DOI] [PubMed] [Google Scholar]
  297. Li X, Hoffman AF, Peng XQ, Lupica CR, Gardner EL, Xi ZX. Attenuation of basal and cocaine-enhanced locomotion and nucleus accumbens dopamine in cannabinoid CB1-receptor-knockout mice. Psychopharmacology (Berl) 2009;204(1):1–11. doi: 10.1007/s00213-008-1432-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Lichtman AH, Leung D, Shelton CC, Saghatelian A, Hardouin C, Boger DL, Cravatt BF. Reversible inhibitors of fatty acid amide hydrolase that promote analgesia: evidence for an unprecedented combination of potency and selectivity. J Pharmacol Exp Ther. 2004;311(2):441–448. doi: 10.1124/jpet.104.069401. [DOI] [PubMed] [Google Scholar]
  299. Lisboa SF, Resstel LB, Aguiar DC, Guimaraes FS. Activation of cannabinoid CB1 receptors in the dorsolateral periaqueductal gray induces anxiolytic effects in rats submitted to the Vogel conflict test. Eur J Pharmacol. 2008;593(1-3):73–78. doi: 10.1016/j.ejphar.2008.07.032. [DOI] [PubMed] [Google Scholar]
  300. Liu J, Wang L, Harvey-White J, Osei-Hyiaman D, Razdan R, Gong Q, Chan AC, Zhou Z, Huang BX, Kim HY, Kunos G. A biosynthetic pathway for anandamide. Proc Natl Acad Sci U S A. 2006;103(36):13345–13350. doi: 10.1073/pnas.0601832103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. Liu J, Wang L, Harvey-White J, Huang BX, Kim HY, Luquet S, Palmiter RD, Krystal G, Rai R, Mahadevan A, Razdan RK, Kunos G. Multiple pathways involved in the biosynthesis of anandamide. Neuropharmacology. 2008;54(1):1–7. doi: 10.1016/j.neuropharm.2007.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  302. Liu QS, Pu L, Poo MM. Repeated cocaine exposure in vivo facilitates LTP induction in midbrain dopamine neurons. Nature. 2005;437(7061):1027–1031. doi: 10.1038/nature04050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Long JZ, Nomura DK, Cravatt BF. Characterization of monoacylglycerol lipase inhibition reveals differences in central and peripheral endocannabinoid metabolism. Chem Biol. 2009a;16(7):744–753. doi: 10.1016/j.chembiol.2009.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  304. Long JZ, Nomura DK, Vann RE, Walentiny DM, Booker L, Jin X, Burston JJ, Sim-Selley LJ, Lichtman AH, Wiley JL, Cravatt BF. Dual blockade of FAAH and MAGL identifies behavioral processes regulated by endocannabinoid crosstalk in vivo. Proc Natl Acad Sci U S A. 2009b;106(48):20270–20275. doi: 10.1073/pnas.0909411106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Long JZ, Li W, Booker L, Burston JJ, Kinsey SG, Schlosburg JE, Pavon FJ, Serrano AM, Selley DE, Parsons LH, Lichtman AH, Cravatt BF. Selective blockade of 2-arachidonoylglycerol hydrolysis produces cannabinoid behavioral effects. Nat Chem Biol. 2009c;5(1):37–44. doi: 10.1038/nchembio.129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Lopez-Moreno JA, Gonzalez-Cuevas G, Rodriguez de Fonseca F, Navarro M. Long-lasting increase of alcohol relapse by the cannabinoid receptor agonist WIN 55,212-2 during alcohol deprivation. J Neurosci. 2004;24(38):8245–8252. doi: 10.1523/JNEUROSCI.2179-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Lovinger DM. Presynaptic modulation by endocannabinoids. Handb Exp Pharmaco. 2008;184:435–477. doi: 10.1007/978-3-540-74805-2_14. [DOI] [PubMed] [Google Scholar]
  308. Luchicchi A, Lecca S, Carta S, Pillolla G, Muntoni AL, Yasar S, Goldberg SR, Pistis M. Effects of fatty acid amide hydrolase inhibition on neuronal responses to nicotine, cocaine and morphine in the nucleus accumbens shell and ventral tegmental area: involvement of PPAR-alpha nuclear receptors. Addict Biol. 2010 doi: 10.1111/j.1369-1600.2010.00222.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  309. Lutz B. On-demand activation of the endocannabinoid system in the control of neuronal excitability and epileptiform seizures. Biochem Pharmacol. 2004;68(9):1691–1698. doi: 10.1016/j.bcp.2004.07.007. [DOI] [PubMed] [Google Scholar]
  310. Lutz B. Endocannabinoid signals in the control of emotion. Curr Opin Pharmacol. 2009;9(1):46–52. doi: 10.1016/j.coph.2008.12.001. [DOI] [PubMed] [Google Scholar]
  311. Lynch WJ, Kushner MG, Rawleigh JM, Fiszdon J, Carroll ME. The effects of restraint stress on voluntary ethanol consumption in rats. Exp Clin Psychopharmacol. 1999;7(4):318–323. doi: 10.1037//1064-1297.7.4.318. [DOI] [PubMed] [Google Scholar]
  312. Maccarrone M, Dainese E, Oddi S. Intracellular trafficking of anandamide: new concepts for signaling. Trends Biochem Sci. 2010;35(11):601–608. doi: 10.1016/j.tibs.2010.05.008. [DOI] [PubMed] [Google Scholar]
  313. Mackie K, Hille B. Cannabinoids inhibit N-type calcium channels in neuroblastoma-glioma cells. Proc Natl Acad Sci U S A. 1992;89(9):3825–3829. doi: 10.1073/pnas.89.9.3825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Mackie K, Lai Y, Westenbroek R, Mitchell R. Cannabinoids activate an inwardly rectifying potassium conductance and inhibit Q-type calcium currents in AtT20 cells transfected with rat brain cannabinoid receptor. J Neurosci. 1995;15(10):6552–6561. doi: 10.1523/JNEUROSCI.15-10-06552.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Malcher-Lopes R, Di S, Marcheselli VS, Weng FJ, Stuart CT, Bazan NG, Tasker JG. Opposing crosstalk between leptin and glucocorticoids rapidly modulates synaptic excitation via endocannabinoid release. J Neurosci. 2006;26(24):6643–6650. doi: 10.1523/JNEUROSCI.5126-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Maldonado R, Valverde O. Participation of the opioid system in cannabinoid-induced antinociception and emotional-like responses. Eur Neuropsychopharmacol. 2003;13(6):401–410. doi: 10.1016/j.euroneuro.2003.08.001. [DOI] [PubMed] [Google Scholar]
  317. Malinen H, Hyytia P. Ethanol self-administration is regulated by CB1 receptors in the nucleus accumbens and ventral tegmental area in alcohol-preferring AA rats. Alcohol Clin Exp Res. 2008;32(11):1976–1983. doi: 10.1111/j.1530-0277.2008.00786.x. [DOI] [PubMed] [Google Scholar]
  318. Malinen H, Lehtonen M, Hyytia P. Modulation of brain endocannabinoid levels by voluntary alcohol consumption in alcohol-preferring AA rats. Alcohol Clin Exp Res. 2009;33(10):1711–1720. doi: 10.1111/j.1530-0277.2009.01008.x. [DOI] [PubMed] [Google Scholar]
  319. Mallet PE, Beninger RJ. Delta9-tetrahydrocannabinol, but not the endogenous cannabinoid receptor ligand anandamide, produces conditioned place avoidance. Life Sci. 1998;62(26):2431–2439. doi: 10.1016/s0024-3205(98)00226-4. [DOI] [PubMed] [Google Scholar]
  320. Mansour A, Fox CA, Akil H, Watson SJ. Opioid-receptor mRNA expression in the rat CNS: anatomical and functional implications. Trends Neurosci. 1995;18(1):22–29. doi: 10.1016/0166-2236(95)93946-u. [DOI] [PubMed] [Google Scholar]
  321. Manzanares J, Corchero J, Fuentes JA. Opioid and cannabinoid receptor-mediated regulation of the increase in adrenocorticotropin hormone and corticosterone plasma concentrations induced by central administration of delta(9)-tetrahydrocannabinol in rats. Brain Res. 1999a;839(1):173–179. doi: 10.1016/s0006-8993(99)01756-4. [DOI] [PubMed] [Google Scholar]
  322. Manzanares J, Ortiz S, Oliva JM, Perez-Rial S, Palomo T. Interactions between cannabinoid and opioid receptor systems in the mediation of ethanol effects. Alcohol Alcohol. 2005;40(1):25–34. doi: 10.1093/alcalc/agh112. [DOI] [PubMed] [Google Scholar]
  323. Manzanares J, Corchero J, Romero J, Fernandez-Ruiz JJ, Ramos JA, Fuentes JA. Pharmacological and biochemical interactions between opioids and cannabinoids. Trends Pharmacol Sci. 1999b;20(7):287–294. doi: 10.1016/s0165-6147(99)01339-5. [DOI] [PubMed] [Google Scholar]
  324. Manzanedo C, Aguilar MA, Rodriguez-Arias M, Navarro M, Minarro J. Cannabinoid agonist-induced sensitisation to morphine place preference in mice. Neuroreport. 2004;15(8):1373–1377. doi: 10.1097/01.wnr.0000126217.87116.8c. [DOI] [PubMed] [Google Scholar]
  325. Marchalant Y, Brothers HM, Norman GJ, Karelina K, DeVries AC, Wenk GL. Cannabinoids attenuate the effects of aging upon neuroinflammation and neurogenesis. Neurobiol Dis. 2009;34(2):300–307. doi: 10.1016/j.nbd.2009.01.014. [DOI] [PubMed] [Google Scholar]
  326. Marco EM, Granstrem O, Moreno E, Llorente R, Adriani W, Laviola G, Viveros MP. Subchronic nicotine exposure in adolescence induces long-term effects on hippocampal and striatal cannabinoid-CB1 and mu-opioid receptors in rats. Eur J Pharmacol. 2007;557(1):37–43. doi: 10.1016/j.ejphar.2006.11.013. [DOI] [PubMed] [Google Scholar]
  327. Marrs WR, Blankman JL, Horne EA, Thomazeau A, Lin YH, Coy J, Bodor AL, Muccioli GG, Hu SS-J, Woodruff G, Fung S, Lafourcade M, Alexander JP, Long JZ, Li W, Xu C, Moller T, Mackie K, Manzoni OJ, Cravatt BF, Stella N. The serine hydrolase ABHD6 controls the accumulation and efficacy of 2-AG at cannabinoid receptors. Nat Neurosci. 2010 doi: 10.1038/nn.2601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Marsicano G, Lutz B. Neuromodulatory functions of the endocannabinoid system. J Endocrinol Invest. 2006;29(3 Suppl):27–46. [PubMed] [Google Scholar]
  329. Marsicano G, Wotjak CT, Azad SC, Bisogno T, Rammes G, Cascio MG, Hermann H, Tang J, Hofmann C, Zieglgansberger W, Di Marzo V, Lutz B. The endogenous cannabinoid system controls extinction of aversive memories. Nature. 2002;418(6897):530–534. doi: 10.1038/nature00839. [DOI] [PubMed] [Google Scholar]
  330. Martin BR, Compton DR, Thomas BF, Prescott WR, Little PJ, Razdan RK, Johnson MR, Melvin LS, Mechoulam R, Ward SJ. Behavioral, biochemical, and molecular modeling evaluations of cannabinoid analogs. Pharmacol Biochem Behav. 1991;40(3):471–478. doi: 10.1016/0091-3057(91)90349-7. [DOI] [PubMed] [Google Scholar]
  331. Martin M, Ledent C, Parmentier M, Maldonado R, Valverde O. Cocaine, but not morphine, induces conditioned place preference and sensitization to locomotor responses in CB1 knockout mice. Eur J Neurosci. 2000;12(11):4038–4046. doi: 10.1046/j.1460-9568.2000.00287.x. [DOI] [PubMed] [Google Scholar]
  332. Martin M, Ledent C, Parmentier M, Maldonado R, Valverde O. Involvement of CB1 cannabinoid receptors in emotional behaviour. Psychopharmacology (Berl) 2002;159(4):379–387. doi: 10.1007/s00213-001-0946-5. [DOI] [PubMed] [Google Scholar]
  333. Martin-Fardon R, Zorrilla EP, Ciccocioppo R, Weiss F. Role of innate and drug-induced dysregulation of brain stress and arousal systems in addiction: Focus on corticotropin-releasing factor, nociceptin/orphanin FQ, and orexin/hypocretin. Brain Res. 2010;1314:145–161. doi: 10.1016/j.brainres.2009.12.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. 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(3):R1–2. doi: 10.1016/s0014-2999(99)00656-1. [DOI] [PubMed] [Google Scholar]
  335. Mascia P, Pistis M, Justinova Z, Panlilio LV, Luchicchi A, Lecca S, Scherma M, Fratta W, Fadda P, Barnes C, Redhi GH, Yasar S, Le Foll B, Tanda G, Piomelli D, Goldberg SR. Blockade of nicotine reward and reinstatement by activation of alpha-type peroxisome proliferator-activated receptors. Biol Psychiatry. 2011;69(7):633–641. doi: 10.1016/j.biopsych.2010.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  336. Massi P, Vaccani A, Romorini S, Parolaro D. Comparative characterization in the rat of the interaction between cannabinoids and opiates for their immunosuppressive and analgesic effects. J Neuroimmunol. 2001;117(1-2):116–124. doi: 10.1016/s0165-5728(01)00323-x. [DOI] [PubMed] [Google Scholar]
  337. Mato S, Robbe D, Puente N, Grandes P, Manzoni OJ. Presynaptic homeostatic plasticity rescues long-term depression after chronic Delta 9-tetrahydrocannabinol exposure. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2005;25(50):11619–11627. doi: 10.1523/JNEUROSCI.2294-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  338. Mato S, Chevaleyre V, Robbe D, Pazos A, Castillo PE, Manzoni OJ. A single in-vivo exposure to delta 9THC blocks endocannabinoid-mediated synaptic plasticity. Nat Neurosci. 2004;7(6):585–586. doi: 10.1038/nn1251. [DOI] [PubMed] [Google Scholar]
  339. Matsuda LA, Bonner TI, Lolait SJ. Localization of cannabinoid receptor mRNA in rat brain. J Comp Neurol. 1993;327(4):535–550. doi: 10.1002/cne.903270406. [DOI] [PubMed] [Google Scholar]
  340. Matthes HW, Maldonado R, Simonin F, Valverde O, Slowe S, Kitchen I, Befort K, Dierich A, Le Meur M, Dolle P, Tzavara E, Hanoune J, Roques BP, Kieffer BL. Loss of morphine-induced analgesia, reward effect and withdrawal symptoms in mice lacking the mu-opioid-receptor gene. Nature. 1996;383(6603):819–823. doi: 10.1038/383819a0. [DOI] [PubMed] [Google Scholar]
  341. Matthews RT, German DC. Electrophysiological evidence for excitation of rat ventral tegmental area dopamine neurons by morphine. Neuroscience. 1984;11(3):617–625. doi: 10.1016/0306-4522(84)90048-4. [DOI] [PubMed] [Google Scholar]
  342. McBride WJ, Murphy JM, Ikemoto S. Localization of brain reinforcement mechanisms: intracranial self-administration and intracranial place-conditioning studies. Behav Brain Res. 1999;101(2):129–152. doi: 10.1016/s0166-4328(99)00022-4. [DOI] [PubMed] [Google Scholar]
  343. McElligott ZA, Winder DG. Modulation of glutamatergic synaptic transmission in the bed nucleus of the stria terminalis. Prog Neuropsychopharmacol Biol Psychiatry. 2009;33(8):1329–1335. doi: 10.1016/j.pnpbp.2009.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  344. McGregor IS, Dastur FN, McLellan RA, Brown RE. Cannabinoid modulation of rat pup ultrasonic vocalizations. Eur J Pharmacol. 1996;313(1-2):43–49. doi: 10.1016/0014-2999(96)00511-0. [DOI] [PubMed] [Google Scholar]
  345. McGregor IS, Dam KD, Mallet PE, Gallate JE. Delta9-THC reinstates beer- and sucrose-seeking behaviour in abstinent rats: comparison with midazolam, food deprivation and predator odour. Alcohol Alcohol. 2005;40(1):35–45. doi: 10.1093/alcalc/agh113. [DOI] [PubMed] [Google Scholar]
  346. McKinney MK, Cravatt BF. Structure and function of fatty acid amide hydrolase. Annu Rev Biochem. 2005;74:411–432. doi: 10.1146/annurev.biochem.74.082803.133450. [DOI] [PubMed] [Google Scholar]
  347. McLaughlin RJ, Hill MN, Morrish AC, Gorzalka BB. Local enhancement of cannabinoid CB1 receptor signalling in the dorsal hippocampus elicits an antidepressant-like effect. Behav Pharmacol. 2007;18(5-6):431–438. doi: 10.1097/FBP.0b013e3282ee7b44. [DOI] [PubMed] [Google Scholar]
  348. McLellan AT, Lewis DC, O’Brien CP, Kleber HD. Drug dependence, a chronic medical illness: implications for treatment, insurance, and outcomes evaluation. Jama. 2000;284(13):1689–1695. doi: 10.1001/jama.284.13.1689. [DOI] [PubMed] [Google Scholar]
  349. McMahon LR. Characterization of cannabinoid agonists and apparent pA2 analysis of cannabinoid antagonists in rhesus monkeys discriminating Delta9-tetrahydrocannabinol. J Pharmacol Exp Ther. 2006;319(3):1211–1218. doi: 10.1124/jpet.106.107110. [DOI] [PubMed] [Google Scholar]
  350. McMahon LR, Ginsburg BC, Lamb RJ. Cannabinoid agonists differentially substitute for the discriminative stimulus effects of Delta(9)-tetrahydrocannabinol in C57BL/6J mice. Psychopharmacology (Berl) 2008;198(4):487–495. doi: 10.1007/s00213-007-0900-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  351. 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(1):83–90. doi: 10.1016/0006-2952(95)00109-d. [DOI] [PubMed] [Google Scholar]
  352. Melis M, Pillolla G, Luchicchi A, Muntoni AL, Yasar S, Goldberg SR, Pistis M. Endogenous fatty acid ethanolamides suppress nicotine-induced activation of mesolimbic dopamine neurons through nuclear receptors. J Neurosci. 2008;28(51):13985–13994. doi: 10.1523/JNEUROSCI.3221-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  353. Melis M, Carta S, Fattore L, Tolu S, Yasar S, Goldberg SR, Fratta W, Maskos U, Pistis M. Peroxisome Proliferator-Activated Receptors-Alpha Modulate Dopamine Cell Activity Through Nicotinic Receptors. Biol Psychiatry. 2010 doi: 10.1016/j.biopsych.2010.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  354. Mendelson JH, Mello NK. Management of cocaine abuse and dependence. N Engl J Med. 1996;334(15):965–972. doi: 10.1056/NEJM199604113341507. [DOI] [PubMed] [Google Scholar]
  355. Merritt LL, Martin BR, Walters C, Lichtman AH, Damaj MI. The endogenous cannabinoid system modulates nicotine reward and dependence. J Pharmacol Exp Ther. 2008;326(2):483–492. doi: 10.1124/jpet.108.138321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  356. Micale V, Cristino L, Tamburella A, Petrosino S, Leggio GM, Drago F, Di Marzo V. Anxiolytic effects in mice of a dual blocker of fatty acid amide hydrolase and transient receptor potential vanilloid type-1 channels. Neuropsychopharmacology. 2009;34(3):593–606. doi: 10.1038/npp.2008.98. [DOI] [PubMed] [Google Scholar]
  357. Miller WR, Harris RJ. A simple scale of Gorski’s warning signs for relapse. J Stud Alcohol. 2000;61(5):759–765. doi: 10.15288/jsa.2000.61.759. [DOI] [PubMed] [Google Scholar]
  358. Moise AM, Eisenstein SA, Astarita G, Piomelli D, Hohmann AG. An endocannabinoid signaling system modulates anxiety-like behavior in male Syrian hamsters. Psychopharmacology (Berl) 2008;200(3):333–346. doi: 10.1007/s00213-008-1209-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  359. Moreira FA, Aguiar DC, Guimaraes FS. Anxiolytic-like effect of cannabinoids injected into the rat dorsolateral periaqueductal gray. Neuropharmacology. 2007;52(3):958–965. doi: 10.1016/j.neuropharm.2006.10.013. [DOI] [PubMed] [Google Scholar]
  360. Moreira FA, Kaiser N, Monory K, Lutz B. Reduced anxiety-like behaviour induced by genetic and pharmacological inhibition of the endocannabinoid-degrading enzyme fatty acid amide hydrolase (FAAH) is mediated by CB1 receptors. Neuropharmacology. 2008;54(1):141–150. doi: 10.1016/j.neuropharm.2007.07.005. [DOI] [PubMed] [Google Scholar]
  361. Morrish AC, Hill MN, Riebe CJ, Gorzalka BB. Protracted cannabinoid administration elicits antidepressant behavioral responses in rats: role of gender and noradrenergic transmission. Physiol Behav. 2009;98(1-2):118–124. doi: 10.1016/j.physbeh.2009.04.023. [DOI] [PubMed] [Google Scholar]
  362. Munro S, Thomas KL, Abu-Shaar M. Molecular characterization of a peripheral receptor for cannabinoids. Nature. 1993;365(6441):61–65. doi: 10.1038/365061a0. [DOI] [PubMed] [Google Scholar]
  363. 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(2):243–253. doi: 10.1016/j.neuropharm.2003.09.002. [DOI] [PubMed] [Google Scholar]
  364. Naderi N, Haghparast A, Saber-Tehrani A, Rezaii N, Alizadeh AM, Khani A, Motamedi F. Interaction between cannabinoid compounds and diazepam on anxiety-like behaviour of mice. Pharmacol Biochem Behav. 2008;89(1):64–75. doi: 10.1016/j.pbb.2007.11.001. [DOI] [PubMed] [Google Scholar]
  365. Naidu PS, Varvel SA, Ahn K, Cravatt BF, Martin BR, Lichtman AH. Evaluation of fatty acid amide hydrolase inhibition in murine models of emotionality. Psychopharmacology (Berl) 2007;192(1):61–70. doi: 10.1007/s00213-006-0689-4. [DOI] [PubMed] [Google Scholar]
  366. Navarro M, Hernandez E, Munoz RM, del Arco I, Villanua MA, Carrera MR, Rodriguez de Fonseca F. Acute administration of the CB1 cannabinoid receptor antagonist SR 141716A induces anxiety-like responses in the rat. Neuroreport. 1997;8(2):491–496. doi: 10.1097/00001756-199701200-00023. [DOI] [PubMed] [Google Scholar]
  367. Navarro M, Chowen J, Rocio ACM, del Arco I, Villanua MA, Martin Y, Roberts AJ, Koob GF, de Fonseca FR. CB1 cannabinoid receptor antagonist-induced opiate withdrawal in morphine-dependent rats. Neuroreport. 1998;9(15):3397–3402. doi: 10.1097/00001756-199810260-00012. [DOI] [PubMed] [Google Scholar]
  368. 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(14):5344–5350. doi: 10.1523/JNEUROSCI.21-14-05344.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  369. Nestler EJ. From neurobiology to treatment: progress against addiction. Nat Neurosci. 2002;5(Suppl):1076–1079. doi: 10.1038/nn945. [DOI] [PubMed] [Google Scholar]
  370. Nissen SE, Nicholls SJ, Wolski K, Rodes-Cabau J, Cannon CP, Deanfield JE, Despres JP, Kastelein JJ, Steinhubl SR, Kapadia S, Yasin M, Ruzyllo W, Gaudin C, Job B, Hu B, Bhatt DL, Lincoff AM, Tuzcu EM. Effect of rimonabant on progression of atherosclerosis in patients with abdominal obesity and coronary artery disease: the STRADIVARIUS randomized controlled trial. Jama. 2008;299(13):1547–1560. doi: 10.1001/jama.299.13.1547. [DOI] [PubMed] [Google Scholar]
  371. Niyuhire F, Varvel SA, Thorpe AJ, Stokes RJ, Wiley JL, Lichtman AH. The disruptive effects of the CB1 receptor antagonist rimonabant on extinction learning in mice are task-specific. Psychopharmacology (Berl) 2007;191(2):223–231. doi: 10.1007/s00213-006-0650-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  372. Nunes EV, Levin FR. Treatment of depression in patients with alcohol or other drug dependence: a meta-analysis. Jama. 2004a;291(15):1887–1896. doi: 10.1001/jama.291.15.1887. [DOI] [PubMed] [Google Scholar]
  373. Nunes EV, Sullivan MA, Levin FR. Treatment of depression in patients with opiate dependence. Biol Psychiatry. 2004b;56(10):793–802. doi: 10.1016/j.biopsych.2004.06.037. [DOI] [PubMed] [Google Scholar]
  374. O’Brien CP. A range of research-based pharmacotherapies for addiction. Science. 1997;278(5335):66–70. doi: 10.1126/science.278.5335.66. [DOI] [PubMed] [Google Scholar]
  375. O’Brien CP. Review. Evidence-based treatments of addiction. Philos Trans R Soc Lond B Biol Sci. 2008;363(1507):3277–3286. doi: 10.1098/rstb.2008.0105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  376. O’Brien CP, McLellan AT. Myths about the treatment of addiction. Lancet. 1996;347(8996):237–240. doi: 10.1016/s0140-6736(96)90409-2. [DOI] [PubMed] [Google Scholar]
  377. O’Brien CP, Gardner EL. Critical assessment of how to study addiction and its treatment: human and non-human animal models. Pharmacol Ther. 2005;108(1):18–58. doi: 10.1016/j.pharmthera.2005.06.018. [DOI] [PubMed] [Google Scholar]
  378. O’Brien CP, Childress AR, McLellan AT, Ehrman R. Classical conditioning in drug-dependent humans. Ann N Y Acad Sci. 1992;654:400–415. doi: 10.1111/j.1749-6632.1992.tb25984.x. [DOI] [PubMed] [Google Scholar]
  379. O’Brien CP, Childress AR, Ehrman R, Robbins SJ. Conditioning factors in drug abuse: can they explain compulsion? J Psychopharmacol. 1998;12(1):15–22. doi: 10.1177/026988119801200103. [DOI] [PubMed] [Google Scholar]
  380. O’Sullivan SE. Cannabinoids go nuclear: evidence for activation of peroxisome proliferator-activated receptors. Br J Pharmacol. 2007;152(5):576–582. doi: 10.1038/sj.bjp.0707423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  381. Okamoto Y, Morishita J, Tsuboi K, Tonai T, Ueda N. Molecular characterization of a phospholipase D generating anandamide and its congeners. J Biol Chem. 2004;279(7):5298–5305. doi: 10.1074/jbc.M306642200. [DOI] [PubMed] [Google Scholar]
  382. Onaivi ES, Chakrabarti A, Gwebu ET, Chaudhuri G. Neurobehavioral effects of delta 9-THC and cannabinoid (CB1) receptor gene expression in mice. Behav Brain Res. 1995;72(1-2):115–125. doi: 10.1016/0166-4328(96)00139-8. [DOI] [PubMed] [Google Scholar]
  383. 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;1074:514–536. doi: 10.1196/annals.1369.052. [DOI] [PubMed] [Google Scholar]
  384. Onaivi ES, Ishiguro H, Gong JP, Patel S, Meozzi PA, Myers L, Perchuk A, Mora Z, Tagliaferro PA, Gardner E, Brusco A, Akinshola BE, Hope B, Lujilde J, Inada T, Iwasaki S, Macharia D, Teasenfitz L, Arinami T, Uhl GR. Brain neuronal CB2 cannabinoid receptors in drug abuse and depression: from mice to human subjects. PLoS One. 2008;3(2):e1640. doi: 10.1371/journal.pone.0001640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  385. Orio L, Edwards S, George O, Parsons LH, Koob GF. A role for the endocannabinoid system in the increased motivation for cocaine in extended-access conditions. J Neurosci. 2009;29(15):4846–4857. doi: 10.1523/JNEUROSCI.0563-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  386. Overton HA, Babbs AJ, Doel SM, Fyfe MC, Gardner LS, Griffin G, Jackson HC, Procter MJ, Rasamison CM, Tang-Christensen M, Widdowson PS, Williams GM, Reynet C. Deorphanization of a G protein-coupled receptor for oleoylethanolamide and its use in the discovery of small-molecule hypophagic agents. Cell Metab. 2006;3(3):167–175. doi: 10.1016/j.cmet.2006.02.004. [DOI] [PubMed] [Google Scholar]
  387. Oviedo A, Glowa J, Herkenham M. Chronic cannabinoid administration alters cannabinoid receptor binding in rat brain: a quantitative autoradiographic study. Brain Res. 1993;616(1-2):293–302. doi: 10.1016/0006-8993(93)90220-h. [DOI] [PubMed] [Google Scholar]
  388. Pamplona FA, Bitencourt RM, Takahashi RN. Short- and long-term effects of cannabinoids on the extinction of contextual fear memory in rats. Neurobiol Learn Mem. 2008;90(1):290–293. doi: 10.1016/j.nlm.2008.04.003. [DOI] [PubMed] [Google Scholar]
  389. Pamplona FA, Prediger RD, Pandolfo P, Takahashi RN. The cannabinoid receptor agonist WIN 55,212-2 facilitates the extinction of contextual fear memory and spatial memory in rats. Psychopharmacology (Berl) 2006;188(4):641–649. doi: 10.1007/s00213-006-0514-0. [DOI] [PubMed] [Google Scholar]
  390. Pan B, Hillard CJ, Liu QS. Endocannabinoid signaling mediates cocaine-induced inhibitory synaptic plasticity in midbrain dopamine neurons. J Neurosci. 2008;28(6):1385–1397. doi: 10.1523/JNEUROSCI.4033-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  391. Pandey R, Mousawy K, Nagarkatti M, Nagarkatti P. Endocannabinoids and immune regulation. Pharmacol Res. 2009;60(2):85–92. doi: 10.1016/j.phrs.2009.03.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  392. Patel S, Hillard CJ. Pharmacological evaluation of cannabinoid receptor ligands in a mouse model of anxiety: further evidence for an anxiolytic role for endogenous cannabinoid signaling. J Pharmacol Exp Ther. 2006;318(1):304–311. doi: 10.1124/jpet.106.101287. [DOI] [PubMed] [Google Scholar]
  393. Patel S, Rademacher DJ, Hillard CJ. Differential regulation of the endocannabinoids anandamide and 2-arachidonylglycerol within the limbic forebrain by dopamine receptor activity. J Pharmacol Exp Ther. 2003;306(3):880–888. doi: 10.1124/jpet.103.054270. [DOI] [PubMed] [Google Scholar]
  394. Patel S, Cravatt BF, Hillard CJ. Synergistic interactions between cannabinoids and environmental stress in the activation of the central amygdala. Neuropsychopharmacology. 2005a;30(3):497–507. doi: 10.1038/sj.npp.1300535. [DOI] [PubMed] [Google Scholar]
  395. Patel S, Roelke CT, Rademacher DJ, Hillard CJ. Inhibition of restraint stress-induced neural and behavioural activation by endogenous cannabinoid signalling. Eur J Neurosci. 2005b;21(4):1057–1069. doi: 10.1111/j.1460-9568.2005.03916.x. [DOI] [PubMed] [Google Scholar]
  396. Patel S, Roelke CT, Rademacher DJ, Cullinan WE, Hillard CJ. Endocannabinoid signaling negatively modulates stress-induced activation of the hypothalamic-pituitary-adrenal axis. Endocrinology. 2004;145(12):5431–5438. doi: 10.1210/en.2004-0638. [DOI] [PubMed] [Google Scholar]
  397. Paulus MP. Neural basis of reward and craving--a homeostatic point of view. Dialogues Clin Neurosci. 2007;9(4):379–387. doi: 10.31887/DCNS.2007.9.4/mpaulus. [DOI] [PMC free article] [PubMed] [Google Scholar]
  398. Perera TD, Park S, Nemirovskaya Y. Cognitive role of neurogenesis in depression and antidepressant treatment. Neuroscientist. 2008;14(4):326–338. doi: 10.1177/1073858408317242. [DOI] [PubMed] [Google Scholar]
  399. 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(3):368–377. doi: 10.1007/s00213-005-0195-0. [DOI] [PubMed] [Google Scholar]
  400. Pertwee R, Griffin G, Hanus L, Mechoulam R. Effects of two endogenous fatty acid ethanolamides on mouse vasa deferentia. Eur J Pharmacol. 1994;259(2):115–120. doi: 10.1016/0014-2999(94)90499-5. [DOI] [PubMed] [Google Scholar]
  401. Pertwee RG. Inverse agonism and neutral antagonism at cannabinoid CB1 receptors. Life Sci. 2005;76(12):1307–1324. doi: 10.1016/j.lfs.2004.10.025. [DOI] [PubMed] [Google Scholar]
  402. Pertwee RG. Receptors and channels targeted by synthetic cannabinoid receptor agonists and antagonists. Curr Med Chem. 2010;17(14):1360–1381. doi: 10.2174/092986710790980050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  403. Petrosino S, Di Marzo V. FAAH and MAGL inhibitors: therapeutic opportunities from regulating endocannabinoid levels. Curr Opin Investig Drugs. 2010;11(1):51–62. [PubMed] [Google Scholar]
  404. Piazza PV, Le Moal M. The role of stress in drug self-administration. Trends Pharmacol Sci. 1998;19(2):67–74. doi: 10.1016/s0165-6147(97)01115-2. [DOI] [PubMed] [Google Scholar]
  405. Picciotto MR, Caldarone BJ, King SL, Zachariou V. Nicotinic receptors in the brain. Links between molecular biology and behavior. Neuropsychopharmacology. 2000;22(5):451–465. doi: 10.1016/S0893-133X(99)00146-3. [DOI] [PubMed] [Google Scholar]
  406. Pierce RC, Kumaresan V. The mesolimbic dopamine system: the final common pathway for the reinforcing effect of drugs of abuse? Neurosci Biobehav Rev. 2006;30(2):215–238. doi: 10.1016/j.neubiorev.2005.04.016. [DOI] [PubMed] [Google Scholar]
  407. Piomelli D. The molecular logic of endocannabinoid signalling. Nat Rev Neurosci. 2003;4(11):873–884. doi: 10.1038/nrn1247. [DOI] [PubMed] [Google Scholar]
  408. Ponce G, Hoenicka J, Rubio G, Ampuero I, Jimenez-Arriero MA, Rodriguez-Jimenez R, Palomo T, Ramos JA. Association between cannabinoid receptor gene (CNR1) and childhood attention deficit/hyperactivity disorder in Spanish male alcoholic patients. Mol Psychiatry. 2003;8(5):466–467. doi: 10.1038/sj.mp.4001278. [DOI] [PubMed] [Google Scholar]
  409. Porter AC, Sauer JM, Knierman MD, Becker GW, Berna MJ, Bao J, Nomikos GG, Carter P, Bymaster FP, Leese AB, Felder CC. Characterization of a novel endocannabinoid, virodhamine, with antagonist activity at the CB1 receptor. J Pharmacol Exp Ther. 2002;301(3):1020–1024. doi: 10.1124/jpet.301.3.1020. [DOI] [PubMed] [Google Scholar]
  410. Preuss UW, Koller G, Zill P, Bondy B, Soyka M. Alcoholism-related phenotypes and genetic variants of the CB1 receptor. Eur Arch Psychiatry Clin Neurosci. 2003;253(6):275–280. doi: 10.1007/s00406-003-0440-7. [DOI] [PubMed] [Google Scholar]
  411. Price TJ, Patwardhan A, Akopian AN, Hargreaves KM, Flores CM. Modulation of trigeminal sensory neuron activity by the dual cannabinoid-vanilloid agonists anandamide, N-arachidonoyl-dopamine and arachidonyl-2-chloroethylamide. Br J Pharmacol. 2004;141(7):1118–1130. doi: 10.1038/sj.bjp.0705711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  412. Pugh G, Jr., Mason DJ, Jr., Combs V, Welch SP. Involvement of dynorphin B in the antinociceptive effects of the cannabinoid CP55,940 in the spinal cord. J Pharmacol Exp Ther. 1997;281(2):730–737. [PubMed] [Google Scholar]
  413. Quirk GJ, Gehlert DR. Inhibition of the amygdala: key to pathological states? Ann N Y Acad Sci. 2003;985:263–272. doi: 10.1111/j.1749-6632.2003.tb07087.x. [DOI] [PubMed] [Google Scholar]
  414. Racz I, Bilkei-Gorzo A, Toth ZE, Michel K, Palkovits M, Zimmer A. A critical role for the cannabinoid CB1 receptors in alcohol dependence and stress-stimulated ethanol drinking. J Neurosci. 2003;23(6):2453–2458. doi: 10.1523/JNEUROSCI.23-06-02453.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  415. Rademacher DJ, Hillard CJ. Interactions between endocannabinoids and stress-induced decreased sensitivity to natural reward. Prog Neuropsychopharmacol Biol Psychiatry. 2007;31(3):633–641. doi: 10.1016/j.pnpbp.2006.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  416. Rademacher DJ, Meier SE, Shi L, Ho WS, Jarrahian A, Hillard CJ. Effects of acute and repeated restraint stress on endocannabinoid content in the amygdala, ventral striatum, and medial prefrontal cortex in mice. Neuropharmacology. 2008;54(1):108–116. doi: 10.1016/j.neuropharm.2007.06.012. [DOI] [PubMed] [Google Scholar]
  417. Re G, Barbero R, Miolo A, Di Marzo V. Palmitoylethanolamide, endocannabinoids and related cannabimimetic compounds in protection against tissue inflammation and pain: potential use in companion animals. Vet J. 2007;173(1):21–30. doi: 10.1016/j.tvjl.2005.10.003. [DOI] [PubMed] [Google Scholar]
  418. Reich CG, Mohammadi MH, Alger BE. Endocannabinoid modulation of fear responses: learning and state-dependent performance effects. J Psychopharmacol. 2008;22(7):769–777. doi: 10.1177/0269881107083999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  419. 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;780:168–175. doi: 10.1111/j.1749-6632.1996.tb15121.x. [DOI] [PubMed] [Google Scholar]
  420. Rice OV, Gordon N, Gifford AN. Conditioned place preference to morphine in cannabinoid CB1 receptor knockout mice. Brain Res. 2002;945(1):135–138. doi: 10.1016/s0006-8993(02)02890-1. [DOI] [PubMed] [Google Scholar]
  421. Richman JA, Flaherty JA, Rospenda KM. Perceived workplace harassment experiences and problem drinking among physicians: broadening the stress/alienation paradigm. Addiction. 1996;91(3):391–403. [PubMed] [Google Scholar]
  422. Riddle EL, Fleckenstein AE, Hanson GR. Role of monoamine transporters in mediating psychostimulant effects. Aaps J. 2005;7(4):E847–851. doi: 10.1208/aapsj070481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  423. Ritz MC, Lamb RJ, Goldberg SR, Kuhar MJ. Cocaine receptors on dopamine transporters are related to self-administration of cocaine. Science. 1987;237(4819):1219–1223. doi: 10.1126/science.2820058. [DOI] [PubMed] [Google Scholar]
  424. Rodd ZA, Melendez RI, Bell RL, Kuc KA, Zhang Y, Murphy JM, McBride WJ. Intracranial self-administration of ethanol within the ventral tegmental area of male Wistar rats: evidence for involvement of dopamine neurons. J Neurosci. 2004;24(5):1050–1057. doi: 10.1523/JNEUROSCI.1319-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  425. Rodd-Henricks ZA, McKinzie DL, Crile RS, Murphy JM, McBride WJ. Regional heterogeneity for the intracranial self-administration of ethanol within the ventral tegmental area of female Wistar rats. Psychopharmacology (Berl) 2000;149(3):217–224. doi: 10.1007/s002139900347. [DOI] [PubMed] [Google Scholar]
  426. Rodriguez de Fonseca F, Navarro M, Gomez R, Escuredo L, Nava F, Fu J, Murillo-Rodriguez E, Giuffrida A, LoVerme J, Gaetani S, Kathuria S, Gall C, Piomelli D. An anorexic lipid mediator regulated by feeding. Nature. 2001;414(6860):209–212. doi: 10.1038/35102582. [DOI] [PubMed] [Google Scholar]
  427. 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(3):823–833. doi: 10.1523/JNEUROSCI.21-03-00823.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  428. Romero J, Garcia L, Fernandez-Ruiz JJ, Cebeira M, Ramos JA. Changes in rat brain cannabinoid binding sites after acute or chronic exposure to their endogenous agonist, anandamide, or to delta 9-tetrahydrocannabinol. Pharmacol Biochem Behav. 1995;51(4):731–737. doi: 10.1016/0091-3057(95)00023-p. [DOI] [PubMed] [Google Scholar]
  429. Romero J, Berrendero F, Garcia-Gil L, Ramos JA, Fernandez-Ruiz JJ. Cannabinoid receptor and WIN-55,212-2-stimulated [35S]GTP gamma S binding and cannabinoid receptor mRNA levels in the basal ganglia and the cerebellum of adult male rats chronically exposed to delta 9-tetrahydrocannabinol. J Mol Neurosci. 1998;11(2):109–119. doi: 10.1385/JMN:11:2:109. [DOI] [PubMed] [Google Scholar]
  430. Roohbakhsh A, Moghaddam AH, Massoudi R, Zarrindast MR. Role of dorsal hippocampal cannabinoid receptors and nitric oxide in anxiety like behaviours in rats using the elevated plus-maze test. Clin Exp Pharmacol Physiol. 2007;34(3):223–229. doi: 10.1111/j.1440-1681.2007.04576.x. [DOI] [PubMed] [Google Scholar]
  431. Rospenda KM, Richman JA, Wislar JS, Flaherty JA. Chronicity of sexual harassment and generalized work-place abuse: effects on drinking outcomes. Addiction. 2000;95(12):1805–1820. doi: 10.1046/j.1360-0443.2000.9512180510.x. [DOI] [PubMed] [Google Scholar]
  432. Ross S, Peselow E. The neurobiology of addictive disorders. Clin Neuropharmacol. 2009;32(5):269–276. doi: 10.1097/wnf.0b013e3181a9163c. [DOI] [PubMed] [Google Scholar]
  433. Rothman RB, Baumann MH. Monoamine transporters and psychostimulant drugs. Eur J Pharmacol. 2003;479(1-3):23–40. doi: 10.1016/j.ejphar.2003.08.054. [DOI] [PubMed] [Google Scholar]
  434. Rouge-Pont F, Deroche V, Le Moal M, Piazza PV. Individual differences in stress-induced dopamine release in the nucleus accumbens are influenced by corticosterone. Eur J Neurosci. 1998;10(12):3903–3907. doi: 10.1046/j.1460-9568.1998.00438.x. [DOI] [PubMed] [Google Scholar]
  435. Rouge-Pont F, Marinelli M, Le Moal M, Simon H, Piazza PV. Stress-induced sensitization and glucocorticoids. II. Sensitization of the increase in extracellular dopamine induced by cocaine depends on stress-induced corticosterone secretion. J Neurosci. 1995;15(11):7189–7195. doi: 10.1523/JNEUROSCI.15-11-07189.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  436. Rubino T, Realini N, Braida D, Alberio T, Capurro V, Vigano D, Guidali C, Sala M, Fasano M, Parolaro D. The depressive phenotype induced in adult female rats by adolescent exposure to THC is associated with cognitive impairment and altered neuroplasticity in the prefrontal cortex. Neurotox Res. 2009;15(4):291–302. doi: 10.1007/s12640-009-9031-3. [DOI] [PubMed] [Google Scholar]
  437. Rubino T, Realini N, Castiglioni C, Guidali C, Vigano D, Marras E, Petrosino S, Perletti G, Maccarrone M, Di Marzo V, Parolaro D. Role in anxiety behavior of the endocannabinoid system in the prefrontal cortex. Cereb Cortex. 2008;18(6):1292–1301. doi: 10.1093/cercor/bhm161. [DOI] [PubMed] [Google Scholar]
  438. Rubio M, McHugh D, Fernandez-Ruiz J, Bradshaw H, Walker JM. Short-term exposure to alcohol in rats affects brain levels of anandamide, other N-acylethanolamines and 2-arachidonoyl-glycerol. Neurosci Lett. 2007;421(3):270–274. doi: 10.1016/j.neulet.2007.05.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  439. Russo R, Loverme J, La Rana G, Compton TR, Parrott J, Duranti A, Tontini A, Mor M, Tarzia G, Calignano A, Piomelli D. The fatty acid amide hydrolase inhibitor URB597 (cyclohexylcarbamic acid 3′-carbamoylbiphenyl-3-yl ester) reduces neuropathic pain after oral administration in mice. J Pharmacol Exp Ther. 2007;322(1):236–242. doi: 10.1124/jpet.107.119941. [DOI] [PubMed] [Google Scholar]
  440. Rutkowska M, Jachimczuk O. Antidepressant--like properties of ACEA (arachidonyl-2-chloroethylamide), the selective agonist of CB1 receptors. Acta Pol Pharm. 2004;61(2):165–167. [PubMed] [Google Scholar]
  441. Rutkowska M, Jamontt J, Gliniak H. Effects of cannabinoids on the anxiety-like response in mice. Pharmacol Rep. 2006;58(2):200–206. [PubMed] [Google Scholar]
  442. Ryberg E, Larsson N, Sjogren S, Hjorth S, Hermansson NO, Leonova J, Elebring T, Nilsson K, Drmota T, Greasley PJ. The orphan receptor GPR55 is a novel cannabinoid receptor. Br J Pharmacol. 2007;152(7):1092–1101. doi: 10.1038/sj.bjp.0707460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  443. Saario SM, Laitinen JT. Monoglyceride lipase as an enzyme hydrolyzing 2-arachidonoylglycerol. Chem Biodivers. 2007;4(8):1903–1913. doi: 10.1002/cbdv.200790158. [DOI] [PubMed] [Google Scholar]
  444. Santos CJ, Stern CA, Bertoglio LJ. Attenuation of anxiety-related behaviour after the antagonism of transient receptor potential vanilloid type 1 channels in the rat ventral hippocampus. Behav Pharmacol. 2008;19(4):357–360. doi: 10.1097/FBP.0b013e3283095234. [DOI] [PubMed] [Google Scholar]
  445. Schacht JP, Selling RE, Hutchison KE. Intermediate cannabis dependence phenotypes and the FAAH C385A variant: an exploratory analysis. Psychopharmacology (Berl) 2009;203(3):511–517. doi: 10.1007/s00213-008-1397-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  446. Scherma M, Medalie J, Fratta W, Vadivel SK, Makriyannis A, Piomelli D, Mikics E, Haller J, Yasar S, Tanda G, Goldberg SR. The endogenous cannabinoid anandamide has effects on motivation and anxiety that are revealed by fatty acid amide hydrolase (FAAH) inhibition. Neuropharmacology. 2008a;54(1):129–140. doi: 10.1016/j.neuropharm.2007.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  447. Scherma M, Panlilio LV, Fadda P, Fattore L, Gamaleddin I, Le Foll B, Justinova Z, Mikics E, Haller J, Medalie J, Stroik J, Barnes C, Yasar S, Tanda G, Piomelli D, Fratta W, Goldberg SR. Inhibition of anandamide hydrolysis by cyclohexyl carbamic acid 3′-carbamoyl-3-yl ester (URB597) reverses abuse-related behavioral and neurochemical effects of nicotine in rats. J Pharmacol Exp Ther. 2008b;327(2):482–490. doi: 10.1124/jpet.108.142224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  448. Schlosburg JE, Carlson BL, Ramesh D, Abdullah RA, Long JZ, Cravatt BF, Lichtman AH. Inhibitors of endocannabinoid-metabolizing enzymes reduce precipitated withdrawal responses in THC-dependent mice. Aaps J. 2009;11(2):342–352. doi: 10.1208/s12248-009-9110-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  449. Schlosburg JE, Blankman JL, Long JZ, Nomura DK, Pan B, Kinsey SG, Nguyen PT, Ramesh D, Booker L, Burston JJ, Thomas EA, Selley DE, Sim-Selley LJ, Liu QS, Lichtman AH, Cravatt BF. Chronic monoacylglycerol lipase blockade causes functional antagonism of the endocannabinoid system. Nat Neurosci. 2010;13(9):1113–1119. doi: 10.1038/nn.2616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  450. Schmidt LG, Samochowiec J, Finckh U, Fiszer-Piosik E, Horodnicki J, Wendel B, Rommelspacher H, Hoehe MR. Association of a CB1 cannabinoid receptor gene (CNR1) polymorphism with severe alcohol dependence. Drug Alcohol Depend. 2002;65(3):221–224. doi: 10.1016/s0376-8716(01)00164-8. [DOI] [PubMed] [Google Scholar]
  451. See RE. Neural substrates of cocaine-cue associations that trigger relapse. Eur J Pharmacol. 2005;526(1-3):140–146. doi: 10.1016/j.ejphar.2005.09.034. [DOI] [PubMed] [Google Scholar]
  452. Serra S, Brunetti G, Pani M, Vacca G, Carai MA, Gessa GL, Colombo G. Blockade by the cannabinoid CB(1) receptor antagonist, SR 141716, of alcohol deprivation effect in alcohol-preferring rats. Eur J Pharmacol. 2002;443(1-3):95–97. doi: 10.1016/s0014-2999(02)01594-7. [DOI] [PubMed] [Google Scholar]
  453. Serra S, Carai MA, Brunetti G, Gomez R, Melis S, Vacca G, Colombo G, Gessa GL. The cannabinoid receptor antagonist SR 141716 prevents acquisition of drinking behavior in alcohol-preferring rats. Eur J Pharmacol. 2001;430(2-3):369–371. doi: 10.1016/s0014-2999(01)01379-6. [DOI] [PubMed] [Google Scholar]
  454. Shaham Y, Hope BT. The role of neuroadaptations in relapse to drug seeking. Nat Neurosci. 2005;8(11):1437–1439. doi: 10.1038/nn1105-1437. [DOI] [PubMed] [Google Scholar]
  455. Shaham Y, Erb S, Stewart J. Stress-induced relapse to heroin and cocaine seeking in rats: a review. Brain Res Brain Res Rev. 2000;33(1):13–33. doi: 10.1016/s0165-0173(00)00024-2. [DOI] [PubMed] [Google Scholar]
  456. Shalev U, Grimm JW, Shaham Y. Neurobiology of relapse to heroin and cocaine seeking: a review. Pharmacol Rev. 2002;54(1):1–42. doi: 10.1124/pr.54.1.1. [DOI] [PubMed] [Google Scholar]
  457. Sharir H, Abood ME. Pharmacological characterization of GPR55, a putative cannabinoid receptor. Pharmacol Ther. 2010;126(3):301–313. doi: 10.1016/j.pharmthera.2010.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  458. Shearman LP, Rosko KM, Fleischer R, Wang J, Xu S, Tong XS, Rocha BA. Antidepressant-like and anorectic effects of the cannabinoid CB1 receptor inverse agonist AM251 in mice. Behav Pharmacol. 2003;14(8):573–582. doi: 10.1097/00008877-200312000-00001. [DOI] [PubMed] [Google Scholar]
  459. Shoaib M. The cannabinoid antagonist AM251 attenuates nicotine self-administration and nicotine-seeking behaviour in rats. Neuropharmacology. 2008;54(2):438–444. doi: 10.1016/j.neuropharm.2007.10.011. [DOI] [PubMed] [Google Scholar]
  460. Sidhpura N, Parsons LH. Endocannabinoid-mediated synaptic plasticity and addiction-related behavior. Neuropharmacology. 2011 doi: 10.1016/j.neuropharm.2011.05.034. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  461. Silagy C, Lancaster T, Stead L, Mant D, Fowler G. Nicotine replacement therapy for smoking cessation. Cochrane Database Syst Rev. 2004;(3) doi: 10.1002/14651858.CD000146.pub2. CD000146. [DOI] [PubMed] [Google Scholar]
  462. Simon GM, Cravatt BF. Endocannabinoid biosynthesis proceeding through glycerophospho-N-acyl ethanolamine and a role for alpha/beta-hydrolase 4 in this pathway. J Biol Chem. 2006;281(36):26465–26472. doi: 10.1074/jbc.M604660200. [DOI] [PubMed] [Google Scholar]
  463. Simon GM, Cravatt BF. Anandamide biosynthesis catalyzed by the phosphodiesterase GDE1 and detection of glycerophospho-N-acyl ethanolamine precursors in mouse brain. J Biol Chem. 2008;283(14):9341–9349. doi: 10.1074/jbc.M707807200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  464. 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(2):244–253. doi: 10.1016/j.brainres.2004.08.027. [DOI] [PubMed] [Google Scholar]
  465. Sinha R, Catapano D, O’Malley S. Stress-induced craving and stress response in cocaine dependent individuals. Psychopharmacology (Berl) 1999;142(4):343–351. doi: 10.1007/s002130050898. [DOI] [PubMed] [Google Scholar]
  466. Sipe JC, Chiang K, Gerber AL, Beutler E, Cravatt BF. A missense mutation in human fatty acid amide hydrolase associated with problem drug use. Proc Natl Acad Sci U S A. 2002;99(12):8394–8399. doi: 10.1073/pnas.082235799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  467. Smith PB, Compton DR, Welch SP, Razdan RK, Mechoulam R, Martin BR. The pharmacological activity of anandamide, a putative endogenous cannabinoid, in mice. J Pharmacol Exp Ther. 1994;270(1):219–227. [PubMed] [Google Scholar]
  468. 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(2):408–419. doi: 10.1111/j.1471-4159.2006.03880.x. [DOI] [PubMed] [Google Scholar]
  469. Solinas M, Tanda G, Wertheim CE, Goldberg SR. Dopaminergic augmentation of delta-9-tetrahydrocannabinol (THC) discrimination: possible involvement of D(2)-induced formation of anandamide. Psychopharmacology (Berl) 2010;209(2):191–202. doi: 10.1007/s00213-010-1789-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  470. 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(1):93–102. doi: 10.1124/jpet.102.047928. [DOI] [PubMed] [Google Scholar]
  471. 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(11):2046–2057. doi: 10.1038/sj.npp.1300754. [DOI] [PubMed] [Google Scholar]
  472. 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(1):370–380. doi: 10.1124/jpet.106.114124. [DOI] [PubMed] [Google Scholar]
  473. 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(9):1670–1680. doi: 10.1038/sj.npp.1300707. [DOI] [PubMed] [Google Scholar]
  474. Spanagel R, Holter SM. Pharmacological validation of a new animal model of alcoholism. J Neural Transm. 2000;107(6):669–680. doi: 10.1007/s007020070068. [DOI] [PubMed] [Google Scholar]
  475. 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(3):343–350. doi: 10.1038/sj.bjp.0705932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  476. Steiner MA, Wanisch K, Monory K, Marsicano G, Borroni E, Bachli H, Holsboer F, Lutz B, Wotjak CT. Impaired cannabinoid receptor type 1 signaling interferes with stress-coping behavior in mice. Pharmacogenomics J. 2008;8(3):196–208. doi: 10.1038/sj.tpj.6500466. [DOI] [PubMed] [Google Scholar]
  477. Stella N. Endocannabinoid signaling in microglial cells. Neuropharmacology. 2009;56(Suppl 1):244–253. doi: 10.1016/j.neuropharm.2008.07.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  478. Stella N, Piomelli D. Receptor-dependent formation of endogenous cannabinoids in cortical neurons. Eur J Pharmacol. 2001;425(3):189–196. doi: 10.1016/s0014-2999(01)01182-7. [DOI] [PubMed] [Google Scholar]
  479. Stella N, Schweitzer P, Piomelli D. A second endogenous cannabinoid that modulates long-term potentiation. Nature. 1997;388(6644):773–778. doi: 10.1038/42015. [DOI] [PubMed] [Google Scholar]
  480. Stewart J. Pathways to relapse: the neurobiology of drug- and stress-induced relapse to drug-taking. J Psychiatry Neurosci. 2000;25(2):125–136. [PMC free article] [PubMed] [Google Scholar]
  481. Stewart SH, Kushner MG. Introduction to the Special Issue on “Anxiety Sensitivity and Addictive Behaviors”. Addict Behav. 2001;26(6):775–785. doi: 10.1016/s0306-4603(01)00236-2. [DOI] [PubMed] [Google Scholar]
  482. 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(1):89–97. doi: 10.1006/bbrc.1995.2437. [DOI] [PubMed] [Google Scholar]
  483. Sulzer D, Rayport S. Amphetamine and other psychostimulants reduce pH gradients in midbrain dopaminergic neurons and chromaffin granules: a mechanism of action. Neuron. 1990;5(6):797–808. doi: 10.1016/0896-6273(90)90339-h. [DOI] [PubMed] [Google Scholar]
  484. Suzuki A, Josselyn SA, Frankland PW, Masushige S, Silva AJ, Kida S. Memory reconsolidation and extinction have distinct temporal and biochemical signatures. J Neurosci. 2004;24(20):4787–4795. doi: 10.1523/JNEUROSCI.5491-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  485. Swanson CJ, Baker DA, Carson D, Worley PF, Kalivas PW. Repeated cocaine administration attenuates group I metabotropic glutamate receptor-mediated glutamate release and behavioral activation: a potential role for Homer. J Neurosci. 2001;21(22):9043–9052. doi: 10.1523/JNEUROSCI.21-22-09043.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  486. Takahashi E, Katayama M, Niimi K, Itakura C. Additive subthreshold dose effects of cannabinoid CB(1) receptor antagonist and selective serotonin reuptake inhibitor in antidepressant behavioral tests. Eur J Pharmacol. 2008;589(1-3):149–156. doi: 10.1016/j.ejphar.2008.05.020. [DOI] [PubMed] [Google Scholar]
  487. 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(5321):2048–2050. doi: 10.1126/science.276.5321.2048. [DOI] [PubMed] [Google Scholar]
  488. Tanda G, Munzar P, Goldberg SR. Self-administration behavior is maintained by the psychoactive ingredient of marijuana in squirrel monkeys. Nat Neurosci. 2000;3(11):1073–1074. doi: 10.1038/80577. [DOI] [PubMed] [Google Scholar]
  489. Terzian AL, Aguiar DC, Guimaraes FS, Moreira FA. Modulation of anxiety-like behaviour by Transient Receptor Potential Vanilloid Type 1 (TRPV1) channels located in the dorsolateral periaqueductal gray. Eur Neuropsychopharmacol. 2009;19(3):188–195. doi: 10.1016/j.euroneuro.2008.11.004. [DOI] [PubMed] [Google Scholar]
  490. 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(2):206–213. doi: 10.1016/j.bbr.2005.06.021. [DOI] [PubMed] [Google Scholar]
  491. Thomas MJ, Kalivas PW, Shaham Y. Neuroplasticity in the mesolimbic dopamine system and cocaine addiction. Br J Pharmacol. 2008;154(2):327–342. doi: 10.1038/bjp.2008.77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  492. Tidey JW, Miczek KA. Acquisition of cocaine self-administration after social stress: role of accumbens dopamine. Psychopharmacology (Berl) 1997;130(3):203–212. doi: 10.1007/s002130050230. [DOI] [PubMed] [Google Scholar]
  493. Torregrossa MM, Corlett PR, Taylor JR. Aberrant learning and memory in addiction. Neurobiol Learn Mem. 2011 doi: 10.1016/j.nlm.2011.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  494. Tournier M, Sorbara F, Gindre C, Swendsen JD, Verdoux H. Cannabis use and anxiety in daily life: a naturalistic investigation in a non-clinical population. Psychiatry Res. 2003;118(1):1–8. doi: 10.1016/s0165-1781(03)00052-0. [DOI] [PubMed] [Google Scholar]
  495. Twitchell W, Brown S, Mackie K. Cannabinoids inhibit N- and P/Q-type calcium channels in cultured rat hippocampal neurons. J Neurophysiol. 1997;78(1):43–50. doi: 10.1152/jn.1997.78.1.43. [DOI] [PubMed] [Google Scholar]
  496. Tyndale RF, Payne JI, Gerber AL, Sipe JC. The fatty acid amide hydrolase C385A (P129T) missense variant in cannabis users: studies of drug use and dependence in Caucasians. Am J Med Genet B Neuropsychiatr Genet. 2007;144B(5):660–666. doi: 10.1002/ajmg.b.30491. [DOI] [PubMed] [Google Scholar]
  497. Tzavara ET, Davis RJ, Perry KW, Li X, Salhoff C, Bymaster FP, Witkin JM, Nomikos GG. The CB1 receptor antagonist SR141716A selectively increases monoaminergic neurotransmission in the medial prefrontal cortex: implications for therapeutic actions. Br J Pharmacol. 2003;138(4):544–553. doi: 10.1038/sj.bjp.0705100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  498. Uriguen L, Perez-Rial S, Ledent C, Palomo T, Manzanares J. Impaired action of anxiolytic drugs in mice deficient in cannabinoid CB1 receptors. Neuropharmacology. 2004;46(7):966–973. doi: 10.1016/j.neuropharm.2004.01.003. [DOI] [PubMed] [Google Scholar]
  499. Uriguen L, Berrendero F, Ledent C, Maldonado R, Manzanares J. Kappa- and delta-opioid receptor functional activities are increased in the caudate putamen of cannabinoid CB1 receptor knockout mice. Eur J Neurosci. 2005;22(8):2106–2110. doi: 10.1111/j.1460-9568.2005.04372.x. [DOI] [PubMed] [Google Scholar]
  500. 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(2):564–578. doi: 10.1038/sj.bjp.0704479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  501. Valverde O, Ledent C, Beslot F, Parmentier M, Roques BP. Reduction of stress-induced analgesia but not of exogenous opioid effects in mice lacking CB1 receptors. Eur J Neurosci. 2000;12(2):533–539. doi: 10.1046/j.1460-9568.2000.00929.x. [DOI] [PubMed] [Google Scholar]
  502. 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(9):1816–1824. doi: 10.1046/j.0953-816x.2001.01558.x. [DOI] [PubMed] [Google Scholar]
  503. Valverde O, Aso E, Ozaita A, Valdizan E, Pazos A, Maldonado R. Depressive-like phenotype in mice lacking cannabinoid CB1 receptors. European Journal of Neuropscyhopharmacology. 2006;16:S302. [Google Scholar]
  504. 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(5746):329–332. doi: 10.1126/science.1115740. [DOI] [PubMed] [Google Scholar]
  505. Vann RE, Warner JA, Bushell K, Huffman JW, Martin BR, Wiley JL. Discriminative stimulus properties of delta9-tetrahydrocannabinol (THC) in C57Bl/6J mice. Eur J Pharmacol. 2009;615(1-3):102–107. doi: 10.1016/j.ejphar.2009.05.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  506. Varma N, Carlson GC, Ledent C, Alger BE. Metabotropic glutamate receptors drive the endocannabinoid system in hippocampus. J Neurosci. 2001;21(24):RC188. doi: 10.1523/JNEUROSCI.21-24-j0003.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  507. Vela G, Ruiz-Gayo M, Fuentes JA. Anandamide decreases naloxone-precipitated withdrawal signs in mice chronically treated with morphine. Neuropharmacology. 1995;34(6):665–668. doi: 10.1016/0028-3908(95)00032-2. [DOI] [PubMed] [Google Scholar]
  508. Vengeliene V, Siegmund S, Singer MV, Sinclair JD, Li TK, Spanagel R. A comparative study on alcohol-preferring rat lines: effects of deprivation and stress phases on voluntary alcohol intake. Alcohol Clin Exp Res. 2003;27(7):1048–1054. doi: 10.1097/01.ALC.0000075829.81211.0C. [DOI] [PubMed] [Google Scholar]
  509. Vigano D, Rubino T, Parolaro D. Molecular and cellular basis of cannabinoid and opioid interactions. Pharmacol Biochem Behav. 2005;81(2):360–368. doi: 10.1016/j.pbb.2005.01.021. [DOI] [PubMed] [Google Scholar]
  510. 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(7):1849–1857. doi: 10.1111/j.1460-9568.2004.03645.x. [DOI] [PubMed] [Google Scholar]
  511. Vigano D, Cascio M. Grazia, 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(6):1160–1167. doi: 10.1038/sj.npp.1300117. [DOI] [PubMed] [Google Scholar]
  512. Vinklerova J, Novakova J, Sulcova A. Inhibition of methamphetamine self-administration in rats by cannabinoid receptor antagonist AM 251. J Psychopharmacol. 2002;16(2):139–143. doi: 10.1177/026988110201600204. [DOI] [PubMed] [Google Scholar]
  513. Vinod KY, Yalamanchili R, Xie S, Cooper TB, Hungund BL. Effect of chronic ethanol exposure and its withdrawal on the endocannabinoid system. Neurochem Int. 2006;49(6):619–625. doi: 10.1016/j.neuint.2006.05.002. [DOI] [PubMed] [Google Scholar]
  514. Vinod KY, Sanguino E, Yalamanchili R, Manzanares J, Hungund BL. Manipulation of fatty acid amide hydrolase functional activity alters sensitivity and dependence to ethanol. J Neurochem. 2008;104(1):233–243. doi: 10.1111/j.1471-4159.2007.04956.x. [DOI] [PubMed] [Google Scholar]
  515. Vinod KY, Maccioni P, Garcia-Gutierrez MS, Femenia T, Xie S, Carai MA, Manzanares J, Cooper TB, Hungund BL, Colombo G. Innate difference in the endocannabinoid signaling and its modulation by alcohol consumption in alcohol-preferring sP rats. Addict Biol. 2011 doi: 10.1111/j.1369-1600.2010.00299.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  516. Viscomi MT, Oddi S, Latini L, Pasquariello N, Florenzano F, Bernardi G, Molinari M, Maccarrone M. Selective CB2 receptor agonism protects central neurons from remote axotomy-induced apoptosis through the PI3K/Akt pathway. J Neurosci. 2009;29(14):4564–4570. doi: 10.1523/JNEUROSCI.0786-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  517. Viveros MP, Marco EM, File SE. Endocannabinoid system and stress and anxiety responses. Pharmacol Biochem Behav. 2005;81(2):331–342. doi: 10.1016/j.pbb.2005.01.029. [DOI] [PubMed] [Google Scholar]
  518. Viveros MP, de Fonseca FR, Bermudez-Silva FJ, McPartland JM. Critical role of the endocannabinoid system in the regulation of food intake and energy metabolism, with phylogenetic, developmental, and pathophysiological implications. Endocr Metab Immune Disord Drug Targets. 2008;8(3):220–230. doi: 10.2174/187153008785700082. [DOI] [PubMed] [Google Scholar]
  519. Vlachou S, Nomikos GG, Panagis G. WIN 55,212-2 decreases the reinforcing actions of cocaine through CB1 cannabinoid receptor stimulation. Behav Brain Res. 2003;141(2):215–222. doi: 10.1016/s0166-4328(02)00370-4. [DOI] [PubMed] [Google Scholar]
  520. Vlachou S, Nomikos GG, Panagis G. Effects of endocannabinoid neurotransmission modulators on brain stimulation reward. Psychopharmacology (Berl) 2006;188(3):293–305. doi: 10.1007/s00213-006-0506-0. [DOI] [PubMed] [Google Scholar]
  521. Vlachou S, Stamatopoulou F, Nomikos GG, Panagis G. Enhancement of endocannabinoid neurotransmission through CB1 cannabinoid receptors counteracts the reinforcing and psychostimulant effects of cocaine. Int J Neuropsychopharmacol. 2008;11(7):905–923. doi: 10.1017/S1461145708008717. [DOI] [PubMed] [Google Scholar]
  522. Volkow ND, Fowler JS, Wang GJ, Swanson JM, Telang F. Dopamine in drug abuse and addiction: results of imaging studies and treatment implications. Arch Neurol. 2007;64(11):1575–1579. doi: 10.1001/archneur.64.11.1575. [DOI] [PubMed] [Google Scholar]
  523. Voltaire-Carlsson A, Hiltunen AJ, Koechling UM, Borg S. Effects of long-term abstinence on psychological functioning: a prospective longitudinal analysis comparing alcohol-dependent patients and healthy volunteers. Alcohol. 1996;13(5):415–421. doi: 10.1016/0741-8329(96)81678-8. [DOI] [PubMed] [Google Scholar]
  524. Wade MR, Degroot A, Nomikos GG. Cannabinoid CB1 receptor antagonism modulates plasma corticosterone in rodents. Eur J Pharmacol. 2006;551(1-3):162–167. doi: 10.1016/j.ejphar.2006.08.083. [DOI] [PubMed] [Google Scholar]
  525. 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(3):1393–1398. doi: 10.1073/pnas.0336351100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  526. Ward SJ, Rosenberg M, Dykstra LA, Walker EA. The CB1 antagonist rimonabant (SR141716) blocks cue-induced reinstatement of cocaine seeking and other context and extinction phenomena predictive of relapse. Drug Alcohol Depend. 2009;105(3):248–255. doi: 10.1016/j.drugalcdep.2009.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  527. Weiss F. Neurobiology of craving, conditioned reward and relapse. Curr Opin Pharmacol. 2005;5(1):9–19. doi: 10.1016/j.coph.2004.11.001. [DOI] [PubMed] [Google Scholar]
  528. Weitlauf C, Egli RE, Grueter BA, Winder DG. High-frequency stimulation induces ethanol-sensitive long-term potentiation at glutamatergic synapses in the dorsolateral bed nucleus of the stria terminalis. J Neurosci. 2004;24(25):5741–5747. doi: 10.1523/JNEUROSCI.1181-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  529. Werling LL, Reed SC, Wade D, Izenwasser S. Chronic nicotine alters cannabinoid-mediated locomotor activity and receptor density in periadolescent but not adult male rats. Int J Dev Neurosci. 2009;27(3):263–269. doi: 10.1016/j.ijdevneu.2008.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  530. White NM. Addictive drugs as reinforcers: multiple partial actions on memory systems. Addiction. 1996;91(7):921–949. discussion 951-965. [PubMed] [Google Scholar]
  531. Whyte LS, Ryberg E, Sims NA, Ridge SA, Mackie K, Greasley PJ, Ross RA, Rogers MJ. The putative cannabinoid receptor GPR55 affects osteoclast function in vitro and bone mass in vivo. Proc Natl Acad Sci U S A. 2009;106(38):16511–16516. doi: 10.1073/pnas.0902743106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  532. Wiley JL, Huffman JW, Balster RL, Martin BR. Pharmacological specificity of the discriminative stimulus effects of delta 9-tetrahydrocannabinol in rhesus monkeys. Drug Alcohol Depend. 1995;40(1):81–86. doi: 10.1016/0376-8716(95)01193-5. [DOI] [PubMed] [Google Scholar]
  533. 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(3):173–179. doi: 10.1037/1064-1297.12.3.173. [DOI] [PubMed] [Google Scholar]
  534. Wilson RI, Nicoll RA. Endocannabinoid signaling in the brain. Science. 2002;296(5568):678–682. doi: 10.1126/science.1063545. [DOI] [PubMed] [Google Scholar]
  535. Wise RA, Morales M. A ventral tegmental CRF-glutamate-dopamine interaction in addiction. Brain Res. 2010;1314:38–43. doi: 10.1016/j.brainres.2009.09.101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  536. Xi ZX, Stein EA. Increased mesolimbic GABA concentration blocks heroin self-administration in the rat. J Pharmacol Exp Ther. 2000;294(2):613–619. [PubMed] [Google Scholar]
  537. Xi ZX, Stein EA. GABAergic mechanisms of opiate reinforcement. Alcohol Alcohol. 2002;37(5):485–494. doi: 10.1093/alcalc/37.5.485. [DOI] [PubMed] [Google Scholar]
  538. 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(33):8531–8536. doi: 10.1523/JNEUROSCI.0726-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  539. Xi ZX, Spiller K, Pak AC, Gilbert J, Dillon C, Li X, Peng XQ, Gardner EL. Cannabinoid CB1 receptor antagonists attenuate cocaine’s rewarding effects: experiments with self-administration and brain-stimulation reward in rats. Neuropsychopharmacology. 2008;33(7):1735–1745. doi: 10.1038/sj.npp.1301552. [DOI] [PubMed] [Google Scholar]
  540. Yahyavi-Firouz-Abadi N, See RE. Anti-relapse medications: preclinical models for drug addiction treatment. Pharmacol Ther. 2009;124(2):235–247. doi: 10.1016/j.pharmthera.2009.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  541. Yin H, Chu A, Li W, Wang B, Shelton F, Otero F, Nguyen DG, Caldwell JS, Chen YA. Lipid G protein-coupled receptor ligand identification using beta-arrestin PathHunter assay. J Biol Chem. 2009;284(18):12328–12338. doi: 10.1074/jbc.M806516200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  542. Yin HH, Ostlund SB, Balleine BW. Reward-guided learning beyond dopamine in the nucleus accumbens: the integrative functions of cortico-basal ganglia networks. Eur J Neurosci. 2008;28(8):1437–1448. doi: 10.1111/j.1460-9568.2008.06422.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  543. Zhang PW, Ishiguro H, Ohtsuki T, Hess J, Carillo F, Walther D, Onaivi ES, Arinami T, Uhl GR. Human cannabinoid receptor 1: 5′ exons, candidate regulatory regions, polymorphisms, haplotypes and association with polysubstance abuse. Mol Psychiatry. 2004;9(10):916–931. doi: 10.1038/sj.mp.4001560. [DOI] [PubMed] [Google Scholar]
  544. Zuo L, Kranzler HR, Luo X, Covault J, Gelernter J. CNR1 variation modulates risk for drug and alcohol dependence. Biol Psychiatry. 2007;62(6):616–626. doi: 10.1016/j.biopsych.2006.12.004. [DOI] [PubMed] [Google Scholar]
  545. Zuo L, Kranzler HR, Luo X, Yang BZ, Weiss R, Brady K, Poling J, Farrer L, Gelernter J. Interaction between two independent CNR1 variants increases risk for cocaine dependence in European Americans: a replication study in family-based sample and population-based sample. Neuropsychopharmacology. 2009;34(6):1504–1513. doi: 10.1038/npp.2008.206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  546. Zygmunt PM, Petersson J, Andersson DA, Chuang H, Sorgard M, Di Marzo V, Julius D, Hogestatt ED. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature. 1999;400(6743):452–457. doi: 10.1038/22761. [DOI] [PubMed] [Google Scholar]

RESOURCES