Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2014 Jul 1;172(2):297–310. doi: 10.1111/bph.12618

The role of δ-opioid receptors in learning and memory underlying the development of addiction

Paul Klenowski 1, Michael Morgan 1, Selena E Bartlett 1
PMCID: PMC4292947  PMID: 24641428

Abstract

Opioids are important endogenous ligands that exist in both invertebrates and vertebrates and signal by activation of opioid receptors to produce analgesia and reward or pleasure. The μ-opioid receptor is the best known of the opioid receptors and mediates the acute analgesic effects of opiates, while the δ-opioid receptor (DOR) has been less well studied and has been linked to effects that follow from chronic use of opiates such as stress, inflammation and anxiety. Recently, DORs have been shown to play an essential role in emotions and increasing evidence points to a role in learning actions and outcomes. The process of learning and memory in addiction has been proposed to involve strengthening of specific brain circuits when a drug is paired with a context or environment. The DOR is highly expressed in the hippocampus, amygdala, striatum and other basal ganglia structures known to participate in learning and memory. In this review, we will focus on the role of the DOR and its potential role in learning and memory underlying the development of addiction.

LINKED ARTICLES

This article is part of a themed section on Opioids: New Pathways to Functional Selectivity. To view the other articles in this section visit http://dx.doi.org/10.1111/bph.2015.172.issue-2

Keywords: addiction, δ-opioid receptor, enkephalin, ethanol, learning, memory, reward

Introduction

Opioid receptors are members of the class A GPCR subfamily and include the δ-(DOR), μ-(MOR) and κ-(KOR) opioid receptor subtypes (see Rachinger-Adam et al., 2011; Feng et al., 2012; Pradhan et al., 2012; Lamberts and Traynor, 2013; nomenclature follows Alexander et al., 2013). The endogenous ligands that bind DORs are the enkephalins (Leu and Met). DOR activation leads to the activation and dissociation of Gi/Go-associated subunits and has subsequent effects on a number of effectors, including inhibition of cAMP production (Law and Loh, 1993), stimulation of signalling kinases such as ERK and src (Gutstein et al., 1997; Kramer et al., 2000), induction of β-arrestin (Cheng et al., 1998), inhibition of voltage-gated Ca2+ channels (Piros et al., 1996), the opening of inward rectifying K+ channels (North et al., 1987), and ultimately inhibition of neuronal cell activation through reduced firing or neurotransmitter release. After activation, DORs are internalized and degraded in lysosomes, desensitizing the cell to further activation by an agonist (Pradhan et al., 2009). This is unlike the MORs, which upon activation are also internalized but then recycled to the cell surface (Sternini et al., 1996; Whistler et al., 1999).

There is a significant debate about DOR pharmacology (see Table 1 for a list of DOR compounds discussed in this review) as DOR agonists have differential behavioural effects in in vivo experiments and, in vitro, DOR pharmacology changes when the receptor is co-expressed with either MOR (Sofuoglu et al., 1993; Stewart and Hammond, 1993; Jordan and Devi, 1999; Decaillot et al., 2008; van Rijn and Whistler, 2009; van Rijn et al., 2010b; 2012; Dietis et al., 2011) or KOR (Jordan and Devi, 1999; Gomes et al., 2000). Although only one gene has been cloned for DOR (Evans et al., 1992; Kieffer et al., 1992), it has been proposed that there are two DOR subtypes, DOR1 and DOR2, based on the inability of some DOR antagonists to block the effect of DOR agonists. In animal models of ethanol dependence, the DOR subtypes have opposing effects, where the administration of the DOR1 agonist, D-penicillamine(2,5)-enkephalin (DPDPE), and the DOR2 antagonist, naltriben, decreased ethanol consumption in rats and mice, while the DOR2 agonist SNC80 increased ethanol consumption (Krishnan-Sarin et al., 1995b; van Rijn and Whistler, 2009; van Rijn et al., 2010a; Nielsen et al., 2012b). Administration of the DOR1 agonist TAN-67 to either MOR-/- or DOR-/-, but not KOR-/- mice, did not change the level of ethanol consumption, providing evidence that functional MORs were required for DOR1 activity (see van Rijn et al., 2013). In addition, DORs are primarily located intracellularly (Arvidsson et al., 1995; Kalyuzhny et al., 1996; Kalyuzhny and Wessendorf, 1998; Cahill et al., 2001a; Commons et al., 2001) and are translocated to the plasma membrane of neurons in the dorsal spinal cord cell following chronic morphine treatment (Cahill et al., 2001b; Morinville et al., 2003). This increases DOR activity and produces DOR-mediated analgesia (Hack et al., 2005). It has been argued that DORs become functional upon translocation to the plasma membrane, as they are primarily localized to dense-core vesicles in axon terminals (Kalyuzhny et al., 1996; Commons et al., 2001). This review will predominantly focus on the DOR however, in some cases we will discuss experiments that involve MORs when they are related to DORs.

Table 1.

Affinity estimates of opioid ligands at MORs, DORs and KORs

Compound Species Tissue/recombinant pKi MOR pKi DOR pKi KOR Reference
ADL5747 Human Recombinant n/a 8.57 n/a Le Bourdonnec et al. (2009)
BW373U86 Rat Brain 7.82 8.74 7.47 Chang et al. (1993)
CTAP Rat Brain 2.1a 5314a n/a Kazmierski et al. (1988)
CTOP Rat Recombinant 9.74 Raynor et al. (1994)
Mouse Recombinant <6.00 <6.00 Raynor et al. (1994)
DALA Rat Brain 8.30 8.80 8.92 Maruyama et al. (1987)
DAMGO Rat Brain 8.97 7.19 Amiche et al. (1989)
Guinea pig Cerebellum <4.70 Amiche et al. (1989)
Deltorphin Rat Brain 5.79 8.62 <4.60 Kreil et al. (1989)
DPDPE Guinea pig Brain 6.61 9.02 4.92 Mosberg et al. (1987)
Naltrexone Rat Vas deferens 9.11 Carroll et al. (1988)
Mouse Recombinant 6.63 Raynor et al. (1994)
Guinea pig Cerebral cortex 9.08 Smith et al. (1989)
Naltriben Rat Recombinant 7.92 Raynor et al. (1994)
Mouse Recombinant 10.89 7.89 Raynor et al. (1994)
Naltrindole Rat Recombinant 7.20 Raynor et al. (1994)
Mouse Recombinant 10.70 7.18 Raynor et al. (1994)
NIH11082 Rat Recombinant 7.99 6.85 7.54 Traynor et al. (2005)
SB-235863 Human Recombinant 6.04 8.32 6.60 Petrillo et al. (2003)
SCN80 Rat Brain 5.12 9.38 n/a Codd et al. (2006)
SDM25N Rat Brain 5.10 8.33 5.42 McLamore et al. (2001)
SoRI9409 Rat Brain 7.29 8.66 Ananthan et al. (1999)
Guinea pig Brain 7.70 Ananthan et al. (1999)
TAN67 Guinea pig Brain 5.63 8.95 5.75 Nagase et al. (1998)
TIPPψ Rat Brain 5.49 9.51 Schiller et al. (1993)
<5 Nevin et al. (1995)

Affinity estimates from rodent and guinea pig tissue or recombinant receptors expressed in transfected cell lines are shown as pKi for MOR, DOR and KOR, except for CTAP, which is reported as IC50 denoted by a.

DORs in addiction

The DOR contributes to the rewarding/reinforcing properties of drugs of abuse by modulating presynaptic release of dopamine in the ventral tegmental area (VTA) (Pentney and Gratton, 1991) and the nucleus accumbens (NAc) (Broekkamp et al., 1979; Latimer et al., 2013; Di Chiara and Imperato, 1988a; b; Spanagel et al., 1990; 1992; Leone et al., 1991; Longoni et al., 1991; Pentney and Gratton, 1991; Devine et al., 1993; Pontieri et al., 1995; Yoshida et al., 1999; Hirose et al., 2005), two key brain regions associated with the development of addiction. While morphine operant self-administration is not directly modulated by DOR, shown using transgenic DOR-/- mice (Le Merrer et al., 2011), DORs in the VTA and NAc play a role in cocaine self-administration, where inhibiting DORs with naltrindole 5′-isothiocyanate in the NAc reduces, and in the VTA increases, cocaine self-administration (Ward and Roberts, 2007). Furthermore, DOR-/- mice have decreased morphine and nicotine-induced conditioned place preference (CPP) and lithium-induced conditioned place aversion (CPA) compared with wild-type mice, but unchanged cannabinoid CPP (Ghozland et al., 2002; Chefer and Shippenberg, 2009; Le Merrer et al., 2011; Berrendero et al., 2012). Furthermore, β-endorphin and pro-enkephalin double knockout mice have reduced CPP for ethanol, indicating that this form of learning may involve an interaction between the MORs and DORs (Tseng et al., 2013). This suggests that the DOR contributes to the acute rewarding properties of morphine, cocaine and nicotine, and that DOR plays an important role in learning paired cue associations (see Figure 1). For a review of the role of opioid receptors in models of drug abuse, see Charbogne et al. (2014).

Figure 1.

Figure 1

Activity of DORs in brain regions associated with learning and memory in addiction including the prefrontal cortex (PFC), hippocampus, NAc, amygdala and the VTA. Arrows indicate sites of neurotransmitter modulation by DORs, which have been shown to be important for drug reward and reinforcement, cue and contextual learning associated with drug use, and negative emotions such as stress and anxiety that contribute to drug relapse.

There have been many studies showing that DORs contribute to the reinforcing properties of ethanol (see Nielsen and Bartlett, 2012) and these studies have contrasting results. However, it has been shown that the administration of non-selective DOR antagonists naltriben and naltrindole reduced ethanol intake in rodents (Froehlich et al., 1991; Le et al., 1993; Krishnan-Sarin et al., 1995a; b; June et al., 1999; Nielsen et al., 2012b). The brain regions responsible for the effect of DOR on ethanol consumption have been shown to be the basolateral amygdala (BLA) (Hyytia and Kiianmaa, 2001), NAc (Hyytia and Kiianmaa, 2001; Barson et al., 2009), paraventricular nucleus (PVN) (Barson et al., 2010) and dorsal striatum (Nielsen et al., 2012b). Additionally, activation of the DOR in both the perifornical lateral hypothalamus by [D-Ala2]-methionine enkephalinamide and the VTA with DPDPE results in a decrease in ethanol consumption (Margolis et al., 2008; Chen et al., 2013). Microinfusion of a DOR selective antagonist H-Tyr-Tic[CH2NH]-Phe-Phe-OH (TIPPψ), into the VTA increases ethanol consumption in low-drinking animals, an effect blocked by the GABAA antagonist bicuculline. This suggests that activation of DOR causes inhibition of GABAergic neurons (Margolis et al., 2008). This effect is likely to be mediated by DOR1, as microinfusion of the selective DOR2 agonist deltorphin had no effect on GABAergic activity or ethanol consumption (Margolis et al., 2008; Mitchell et al., 2014).

In relapse, the DOR plays an important role in both cue and stress-induced reinstatement (Ciccocioppo et al., 2002; Marinelli et al., 2009; Nielsen et al., 2012a). Yohimbine-induced stress and long-term ethanol consumption cause increased GTPγS activity at the cell membrane, indicating a possible increase in trafficking of DOR to the cell membrane (Nielsen et al., 2012a). Brain slices taken from the yohimbine-treated rats had greater activity with TAN-67, a DOR1 selective agonist, over deltorphin, a DOR2 agonist, indicating DOR1 plays a greater role in stress-induced reinstatement (Nielsen et al., 2012a). Furthermore, systemic administration of TAN-67 and naltriben, a DOR2 selective antagonist, both reduced ethanol consumption in mice, suggesting that DOR1 and DOR2 may exert opposing effects on ethanol consumption (van Rijn and Whistler, 2009). Visualization of DOR in DOR-eGFP transgenic mice showed that DORs were located both intracellularly and at the cell surface (Scherrer et al., 2006). Long-term ethanol consumption increased DOR activity in the dorsal striatum of rats (Nielsen et al., 2012a; b), suggesting that long-term ethanol consumption facilitated increased DOR expression at the cell membrane and that the DORs were dynamically regulated by drugs of abuse such as morphine and ethanol. Together, these data indicate that DORs strengthen the association between the drug and the cue or context in which the drug is taken.

It is known that DOR plays a role in reward/reinforcement, and in addition may play a role in the regulation of negative emotions such as anxiety (see Chu Sin Chung and Kieffer, 2013). DOR-/- mice have higher ethanol consumption and anxiety levels than the corresponding wild-type mice (Filliol et al., 2000; Roberts et al., 2001) and alcohol-preferring rats exhibit a 10–20% decrease in DOR binding in the BLA and posterior hippocampus (Strother et al., 2001). Additionally, after 14 days of binge cocaine administration followed by withdrawal, DORs uncouple from their G-protein causing receptor desensitization, leading to greater anxiety and depressive behaviour (Perrine et al., 2008). These results suggest that DOR activation may provide a protective mechanism against the exacerbation of anxious behaviour. Conversely, during states of stress, including opioid withdrawal, repeated immobilization, novel stress such as a single immobilization event, intraperitoneal injection of hypotonic saline or a puff of air, increased enkephalin gene expression is observed in the PVN (Lightman and Young, 1987; Dumont et al., 2000). Furthermore, DOR-stimulated [35S]GTPγS binding is increased in the midbrain of stressed rats (Nielsen et al., 2012a). This is reversed by the DOR antagonist SoRI-9409, which also prevents stress-induced reinstatement of ethanol seeking (Xu et al., 2001; Nielsen et al., 2008; 2012a; Ananthan et al., 2012). Together, this suggests that the DOR modulates negative emotions that reinforce drug seeking in rats (Nielsen et al., 2012a). The DOR-mediated reductions in ethanol consumption appear to result from two possible mechanisms: (i) DOR activation that produces decreased anxiety and (ii) reduced stress caused by inhibition of DOR activity in brain regions associated with stress responses including the PVN and amygdala.

Learning and memory associated with drug addiction

‘Alcoholism is an excellent example of the transition from an easily modulated behavior to a habit of drunkenness’ (Edwards, 2010). The process of learning and memory when a drug is paired with a specific context or environment relies on neural connections among cortical structures, the hippocampus and the amygdala (Pennartz et al., 2011). Furthermore, additional connections between these structures, the striatum and other basal ganglia structures appear to contribute to the formation and storage of procedural memories including skill and habit learning (Pennartz et al., 2011). These brain circuits also encode memories of emotional and reward value to internal and external stimuli, which contributes to predictive and evaluative learning and drives motivational behaviours (Koob and Volkow, 2010). Because DORs are highly expressed within these brain circuits, studies are beginning to uncover their role in memory and learning (Paden et al., 1987; Blackburn et al., 1988; Erbs et al., 2012). This review will discuss how DORs and other opioid receptor subtypes affect neuronal function in learning and memory processes associated with addiction by examining the function of opioid receptor activity in the brain regions and circuits associated with learning and memory.

DORs in the hippocampus in memory formation and retrieval of spatial and contextual associations

Studies with learning protocols such as CPP, cross-maze tasks and a novel object recognition task have demonstrated that the hippocampus plays an important role in memory formation and retrieval of spatial and contextual associations (Ammassari-Teule et al., 1991; Kim and Fanselow, 1992; Packard and McGaugh, 1996; Frankland et al., 1998; Anagnostaras et al., 1999; Bast et al., 2001; Rossato et al., 2004). The expression of DORs in the hippocampus is predominantly in GABAergic neurons, shown using GFP-tagged DOR transgenic mice (Erbs et al., 2012). DORs are mainly localized presynaptically in interneurons that form afferent connections to glutamatergic principal cells (Rezai et al., 2012) and DOR-containing interneurons have been shown to project to the pyramidal neuronal dendritic layers (Svoboda et al., 1999). This suggests that activation of DORs inhibits glutamatergic pyramidal cell firing, and in doing so increases afferent signalling to the pyramidal neuronal dendrites (Svoboda et al., 1999).

Pharmacological and electrophysiological characterization of DORs has been performed in brain regions involved in learning and memory formation including the hippocampus, amygdala and striatum (Jiang and North, 1992; Simmons and Chavkin, 1996; Kang-Park et al., 2007). The physiological effects of DORs within the hippocampus are well defined. These include inhibition of presynaptic neurotransmitter release and increased excitation of pyramidal cells in CA1, CA3 and dentate gyrus regions (Masukawa and Prince, 1982; Duggan and North, 1983; Gruol et al., 1983; Wiesner et al., 1986; Wiesner and Henriksen, 1987; Neumaier et al., 1988; Pang and Rose, 1989), leading to a reduction in evoked and spontaneous GABA-mediated IPSPs (Cohen et al., 1992; Lupica et al., 1992; Lupica, 1995). Because increased pyramidal cell excitability in the hippocampus is thought to facilitate long-term potentiation (LTP), it is conceivable that DOR activation contributes to hippocampal LTP. This is supported by studies showing that the release of enkephalins in nerve fibres of the lateral perforant path and DOR induced LTP in this region (Chavkin et al., 1985; Bramham et al., 1991; Bramham and Sarvey, 1996).

DOR-mediated inhibition of hippocampal GABA release disrupts memory retrieval associated with these learning paradigms and activation of GABAA receptors in the hippocampus impairs memory acquisition and consolidation in contextual learning tasks (Ammassari-Teule et al., 1991; Kim and Fanselow, 1992; Frankland et al., 1998; Anagnostaras et al., 1999; Bast et al., 2001). It has been shown that re-exposure to a context previously paired with morphine stimulates DOR internalization in the CA1, CA3 and dentate gyrus regions of GFP-tagged DOR mice, possibly due to the release of enkephalin (Faget et al., 2012). Additionally, DOR-/- mice display impaired performance in contextual and spatial learning tasks (Le Merrer et al., 2013). One possibility is that DOR-mediated LTP in the hippocampus may be associated with the acquisition and consolidation of this form of declarative memory.

DORs in the amygdala during incentive learning and motivation

The amygdala plays an important role in emotional states including stress, anxiety and depression that contribute to addiction and drug-seeking behaviour (Koob, 2009; Koob and Volkow, 2010). Furthermore, strong evidence implicates the amygdala in incentive learning and motivational behaviours associated with the rewarding effects of addictive substances (Everitt et al., 1999; Robbins and Everitt, 2002). Input and output connections of the BLA appear to play a role in the formation of memories that encode incentive value associated with rewarding stimulus (Everitt et al., 1999; Robbins and Everitt, 2002). The greatest levels of DORs in the amygdala are located in the BLA (Paden et al., 1987). DORs are localized in MOR expressing fibres in the central nucleus of the amygdala (CeA) (Chieng et al., 2006) and are absent from glutamatergic and GABAergic synapses. Following chronic morphine and ethanol dependence, it has been shown that functional DORs are recruited to the synapse leading to inhibition of glutamatergic synaptic currents (Bie et al., 2009a; b). The recruitment of functional DORs in the CeA results from morphine-induced up-regulation of nerve growth factor (Bie et al., 2012). The predominant enkephalinergic afferents extending to the medial CeA emanate from the bed nucleus of the stria terminalis along with other regions of the amygdala (Poulin et al., 2006).

Opiate-dependent enhancement of excitatory input to the BLA involves increased dopaminergic signalling onto pyramidal cells (Li et al., 2011). This mechanism is thought to involve an up-regulation of presynaptic dopamine D1 receptors caused by chronic morphine treatment, which facilitates the excitatory effect of dopamine leading to enhanced glutamate release (Li et al., 2011). Activation of dopamine D1 and D2 receptors within the BLA also increased the activity of neurons within the NAc shell and enhanced the rewarding effects of opiates (Lintas et al., 2012).

Opioid receptor activation also modulates GABAergic transmission in the amygdala (Ford et al., 2006; 2007; Kang-Park et al., 2007; 2009; 2013). All three opioid receptor subtypes have been shown to produce inhibition of GABAergic transmission in dopamine neurons projecting from the VTA to the BLA and reduce GABAergic IPSCs in CeA (Ford et al., 2006; 2007; Kang-Park et al., 2007; 2009; 2013). It is proposed that the inhibitory effects of opioid receptor activation on GABA in the CeA cause a sensitization to the anxiolytic effects of ethanol and contribute to increased ethanol reinforcement (Kang-Park et al., 2007; 2009; 2013).

It has become increasingly recognized that exposure to environmental contexts associated with drug use can lead to relapse. The cortical-hippocampus-amygdala brain circuit is thought to be critical for the development of associations formed between a particular context or environment and a conditioned stimulus (Gruber and McDonald, 2012). The DOR antagonist naltrindole inhibits cue-induced ethanol seeking when administered centrally and reduces ethanol and morphine-induced CPP when microinfused into the CeA (Marinelli et al., 2009). This suggests that DOR activity within the CeA plays an important role in learned associations that are formed during drug-context conditioning paradigms.

Opioid receptors in the striatum in motor and habit learning

The striatum and associated basal ganglia circuitry have key roles in motor learning, controlling instrumental outcome/actions (operant learning) and the development of habitual performance of these learned behaviours (see Graybiel, 2008). In addition to the connections among the ventral striatum (VS), BLA and hippocampus that contribute to declarative memory formation (Parkinson et al., 2000; Corbit et al., 2001; Roitman et al., 2005; Balleine et al., 2007; Ito et al., 2008), results from recent neuropharmacological gene-targeted and electrophysiology studies are beginning to uncover important contributions of the striatum to procedural learning (Yin and Knowlton, 2006; Wickens et al., 2007; Yin et al., 2008; 2009; Grahn et al., 2009; Lovinger, 2010). Further, there are two subregions in the dorsal striatum: the dorsolateral striatum (DLS) and the dorsomedial striatum (DMS). The DLS is involved in habits and integrates sensorimotor information and utilizes a reinforcement process during their acquisition, while the DMS is involved in non-habitual (or goal-directed) actions and more like the executive system is involved in both cognitive and emotional or reward-related processes (Cui et al., 2013). Mechanisms that mediate long-lasting changes in synaptic plasticity over time are critical for the development of motor skill and habit learning.

DORs in the ventral striatum

The VS includes the NAc core and NAc shell and plays a key role in the rewarding effects of abused substances and goal-directed behaviours (Wise and Bozarth, 1985; Koob, 2006). Furthermore, limbic and cortical inputs to the VS contribute to motivational learning and processes that reinforce drug-seeking behaviour (Wise, 2004; Everitt and Robbins, 2005; Day et al., 2007; Da Cunha et al., 2012). The NAc has been recognized for its role in motivational learning associated with goal-directed actions and decision making that involves reward outcome (Wise, 2004; Everitt and Robbins, 2005; Day et al., 2007; Da Cunha et al., 2012). Rather than memory encoding of reward value, recent evidence suggests that the role of the NAc in these behaviours is primarily facilitated by changes in motor output that determine the motivational drive and influence instrumental performance (Stuber et al., 2011). These changes are derived from the processing of sensory input of the rewarding experience within limbic brain regions including the BLA (Mogenson et al., 1980; Shiflett and Balleine, 2010) which then project to the NAc. A shift in motivation or reward incentive is produced when the ‘predicted’ reward value is different from the ‘expected’ value of the reward, determined by internal motivation conditions and prior learning (Schultz et al., 1997; Stuber et al., 2008).

Predominately, the expression of DORs within the rat NAc is presynaptic, suggesting that their contribution to these behaviours primarily involves effects on inhibitory and dopaminergic input into the NAc (Svingos et al., 1998; 1999; Cahill et al., 2001a). Recently, however, studies with GFP-tagged DOR mice have confirmed postsynaptic expression of DORs in GABergic projection neurons and cholinergic interneurons within the NAc shell (Scherrer et al., 2006; Bertran-Gonzalez et al., 2013). The MORs are expressed mainly at the extrasynaptic membrane of dendrites of cholinergic cells and GABAergic medium-sized spiny neurons (MSNs), which form synapses that receive excitatory and GABAergic input to modulate the activity of NAc neurons (Gracy et al., 1997; Svingos et al., 1997; Wang and Pickel, 1998; Ma et al., 2012). The expression of MORs in GABA terminals of NAc cells suggests that these receptors may also modulate the presynaptic release of GABA (Svingos et al., 1997).

Recent work has shown that the MOR agonist, (D-Ala2-MePhe4-Gly(ol)5)enkephalin (DAMGO), produced inhibition of spontaneous EPSCs (sEPSCs) and spontaneous IPSCs (sIPSCs) in MSNs of the VS containing dopamine D1 and D2 receptors. These effects persisted in the presence of tetrodotoxin, and produced depolarization and enhanced intrinsic cell excitability (Ma et al., 2012).

Studies have also demonstrated the ability of MORs and DORs within the VS to modulate drug-seeking behaviour. Infusions of DOR and MOR agonists into the NAc restored cocaine seeking in rats (Simmons and Self, 2009). Also, abstinence from chronic cocaine use in animals leads to an increase in MOR expression within the NAc core, suggesting that opioid receptor signalling is dynamically regulated within the VS and may contribute to drug reward-related learning (Simmons and Self, 2009).

Recent studies have begun to investigate the effects of MORs and DORs in the NAc on learning using Pavlovian instrumental transfer and outcome devaluation tests (Laurent et al., 2012; Bertran-Gonzalez et al., 2013). These tests rely on a form of predictive learning, which is required to assign a value or validity to a stimulus which is paired with a reward. This work suggests that MORs within the NAc core contribute to motivational drive in reward seeking, as observed by reduced sensitivity to outcome devaluation tasks following infusion of the selective MOR antagonist D-Phe-Cys-Tyr-d-Trp-Arg-Thr-Pen-Thr-NH2 (CTAP) (Laurent et al., 2012). Inhibition of DORs in the NAc shell or core did not affect outcome devaluation but did reduce performance during instrumental transfer (Laurent et al., 2012). As previous studies have shown that DOR and MOR agonists promote feeding behaviour in rodents when infused into the NAc shell (Zhang and Kelley, 2000; Zhang et al., 2003; Smith and Berridge, 2007; Ambroggi et al., 2009; Katsuura and Taha, 2010) and also enhance motor reflexes associated with palatable tastes such as sucrose (Pecina and Berridge, 2005; Smith and Berridge, 2007), it was proposed that the observed MOR effects were caused by changes in evaluative learning that alters incentive value, motor output and motivational drive in reward seeking (Laurent et al., 2012). The reduction in performance mediated by DOR antagonism during instrumental transfer was attributed to a deficit in predictive learning required to guide choice based on paired stimulus-reward outcomes. This is also supported by recent work showing that Pavlovian conditioning in mice increases DOR expression within the somatic membrane of cholinergic interneurons in the NAc shell (Bertran-Gonzalez et al., 2013). This effect persisted well after the initial training phase and correlated with the level of conditioned responding. Furthermore, increased DOR expression was accompanied by higher irregular/burst firing in these neurons but no change in action potential frequency (Bertran-Gonzalez et al., 2013). Increased burst firing variability has previously been shown to produce changes in the pattern release of acetylcholine at target MSNs resulting in disinhibition (Ding et al., 2010). This mechanism of DOR-mediated synaptic plasticity may contribute to future actions that guide choice based on stimulus-reward outcomes (Bertran-Gonzalez et al., 2013).

DORs and MORs in the dorsal striatum

The striatum is the largest nucleus of the basal ganglia that is involved in motor control, reinforcement learning and action selection (Alexander and Crutcher, 1990; Graybiel et al., 1994; Redgrave and Gurney, 2006; Wickens et al., 2007). There are two subregions in the dorsal striatum: the DLS and the DMS. The DLS integrates sensorimotor information and the DMS is involved in motivation and reward (Cui et al., 2013). Furthermore, excitatory inputs to the striatum arise from cortical and thalamic structures and inhibitory outputs project to the basal ganglia output neurons (striatonigral or direct pathway) or to the striatopallidal neurons (indirect pathway) (Cui et al., 2013). There has been considerable debate about the precise roles of these two pathways in action selection, with recent evidence suggesting there is coordinated rather than separate activation of the direct and the indirect pathways (Graybiel, 2008).

The striatum plays an important role in learning and memory. Within the striatum, MSNs constitute approximately 95% of all neurons and receive convergent glutamatergic afferents from the cortex and thalamus, as well as dopaminergic afferents from the substantia nigra (reviewed in Graybiel, 2008). As the only output from the striatum, MSNs link these nuclei to the rest of the basal ganglia. The dorsal striatum also plays a role in signalling proximity and value of distant rewards, where it was shown that dopamine transients in the DLS and DMS appeared as rats navigated mazes seeking for rewards (Howe et al., 2013). It is known that endogenous opioids are widely expressed and are present in the striatum in MSNs that are grouped into two populations based on differential expression of dopamine D1 and D2 receptors and the opioid peptides they contain (Blomeley and Bracci, 2011). For example, the D1 receptor MSNs define the direct pathway and express dynorphins, and the D2 receptor MSNs contribute to the indirect pathway and express enkephalins (Gerfen, 1992; Britt and McGehee, 2008; Gertler et al., 2008). The current thinking is that the D1 pathway mediates action initiation and the D2 pathway inhibits alternative actions (Cui et al., 2013). In the mouse striatum, DOR mRNA is predominantly found in cholinergic neurons (Le Moine et al., 1994) and receptor expression is confined to the soma and proximal dendrites (Scherrer et al., 2006; Bertran-Gonzalez et al., 2013). In the rat, DORs were found in the somatodendritic profiles, axon terminals, in the cytoplasm of dendrites and at the nerve terminals of striatal neurons (Wang et al., 2003). A subset of DOR-expressing dendrites also expressed the D1 receptor, indicating that these dopamine receptors may mediate DOR function (Ambrose et al., 2006).

Recent work is beginning to uncover the potential for opioid receptor activity to affect local processing within the dorsal striatum in order to determine the functional consequences on striatal output. Experiments using antidromic stimulation of MSNs or paired recordings to facilitate the presynaptic release of endogenous opioids revealed negative effects on evoked EPSCs stimulated from cortical inputs (Blomeley and Bracci, 2011). The majority of the current studies have been on the MORs. It is hoped that these studies focused will reveal the role of the DORs in these circuits. The opioid-mediated reduction in evoked EPSC amplitude was blocked by the MOR antagonist CTAP but continued in the presence of the DOR antagonist SDM25N, suggesting that these effects are caused by activation of presynaptic MORs (Blomeley and Bracci, 2011). Postsynaptic MOR activation within cholinergic interneurons in the striatum has also been shown to cause inhibition of spontaneous cell firing (Ponterio et al., 2013).

These studies demonstrate the ability of opioids to modulate the activity between neurons within the striatum. The ability of MORs to inhibit the cortical-induced excitation of neighbouring MSNs is thought to involve the release of enkephalin from striatopallidal MSNs (Blomeley and Bracci, 2011). Local inhibition may also involve MOR-mediated reductions in cholinergic interneuron excitability (Ponterio et al., 2013). The consequence of this activity on output nuclei remains to be determined. Although the dorsal striatum is not implicated in the reinforcing effects of abused substances, the recent work has demonstrated the importance of corticostriatal circuits in the shift between goal-directed and habitual learning (Gremel and Costa, 2013). The contribution of these processes to drug addiction is receiving greater attention considering that addicts exhibit compulsive drug-seeking habits (Everitt et al., 2008). The shift between goal-directed and habitual actions is thought to depend on the transfer of information from the DMS to the DLS. Increased activation of striatal presynaptic MORs in opioid addicts and other substance abusers that modulate endogenous opioid activity (i.e. cocaine) produce changes in receptor density and presumably effect receptor efficacy and signalling. This may alter dynamic processes between MSNs and disrupt striatal information processing and/or transfer, resulting in enhanced storage of procedural and contextual memories associated with drug use.

Pharmacotherapeutics

The early DOR agonists were the peptides DPDPE and deltorphin (Mosberg et al., 1983; Kreil et al., 1989; Kramer et al., 1993). The non-peptide compounds BW373U86 and SNC80, followed by TAN-67, were synthesized later and shown to be highly selective for DOR (Calderon et al., 1994; 1997; Fujii et al., 2001). DOR agonists are thought to be effective analgesics with reduced side effects such as addiction liability and respiratory depression, at least in preclinical animal models (Ling et al., 1985; O'Neill et al., 1997; Gallantine and Meert, 2005). A number of other DOR agonists, including but not limited to ADL5747, SB-235863 and NIH 1108, have undergone preclinical testing, largely focused on the treatment of chronic pain, anxiolytic and antidepressant effects (Petrillo et al., 2003; Aceto et al., 2007; Saitoh et al., 2011; 2013; Nozaki et al., 2012). The DOR antagonist naltrindole has been shown to reduce the behavioural and biochemical signs of morphine withdrawal following chronic morphine administration and reduce ethanol self-administration in animals (Krishnan-Sarin et al., 1995b; Suzuki et al., 1995; Hyytia and Kiianmaa, 2001; Aceto et al., 2007; Nielsen et al., 2008). Although there have been no DOR-targeted drugs approved by the FDA to date, the drug class is still at an early stage in their development, and a number of compounds hold promise for further development (Fujii et al., 2013).

Conclusions

DORs have been known for many decades and have been extensively reviewed; however, their role in regulating emotions and learning has put these receptors into a new light for developing novel pharmacotherapeutics. It may be possible that activation of these receptors plays a role in the storage of reward-related memories during drug taking. The data suggest that developing DOR compounds may be useful for reducing the drug cue and/or context learning during the development of addiction. The ability of drugs that modulate DOR activity to reduce the rewarding properties of addictive substances, alleviate symptoms associated with drug withdrawal and attenuate relapse to drug seeking may represent a novel therapeutic strategy for the management of drug addiction.

Acknowledgments

This work was supported by funding from the Australian Research Council to S. E. B.

Glossary

Abbreviations

BLA

basolateral amygdala

CeA

central nucleus of the amygdala

CPA

conditioned place aversion

CPP

conditioned place preference

DLS

dorsolateral striatum

DMS

dorsomedial striatum

DOR

δ-opioid receptor

DPDE

D-penicillamine(2,5)-enkephalin

KOR

κ-opioid receptor

LTP

long-term potentiation

MOR

μ-opioid receptor

MSN

medium-sized spiny neuron

NAc

nucleus accumbens

PVN

paraventricular nucleus

TIPPψ

H-Tyr-Tic[CH2NH]-Phe-Phe-OH

VS

ventral striatum

VTA

ventral tegmental area

Conflict of interest

None.

References

  1. Aceto MD, May EL, Harris LS, Bowman ER, Cook CD. Pharmacological studies with a nonpeptidic, delta-opioid (-)-(1R,5R,9R)-5,9-dimethyl-2′-hydroxy-2-(6-hydroxyhexyl)-6,7-benzomorphan hydrochloride ((-)-NIH 11082) Eur J Pharmacol. 2007;566:88–93. doi: 10.1016/j.ejphar.2007.03.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL, Spedding M, et al. The Concise Guide to PHARMACOLOGY 2013/14: G Protein-Coupled Receptors. Br J Pharmacol. 2013;170:1459–1581. doi: 10.1111/bph.12445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Alexander GE, Crutcher MD. Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci. 1990;13:266–271. doi: 10.1016/0166-2236(90)90107-l. [DOI] [PubMed] [Google Scholar]
  4. Ambroggi F, Turiault M, Milet A, Deroche-Gamonet V, Parnaudeau S, Balado E, et al. Stress and addiction: glucocorticoid receptor in DORaminoceptive neurons facilitates cocaine seeking. Nat Neurosci. 2009;12:247–249. doi: 10.1038/nn.2282. [DOI] [PubMed] [Google Scholar]
  5. Ambrose LM, Gallagher SM, Unterwald EM, Van Bockstaele EJ. DORamine-D1 and delta-opioid receptors co-exist in rat striatal neurons. Neurosci Lett. 2006;399:191–196. doi: 10.1016/j.neulet.2006.02.027. [DOI] [PubMed] [Google Scholar]
  6. Amiche M, Sagan S, Mor A, Delfour A, Nicolas P. Dermenkephalin (Tyr-D-Met-Phe-His-Leu-Met-Asp-NH2): a potent and fully specific agonist for the delta opioid receptor. Mol Pharmacol. 1989;35:774–779. [PubMed] [Google Scholar]
  7. Ammassari-Teule M, Pavone F, Castellano C, McGaugh JL. Amygdala and dorsal hippocampus lesions block the effects of GABAergic drugs on memory storage. Brain Res. 1991;551:104–109. doi: 10.1016/0006-8993(91)90919-m. [DOI] [PubMed] [Google Scholar]
  8. Anagnostaras SG, Maren S, Fanselow MS. Temporally graded retrograde amnesia of contextual fear after hippocampal damage in rats: within-subjects examination. J Neurosci. 1999;19:1106–1114. doi: 10.1523/JNEUROSCI.19-03-01106.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Ananthan S, Kezar HS, 3rd, Carter RL, Saini SK, Rice KC, Wells JL, et al. Synthesis, opioid receptor binding, and biological activities of naltrexone-derived pyrido- and pyrimidomorphinans. J Med Chem. 1999;42:3527–3538. doi: 10.1021/jm990039i. [DOI] [PubMed] [Google Scholar]
  10. Ananthan S, Saini SK, Dersch CM, Xu H, McGlinchey N, Giuvelis D, et al. 14-Alkoxy- and 14-acyloxypyridomorphinans: mu agonist/delta antagonist opioid analgesics with diminished tolerance and dependence side effects. J Med Chem. 2012;55:8350–8363. doi: 10.1021/jm300686p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Arvidsson U, Dado RJ, Riedl M, Lee JH, Law PY, Loh HH, et al. Delta-opioid receptor immunoreactivity: distribution in brainstem and spinal cord, and relationship to biogenic amines and enkephalin. J Neurosci. 1995;15:1215–1235. doi: 10.1523/JNEUROSCI.15-02-01215.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Balleine BW, Delgado MR, Hikosaka O. The role of the dorsal striatum in reward and decision-making. J Neurosci. 2007;27:8161–8165. doi: 10.1523/JNEUROSCI.1554-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Barson JR, Carr AJ, Soun JE, Sobhani NC, Leibowitz SF, Hoebel BG. Opioids in the nucleus accumbens stimulate ethanol intake. Physiol Behav. 2009;98:453–459. doi: 10.1016/j.physbeh.2009.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Barson JR, Carr AJ, Soun JE, Sobhani NC, Rada P, Leibowitz SF, et al. Opioids in the hypothalamic paraventricular nucleus stimulate ethanol intake. Alcohol Clin Exp Res. 2010;34:214–222. doi: 10.1111/j.1530-0277.2009.01084.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bast T, Zhang WN, Feldon J. Hippocampus and classical fear conditioning. Hippocampus. 2001;11:828–831. doi: 10.1002/hipo.1098. [DOI] [PubMed] [Google Scholar]
  16. Berrendero F, Plaza-Zabala A, Galeote L, Flores A, Bura SA, Kieffer BL, et al. Influence of delta-opioid receptors in the behavioral effects of nicotine. Neuropsychopharmacology. 2012;37:2332–2344. doi: 10.1038/npp.2012.88. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bertran-Gonzalez J, Laurent V, Chieng BC, Christie MJ, Balleine BW. Learning-related translocation of delta-opioid receptors on ventral striatal cholinergic interneurons mediates choice between goal-directed actions. J Neurosci. 2013;33:16060–16071. doi: 10.1523/JNEUROSCI.1927-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bie B, Zhu W, Pan ZZ. Ethanol-induced delta-opioid receptor modulation of glutamate synaptic transmission and conditioned place preference in central amygdala. Neuroscience. 2009a;160:348–358. doi: 10.1016/j.neuroscience.2009.02.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Bie B, Zhu W, Pan ZZ. Rewarding morphine-induced synaptic function of delta-opioid receptors on central glutamate synapses. J Pharmacol Exp Ther. 2009b;329:290–296. doi: 10.1124/jpet.108.148908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Bie B, Wang Y, Cai YQ, Zhang Z, Hou YY, Pan ZZ. Upregulation of nerve growth factor in central amygdala increases sensitivity to opioid reward. Neuropsychopharmacology. 2012;37:2780–2788. doi: 10.1038/npp.2012.144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Blackburn TP, Cross AJ, Hille C, Slater P. Autoradiographic localization of delta opiate receptors in rat and human brain. Neuroscience. 1988;27:497–506. doi: 10.1016/0306-4522(88)90283-7. [DOI] [PubMed] [Google Scholar]
  22. Blomeley CP, Bracci E. Opioidergic interactions between striatal projection neurons. J Neurosci. 2011;31:13346–13356. doi: 10.1523/JNEUROSCI.1775-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Bramham CR, Sarvey JM. Endogenous activation of mu and delta-1 opioid receptors is required for long-term potentiation induction in the lateral perforant path: dependence on GABAergic inhibition. J Neurosci. 1996;16:8123–8131. doi: 10.1523/JNEUROSCI.16-24-08123.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bramham CR, Milgram NW, Srebro B. Delta opioid receptor activation is required to induce LTP of synaptic transmission in the lateral perforant path in vivo. Brain Res. 1991;567:42–50. doi: 10.1016/0006-8993(91)91433-2. [DOI] [PubMed] [Google Scholar]
  25. Britt JP, McGehee DS. Presynaptic opioid and nicotinic receptor modulation of DORamine overflow in the nucleus accumbens. J Neurosci. 2008;28:1672–1681. doi: 10.1523/JNEUROSCI.4275-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Broekkamp CL, Phillips AG, Cools AR. Stimulant effects of enkephalin microinjection into the DORaminergic A10 area. Nature. 1979;278:560–562. doi: 10.1038/278560a0. [DOI] [PubMed] [Google Scholar]
  27. Cahill CM, McClellan KA, Morinville A, Hoffert C, Hubatsch D, O'Donnell D, et al. Immunohistochemical distribution of delta opioid receptors in the rat central nervous system: evidence for somatodendritic labeling and antigen-specific cellular compartmentalization. J Comp Neurol. 2001a;440:65–84. doi: 10.1002/cne.1370. [DOI] [PubMed] [Google Scholar]
  28. Cahill CM, Morinville A, Lee MC, Vincent JP, Collier B, Beaudet A. Prolonged morphine treatment targets delta opioid receptors to neuronal plasma membranes and enhances delta-mediated antinociception. J Neurosci. 2001b;21:7598–7607. doi: 10.1523/JNEUROSCI.21-19-07598.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Calderon SN, Rothman RB, Porreca F, Flippen-Anderson JL, McNutt RW, Xu H, et al. Probes for narcotic receptor mediated phenomena. 19. Synthesis of (+)-4-[(alpha R)-alpha-((2S,5R)-4-allyl-2,5-dimethyl-1-piperazinyl)-3- methoxybenzyl]-N,N-diethylbenzamide (SNC 80): a highly selective, nonpeptide delta opioid receptor agonist. J Med Chem. 1994;37:2125–2128. doi: 10.1021/jm00040a002. [DOI] [PubMed] [Google Scholar]
  30. Calderon SN, Rice KC, Rothman RB, Porreca F, Flippen-Anderson JL, Kayakiri H, et al. Probes for narcotic receptor mediated phenomena. 23. Synthesis, opioid receptor binding, and bioassay of the highly selective delta agonist (+)-4-[(alpha R)-alpha-((2S,5R)-4-Allyl-2,5-dimethyl-1-piperazinyl)-3-methoxybenzyl]- N,N-diethylbenzamide (SNC 80) and related novel nonpeptide delta opioid receptor ligands. J Med Chem. 1997;40:695–704. doi: 10.1021/jm960319n. [DOI] [PubMed] [Google Scholar]
  31. Carroll JA, Shaw JS, Wickenden AD. The physiological relevance of low agonist affinity binding at opioid mu-receptors. Br J Pharmacol. 1988;94:625–631. doi: 10.1111/j.1476-5381.1988.tb11569.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Chang KJ, Rigdon GC, Howard JL, McNutt RW. A novel, potent and selective nonpeptidic delta opioid receptor agonist BW373U86. J Pharmacol Exp Ther. 1993;267:852–857. [PubMed] [Google Scholar]
  33. Charbogne P, Kieffer BL, Befort K. 15 years of genetic approaches in vivo for addiction research: opioid receptor and peptide gene knockout in mouse models of drug abuse. Neuropharmacology. 2014;76:204–217. doi: 10.1016/j.neuropharm.2013.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chavkin C, Shoemaker WJ, McGinty JF, Bayon A, Bloom FE. Characterization of the prodynorphin and proenkephalin neuropeptide systems in rat hippocampus. J Neurosci. 1985;5:808–816. doi: 10.1523/JNEUROSCI.05-03-00808.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Chefer VI, Shippenberg TS. Augmentation of morphine-induced sensitization but reduction in morphine tolerance and reward in delta-opioid receptor knockout mice. Neuropsychopharmacology. 2009;34:887–898. doi: 10.1038/npp.2008.128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Chen YW, Barson JR, Chen A, Hoebel BG, Leibowitz SF. Opioids in the perifornical lateral hypothalamus suppress ethanol drinking. Alcohol. 2013;47:31–38. doi: 10.1016/j.alcohol.2012.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Cheng ZJ, Yu QM, Wu YL, Ma L, Pei G. Selective interference of beta-arrestin 1 with kappa and delta but not mu opioid receptor/G protein coupling. J Biol Chem. 1998;273:24328–24333. doi: 10.1074/jbc.273.38.24328. [DOI] [PubMed] [Google Scholar]
  38. Chieng BC, Christie MJ, Osborne PB. Characterization of neurons in the rat central nucleus of the amygdala: cellular physiology, morphology, and opioid sensitivity. J Comp Neurol. 2006;497:910–927. doi: 10.1002/cne.21025. [DOI] [PubMed] [Google Scholar]
  39. Chu Sin Chung P, Kieffer BL. Delta opioid receptors in brain function and diseases. Pharmacol Ther. 2013;140:112–120. doi: 10.1016/j.pharmthera.2013.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ciccocioppo R, Martin-Fardon R, Weiss F. Effect of selective blockade of mu(1) or delta opioid receptors on reinstatement of alcohol-seeking behavior by drug-associated stimuli in rats. Neuropsychopharmacology. 2002;27:391–399. doi: 10.1016/S0893-133X(02)00302-0. [DOI] [PubMed] [Google Scholar]
  41. Codd EE, Carson JR, Colburn RW, Dax SL, Desai-Krieger D, Martinez RP, et al. The novel, orally active, delta opioid RWJ-394674 is biotransformed to the potent mu opioid RWJ-413216. J Pharmacol Exp Ther. 2006;318:1273–1279. doi: 10.1124/jpet.106.104208. [DOI] [PubMed] [Google Scholar]
  42. Cohen GA, Doze VA, Madison DV. Opioid inhibition of GABA release from presynaptic terminals of rat hippocampal interneurons. Neuron. 1992;9:325–335. doi: 10.1016/0896-6273(92)90171-9. [DOI] [PubMed] [Google Scholar]
  43. Commons KG, Beck SG, Rudoy C, Van Bockstaele EJ. Anatomical evidence for presynaptic modulation by the delta opioid receptor in the ventrolateral periaqueductal gray of the rat. J Comp Neurol. 2001;430:200–208. [PubMed] [Google Scholar]
  44. Corbit LH, Muir JL, Balleine BW. The role of the nucleus accumbens in instrumental conditioning: evidence of a functional dissociation between accumbens core and shell. J Neurosci. 2001;21:3251–3260. doi: 10.1523/JNEUROSCI.21-09-03251.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Cui G, Jun SB, Jin X, Pham MD, Vogel SS, Lovinger DM, et al. Concurrent activation of striatal direct and indirect pathways during action initiation. Nature. 2013;494:238–242. doi: 10.1038/nature11846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Da Cunha C, Gomez AA, Blaha CD. The role of the basal ganglia in motivated behavior. Rev Neurosci. 2012;23:747–767. doi: 10.1515/revneuro-2012-0063. [DOI] [PubMed] [Google Scholar]
  47. Day JJ, Roitman MF, Wightman RM, Carelli RM. Associative learning mediates dynamic shifts in DORamine signaling in the nucleus accumbens. Nat Neurosci. 2007;10:1020–1028. doi: 10.1038/nn1923. [DOI] [PubMed] [Google Scholar]
  48. Decaillot FM, Rozenfeld R, Gupta A, Devi LA. Cell surface targeting of mu-delta opioid receptor heterodimers by RTP4. Proc Natl Acad Sci U S A. 2008;105:16045–16050. doi: 10.1073/pnas.0804106105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Devine DP, Leone P, Pocock D, Wise RA. Differential involvement of ventral tegmental mu, delta and kappa opioid receptors in modulation of basal mesolimbic DORamine release: in vivo microdialysis studies. J Pharmacol Exp Ther. 1993;266:1236–1246. [PubMed] [Google Scholar]
  50. Di Chiara G, Imperato A. Drugs abused by humans preferentially increase synaptic DORamine concentrations in the mesolimbic system of freely moving rats. Proc Natl Acad Sci U S A. 1988a;85:5274–5278. doi: 10.1073/pnas.85.14.5274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Di Chiara G, Imperato A. Opposite effects of mu and kappa opiate agonists on DORamine release in the nucleus accumbens and in the dorsal caudate of freely moving rats. J Pharmacol Exp Ther. 1988b;244:1067–1080. [PubMed] [Google Scholar]
  52. Dietis N, Rowbotham DJ, Lambert DG. Opioid receptor subtypes: fact or artifact? Br J Anaesth. 2011;107:8–18. doi: 10.1093/bja/aer115. [DOI] [PubMed] [Google Scholar]
  53. Ding JB, Guzman JN, Peterson JD, Goldberg JA, Surmeier DJ. Thalamic gating of corticostriatal signaling by cholinergic interneurons. Neuron. 2010;67:294–307. doi: 10.1016/j.neuron.2010.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Duggan AW, North RA. Electrophysiology of opioids. Pharmacol Rev. 1983;35:219–281. [PubMed] [Google Scholar]
  55. Dumont EC, Kinkead R, Trottier JF, Gosselin I, Drolet G. Effect of chronic psychogenic stress exposure on enkephalin neuronal activity and expression in the rat hypothalamic paraventricular nucleus. J Neurochem. 2000;75:2200–2211. doi: 10.1046/j.1471-4159.2000.0752200.x. [DOI] [PubMed] [Google Scholar]
  56. Edwards G. The trouble with drink: why ideas matter. Addiction. 2010;105:797–804. doi: 10.1111/j.1360-0443.2009.02773.x. [DOI] [PubMed] [Google Scholar]
  57. Erbs E, Faget L, Scherrer G, Kessler P, Hentsch D, Vonesch JL, et al. Distribution of delta opioid receptor-expressing neurons in the mouse hippocampus. Neuroscience. 2012;221:203–213. doi: 10.1016/j.neuroscience.2012.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Evans CJ, Keith DE, Jr, Morrison H, Magendzo K, Edwards RH. Cloning of a delta opioid receptor by functional expression. Science. 1992;258:1952–1955. doi: 10.1126/science.1335167. [DOI] [PubMed] [Google Scholar]
  59. Everitt BJ, Robbins TW. Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nat Neurosci. 2005;8:1481–1489. doi: 10.1038/nn1579. [DOI] [PubMed] [Google Scholar]
  60. 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]
  61. Everitt BJ, Belin D, Economidou D, Pelloux Y, Dalley JW, Robbins TW. Review. Neural mechanisms underlying the vulnerability to develop compulsive drug-seeking habits and addiction. Philos Trans R Soc Lond B Biol Sci. 2008;363:3125–3135. doi: 10.1098/rstb.2008.0089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Faget L, Erbs E, Le Merrer J, Scherrer G, Matifas A, Benturquia N, et al. In vivo visualization of delta opioid receptors upon physiological activation uncovers a distinct internalization profile. J Neurosci. 2012;32:7301–7310. doi: 10.1523/JNEUROSCI.0185-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Feng Y, He X, Yang Y, Chao D, Lazarus LH, Xia Y. Current research on opioid receptor function. Curr Drug Targets. 2012;13:230–246. doi: 10.2174/138945012799201612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Filliol D, Ghozland S, Chluba J, Martin M, Matthes HW, Simonin F, et al. Mice deficient for delta- and mu-opioid receptors exhibit opposing alterations of emotional responses. Nat Genet. 2000;25:195–200. doi: 10.1038/76061. [DOI] [PubMed] [Google Scholar]
  65. Ford CP, Mark GP, Williams JT. Properties and opioid inhibition of mesolimbic DORamine neurons vary according to target location. J Neurosci. 2006;26:2788–2797. doi: 10.1523/JNEUROSCI.4331-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Ford CP, Beckstead MJ, Williams JT. Kappa opioid inhibition of somatodendritic DORamine inhibitory postsynaptic currents. J Neurophysiol. 2007;97:883–891. doi: 10.1152/jn.00963.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Frankland PW, Cestari V, Filipkowski RK, McDonald RJ, Silva AJ. The dorsal hippocampus is essential for context discrimination but not for contextual conditioning. Behav Neurosci. 1998;112:863–874. doi: 10.1037//0735-7044.112.4.863. [DOI] [PubMed] [Google Scholar]
  68. Froehlich JC, Zweifel M, Harts J, Lumeng L, Li TK. Importance of delta opioid receptors in maintaining high alcohol drinking. Psychopharmacology (Berl) 1991;103:467–472. doi: 10.1007/BF02244246. [DOI] [PubMed] [Google Scholar]
  69. Fujii H, Kawai K, Kawamura K, Mizusuna A, Onoda Y, Murachi M, et al. Synthesis of optically active TAN-67, a highly selective delta opioid receptor agonist, and investigation of its pharmacological properties. Drug Des Discov. 2001;17:325–330. [PubMed] [Google Scholar]
  70. Fujii H, Takahashi T, Nagase H. Non-peptidic delta opioid receptor agonists and antagonists (2000–2012) Expert Opin Ther Pat. 2013;23:1181–1208. doi: 10.1517/13543776.2013.804066. [DOI] [PubMed] [Google Scholar]
  71. Gallantine EL, Meert TF. A comparison of the antinociceptive and adverse effects of the mu-opioid agonist morphine and the delta-opioid agonist SNC80. Basic Clin Pharmacol Toxicol. 2005;97:39–51. doi: 10.1111/j.1742-7843.2005.pto_97107.x. [DOI] [PubMed] [Google Scholar]
  72. Gerfen CR. The neostriatal mosaic: multiple levels of compartmental organization in the basal ganglia. Annu Rev Neurosci. 1992;15:285–320. doi: 10.1146/annurev.ne.15.030192.001441. [DOI] [PubMed] [Google Scholar]
  73. Gertler TS, Chan CS, Surmeier DJ. Dichotomous anatomical properties of adult striatal medium spiny neurons. J Neurosci. 2008;28:10814–10824. doi: 10.1523/JNEUROSCI.2660-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Ghozland S, Matthes HW, Simonin F, Filliol D, Kieffer BL, Maldonado R. Motivational effects of cannabinoids are mediated by mu-opioid and kappa-opioid receptors. J Neurosci. 2002;22:1146–1154. doi: 10.1523/JNEUROSCI.22-03-01146.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Gomes I, Jordan BA, Gupta A, Trapaidze N, Nagy V, Devi LA. Heterodimerization of mu and delta opioid receptors: a role in opiate synergy. J Neurosci. 2000;20:RC110. doi: 10.1523/JNEUROSCI.20-22-j0007.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Gracy KN, Svingos AL, Pickel VM. Dual ultrastructural localization of mu-opioid receptors and NMDA-type glutamate receptors in the shell of the rat nucleus accumbens. J Neurosci. 1997;17:4839–4848. doi: 10.1523/JNEUROSCI.17-12-04839.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Grahn JA, Parkinson JA, Owen AM. The role of the basal ganglia in learning and memory: neuropsychological studies. Behav Brain Res. 2009;199:53–60. doi: 10.1016/j.bbr.2008.11.020. [DOI] [PubMed] [Google Scholar]
  78. Graybiel AM. Habits, rituals, and the evaluative brain. Annu Rev Neurosci. 2008;31:359–387. doi: 10.1146/annurev.neuro.29.051605.112851. [DOI] [PubMed] [Google Scholar]
  79. Graybiel AM, Aosaki T, Flaherty AW, Kimura M. The basal ganglia and adaptive motor control. Science. 1994;265:1826–1831. doi: 10.1126/science.8091209. [DOI] [PubMed] [Google Scholar]
  80. Gremel CM, Costa RM. Orbitofrontal and striatal circuits dynamically encode the shift between goal-directed and habitual actions. Nat Commun. 2013;4:2264. doi: 10.1038/ncomms3264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Gruber AJ, McDonald RJ. Context, emotion, and the strategic pursuit of goals: interactions among multiple brain systems controlling motivated behavior. Front Behav Neurosci. 2012;6:50. doi: 10.3389/fnbeh.2012.00050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Gruol DL, Chavkin C, Valentino RJ, Siggins GR. Dynorphin-A alters the excitability of pyramidal neurons of the rat hippocampus in vitro. Life Sci. 1983;33(Suppl. 1):533–536. doi: 10.1016/0024-3205(83)90558-1. [DOI] [PubMed] [Google Scholar]
  83. Gutstein HB, Rubie EA, Mansour A, Akil H, Woodgett JR. Opioid effects on mitogen-activated protein kinase signaling cascades. Anesthesiology. 1997;87:1118–1126. doi: 10.1097/00000542-199711000-00016. [DOI] [PubMed] [Google Scholar]
  84. Hack SP, Bagley EE, Chieng BC, Christie MJ. Induction of delta-opioid receptor function in the midbrain after chronic morphine treatment. J Neurosci. 2005;25:3192–3198. doi: 10.1523/JNEUROSCI.4585-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Hirose N, Murakawa K, Takada K, Oi Y, Suzuki T, Nagase H, et al. Interactions among mu- and delta-opioid receptors, especially putative delta1- and delta2-opioid receptors, promote DORamine release in the nucleus accumbens. Neuroscience. 2005;135:213–225. doi: 10.1016/j.neuroscience.2005.03.065. [DOI] [PubMed] [Google Scholar]
  86. Howe MW, Tierney PL, Sandberg SG, Phillips PE, Graybiel AM. Prolonged DORamine signalling in striatum signals proximity and value of distant rewards. Nature. 2013;500:575–579. doi: 10.1038/nature12475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Hyytia P, Kiianmaa K. Suppression of ethanol responding by centrally administered CTOP and naltrindole in AA and Wistar rats. Alcohol Clin Exp Res. 2001;25:25–33. doi: 10.1111/j.1530-0277.2001.tb02123.x. [DOI] [PubMed] [Google Scholar]
  88. Ito R, Robbins TW, Pennartz CM, Everitt BJ. Functional interaction between the hippocampus and nucleus accumbens shell is necessary for the acquisition of appetitive spatial context conditioning. J Neurosci. 2008;28:6950–6959. doi: 10.1523/JNEUROSCI.1615-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Jiang ZG, North RA. Pre- and postsynaptic inhibition by opioids in rat striatum. J Neurosci. 1992;12:356–361. doi: 10.1523/JNEUROSCI.12-01-00356.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Jordan BA, Devi LA. G-protein-coupled receptor heterodimerization modulates receptor function. Nature. 1999;399:697–700. doi: 10.1038/21441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. June HL, McCane SR, Zink RW, Portoghese PS, Li TK, Froehlich JC. The delta 2-opioid receptor antagonist naltriben reduces motivated responding for ethanol. Psychopharmacology (Berl) 1999;147:81–89. doi: 10.1007/s002130051145. [DOI] [PubMed] [Google Scholar]
  92. Kalyuzhny AE, Wessendorf MW. Relationship of mu- and delta-opioid receptors to GABAergic neurons in the central nervous system, including antinociceptive brainstem circuits. J Comp Neurol. 1998;392:528–547. [PubMed] [Google Scholar]
  93. Kalyuzhny AE, Arvidsson U, Wu W, Wessendorf MW. Mu-opioid and delta-opioid receptors are expressed in brainstem antinociceptive circuits: studies using immunocytochemistry and retrograde tract-tracing. J Neurosci. 1996;16:6490–6503. doi: 10.1523/JNEUROSCI.16-20-06490.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Kang-Park M, Kieffer BL, Roberts AJ, Siggins GR, Moore SD. Kappa-opioid receptors in the central amygdala regulate ethanol actions at presynaptic GABAergic sites. J Pharmacol Exp Ther. 2013;346:130–137. doi: 10.1124/jpet.112.202903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Kang-Park MH, Kieffer BL, Roberts AJ, Siggins GR, Moore SD. Presynaptic delta opioid receptors regulate ethanol actions in central amygdala. J Pharmacol Exp Ther. 2007;320:917–925. doi: 10.1124/jpet.106.112722. [DOI] [PubMed] [Google Scholar]
  96. Kang-Park MH, Kieffer BL, Roberts AJ, Roberto M, Madamba SG, Siggins GR, et al. Mu-opioid receptors selectively regulate basal inhibitory transmission in the central amygdala: lack of ethanol interactions. J Pharmacol Exp Ther. 2009;328:284–293. doi: 10.1124/jpet.108.140749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Katsuura Y, Taha SA. Modulation of feeding and locomotion through mu and delta opioid receptor signaling in the nucleus accumbens. Neuropeptides. 2010;44:225–232. doi: 10.1016/j.npep.2009.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Kazmierski W, Wire WS, Lui GK, Knapp RJ, Shook JE, Burks TF, et al. Design and synthesis of somatostatin analogues with topographical properties that lead to highly potent and specific mu opioid receptor antagonists with greatly reduced binding at somatostatin receptors. J Med Chem. 1988;31:2170–2177. doi: 10.1021/jm00119a019. [DOI] [PubMed] [Google Scholar]
  99. Kieffer BL, Befort K, Gaveriaux-Ruff C, Hirth CG. The delta-opioid receptor: isolation of a cDNA by expression cloning and pharmacological characterization. Proc Natl Acad Sci U S A. 1992;89:12048–12052. doi: 10.1073/pnas.89.24.12048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Kim JJ, Fanselow MS. Modality-specific retrograde amnesia of fear. Science. 1992;256:675–677. doi: 10.1126/science.1585183. [DOI] [PubMed] [Google Scholar]
  101. Koob GF. The neurobiology of addiction: a neuroadaptational view relevant for diagnosis. Addiction. 2006;101(Suppl. 1):23–30. doi: 10.1111/j.1360-0443.2006.01586.x. [DOI] [PubMed] [Google Scholar]
  102. Koob GF. Dynamics of neuronal circuits in addiction: reward, antireward, and emotional memory. Pharmacopsychiatry. 2009;42(Suppl. 1):S32–S41. doi: 10.1055/s-0029-1216356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Koob GF, Volkow ND. Neurocircuitry of addiction. Neuropsychopharmacology. 2010;35:217–238. doi: 10.1038/npp.2009.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Kramer HK, Andria ML, Esposito DH, Simon EJ. Tyrosine phosphorylation of the delta-opioid receptor. Evidence for its role in mitogen-activated protein kinase activation and receptor internalization. Biochem Pharmacol. 2000;60:781–792. doi: 10.1016/s0006-2952(00)00400-7. [DOI] [PubMed] [Google Scholar]
  105. Kramer TH, Davis P, Hruby VJ, Burks TF, Porreca F. In vitro potency, affinity and agonist efficacy of highly selective delta opioid receptor ligands. J Pharmacol Exp Ther. 1993;266:577–584. [PubMed] [Google Scholar]
  106. Kreil G, Barra D, Simmaco M, Erspamer V, Erspamer GF, Negri L, et al. Deltorphin, a novel amphibian skin peptide with high selectivity and affinity for delta opioid receptors. Eur J Pharmacol. 1989;162:123–128. doi: 10.1016/0014-2999(89)90611-0. [DOI] [PubMed] [Google Scholar]
  107. Krishnan-Sarin S, Jing SL, Kurtz DL, Zweifel M, Portoghese PS, Li TK, et al. The delta opioid receptor antagonist naltrindole attenuates both alcohol and saccharin intake in rats selectively bred for alcohol preference. Psychopharmacology (Berl) 1995a;120:177–185. doi: 10.1007/BF02246191. [DOI] [PubMed] [Google Scholar]
  108. Krishnan-Sarin S, Portoghese PS, Li TK, Froehlich JC. The delta 2-opioid receptor antagonist naltriben selectively attenuates alcohol intake in rats bred for alcohol preference. Pharmacol Biochem Behav. 1995b;52:153–159. doi: 10.1016/0091-3057(95)00080-g. [DOI] [PubMed] [Google Scholar]
  109. Lamberts JT, Traynor JR. Opioid receptor interacting proteins and the control of opioid signaling. Curr Pharm Des. 2013;19:7333–7347. doi: 10.2174/138161281942140105160625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Latimer LG, Duffy P, Kalivas PW. Mu opioid receptor involvement in enkephalin activation of DORamine neurons in the ventral tegmental area. J Pharmacol Exp Ther. 1987;241:328–337. [PubMed] [Google Scholar]
  111. Laurent V, Leung B, Maidment N, Balleine BW. Mu- and delta-opioid-related processes in the accumbens core and shell differentially mediate the influence of reward-guided and stimulus-guided decisions on choice. J Neurosci. 2012;32:1875–1883. doi: 10.1523/JNEUROSCI.4688-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Law PY, Loh HH. Delta-opioid receptor activates cAMP phosphodiesterase activities in neuroblastoma x glioma NG108-15 hybrid cells. Mol Pharmacol. 1993;43:684–693. [PubMed] [Google Scholar]
  113. Le AD, Poulos CX, Quan B, Chow S. The effects of selective blockade of delta and mu opiate receptors on ethanol consumption by C57BL/6 mice in a restricted access paradigm. Brain Res. 1993;630:330–332. doi: 10.1016/0006-8993(93)90672-a. [DOI] [PubMed] [Google Scholar]
  114. Le Bourdonnec B, Windh RT, Leister LK, Zhou QJ, Ajello CW, Gu M, et al. Spirocyclic delta opioid receptor agonists for the treatment of pain: discovery of N,N-diethyl-3-hydroxy-4-(spiro[chromene-2,4′-piperidine]-4-yl) benzamide (ADL5747) J Med Chem. 2009;52:5685–5702. doi: 10.1021/jm900773n. [DOI] [PubMed] [Google Scholar]
  115. Le Merrer J, Plaza-Zabala A, Del Boca C, Matifas A, Maldonado R, Kieffer BL. Deletion of the delta opioid receptor gene impairs place conditioning but preserves morphine reinforcement. Biol Psychiatry. 2011;69:700–703. doi: 10.1016/j.biopsych.2010.10.021. [DOI] [PubMed] [Google Scholar]
  116. Le Merrer J, Rezai X, Scherrer G, Becker JA, Kieffer BL. Impaired hippocampus-dependent and facilitated striatum-dependent behaviors in mice lacking the delta opioid receptor. Neuropsychopharmacology. 2013;38:1050–1059. doi: 10.1038/npp.2013.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Le Moine C, Kieffer B, Gaveriaux-Ruff C, Befort K, Bloch B. Delta-opioid receptor gene expression in the mouse forebrain: localization in cholinergic neurons of the striatum. Neuroscience. 1994;62:635–640. doi: 10.1016/0306-4522(94)90464-2. [DOI] [PubMed] [Google Scholar]
  118. Leone P, Pocock D, Wise RA. Morphine-DORamine interaction: ventral tegmental morphine increases nucleus accumbens DORamine release. Pharmacol Biochem Behav. 1991;39:469–472. doi: 10.1016/0091-3057(91)90210-s. [DOI] [PubMed] [Google Scholar]
  119. Li Z, Luan W, Chen Y, Chen M, Dong Y, Lai B, et al. Chronic morphine treatment switches the effect of DORamine on excitatory synaptic transmission from inhibition to excitation in pyramidal cells of the basolateral amygdala. J Neurosci. 2011;31:17527–17536. doi: 10.1523/JNEUROSCI.3806-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Lightman SL, Young WS., 3rd Changes in hypothalamic preproenkephalin A mRNA following stress and opiate withdrawal. Nature. 1987;328:643–645. doi: 10.1038/328643a0. [DOI] [PubMed] [Google Scholar]
  121. Ling GS, Spiegel K, Lockhart SH, Pasternak GW. Separation of opioid analgesia from respiratory depression: evidence for different receptor mechanisms. J Pharmacol Exp Ther. 1985;232:149–155. [PubMed] [Google Scholar]
  122. Lintas A, Chi N, Lauzon NM, Bishop SF, Sun N, Tan H, et al. Inputs from the basolateral amygdala to the nucleus accumbens shell control opiate reward magnitude via differential DORamine D1 or D2 receptor transmission. Eur J Neurosci. 2012;35:279–290. doi: 10.1111/j.1460-9568.2011.07943.x. [DOI] [PubMed] [Google Scholar]
  123. Longoni R, Spina L, Mulas A, Carboni E, Garau L, Melchiorri P, et al. D-Ala2)deltorphin II: D1-dependent stereotypies and stimulation of DORamine release in the nucleus accumbens. J Neurosci. 1991;11:1565–1576. doi: 10.1523/JNEUROSCI.11-06-01565.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Lovinger DM. Neurotransmitter roles in synaptic modulation, plasticity and learning in the dorsal striatum. Neuropharmacology. 2010;58:951–961. doi: 10.1016/j.neuropharm.2010.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Lupica CR. Delta and mu enkephalins inhibit spontaneous GABA-mediated IPSCs via a cyclic AMP-independent mechanism in the rat hippocampus. J Neurosci. 1995;15(1 Pt 2):737–749. doi: 10.1523/JNEUROSCI.15-01-00737.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Lupica CR, Proctor WR, Dunwiddie TV. Dissociation of mu and delta opioid receptor-mediated reductions in evoked and spontaneous synaptic inhibition in the rat hippocampus in vitro. Brain Res. 1992;593:226–238. doi: 10.1016/0006-8993(92)91312-3. [DOI] [PubMed] [Google Scholar]
  127. Ma YY, Cepeda C, Chatta P, Franklin L, Evans CJ, Levine MS. Regional and cell-type-specific effects of DAMGO on striatal D1 and D2 DORamine receptor-expressing medium-sized spiny neurons. ASN Neuro. 2012;4:e00077. doi: 10.1042/AN20110063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Margolis EB, Fields HL, Hjelmstad GO, Mitchell JM. Delta-opioid receptor expression in the ventral tegmental area protects against elevated alcohol consumption. J Neurosci. 2008;28:12672–12681. doi: 10.1523/JNEUROSCI.4569-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Marinelli PW, Funk D, Harding S, Li Z, Juzytsch W, Le AD. Roles of opioid receptor subtypes in mediating alcohol-seeking induced by discrete cues and context. Eur J Neurosci. 2009;30:671–678. doi: 10.1111/j.1460-9568.2009.06851.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Maruyama M, Sugino H, Akita K, Hatanaka H. Binding characteristics of [3H]opioid ligands to active opioid binding sites solubilized from rat brain membranes by glycodeoxycholate and NaCl: the recovery of binding activity by dilution. Brain Res. 1987;401:14–22. doi: 10.1016/0006-8993(87)91157-7. [DOI] [PubMed] [Google Scholar]
  131. Masukawa LM, Prince DA. Enkephalin inhibition of inhibitory input to CA1 and CA3 pyramidal neurons in the hippocampus. Brain Res. 1982;249:271–280. doi: 10.1016/0006-8993(82)90061-0. [DOI] [PubMed] [Google Scholar]
  132. McLamore S, Ullrich T, Rothman RB, Xu H, Dersch C, Coop A, et al. Effect of N-alkyl and N-alkenyl substituents in noroxymorphindole, 17-substituted-6,7-dehydro-4,5alpha-epoxy-3,14-dihydroxy-6,7:2′,3′-indolomorphina ns, on opioid receptor affinity, selectivity, and efficacy. J Med Chem. 2001;44:1471–1474. doi: 10.1021/jm000511w. [DOI] [PubMed] [Google Scholar]
  133. Mitchell JM, Margolis EB, Coker AR, Allen DC, Fields HL. Intra-VTA deltorphin, but not DPDPE, induces place preference in ethanol-drinking rats: distinct DOR-1 and DOR-2 mechanisms control ethanol consumption and reward. Alcohol Clin Exp Res. 2014;38:195–203. doi: 10.1111/acer.12246. [DOI] [PubMed] [Google Scholar]
  134. Mogenson GJ, Jones DL, Yim CY. From motivation to action: functional interface between the limbic system and the motor system. Prog Neurobiol. 1980;14:69–97. doi: 10.1016/0301-0082(80)90018-0. [DOI] [PubMed] [Google Scholar]
  135. Morinville A, Cahill CM, Esdaile MJ, Aibak H, Collier B, Kieffer BL, et al. Regulation of delta-opioid receptor trafficking via mu-opioid receptor stimulation: evidence from mu-opioid receptor knock-out mice. J Neurosci. 2003;23:4888–4898. doi: 10.1523/JNEUROSCI.23-12-04888.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Mosberg HI, Hurst R, Hruby VJ, Gee K, Yamamura HI, Galligan JJ, et al. Bis-penicillamine enkephalins possess highly improved specificity toward delta opioid receptors. Proc Natl Acad Sci U S A. 1983;80:5871–5874. doi: 10.1073/pnas.80.19.5871. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Mosberg HI, Omnaas JR, Goldstein A. Structural requirements for delta opioid receptor binding. Mol Pharmacol. 1987;31:599–602. [PubMed] [Google Scholar]
  138. Nagase H, Kawai K, Hayakawa J, Wakita H, Mizusuna A, Matsuura H, et al. Rational drug design and synthesis of a highly selective nonpeptide delta-opioid agonist, (4aS*,12aR*)-4a-(3-hydroxyphenyl)-2-methyl- 1,2,3,4,4a,5,12,12a-octahydropyrido[3,4-b]acridine (TAN-67) Chem Pharm Bull. 1998;46:1695–1702. doi: 10.1248/cpb.46.1695. [DOI] [PubMed] [Google Scholar]
  139. Neumaier JF, Mailheau S, Chavkin C. Opioid receptor-mediated responses in the dentate gyrus and CA1 region of the rat hippocampus. J Pharmacol Exp Ther. 1988;244:564–570. [PubMed] [Google Scholar]
  140. Nevin ST, Toth G, Weltrowska G, Schiller PW, Borsodi A. Synthesis and binding characteristics of tritiated TIPP[psi], a highly specific and stable delta opioid antagonist. Life Sci. 1995;56:PL225–230. doi: 10.1016/0024-3205(95)00012-u. [DOI] [PubMed] [Google Scholar]
  141. Nielsen CK, Bartlett SE. The role of delta opioid receptors in ethanol consumption and seeking: implications for new treatments for alcohol use disorders. In: Contreras CM, editor. Neuroscience – Dealing with Frontiers. Rijeka, Croatia: InTech; 2012. pp. 205–240. [Google Scholar]
  142. Nielsen CK, Simms JA, Pierson HB, Li R, Saini SK, Ananthan S, et al. A novel delta opioid receptor antagonist, SoRI-9409, produces a selective and long-lasting decrease in ethanol consumption in heavy-drinking rats. Biol Psychiatry. 2008;64:974–981. doi: 10.1016/j.biopsych.2008.07.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Nielsen CK, Simms JA, Bito-Onon JJ, Li R, Ananthan S, Bartlett SE. The delta opioid receptor antagonist, SoRI-9409, decreases yohimbine stress-induced reinstatement of ethanol-seeking. Addict Biol. 2012a;17:224–234. doi: 10.1111/j.1369-1600.2010.00295.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Nielsen CK, Simms JA, Li R, Mill D, Yi H, Feduccia AA, et al. Delta-opioid receptor function in the dorsal striatum plays a role in high levels of ethanol consumption in rats. J Neurosci. 2012b;32:4540–4552. doi: 10.1523/JNEUROSCI.5345-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. North RA, Williams JT, Surprenant A, Christie MJ. Mu and delta receptors belong to a family of receptors that are coupled to potassium channels. Proc Natl Acad Sci U S A. 1987;84:5487–5491. doi: 10.1073/pnas.84.15.5487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Nozaki C, Le Bourdonnec B, Reiss D, Windh RT, Little PJ, Dolle RE, et al. Delta-opioid mechanisms for ADL5747 and ADL5859 effects in mice: analgesia, locomotion, and receptor internalization. J Pharmacol Exp Ther. 2012;342:799–807. doi: 10.1124/jpet.111.188987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. O'Neill SJ, Collins MA, Pettit HO, McNutt RW, Chang KJ. Antagonistic modulation between the delta opioid agonist BW373U86 and the mu opioid agonist fentanyl in mice. J Pharmacol Exp Ther. 1997;282:271–277. [PubMed] [Google Scholar]
  148. Packard MG, McGaugh JL. Inactivation of hippocampus or caudate nucleus with lidocaine differentially affects expression of place and response learning. Neurobiol Learn Mem. 1996;65:65–72. doi: 10.1006/nlme.1996.0007. [DOI] [PubMed] [Google Scholar]
  149. Paden CM, Krall S, Lynch WC. Heterogeneous distribution and upregulation of mu, delta and kappa opioid receptors in the amygdala. Brain Res. 1987;418:349–355. doi: 10.1016/0006-8993(87)90102-8. [DOI] [PubMed] [Google Scholar]
  150. Pang K, Rose GM. Differential effects of methionine5-enkephalin on hippocampal pyramidal cells and interneurons. Neuropharmacology. 1989;28:1175–1181. doi: 10.1016/0028-3908(89)90208-6. [DOI] [PubMed] [Google Scholar]
  151. Parkinson JA, Cardinal RN, Everitt BJ. Limbic cortical-ventral striatal systems underlying appetitive conditioning. Prog Brain Res. 2000;126:263–285. doi: 10.1016/S0079-6123(00)26019-6. [DOI] [PubMed] [Google Scholar]
  152. Pecina S, Berridge KC. Hedonic hot spot in nucleus accumbens shell: where do mu-opioids cause increased hedonic impact of sweetness? J Neurosci. 2005;25:11777–11786. doi: 10.1523/JNEUROSCI.2329-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Pennartz CM, Ito R, Verschure PF, Battaglia FP, Robbins TW. The hippocampal-striatal axis in learning, prediction and goal-directed behavior. Trends Neurosci. 2011;34:548–559. doi: 10.1016/j.tins.2011.08.001. [DOI] [PubMed] [Google Scholar]
  154. Pentney RJ, Gratton A. Effects of local delta and mu opioid receptor activation on basal and stimulated DORamine release in striatum and nucleus accumbens of rat: an in vivo electrochemical study. Neuroscience. 1991;45:95–102. doi: 10.1016/0306-4522(91)90106-x. [DOI] [PubMed] [Google Scholar]
  155. Perrine SA, Sheikh IS, Nwaneshiudu CA, Schroeder JA, Unterwald EM. Withdrawal from chronic administration of cocaine decreases delta opioid receptor signaling and increases anxiety- and depression-like behaviors in the rat. Neuropharmacology. 2008;54:355–364. doi: 10.1016/j.neuropharm.2007.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Petrillo P, Angelici O, Bingham S, Ficalora G, Garnier M, Zaratin PF, et al. Evidence for a selective role of the delta-opioid agonist [8R-(4bS*,8aalpha,8abeta, 12bbeta)]7,10-Dimethyl-1-methoxy-11-(2-methylpropyl)oxycarbonyl 5,6,7,8,12,12b-hexahydro-(9H)-4,8-methanobenzofuro[3,2-e]pyrrolo[2,3-g]isoquinoline hydrochloride (SB-235863) in blocking hyperalgesia associated with inflammatory and neuropathic pain responses. J Pharmacol Exp Ther. 2003;307:1079–1089. doi: 10.1124/jpet.103.055590. [DOI] [PubMed] [Google Scholar]
  157. Piros ET, Prather PL, Law PY, Evans CJ, Hales TG. Voltage-dependent inhibition of Ca2 + channels in GH3 cells by cloned mu- and delta-opioid receptors. Mol Pharmacol. 1996;50:947–956. [PubMed] [Google Scholar]
  158. Ponterio G, Tassone A, Sciamanna G, Riahi E, Vanni V, Bonsi P, et al. Powerful inhibitory action of mu opioid receptors (MOR) on cholinergic interneuron excitability in the dorsal striatum. Neuropharmacology. 2013;75C:78–85. doi: 10.1016/j.neuropharm.2013.07.006. [DOI] [PubMed] [Google Scholar]
  159. Pontieri FE, Tanda G, Di Chiara G. Intravenous cocaine, morphine, and amphetamine preferentially increase extracellular DORamine in the ‘shell’ as compared with the ‘core’ of the rat nucleus accumbens. Proc Natl Acad Sci U S A. 1995;92:12304–12308. doi: 10.1073/pnas.92.26.12304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Poulin JF, Chevalier B, Laforest S, Drolet G. Enkephalinergic afferents of the centromedial amygdala in the rat. J Comp Neurol. 2006;496:859–876. doi: 10.1002/cne.20956. [DOI] [PubMed] [Google Scholar]
  161. Pradhan AA, Becker JA, Scherrer G, Tryoen-Toth P, Filliol D, Matifas A, et al. In vivo delta opioid receptor internalization controls behavioral effects of agonists. PLoS ONE. 2009;4:e5425. doi: 10.1371/journal.pone.0005425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Pradhan AA, Smith ML, Kieffer BL, Evans CJ. Ligand-directed signalling within the opioid receptor family. Br J Pharmacol. 2012;167:960–969. doi: 10.1111/j.1476-5381.2012.02075.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Rachinger-Adam B, Conzen P, Azad SC. Pharmacology of peripheral opioid receptors. Curr Opin Anaesthesiol. 2011;24:408–413. doi: 10.1097/ACO.0b013e32834873e5. [DOI] [PubMed] [Google Scholar]
  164. Raynor K, Kong H, Chen Y, Yasuda K, Yu L, Bell GI, et al. Pharmacological characterization of the cloned kappa-, delta-, and mu-opioid receptors. Mol Pharmacol. 1994;45:330–334. [PubMed] [Google Scholar]
  165. Redgrave P, Gurney K. The short-latency DORamine signal: a role in discovering novel actions? Nat Rev Neurosci. 2006;7:967–975. doi: 10.1038/nrn2022. [DOI] [PubMed] [Google Scholar]
  166. Rezai X, Faget L, Bednarek E, Schwab Y, Kieffer BL, Massotte D. Mouse delta opioid receptors are located on presynaptic afferents to hippocampal pyramidal cells. Cell Mol Neurobiol. 2012;32:509–516. doi: 10.1007/s10571-011-9791-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. van Rijn RM, Whistler JL. The delta(1) opioid receptor is a heterodimer that opposes the actions of the delta(2) receptor on alcohol intake. Biol Psychiatry. 2009;66:777–784. doi: 10.1016/j.biopsych.2009.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. van Rijn RM, Brissett DI, Whistler JL. Dual efficacy of delta opioid receptor-selective ligands for ethanol drinking and anxiety. J Pharmacol Exp Ther. 2010a;335:133–139. doi: 10.1124/jpet.110.170969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. van Rijn RM, Whistler JL, Waldhoer M. Opioid-receptor-heteromer-specific trafficking and pharmacology. Curr Opin Pharmacol. 2010b;10:73–79. doi: 10.1016/j.coph.2009.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. van Rijn RM, Brissett DI, Whistler JL. Emergence of functional spinal delta opioid receptors after chronic ethanol exposure. Biol Psychiatry. 2012;71:232–238. doi: 10.1016/j.biopsych.2011.07.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. van Rijn RM, Defriel JN, Whistler JL. Pharmacological traits of delta opioid receptors: pitfalls or opportunities? Psychopharmacology (Berl) 2013;228:1–18. doi: 10.1007/s00213-013-3129-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Robbins TW, Everitt BJ. Limbic-striatal memory systems and drug addiction. Neurobiol Learn Mem. 2002;78:625–636. doi: 10.1006/nlme.2002.4103. [DOI] [PubMed] [Google Scholar]
  173. Roberts AJ, Gold LH, Polis I, McDonald JS, Filliol D, Kieffer BL, et al. Increased ethanol self-administration in delta-opioid receptor knockout mice. Alcohol Clin Exp Res. 2001;25:1249–1256. [PubMed] [Google Scholar]
  174. Roitman MF, Wheeler RA, Carelli RM. Nucleus accumbens neurons are innately tuned for rewarding and aversive taste stimuli, encode their predictors, and are linked to motor output. Neuron. 2005;45:587–597. doi: 10.1016/j.neuron.2004.12.055. [DOI] [PubMed] [Google Scholar]
  175. Rossato JI, Bonini JS, Coitinho AS, Vianna MR, Medina JH, Cammarota M, et al. Retrograde amnesia induced by drugs acting on different molecular systems. Behav Neurosci. 2004;118:563–568. doi: 10.1037/0735-7044.118.3.563. [DOI] [PubMed] [Google Scholar]
  176. Saitoh A, Sugiyama A, Nemoto T, Fujii H, Wada K, Oka J, et al. The novel delta opioid receptor agonist KNT-127 produces antidepressant-like and antinociceptive effects in mice without producing convulsions. Behav Brain Res. 2011;223:271–279. doi: 10.1016/j.bbr.2011.04.041. [DOI] [PubMed] [Google Scholar]
  177. Saitoh A, Sugiyama A, Yamada M, Inagaki M, Oka J, Nagase H, et al. The novel delta opioid receptor agonist KNT-127 produces distinct anxiolytic-like effects in rats without producing the adverse effects associated with benzodiazepines. Neuropharmacology. 2013;67:485–493. doi: 10.1016/j.neuropharm.2012.11.025. [DOI] [PubMed] [Google Scholar]
  178. Scherrer G, Tryoen-Toth P, Filliol D, Matifas A, Laustriat D, Cao YQ, et al. Knockin mice expressing fluorescent delta-opioid receptors uncover G protein-coupled receptor dynamics in vivo. Proc Natl Acad Sci U S A. 2006;103:9691–9696. doi: 10.1073/pnas.0603359103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Schiller PW, Weltrowska G, Nguyen TM, Wilkes BC, Chung NN, Lemieux C. TIPP[psi]: a highly potent and stable pseuDOReptide delta opioid receptor antagonist with extraordinary delta selectivity. J Med Chem. 1993;36:3182–3187. doi: 10.1021/jm00073a020. [DOI] [PubMed] [Google Scholar]
  180. Schultz W, Dayan P, Montague PR. A neural substrate of prediction and reward. Science. 1997;275:1593–1599. doi: 10.1126/science.275.5306.1593. [DOI] [PubMed] [Google Scholar]
  181. Shiflett MW, Balleine BW. At the limbic-motor interface: disconnection of basolateral amygdala from nucleus accumbens core and shell reveals dissociable components of incentive motivation. Eur J Neurosci. 2010;32:1735–1743. doi: 10.1111/j.1460-9568.2010.07439.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Simmons D, Self DW. Role of mu- and delta-opioid receptors in the nucleus accumbens in cocaine-seeking behavior. Neuropsychopharmacology. 2009;34:1946–1957. doi: 10.1038/npp.2009.28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Simmons ML, Chavkin C. Endogenous opioid regulation of hippocampal function. Int Rev Neurobiol. 1996;39:145–196. doi: 10.1016/s0074-7742(08)60666-2. [DOI] [PubMed] [Google Scholar]
  184. Smith JA, Hunter JC, Hill RG, Hughes J. A kinetic analysis of kappa-opioid agonist binding using the selective radioligand [3H]U69593. J Neurochem. 1989;53:27–36. doi: 10.1111/j.1471-4159.1989.tb07291.x. [DOI] [PubMed] [Google Scholar]
  185. Smith KS, Berridge KC. Opioid limbic circuit for reward: interaction between hedonic hotspots of nucleus accumbens and ventral pallidum. J Neurosci. 2007;27:1594–1605. doi: 10.1523/JNEUROSCI.4205-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Sofuoglu M, Portoghese PS, Takemori AE. 7-Benzylidenenaltrexone (BNTX): a selective delta 1 opioid receptor antagonist in the mouse spinal cord. Life Sci. 1993;52:769–775. doi: 10.1016/0024-3205(93)90240-4. [DOI] [PubMed] [Google Scholar]
  187. Spanagel R, Herz A, Shippenberg TS. The effects of opioid peptides on DORamine release in the nucleus accumbens: an in vivo microdialysis study. J Neurochem. 1990;55:1734–1740. doi: 10.1111/j.1471-4159.1990.tb04963.x. [DOI] [PubMed] [Google Scholar]
  188. Spanagel R, Herz A, Shippenberg TS. Opposing tonically active endogenous opioid systems modulate the mesolimbic DORaminergic pathway. Proc Natl Acad Sci U S A. 1992;89:2046–2050. doi: 10.1073/pnas.89.6.2046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Sternini C, Spann M, Anton B, Keith DE, Jr, Bunnett NW, von Zastrow M, et al. Agonist-selective endocytosis of mu opioid receptor by neurons in vivo. Proc Natl Acad Sci U S A. 1996;93:9241–9246. doi: 10.1073/pnas.93.17.9241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Stewart PE, Hammond DL. Evidence for delta opioid receptor subtypes in rat spinal cord: studies with intrathecal naltriben, cyclic[D-Pen2, D-Pen5] enkephalin and [D-Ala2, Glu4]deltorphin. J Pharmacol Exp Ther. 1993;266:820–828. [PubMed] [Google Scholar]
  191. Strother WN, Chernet EJ, Lumeng L, Li TK, McBride WJ. Regional central nervous system densities of delta-opioid receptors in alcohol-preferring P, alcohol-nonpreferring NP, and unselected Wistar rats. Alcohol. 2001;25:31–38. doi: 10.1016/s0741-8329(01)00162-8. [DOI] [PubMed] [Google Scholar]
  192. Stuber GD, Klanker M, de Ridder B, Bowers MS, Joosten RN, Feenstra MG, et al. Reward-predictive cues enhance excitatory synaptic strength onto midbrain DORamine neurons. Science. 2008;321:1690–1692. doi: 10.1126/science.1160873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Stuber GD, Sparta DR, Stamatakis AM, van Leeuwen WA, Hardjoprajitno JE, Cho S, et al. Excitatory transmission from the amygdala to nucleus accumbens facilitates reward seeking. Nature. 2011;475:377–380. doi: 10.1038/nature10194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Suzuki T, Tsuji M, Mori T, Misawa M, Nagase H. Effect of naltrindole on the development of physical dependence on morphine in mice: a behavioral and biochemical study. Life Sci. 1995;57:PL247–252. doi: 10.1016/0024-3205(95)02139-a. [DOI] [PubMed] [Google Scholar]
  195. Svingos AL, Moriwaki A, Wang JB, Uhl GR, Pickel VM. Mu-opioid receptors are localized to extrasynaptic plasma membranes of GABAergic neurons and their targets in the rat nucleus accumbens. J Neurosci. 1997;17:2585–2594. doi: 10.1523/JNEUROSCI.17-07-02585.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Svingos AL, Clarke CL, Pickel VM. Cellular sites for activation of delta-opioid receptors in the rat nucleus accumbens shell: relationship with Met5-enkephalin. J Neurosci. 1998;18:1923–1933. doi: 10.1523/JNEUROSCI.18-05-01923.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Svingos AL, Clarke CL, Pickel VM. Localization of the delta-opioid receptor and DORamine transporter in the nucleus accumbens shell: implications for opiate and psychostimulant cross-sensitization. Synapse. 1999;34:1–10. doi: 10.1002/(SICI)1098-2396(199910)34:1<1::AID-SYN1>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  198. Svoboda KR, Adams CE, Lupica CR. Opioid receptor subtype expression defines morphologically distinct classes of hippocampal interneurons. J Neurosci. 1999;19:85–95. doi: 10.1523/JNEUROSCI.19-01-00085.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Traynor JR, Fantegrossi W, Woods JH. Evaluation of compounds for opioid activity. In: WL Dewey., editor. Problems of Drug Dependence, 2004: Proceedings of the 66th Annual Scientific Meeting. Rockville, MD: U.S. Department of Health and Human Services; 2005. pp. 131–159. [Google Scholar]
  200. Tseng A, Nguyen K, Hamid A, Garg M, Marquez P, Lutfy K. The role of endogenous beta-endorphin and enkephalins in ethanol reward. Neuropharmacology. 2013;73C:290–300. doi: 10.1016/j.neuropharm.2013.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Wang H, Pickel VM. Dendritic spines containing mu-opioid receptors in rat striatal patches receive asymmetric synapses from prefrontal corticostriatal afferents. J Comp Neurol. 1998;396:223–237. [PubMed] [Google Scholar]
  202. Wang H, Cuzon VC, Pickel VM. Ultrastructural localization of delta-opioid receptors in the rat caudate-putamen nucleus during postnatal development: relation to synaptogenesis. J Comp Neurol. 2003;467:343–353. doi: 10.1002/cne.10920. [DOI] [PubMed] [Google Scholar]
  203. Ward SJ, Roberts DC. Microinjection of the delta-opioid receptor selective antagonist naltrindole 5′-isothiocyanate site specifically affects cocaine self-administration in rats responding under a progressive ratio schedule of reinforcement. Behav Brain Res. 2007;182:140–144. doi: 10.1016/j.bbr.2007.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Whistler JL, Chuang HH, Chu P, Jan LY, von Zastrow M. Functional dissociation of mu opioid receptor signaling and endocytosis: implications for the biology of opiate tolerance and addiction. Neuron. 1999;23:737–746. doi: 10.1016/s0896-6273(01)80032-5. [DOI] [PubMed] [Google Scholar]
  205. Wickens JR, Horvitz JC, Costa RM, Killcross S. DORaminergic mechanisms in actions and habits. J Neurosci. 2007;27:8181–8183. doi: 10.1523/JNEUROSCI.1671-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Wiesner JB, Henriksen SJ. Enkephalin enhances responsiveness to perforant path input while decreasing spontaneous activity in the dentate gyrus. Neurosci Lett. 1987;74:95–101. doi: 10.1016/0304-3940(87)90058-9. [DOI] [PubMed] [Google Scholar]
  207. Wiesner JB, Henriksen SJ, Bloom FE. Opioid enhancement of perforant path transmission: effect of an enkephalin analog on inhibition and facilitation in the dentate gyrus. Brain Res. 1986;399:404–408. doi: 10.1016/0006-8993(86)91537-4. [DOI] [PubMed] [Google Scholar]
  208. Wise RA. DORamine, learning and motivation. Nat Rev Neurosci. 2004;5:483–494. doi: 10.1038/nrn1406. [DOI] [PubMed] [Google Scholar]
  209. Wise RA, Bozarth MA. Brain mechanisms of drug reward and euphoria. Psychiatr Med. 1985;3:445–460. [PubMed] [Google Scholar]
  210. Xu H, Lu YF, Rice KC, Ananthan S, Rothman RB. SoRI 9409, a non-peptide opioid mu receptor agonist/delta receptor antagonist, fails to stimulate [35S]-GTP-gamma-S binding at cloned opioid receptors. Brain Res Bull. 2001;55:507–511. doi: 10.1016/s0361-9230(01)00550-0. [DOI] [PubMed] [Google Scholar]
  211. Yin HH, Knowlton BJ. The role of the basal ganglia in habit formation. Nat Rev Neurosci. 2006;7:464–476. doi: 10.1038/nrn1919. [DOI] [PubMed] [Google Scholar]
  212. Yin HH, Ostlund SB, Balleine BW. Reward-guided learning beyond DORamine in the nucleus accumbens: the integrative functions of cortico-basal ganglia networks. Eur J Neurosci. 2008;28:1437–1448. doi: 10.1111/j.1460-9568.2008.06422.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Yin HH, Mulcare SP, Hilario MR, Clouse E, Holloway T, Davis MI, et al. Dynamic reorganization of striatal circuits during the acquisition and consolidation of a skill. Nat Neurosci. 2009;12:333–341. doi: 10.1038/nn.2261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Yoshida Y, Koide S, Hirose N, Takada K, Tomiyama K, Koshikawa N, et al. Fentanyl increases DORamine release in rat nucleus accumbens: involvement of mesolimbic mu- and delta-2-opioid receptors. Neuroscience. 1999;92:1357–1365. doi: 10.1016/s0306-4522(99)00046-9. [DOI] [PubMed] [Google Scholar]
  215. Zhang M, Kelley AE. Enhanced intake of high-fat food following striatal mu-opioid stimulation: microinjection mapping and fos expression. Neuroscience. 2000;99:267–277. doi: 10.1016/s0306-4522(00)00198-6. [DOI] [PubMed] [Google Scholar]
  216. Zhang M, Balmadrid C, Kelley AE. Nucleus accumbens opioid, GABaergic, and DORaminergic modulation of palatable food motivation: contrasting effects revealed by a progressive ratio study in the rat. Behav Neurosci. 2003;117:202–211. doi: 10.1037/0735-7044.117.2.202. [DOI] [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES