Sunteți pe pagina 1din 12

Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


journal homepage: www.elsevier.com/locate/neubiorev

Review

Neurobiological mechanisms involved in nicotine dependence and reward:


Participation of the endogenous opioid system
Fernando Berrendero a , Patricia Robledo a,b , José Manuel Trigo a ,
Elena Martín-García a , Rafael Maldonado a,∗
a
Laboratory of Neuropharmacology, Department of Experimental and Health Sciences, University Pompeu Fabra, PRBB, C/Doctor Aiguader 88, 08003 Barcelona, Spain
b
Municipal Institute of Medical Research (IMIM), Barcelona, Spain

a r t i c l e i n f o a b s t r a c t

Keywords: Nicotine is the primary component of tobacco that maintains the smoking habit and develops addiction.
Addiction The adaptive changes of nicotinic acetylcholine receptors produced by repeated exposure to nicotine play
Enkephalin a crucial role in the establishment of dependence. However, other neurochemical systems also participate
Dynorphin
in the addictive effects of nicotine including glutamate, cannabinoids, GABA and opioids. This review
Beta-endorphin
will cover the involvement of these neurotransmitters in nicotine addictive properties, with a special
Tolerance
Withdrawal emphasis on the endogenous opioid system. Thus, endogenous enkephalins and beta-endorphins acting
Naltrexone on mu-opioid receptors are involved in nicotine-rewarding effects, whereas opioid peptides derived
Opioid receptor from prodynorphin participate in nicotine aversive responses. An up-regulation of mu-opioid receptors
has been reported after chronic nicotine treatment that could counteract the development of nicotine
tolerance, whereas the down-regulation induced on kappa-opioid receptors seems to facilitate nicotine
tolerance. Endogenous enkephalins acting on mu-opioid receptors also play a role in the development of
physical dependence to nicotine. In agreement with these actions of the endogenous opioid system, the
opioid antagonist naltrexone has shown to be effective for smoking cessation in certain sub-populations
of smokers.
© 2010 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
2. Nicotinic acetylcholine receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
3. Behavioral models to evaluate nicotine addictive properties in animals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
4. Neurotransmitters involved in nicotine addictive properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.1. Involvement of different nAChR subunits in nicotine reward, motivation and reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.2. Involvement of glutamatergic receptors in nicotine reward, motivation and reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.3. Involvement of cannabinoid receptors in nicotine reward, motivation and reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.4. Other neurotransmitters involved in nicotine reward, motivation and reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.5. Neurotransmitters involved in nicotine tolerance and withdrawal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.6. Neurotransmitters involved in nicotine relapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
5. The endogenous opioid system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
6. Effects of nicotine administration on the endogenous opioid system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7. The role of the endogenous opioid system in nicotine reward, motivation and reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
8. The role of the endogenous opioid system in the development of nicotine tolerance and physical dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
9. Efficacy of opioid antagonists in the treatment for smoking cessation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
10. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228

∗ Corresponding author. Tel.: +34 93 3160824; fax: +34 93 3160901.


E-mail address: rafael.maldonado@upf.edu (R. Maldonado).

0149-7634/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.neubiorev.2010.02.006
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 221

1. Introduction this addictive behavior (Koob and Le Moal, 2008). Several behav-
ioral models have evaluated the development of tolerance and
Nicotine is the main addictive component of tobacco that physical dependence after repeated nicotine administration. How-
maintains the smoking habit. Moreover, the reinforcing effects of ever, the development of tolerance and physical dependence that
nicotine are thought to be the primary reason why humans inhale is concurrent to the nicotine addictive process does not seem
tobacco smoke (Shoaib, 2006). Tobacco consumption is one of the to be etiologically related to nicotine addiction (Volkow and Li,
main public health problems worldwide and represents a leading 2005). The negative consequences of nicotine withdrawal that are
cause of preventable deaths in most developed countries. Tobacco closely related to the maintenance of the addictive process can
smoke contains over 4000 chemicals, many of which have marked be evaluated in rodents by measuring several emotional symp-
irritant properties, are known carcinogens and can enhance the toms such as increased anxiety, aversive effects and reward deficits
addictive effects of nicotine (Brennan et al., 2009). Tobacco use (Jackson et al., 2008; Johnson et al., 2008). Thus, the aversive
causes five million deaths per year, and if the present trend con- manifestations of withdrawal are mainly evaluated in rodents by
tinues, 10 million smokers per year are predicted to die by 2025 using the place conditioning paradigm, whereas the associated
(Hatsukami et al., 2008). Although most smokers wish to stop, few reward deficits are currently evaluated using intracranial self-
succeed. Indeed, relapse rates are as high as 80% one year after stimulation techniques (Jackson et al., 2009; Bruijnzeel et al., 2009).
the quit date, even with the help of the available medications Both behavioral paradigms have been extensively used to eval-
and additional non-pharmacological therapies (Dwoskin et al., uate nicotine-rewarding effects (Castañé et al., 2002; Hollander
2009). et al., 2008). In the place conditioned paradigms, the subjective
effects of the drug (or its withdrawal) are repeatedly paired to a
previously neutral spatial environmental stimulus. Through this
2. Nicotinic acetylcholine receptors
repeated association process, the environment acts as a condi-
tioned stimulus. The animal will then prefer (conditioned place
Nicotine produces its pharmacological effects through the acti-
preference) or avoid (conditioned place aversion) this conditioned
vation of nicotinic acetylcholine receptors (nAChRs), which are
stimulus, depending on the rewarding or aversive effects pro-
widely distributed through the central nervous system (CNS). These
duced by the presence of the drug or by its withdrawal. Although
receptors are ligand-gated ion channels with a central cation pore
a conditioned approach/avoidance towards specific stimuli can
composed of five subunits, which can either be homomeric or
also occur in humans as a result of drug consumption (Childs and
heteromeric (Millar and Gotti, 2009). So far, 12 different neu-
de Wit, 2009), the place conditioning paradigms are not primar-
ronal nAChR subunits have been identified: ␣2 –␣10 and ␤2 –␤4 .
ily intended to model any particular feature of human behavior.
The most abundant nAChR subtypes in the nervous system are
These paradigms mainly represent an indirect assessment of the
homomeric ␣7 and heteromeric ␣4 ␤2 (Millar and Gotti, 2009).
rewarding or aversive effects of a drug or its withdrawal by
The nAChRs play a modulatory role on the activity of multiple
measuring the response of the animal towards the conditioned
neurotransmitters, as they have been detected on presynaptic
stimulus. Two different phases (acquisition and expression) of
terminals, cell bodies and dendrites of many neuronal subtypes
the place conditioning that have different psychological implica-
(Dajas-Bailador and Wonnacott, 2004). The activation of nAChRs by
tions are evaluated in this paradigm. Indeed, acquisition seems
nicotine increases the release of most neurotransmitters including
to be related to reward and learning whereas expression would
dopamine (Pontieri et al., 1996), noradrenaline (Clarke and Reuben,
be more related to incentive motivation, memory recall or sign-
1996), acetylcholine (Wilkie et al., 1993), glutamate (McGehee et
tracking.
al., 1995) and GABA (Yang et al., 1996). As a consequence, nico-
Intracranial electric self-stimulation procedures have been
tine modifies a large number of physiological processes such as
widely used to study the effects of drugs of abuse on the activ-
locomotion, nociception, anxiety, learning and memory, as well
ity of the reward circuits (Sanchis-Segura and Spanagel, 2006). In
as produces several behavioral responses directly related to its
this paradigm, animals are trained to maintain an operant behav-
addictive properties, including rewarding effects and physical
ior to obtain an electric pulse through an electrode placed in a
dependence (Decker et al., 1995). Due to the limited specificity
reward-related brain site, most frequently the lateral hypothala-
of the available nAChR agonists and antagonists, the generation
mic area. The threshold of the minimal current needed to promote
of mice with genetic nAChR subunit modifications has provided a
intracranial electric self-stimulation is usually estimated. A drug
useful experimental approach to assess the specific contribution of
that stimulates the reward circuit will decrease this threshold,
these subunits in the pharmacological actions induced by nicotine
which would be related to its rewarding properties, whereas a drug
(Mineur and Picciotto, 2008). In addition, novel strategies have also
or a state of withdrawal producing aversive effects will enhance the
been developed to re-express nAChR subunits in a region-specific
minimal current required to maintain the self-stimulation (Markou
manner using lentiviral vectors in a knockout background, which
and Koob, 1993).
has further advanced the knowledge of these receptors (Maskos,
Another animal model that has been widely used to directly
2007).
evaluate the reinforcing properties of nicotine is the operant self-
administration procedure, which is a valid and reliable model of
3. Behavioral models to evaluate nicotine addictive drug consumption in humans (Sanchis-Segura and Spanagel, 2006).
properties in animals High rates of operant responding for obtaining intravenous nico-
tine infusions have been described in various species (Goldberg et
Nicotine addiction is a chronic brain disorder characterized by al., 1981; Risner and Goldberg, 1983; Corrigall and Coen, 1989;
compulsive tobacco use, loss of control over tobacco consump- Shoaib et al., 1997; Stolerman et al., 1999; Martín-García et al.,
tion despite its harmful effects, the appearance of withdrawal 2009), confirming the abuse liability of this drug (Stolerman and
symptoms upon cessation of tobacco smoking, and relapse after Jarvis, 1995). Operant models are based on the learning contin-
periods of abstinence (McLellan et al., 2000). Similar to other gency defined as “positive reinforcement”. In these models, the
addictive processes, the initiation of nicotine addiction has been drug serves as a positive reinforcer that is delivered in a con-
related to its capacity to induce reinforcing effects. On the other tingent manner upon completion of the scheduled requirements
hand, the negative consequences of nicotine abstinence have a cru- (Sanchis-Segura and Spanagel, 2006). The operant chambers are
cial motivational significance for the maintenance and relapse of equipped with manipulandi that transmit the operant response
222 F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

as well as devices that deliver the drug (reinforcer). The response et al., 2008). In addition, ␣4 knockout mice do not self-administer
in the active manipulandum is linked to the delivery of the drug, nicotine (Pons et al., 2008) and fail to show the nicotine-dependent
while the response in the inactive manipulandum results in the enhancement of dopamine release in the ventral striatum (Marubio
delivery of the drug vehicle or lacks any programmed consequence. et al., 2003), whereas selective activation of ␣4 containing nAChRs
The fixed ratio and progressive ratio schedule reinforcement pro- is sufficient to induce nicotine locomotor sensitization and condi-
grams are commonly used. Under a fixed ratio schedule, the drug tioned place preference (Tapper et al., 2004). The ␣6 subunit located
is delivered each time a pre-selected number of responses are in the VTA also plays a role in the reinforcing effects of nicotine
completed. For nicotine self-administration, the route of drug (Pons et al., 2008). This result is consistent with a recent study
delivery is usually intravenous, and the number of responses showing a blockade of nicotine conditioned place preference in
required to obtain the drug is generally kept low, most frequently mice pre-treated with an ␣6 subunit selective antagonist (Jackson
at a fixed ratio of one (continuous reinforcement, i.e., one nico- et al., 2009). The precise role of the ␣7 homomeric nAChRs in nico-
tine infusion after each response in the active manipulandum). tine reinforcing effects remains unclear because conflicting results
Under the progressive ratio schedule, the response requirement have been obtained in mutant mice lacking this subunit and in
to deliver the drug escalates according to an arithmetic pro- rodents injected with the relatively selective ␣7 nAChR antago-
gression. The common index of performance evaluated in this nist methyllycaconitine (Pons et al., 2008; Markou and Paterson,
schedule is the breaking point defined as the highest number of 2001; Walters et al., 2006). Along with research in animal models,
responses that the animal accomplishes to obtain a single infusion recent genome-wide association studies in humans have revealed
of drug, which provides information about its motivation for the a clear linkage between genetic variations in the nAChRs and the
drug. risk for nicotine dependence (Bierut, 2009). Thus, the region on
chromosome 15 that includes the family of ␣5 –␣3 –␤4 nAChR genes
has been associated with the development of nicotine dependence
4. Neurotransmitters involved in nicotine addictive
(Thorgeirsson et al., 2008; Berrettini et al., 2008) and lung cancer
properties
(Thorgeirsson et al., 2008; Hung et al., 2008; Amos et al., 2008).
The connection between the genetic variant at chromosome 15 and
The neurobiology of nicotine addiction is a complex behavioral
lung cancer is thought to be either direct (Hung et al., 2008; Amos et
phenomenon in which various transmitter systems are involved.
al., 2008) or mediated through a modification of smoking behavior
This section of the review summarizes the involvement of nicotinic
(Thorgeirsson et al., 2008).
acetylcholine receptors and several heterologous neurochemical
systems in the addictive effects of nicotine. A particular empha-
sis is then devoted to the role played by the endogenous opioid 4.2. Involvement of glutamatergic receptors in nicotine reward,
system in the different processes that contribute to nicotine addic- motivation and reinforcement
tion.
Nicotine stimulates the release of glutamate through activa-
4.1. Involvement of different nAChR subunits in nicotine reward, tion of nAChRs located on glutamatergic terminals in several brain
motivation and reinforcement regions including the VTA (Fu et al., 2000). Furthermore, consid-
erable evidence indicates that glutamate, through activation of
The mesocorticolimbic system mediates the rewarding proper- ionotropic and metabotropic glutamate receptors located on post-
ties of most drugs of abuse, including nicotine (Koob and Le Moal, synaptic dopamine neurons, is critically involved in the reinforcing
2008). An important component of this system is the dopaminergic properties of nicotine (Liechti and Markou, 2008). Thus, nicotine-
projection from the ventral tegmental area (VTA) to the frontal cor- induced dopamine release in the NAc is blocked by systemic
tex and limbic structures, such as the nucleus accumbens (NAc) and administration of NMDA and AMPA receptor antagonists (Kosowski
amygdala. However, other additional brain areas that interact with et al., 2004). In addition, the NMDA receptor antagonist, LY235959,
these mesolimbic structures also play an important role in acute administered systemically or directly into the VTA or the central
drug reward and chronic changes in reward associated with addic- amygdala, decreases intravenous nicotine self-administration in
tion. These regions include the amygdala and related structures rats (Kenny et al., 2009). Several studies have also involved post-
of the so-called ‘extended amygdala’, hippocampus, hypothalamus synaptic mGlu5 and presynaptic mGlu2/3 receptors in nicotine
and several regions of the frontal cortex, among others (Nestler, reinforcing effects. Thus, the mGlu5 receptor antagonist, MPEP,
2005). Nicotine administration increases dopamine transmission in decreases nicotine self-administration in rats and mice (Paterson
the NAc and other limbic structures (Di Chiara and Imperato, 1988) et al., 2003). The same antagonist also diminished the motiva-
by direct stimulation of nAChRs within the VTA (Nisell et al., 1994). tion to work for nicotine infusions as revealed by the reduction
A great body of evidence suggests that nAChRs containing ␣4 ␤2 in the breakpoint on a progressive ratio schedule of reinforcement
subunits located on dopamine cell bodies contribute to the final (Paterson and Markou, 2005). The administration of the mGlu2/3
activation of VTA dopamine neurons (Mansvelder and McGehee, agonist LY379268 either systemically or intra-VTA also decreases
2003). Indeed, local and systemic administration of the selective nicotine self-administration in rats (Liechti et al., 2007). This last
␣4 ␤2 antagonist dihydro-beta-erythroidine (DH␤E) blocks nicotine result is in accordance with previous studies showing that presy-
self-administration in rodents (Grottick et al., 2000). In agreement, naptic mGlu2/3 receptors negatively modulate glutamate release
mice with the ␤2 subunit knocked out do not self-administer nico- (Schoepp et al., 2003). Cholinergic and glutamatergic inputs from
tine (Pons et al., 2008; Picciotto et al., 1998), and nicotine does not the pedunculopontine tegmental nucleus (PPTg) to the VTA seem to
enhance dopamine levels in the NAc in these mutants (Picciotto et play a crucial role in nicotine reward and reinforcement because a
al., 1998). The specific location of nAChRs containing the ␤2 subunit complete lesion of the PPTg reduces nicotine self-administration
in the VTA plays a crucial role in the mediation of nicotine reinforce- behavior (Lança et al., 2000; Picciotto and Corrigall, 2002) and
ment as demonstrated by studies re-expressing this subunit by blocks conditioned place preference induced by the intra-VTA
injecting a lentiviral vector into the VTA of mice carrying ␤2 subunit injection of nicotine (Laviolette et al., 2002). In contrast with these
deletions (Maskos et al., 2005). In contrast, re-expressing the ␤2 results, intravenous nicotine self-administration was significantly
subunit in the substantia nigra pars compacta instead of the VTA of elevated after injuring the posterior, but not anterior, portion of the
␤2 knockout mice did not yield nicotine self-administration (Pons PPTg (Alderson et al., 2006).
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 223

4.3. Involvement of cannabinoid receptors in nicotine reward, 4.5. Neurotransmitters involved in nicotine tolerance and
motivation and reinforcement withdrawal

A considerable number of studies demonstrate that the endo- The different neuronal adaptations occurring following
cannabinoid system is also involved in the rewarding and repeated exposure to nicotine, including desensitization and
reinforcing effects of nicotine (Maldonado et al., 2006; Scherma et up-regulation of nAChRs (Quick and Lester, 2002), are involved
al., 2008). Indeed, the selective CB1 receptor antagonist rimonabant in the development of nicotine tolerance and the appearance of
reduces nicotine self-administration in rats (Cohen et al., 2002) and a withdrawal syndrome following smoking cessation. Chronic
nicotine-induced conditioned place preference in rats and mice nicotine administration produces tolerance to most of its phar-
(Le Foll and Goldberg, 2004; Merritt et al., 2008). In addition, in macological effects (Benowitz, 2008), such as hypolocomotion,
vivo microdialysis and voltammetry studies have revealed that convulsive effects, hypothermia and antinociception, whereas
rimonabant pre-treatment blocks nicotine-enhanced dopamine no tolerance has been observed to its effects on memory and
extracellular levels in the NAc (Cohen et al., 2002; Cheer et al., attention (Marks et al., 1986; Miner and Collins, 1988; Collins et
2007) and in the bed nucleus of the stria terminalis (Cheer et al., 1988; Damaj and Martin, 1996; Benowitz, 2008). In humans,
al., 2007). Nicotine conditioned place preference was also absent cessation of tobacco intake precipitates both somatic and affective
in knockout mice lacking CB1 receptors (Castañé et al., 2002; symptoms of withdrawal, which may include severe craving for
Merritt et al., 2008). Based on the behavioral and biochemical nicotine, irritability, anxiety, loss of concentration, restlessness,
results obtained in rodents, several clinical trials were devel- decreased heart rate, depressed mood, impatience, insomnia,
oped to evaluate the efficacy of rimonabant for smoking cessation increased appetite and weight gain (Hughes and Hatsukami,
(STRATUS, studies with rimonabant and tobacco use) (Cahill and 1986; Hughes, 2007). In rodents, nicotine withdrawal is also
Ussher, 2007). Rimonabant significantly reduced smoking in two characterized by the manifestation of both somatic signs and affec-
clinical trials (STRATUS-NORTH AMERICA and STRATUS-WORLD tive changes similar to those observed in humans. The somatic
WIDE), although this effect was not significant in the STRATUS- signs include teeth chattering, palpebral ptosis, tremor, wet
EUROPE trial. The different clinical trials performed with this dog shakes, changes in locomotor activity and other behavioral
cannabinoid antagonist on smoking cessation, obesity and type signs (Malin et al., 1992). The mechanisms and brain regions
II diabetes have reported several gastrointestinal and psychiatric underlying nicotine physical dependence have not yet been
side effects including nausea, anxiety and depression. Due to these clarified, although an involvement of the medial habenula and
psychiatric side effects, the European Medicines Agency (EMEA) the interpeduncular nucleus has been recently reported (Salas
recommended the suspension of rimonabant marketing on October et al., 2009). The affective changes associated with nicotine
23rd, 2008. withdrawal include increased anxiety-like behavior, aversive
effects revealed in the place conditioning paradigm (Balerio et
al., 2004; Jackson et al., 2008), and reward deficits demonstrated
by intracranial self-stimulation techniques (Johnson et al., 2008).
4.4. Other neurotransmitters involved in nicotine reward, Similar to other drugs of abuse, hyperactivity of corticotropin-
motivation and reinforcement releasing-factor (CRF) neurons (Bruijnzeel et al., 2007) and c-fos
induction in the central amygdala (Panagis et al., 2000), as well
Other neurotransmitters are also involved in the rewarding as decreased dopaminergic activity in the NAc (Hildebrand et
and reinforcing properties of nicotine. The serotonergic (5-HT) al., 1999), have been related to the negative affective states
system, mainly through the 5-HT2c receptor subtype, seems to associated with nicotine withdrawal. In addition, glutamate also
play a role in nicotine reward by exerting an inhibitory influ- contributes to the modulation of the negative affective symptoms
ence on dopaminergic activity in the VTA (Di Matteo et al., 1999). of nicotine abstinence. Thus, mGlu2/3 receptor antagonists, which
Thus, the 5-HT2c agonist, Ro 60-0175 reduces intravenous nico- increase extracellular glutamate in the NAc, attenuate reward
tine self-administration (Grottick et al., 2001), although the same deficits associated with nicotine withdrawal in rodents and could
compound also attenuated the response to food. In contrast, no also alleviate the depressive-like symptoms related to nicotine
modification of nicotine-induced conditioned place preference was abstinence in humans (Kenny et al., 2003; Liechti and Markou,
observed by the 5-HT2c agonist, WAY 161503 in a recent report 2008).
(Hayes et al., 2009). On the other hand, tobacco smoke con-
tains monoamine oxidase (MAO) inhibitors, which are thought 4.6. Neurotransmitters involved in nicotine relapse
to enhance the reinforcing effects of nicotine. Behavioral stud-
ies confirmed that nicotine self-administration was facilitated Animal models of relapse have shown that nicotine-seeking can
in rats pre-treated with the irreversible and non-selective MAO be triggered by nicotine-associated (conditioned) cues (Martín-
inhibitor tranylcypromine (Villégier et al., 2006, 2007). Recently, García et al., 2009), stressors (Bilkei-Gorzo et al., 2008) (e.g.,
the hypothalamic neuropeptides hypocretins acting in the insula mild footshocks), and re-exposure to nicotine (Dravolina et al.,
have also been involved in nicotine reward (Hollander et al., 2007), which are the same events that trigger nicotine craving and
2008). Dopamine neurons in the VTA are under inhibitory con- relapse in humans. The neurobiological mechanisms involved in
trol of GABAergic inputs that also participate in nicotine-rewarding the processes underlying relapse to nicotine-seeking are poorly
effects. Hence, the administration of the GABA-B receptor agonists, understood. Recent studies have reported that GABA, glutamate
baclofen and CGP44532, as well as several GABA-B receptor posi- and endocannabinoids could play a role in this phenomenon. Thus,
tive allosteric modulators, decrease nicotine self-administration in the acute administration of the GABA-B receptor agonist, CGP44532
rats (Paterson et al., 2004, 2008). Baclofen also inhibits the expres- decreases cue-induced reinstatement of nicotine-seeking behavior
sion of nicotine-induced conditioned place preference in rats (Le (Paterson et al., 2005). In agreement, the GABA-B agonist baclofen
Foll et al., 2008). Although GABA neurons are acutely activated by prevents the reinstatement of nicotine-seeking and conditioned
nicotine, following repeated exposure to this drug, ␣4 ␤2 nAChRs place-preference triggered by nicotine priming in rats (Fattore et al.,
located on GABA cells tend to desensitize (Mansvelder et al., 2002), 2009). Glutamate is also involved in this process since the systemic
contributing to the final activation of mesolimbic dopamine neu- administration of the mGlu5 receptor antagonist, MPEP (Bespalov
rons. et al., 2005) or the mGlu2/3 receptor agonist, LY-379268 (Liechti et
224 F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

al., 2007) decreases cue-induced reinstatement of nicotine-seeking opposing their rewarding effects (see Shippenberg et al., 2007 for
in rats. review). In agreement, KOR agonists induce dysphoria in humans
Recent evidence suggests the involvement of the endocannabi- and aversive effects in animal models (Hasebe et al., 2004).
noid system in relapse to nicotine-seeking behavior (De Vries
and Schoffelmeer, 2005). Pre-treatment with the CB1 antago-
6. Effects of nicotine administration on the endogenous
nist, rimonabant attenuates the reinstatement of nicotine-seeking
opioid system
behavior induced by nicotine-associated cues in rats (Cohen et al.,
2005). This cannabinoid antagonist also reduces reinstatement of
The acute administration of nicotine promotes the release of
operant nicotine-seeking behavior induced by associated cues (De
endogenous opioid peptides and induces changes in the expres-
Vries et al., 2005) and reinstatement of nicotine conditioned place-
sion of these peptides. Accordingly, an increase in met-enkephalin
preference provoked by nicotine priming (Biala et al., 2009). The
(Dhatt et al., 1995) and dynorphin (Isola et al., 2009) immunore-
cannabinoid antagonist, AM251 also dose-dependently reduces
activities was found in the striatum of mice after acute nicotine
reinstatement produced by the combination of nicotine-associated
injection. An enhancement of ␤-endorphin secretion was also
cues and a nicotine priming dose (Shoaib, 2008).
shown in vitro after nicotine exposure (Marty et al., 1985), although
in vivo microdialysis did not reveal any changes in the rat NAc
after acute nicotine administration (Olive et al., 2001). In humans,
5. The endogenous opioid system
increased peripheral concentrations of ␤-endorphin have been
observed in persons who smoked fewer than 10 cigarettes per
Multiple studies in animal models and humans indicate that the
day (del Arbol et al., 2000). However, changes in the peripheral
endogenous opioid system is an important neurobiological sub-
levels of ␤-endorphin do not seem to be directly related to the
strate for the addictive properties of nicotine. The endogenous
central activity of the endogenous opioid system and have been
opioid peptides and receptors are largely distributed within the
more likely related to a peripheral response to stress (Rasmussen
CNS and are also present in several peripheral tissues. This wide
and Farr, 2009; Spaziante et al., 1990). On the other hand, acute
distribution is related to the important functions played by this
nicotine administration increases PENK mRNA expression in brain
neuromodulatory system in the control of several physiological
regions related to drug-dependence and withdrawal such as stria-
responses including nociception, emotional behavior, learning and
tum, amygdala, periaqueductal gray matter, locus coeruleus, raphe
memory, and reward processes (Bodnar, 2008). Three different sub-
and hippocampus (Dhatt et al., 1995; Houdi et al., 1991, 1998).
types of opioid receptors, mu (MOR), delta (DOR) and kappa (KOR),
PDYN mRNA expression was also increased in mouse striatum and
have been identified, cloned and characterized at the molecular,
NAc after the administration of a single dose of nicotine (Isola et al.,
biochemical and pharmacological level (Kieffer and Evans, 2009).
2009).
A fourth member of the opioid peptide receptor family, the N/OFQ
The effects of repeated exposure to nicotine on the differ-
receptor, was cloned in 1994 (Mollereau et al., 1994). Three families
ent components of the endogenous opioid system have also been
of endogenous opioid peptides derived from either proopiome-
investigated. Thus, chronic nicotine administration decreases PENK
lanocortin (POMC), proenkephalin (PENK) or prodynorphin (PDYN)
mRNA expression in striatum and hippocampus of rats (Houdi et
have also been identified and cloned (Xue and Domino, 2008). These
al., 1998) and up-regulates MOR in the striatum (Wewers et al.,
precursors generate several active peptides including ␤-endorphin,
1999). Conversely, a recent study shows down-regulation of MOR
met- and leu-enkephalins, and dynorphins, respectively (Kieffer
in the hippocampus and striatum following subchronic nicotine in
and Gavériaux-Ruff, 2002) that exhibit different affinities for each
rats (Marco et al., 2007). In C57BL/6J mice, repeated administra-
opioid receptor. ␤-endorphin binds with a higher affinity to MOR
tion of nicotine decreases MOR levels in the caudate-putamen, as
than DOR or KOR, and it is a main endogenous ligand for MOR.
well as in the core and shell of the NAc (Galeote et al., 2006). On
The affinity of met- and leu-enkephalin for DOR is 20-fold greater
the contrary, no changes in number, affinity or functional activity
than that for MOR, and dynorphins are presumed to be the main
of MOR were found following chronic nicotine exposure in NMRI
endogenous ligands for KOR (Roth-Deri et al., 2008). More recently,
mice (Vihavainen et al., 2008). The contradictory results found in
another endogenous opioid peptide named nociceptin/orphanin FQ
these studies might be attributed to the different routes of nico-
(N/OFQ) (Meunier et al., 1995; Reinscheid et al., 1995) has been
tine administration (subcutaneous vs. oral), doses (1.75 mg/kg vs.
discovered and proposed as the endogenous ligand of the N/OFQ
50–500 ␮g/ml), and strain of mice (C57BL/6J vs. NMRI) used.
receptor. Two putative endogenous opioid peptides, endomorphin-
1 and -2, have also been proposed as MOR selective endogenous
opioids (Zadina et al., 1997). However, no gene, precursor pro- 7. The role of the endogenous opioid system in nicotine
tein, or other mechanism for their endogenous synthesis has been reward, motivation and reinforcement
identified.
Opioid receptors and their endogenous ligands play an impor- Neurochemical studies in rodents have shown that nicotine-
tant role in brain reward processes, and modulate the behavioral induced dopamine release in the NAc is modulated by the activation
effects of nicotine and other drugs of abuse. Accordingly, opioid of MOR in the VTA (Tanda and Di Chiara, 1998). Moreover, the
receptors and their endogenous ligands are highly expressed in increase of dopamine in the NAc induced by nicotine can be blocked
areas of the brain involved in reward such as the VTA, NAc, pre- by the administration of the MOR antagonist/KOR agonist, cycla-
frontal cortex and extended amygdala (Delfs et al., 1994; Mansour zocine (Maisonneuve and Glick, 1999). Therefore, nicotine-induced
et al., 1995). Opioid receptor agonists also induce rewarding and/or dopamine release in the NAc depends on the integrity of the
aversive effects. Thus, morphine and other MOR agonists are endogenous opioid system. Accordingly, the enhancement of extra-
self-administered by humans and experimental animals whereas cellular levels of dopamine induced in the NAc by acute nicotine
systemic administration of the preferential MOR antagonists nalox- administration was decreased in mice lacking PENK (Berrendero et
one and naltrexone produces dysphoria in humans and conditioned al., 2005). In addition, a recent study has highlighted the existence
aversive effects in animals (Shippenberg et al., 2008). DOR agonists of a mechanistic overlap between the effects of nicotine and opiates
also induce rewarding effects in animal models (Suzuki et al., 1997; within the dopamine reward pathway (Britt and McGehee, 2008).
Hutcheson et al., 2001). On the contrary, the dynorphin/KOR sys- Pharmacological studies have also provided evidence for the
tem participates in the addictive properties of drugs of abuse by involvement of the endogenous opioid system, and more specifi-
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 225

Table 1
Changes on nicotine addictive properties in opioid receptor or peptide precursor knockout mice.

Knockout mice Behavioral/molecular response Effect Reference

␮-Opioid receptor Conditioned place preference Attenuation Berrendero et al. (2002)


Physical dependence Attenuation Berrendero et al. (2002)
CREB phosphorylation Suppression Walters et al. (2005)
Sensitization Suppression Yoo et al. (2004)
Neuronal nitric oxide synthase expression Attenuation Yoo et al. (2005)
Tolerance Faster development Galeote et al. (2006)

Pre-proenkephalin gene Conditioned place preference Attenuation Berrendero et al. (2005)


Physical dependence Attenuation Berrendero et al. (2005)
DA release in NAc Attenuation Berrendero et al. (2005)

␤-Endorphin Conditioned place preference Attenuation Trigo et al. (2009)


Physical dependence No change Trigo et al. (2009)

Prodynorphin Conditioned place preference No change Galeote et al. (2009)


Physical dependence No change Galeote et al. (2009)
Self-administration Enhanced sensitivity and acquisition Galeote et al. (2009)

Nociceptin Two-bottle choice Increased intake Sakoori and Murphy (2009)


Conditioned place aversion Manifestation Sakoori and Murphy (2009)

cally for MOR in nicotine-rewarding properties (Table 2). Thus, the has provided new information about the specific involvement of
administration of the MOR antagonist, glycyl-glutamine blocked each component of the endogenous opioid system in nicotine
nicotine conditioned place preference (Göktalay et al., 2006). reward. However, additional studies using conditional knockout
Furthermore, the effectiveness of lobeline, a potential pharma- mice with spatially and/or temporally restricted deletions will be
cotherapy for nicotine addiction, in reducing nicotine evoked required to exclude the possible influence of any compensatory
dopamine overflow from striatum of rats (Miller et al., 2000) seems change and to clarify the specific neural pathways involved in these
to be mediated by its ability to antagonize MOR (Miller et al., 2007). responses.
On the other hand, the role of DOR in nicotine-rewarding effects has The integrity of the endogenous opioid system is also criti-
not yet been clarified, and this receptor does not seem to participate cal for the expression of the molecular responses related to the
in nicotine sensitization since the administration of the DOR antag- rewarding properties of nicotine. Thus, CREB phosphorylation in
onist naltrindole did not modify locomotor sensitization induced by response to either nicotine administration or its associated cues
nicotine (Heidbreder et al., 1996). Nevertheless, the administration was inhibited in knockout mice lacking MOR (Walters et al., 2005).
of KOR agonists blocks the expression of nicotine sensitization, and The same inhibition of CREB phosphorylation was observed in
this effect was reversed by the KOR antagonist nor-binaltorphimine wild-type mice following the administration of naloxone, which
(Hahn et al., 2000). These pharmacological studies confirm the par- blocked the behavioral expression of nicotine conditioned place
ticipation of the endogenous opioid system in nicotine reward. preference (Walters et al., 2005). In agreement, ␤-arrestin-2, which
However, the useful data provided by these pharmacological tools regulates the activity of several receptors including MOR (Bohn
are limited by the selectivity of the ligands and the lack of informa- et al., 1999; Oakley et al., 2000; Zheng et al., 2008), partici-
tion about the neural pathways involved when using the systemic pates in the induction and expression of nicotine sensitization and
route. the effects induced by nicotine on the BDNF content in the NAc
Studies using genetically modified mice have clarified the dif- (Correll et al., 2009). Further supporting these data, genome-wide
ferential involvement of the diverse opioid peptides and receptors association studies have revealed a significant association among
in the rewarding properties of nicotine (Table 1). Thus, nicotine- ␤-arrestin-1 and -2 gene variants with nicotine dependence in a
induced conditioned place preference was blunted in mice lacking subpopulation of smokers of European origin in America (Sun et
either the MOR or PENK gene (Berrendero et al., 2002, 2005) and al., 2008).
was attenuated in mice lacking ␤-endorphin (Trigo et al., 2009), Human imaging studies using positron emission tomography
indicating that the activation of MOR by endogenous enkephalins have shown that nicotine activates MORs in the rostral and dorsal
and ␤-endorphin is required to obtain this nicotine effect. Nev- anterior cingulate cortex (Scott et al., 2007). These brain struc-
ertheless, the use of these conventional knockout mice does not tures are closely related to the anticipation of reward (Kilts et
allow to distinguish effects on the acquisition and expression of the al., 2001) and the discrimination between potentially rewarding
place conditioning which could have different psychological impli- and non-rewarding outcomes (Knutson et al., 2003). In agreement,
cations. On the other hand, the presence of MOR is also required association analysis in humans has revealed that some polymor-
for the development and expression of behavioral sensitization phisms in the MOR gene are involved in smoking initiation and
to nicotine in mice (Yoo et al., 2004). The kappa/dynorphin sys- dependence (Zhang et al., 2006), and smoking reward has also been
tem seems to play an opposite role in nicotine reward. Thus, mice associated with MOR genetic variance (Perkins et al., 2008). There-
lacking PDYN showed an enhanced sensitivity to nicotine self- fore, the efficacy of naltrexone on smoking cessation in humans
administration, which suggests that opioid peptides derived from (see Section 9) might be due to the ability of this opioid antagonist
PDYN would mediate nicotine aversive effects (Galeote et al., 2009). to reduce the reinforcing effects of nicotine (Rukstalis et al., 2005;
Accordingly, opioid peptides derived from PDYN also participate Wewers et al., 1998).
in the aversive effects of other drugs of abuse, such as cocaine
(Shippenberg et al., 2007) and cannabinoids (Mendizábal et al., 8. The role of the endogenous opioid system in the
2006). Recent findings suggest that the opioid peptide N/OFQ is development of nicotine tolerance and physical dependence
involved in nicotine-rewarding effects. Indeed, mice lacking N/OFQ
show an increase in voluntary nicotine intake and a decrease The participation of the endogenous opioid system in the devel-
in behavioral sensitization to low doses of nicotine (Sakoori and opment of tolerance to the pharmacological effects of nicotine
Murphy, 2009). The use of these conventional knockout mice has been shown using pharmacological and genetic tools. Indeed,
226 F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

Table 2
Effects of pharmacological interventions on the opioid system on nicotine addictive properties.

Drug Behavioral/molecular response Dose Effect Animal Reference

Morphine Physical dependence expression 4.5 mg/kg, s.c. Suppression Rat Malin et al. (1993)
Physical dependence development 3 mg/kg, s.c. Suppression Rat Ise et al. (2000)
Cross-tolerance 12.5 or 25 mg/kg, s.c. Manifestation Rat Zarrindast et al. (2003)
Tolerance 50 mg/kg s.c. Enhanced Mice Biala and Weglinska (2006)

Naloxone Withdrawal syndrome 4.5 mg/kg, s.c. Manifestation Rat Malin et al. (1993)
Conditioned place preference 0.5, 1 and 2 mg/kg, i.p. Suppression Mice Zarrindast et al. (2003)
Conditioned place aversion 0.5 and 1 mg/kg, i.p. Attenuation Mice Zarrindast et al. (2003)
Physical dependence expression 3 mg/kg, s.c. Manifestation Mice Balerio et al. (2004)
Withdrawal syndrome 1 mg/kg, i.p. Manifestation Mice Biala et al. (2005)
Conditioned place preference 1 mg/kg, s.c. Suppression Mice Walters et al. (2005)
CREB phosphorylation 1 mg/kg, s.c. Suppression Mice Walters et al. (2005)

TAN-67 Physical dependence 56 mg/kg, s.c. Suppression Rat Ise et al. (2000)
Cyclazocine DA release in NAc 0.5 mg/kg, i.p. Suppression Rat Maisonneuve and Glick (1999)
Glycyl-glutamine Conditioned place preference 100 nmol i.c.v. Suppression Rat Göktalay et al. (2006)
Physical dependence expression 100 nmol i.c.v. Attenuation Rat Göktalay et al. (2006)
Naltrindole Sensitization 0.3, 1.0 mg/kg s.c. No change Rat Heidbreder et al. (1996)

U50,488H Hyperlocomotion 3 mg/kg s.c. Suppression Rat Hahn et al. (2000)


Conditioned place aversion 1.0 mg/kg, s.c. Attenuation Rat Ise et al. (2002)
Cross-tolerance 10 mg/kg, s.c. Manifestation Mice Galeote et al. (2008)

U69,593 Hyperlocomotion 0.32 mg/kg i.p. Suppression Rat Hahn et al. (2000)
CI-977 Hyperlocomotion 0.02 mg/kg s.c. Suppression Rat Hahn et al. (2000)
TRK-820 Conditioned place aversion 0.03 mg/kg, s.c. Attenuation Rat Ise et al. (2002)

cross-tolerance between the antinociceptive effects of nicotine and withdrawal and the somatic manifestations of withdrawal to both
morphine has been reported (Biala and Weglinska, 2006). In addi- drugs of abuse share several common signs, such as abdominal
tion, an increase in MOR activation of G-proteins has been found constriction (writhes), facial fasciculation, eyeblink, ptosis, escape
in the spinal cord during chronic nicotine treatment in tolerant attempts, foot licks, genital grooming, shakes, tremors, scratches,
mice, and MOR knockout mice developed faster tolerance to nico- teeth chattering and difficult breathing (Malin et al., 1993, 1996;
tine antinociceptive effects than wild-type animals (Galeote et al., Hildebrand et al., 1997). These pharmacological results suggest
2006). These results suggest that the increased activation of MORs the existence of common mechanisms in the regulation of opi-
during chronic nicotine administration could be an adaptive mech- oid and nicotine physical dependence. Naloxone was even more
anism to counteract the establishment of tolerance to nicotine effective than the nAChR antagonist mecamylamine in precipitat-
antinociception. KORs have also been reported to participate in the ing the negative emotional manifestations of nicotine withdrawal
development of nicotine tolerance. Thus, a decrease in the density measured in the place conditioning paradigm in mice (Balerio et
of KORs has been shown in all layers of the spinal cord in nico- al., 2004). However, the opioid antagonists naloxone (Tome et al.,
tine tolerant mice, and bi-directional cross-tolerance between the 2001) and naltrexone (Almeida et al., 2000) also show some affin-
antinociceptive effects of nicotine and the KOR agonist U50,488H ity to nAChRs, which could have some influence in the behavioral
has been reported during chronic treatment with these compounds effects induced by these antagonists in nicotine-dependent animals
(Galeote et al., 2008). This down-regulation of KOR in the spinal and humans. On the other hand, the administration of MOR, DOR or
cord seems to contribute to the development of nicotine tolerance KOR agonists reduces the capability of mecamylamine to produce
since these receptors are necessary for a complete manifestation of aversive manifestations of nicotine withdrawal in the place condi-
nicotine antinociception (Galeote et al., 2008). Therefore, the adap- tioning paradigm (Ise et al., 2000, 2002). A recent study has shown
tive changes occurring during chronic nicotine administration in an increase in PDYN expression in the NAc in nicotine-abstinent
MORs and KORs seem to play an opposite role in the development rodents (Isola et al., 2008) that could be related to the emergence
of tolerance to nicotine antinociception. of the negative affective states observed during withdrawal. In
Chronic nicotine exposure induces adaptive changes in the contrast to the attenuation of mecamylamine-precipitated nico-
CNS that lead to physical dependence and withdrawal mani- tine withdrawal aversion by KOR agonists (Ise et al., 2002), this
festations. Both human and animal studies have reported the last study suggests that KOR antagonists might be considered for
involvement of the endogenous opioid system in the manifes- the amelioration of the negative affective states related to nicotine
tations of nicotine withdrawal. In human studies, the opioid abstinence.
antagonist naloxone has been found to induce somatic signs of Studies using knockout mice have confirmed the role of the
nicotine withdrawal in heavy chronic smokers (Sutherland et al., endogenous opioid system in the physical manifestations of
1995; Krishnan-Sarin et al., 1999). The administration of opioid nicotine withdrawal. Thus, the somatic expression of nicotine
antagonists in human smokers currently causes an initial com- abstinence precipitated by mecamylamine in dependent mice was
pensatory increase in nicotine intake to overcome the blockade significantly attenuated in mice lacking the MOR (Berrendero et al.,
of some of its behavioral effects (Sutherland et al., 1995), sim- 2002) or the PENK gene (Berrendero et al., 2005). In contrast, the
ilar to what occurs with the co-administration of the nicotine absence of the PDYN (Galeote et al., 2009) or ␤-endorphin (Trigo et
antagonist mecamylamine (Pomerleau, 1998). In animal studies, al., 2009) gene did not modify the severity of nicotine withdrawal.
naloxone administration has been reported to precipitate a nico- These results indicate that endogenous enkephalins and the activa-
tine withdrawal syndrome in dependent mice (Balerio et al., 2004; tion of MOR are necessary for the manifestation of the somatic signs
Biala et al., 2005) and rats (Malin et al., 1993), whereas mor- of nicotine abstinence. In agreement, an increase in PENK mRNA
phine attenuates the severity of the spontaneous (Malin et al., levels has been reported in the striatum and hippocampus of rats
1993) or mecamylamine-precipitated nicotine withdrawal syn- during nicotine withdrawal (Houdi et al., 1998) as well as in the
drome (Ise et al., 2000). Nicotine abstinence resembles opiate striatum of nicotine-abstinent mice (Isola et al., 2002).
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 227

9. Efficacy of opioid antagonists in the treatment for specific genetic variations in these sub-populations, including the
smoking cessation MOR gene variant A118G, which has been linked to the relative
reinforcing value of nicotine (Lerman et al., 2004; Ray et al., 2006).
Numerous clinical trials have been performed to evaluate the In addition, the study of DOR and KOR ligands as possible therapies
effectiveness of the non-selective MOR antagonists, naloxone and for smoking cessation could also be of interest given the recog-
naltrexone, for smoking cessation in humans. However, the results nized role of these receptors in the behavioral responses induced
of early studies were not conclusive since only moderate or even by nicotine (Balerio et al., 2005; Hasebe et al., 2004).
negative effects were observed (Karras and Kane, 1980; Gorelick
et al., 1989; Nemeth-Coslett and Griffiths, 1986; Sutherland et al.,
1995; Covey et al., 1999; Wewers et al., 1998). Recent studies 10. Concluding remarks
have almost exclusively concentrated on naltrexone as a suit-
able opioid antagonist in smoking cessation treatment due to its Nicotine addiction is a complex disorder involving adaptive
longer acting effects. Thus, several clinical trials have shown that changes on nAChRs and alterations in other neurochemical sys-
the acute reinforcing value of nicotine is reduced by intermediate tems that modulate the behavioral responses to nicotine. The
doses of naltrexone (50 mg) during short-term nicotine abstinence glutamate, GABA and cannabinoid system are involved in the ini-
using cigarette-smoking choice paradigms (King and Meyer, 2000; tiation of nicotine addiction as they participate in the rewarding
Rukstalis et al., 2005). This effect was confirmed in a placebo- properties of this drug. Some of these neurochemical systems
controlled laboratory study, where subjects received one dose of also participate in the development of nicotine tolerance and
50 mg and 4 h later were given the option of smoking a cigarette at dependence, as well as relapse to nicotine seeking. The endoge-
half-hour intervals. In this study, naltrexone significantly decreased nous opioid system also plays a crucial role in several aspects of
the number of first and second cigarettes chosen (Epstein and the addictive process induced by nicotine. Hence, acute nicotine
King, 2004). Studies evaluating the potential efficacy of naltrex- administration enhances the levels of endogenous opioid peptides
one (50 mg for 8 weeks) as a co-adjunct of nicotine replacement in different brain structures, which has been related to several
therapy in smoking cessation show higher continuous abstinence nicotine pharmacological effects, including reward. Chronic nico-
rates for naltrexone (56%) as compared to placebo (31%) treatment tine exposure and withdrawal also modify endogenous opioid
(Krishnan-Sarin et al., 2003). However, these findings were not sup- peptide content in different brain structures in addition to the
ported in a randomized partially-blind 2 × 2 factorial design trial activity of MORs and KORs in the spinal cord. These changes pos-
since naltrexone at the dose of 50 mg for 12 weeks did not show a sibly mediate the development of nicotine tolerance and physical
significant effect on smoking cessation or craving as compared to dependence. Studies using pharmacological tools and genetically
placebo (Wong et al., 1999). These discrepancies could be due to modified mice have revealed that MOR is a crucial component for
a differential effectiveness of naltrexone depending on the specific nicotine-rewarding effects and for the development of nicotine
individual differences of smokers. In this sense, the dose of 50 mg tolerance and dependence. Endogenous enkephalins and beta-
of naltrexone appears to be more effective for smoking cessation endorphins are responsible for these rewarding effects of nicotine,
in female smokers when given alone (King et al., 2006) or in com- and enkephalins also participate in the development of nicotine
bination with nicotine replacement therapy (Byars et al., 2005). In physical dependence. The PDYN/KOR system plays an opposite
addition, treatment with this dose of naltrexone was related to bet- role to the other components of the endogenous opioid system in
ter rates of smoking cessation than placebo in smokers reporting modulating nicotine effects. Indeed, endogenous opioid peptides
higher rates of depressive symptoms (Walsh et al., 2008). These derived from PDYN are involved in nicotine aversive effects, which
studies suggest that individual differences regarding gender or the would counteract the rewarding responses of nicotine mediated by
incidence of mood disorders, such as depression/anxiety, should be enkephalins and beta-endorphins. KOR also plays an opposite role
taken into account in future clinical studies evaluating the response to MOR in the development of nicotine tolerance. Recent genetic
to naltrexone treatment. studies revealing the association between genetic variations of
One of the mayor issues in nicotine addiction is the high inci- MOR and the risk for smoking initiation and dependence sustain the
dence of relapse, mostly induced by environmental cues that have results of animal studies and provide an additional finding to sup-
been associated with cigarette smoking. Studies evaluating the port the involvement of MORs in the different aspects of nicotine
effects of naltrexone (50 mg single dose) in preventing nicotine addiction.
relapse have also shown mixed results. One trial shows blunt- Most of these animal studies have focused on the involvement of
ing of the urge to smoke and no negative affect following the the different components of the endogenous opioid system in nico-
smoking cues in subjects treated with naltrexone (Hutchison et al., tine tolerance, dependence and reward. However, addiction is not
1999), while another study found reduced cue-elicited withdrawal just the self-administration of drugs but rather constitutes a relaps-
symptoms, but no significant effects on the urges to smoke or ing disorder characterized by compulsive drug use despite adverse
deprivation-induced withdrawal prior to cue exposure (Rohsenow consequences for the user (Diagnostic and Statistical Manual of
et al., 2007). In addition, because naltrexone has some efficacy for Mental Disorders (DSM) of the American Psychiatric Association,
alcohol addiction, its effects on concomitant smoking and drinking 2000). New animal models are now available to evaluate other
behaviors have been examined. Thus, naltrexone (50 mg for 3 days) behavioral responses more directly related to the nicotine addictive
decreases the progression of craving for cigarettes during alco- process, such as reinstatement to nicotine seeking. The pharmaco-
hol intoxication in a laboratory double-blind placebo-controlled logical and genetic tools described in the present review, combined
design study (Ray et al., 2007). Conversely, another study found that with the use of these new animal models, will soon provide a bet-
low doses of naltrexone (25 mg) plus nicotine replacement therapy ter understanding of the neurobiological mechanisms involved in
reduces the risk of hazardous drinking in a placebo-controlled trial nicotine addiction. Furthermore, behavioral models of compulsive
for smoking cessation (O’Malley et al., 2008). In summary, mixed drug seeking that resemble the main diagnosis criteria for addiction
results have been obtained with respect to the effects of naltrexone in humans have been recently described in rodents (Deroche-
with or without nicotine replacement therapy on smoking ces- Gamonet et al., 2004; Vanderschuren and Everitt, 2004). However,
sation, probably due to differential effectiveness of naltrexone in these models were substantiated for cocaine consumption and are
different sub-populations of smokers. Future efforts for identifying not still available for nicotine. The future validation of these new
more suitable opioid targets may focus on evaluating the role of models for nicotine addiction will represent a definitive advance
228 F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

in the knowledge of the substrates involved in this complex patho- Biala, G., Budzynska, B., Staniak, N., 2009. Effects of rimonabant on the reinstatement
logical behavior. of nicotine-conditioned place preference by drug priming in rats. Behav. Brain
Res. 202, 260–265.
In humans, the clinical trials performed with the opioid antag- Biala, G., Weglinska, B., 2006. On the mechanism of cross-tolerance between
onist naltrexone have reported controversial results about its morphine- and nicotine-induced antinociception: involvement of calcium chan-
effectiveness in smoking cessation. These discrepancies could be nels. Prog. Neuropsychopharmacol. Biol. Psychiatry 30, 15–21.
Biala, G., Budzynska, B., Kruk, M., 2005. Naloxone precipitates nicotine abstinence
due to a differential effect of naltrexone depending on specific syndrome and attenuates nicotine-induced antinociception in mice. Pharmacol.
individual factors of smokers. In this sense, the rates of smoking ces- Rep. 57, 755–760.
sation after naltrexone treatment were better in smokers reporting Bierut, L.J., 2009. Nicotine dependence and genetic variation in the nicotinic recep-
tors. Drug Alcohol Depend. 104 (Suppl. 1), S64–69.
higher rates of depressive symptoms and females than in other sub- Bilkei-Gorzo, A., Rácz, I., Michel, K., Darvas, M., Maldonado, R., Zimmer, A., 2008. A
populations. New clinical trials focused on the specific populations common genetic predisposition to stress sensitivity and stress-induced nicotine
that present a better response to naltrexone will be necessary to craving. Biol. Psychiatry 63, 164–171.
Bodnar, R.J., 2008. Endogenous opiates and behavior: 2007. Peptides 29, 2292–2375.
clearly define the therapeutic value of this opioid antagonist for
Bohn, L.M., Lefkowitz, R.J., Gainetdinov, R.R., Peppel, K., Caron, M.G., Lin, F.T.,
smoking cessation, and further genetic studies can also be useful 1999. Enhanced morphine analgesia in mice lacking betaarrestin 2. Science 24,
to better identify these specific populations. The elucidation of the 2495–2498.
roles played by each specific component of the endogenous opi- Brennan, K., Lea, R., Fitzmaurice, P., Truman, P., 2009. Nicotinic receptors and stages
of nicotine dependence. J. Psychopharmacol. (epub ahead of print).
oid system in nicotine addiction represents an important advance Britt, J.P., McGehee, D.S., 2008. Presynaptic opioid and nicotinic receptor modulation
in the identification of new therapeutic targets for this disorder, of dopamine overflow in the nucleus accumbens. J. Neurosci. 28, 1672–1681.
which constitutes one of the main causes of preventable deaths in Bruijnzeel, A.W., Prado, M., Isaac, S., 2009. Corticotropin-releasing factor-1 recep-
tor activation mediates nicotine withdrawal-induced deficit in brain reward
developed countries. function and stress-induced relapse. Biol. Psychiatry 66, 110–117.
Bruijnzeel, A.W., Zislis, G., Wilson, C., Gold, M.S., 2007. Antagonism of CRF recep-
tors prevents the deficit in brain reward function associated with precipitated
Acknowledgements nicotine withdrawal in rats. Neuropsychopharmacology 32, 955–963.
Byars, J.A., Frost-Pineda, K., Jacobs, W.S., Gold, M.S., 2005. Naltrexone augments the
This work was supported by the Spanish “Ministerio de Cien- effects of nicotine replacement therapy in female smokers. J. Addict. Dis. 24,
49–60.
cia e Innovación” (#SAF2007-64062) and “Instituto de Salud Carlos Cahill, K., Ussher, M., 2007. Cannabinoid type 1 receptor antagonists (rimonabant)
III” (#RD06/001/001, PI070709 and PI070559), the Catalan Govern- for smoking cessation. Cochrane Database Syst. Rev. 4, CD005353.
ment (SGR2009-00131), the ICREA Foundation (ICREA Academia- Castañé, A., Valjent, E., Ledent, C., Parmentier, M., Maldonado, R., Valverde, O., 2002.
Lack of CB1 cannabinoid receptors modifies nicotine behavioural responses, but
2008), NIDA (1R01-DA01 6768-0111) and the DG Research of
not nicotine abstinence. Neuropharmacology 43, 857–867.
the European Commission (NEWMOOD LSHM-CT-2004-503474; Clarke, P.B.S., Reuben, M., 1996. Release of [3H]-noradrenaline from rat hippocampal
GENADDICT, #LSHM-CT-2004-05166; and PHECOMP, #LSHM-CT- synaptosomes by nicotine: mediation by different nicotinic receptor subtypes
from striatal [3H]-dopamine release. Br. J. Pharmacol. 117, 595–606.
2007-037669). J.M.T and E.M.G. are post-doctoral fellows of
Cohen, C., Perrault, G., Voltz, C., Steinberg, R., Soubrié, P., 2002. SR141716, a
Instituto de Salud Carlos III “Contratos posdoctorales de perfec- central cannabinoid (CB(1)) receptor antagonist, blocks the motivational and
cionamiento Sara Borrell”. dopamine-releasing effects of nicotine in rats. Behav. Pharmacol. 13, 451–463.
Cohen, C., Perrault, G., Griebel, G., Soubrié, P., 2005. Nicotine-associated cues main-
tain nicotine-seeking behavior in rats several weeks after nicotine withdrawal:
References reversal by the cannabinoid (CB1) receptor antagonist, rimonabant (SR141716).
Neuropsychopharmacology 30, 145–1145.
Alderson, H.L., Latimer, M.P., Winn, P., 2006. Intravenous self-administration of nico- Collins, A.C., Romm, E., Wehner, J.M., 1988. Nicotine tolerance: an analysis of the
tine is altered by lesions of the posterior, but not anterior, pedunculopontine time course of its development and loss in the rat. Psychopharmacology (Berl.)
tegmental nucleus. Eur. J. Neurosci. 23, 2169–2175. 96, 7–14.
Almeida, L.E., Pereira, E.F., Alkondon, M., Fawcett, W.P., Randall, W.R., Albu- Correll, J.A., Noel, D.M., Sheppard, A.B., Thompson, K.N., Li, Y., Yin, D., Brown, R.W.,
querque, E.X., 2000. The opioid antagonist naltrexone inhibits activity and alters 2009. Nicotine sensitization and analysis of brain-derived neurotrophic factor
expression of alpha7 and alpha4beta2 nicotinic receptors in hippocampal neu- in adolescent beta-arrestin-2 knockout mice. Synapse 63, 510–519.
rons: implications for smoking cessation programs. Neuropharmacology 39, Corrigall, W.A., Coen, K.M., 1989. Nicotine maintains robust self-administration in
2740–2755. rats on a limited-access schedule. Psychopharmacology (Berl.) 99, 473–478.
American Psychiatric Association, 2000. Diagnostic and Statistical Manual of Men- Covey, L.S., Glassman, A.H., Stetner, F., 1999. Naltrexone effects on short-term and
tal Disorders, 4 ed. American Psychiatric Association, Washington, DC (revised long-term smoking cessation. J. Addict. Dis. 18, 31–40.
version). Cheer, J.F., Wassum, K.M., Sombers, L.A., Heien, M.L., Ariansen, J.L., Aragona, B.J.,
Amos, C.I., Wu, X., Broderick, P., Gorlov, I.P., Gu, J., Eisen, T., Dong, Q., Zhang, Q., Gu, Phillips, P.E., Wightman, R.M., 2007. Phasic dopamine release evoked by abused
X., Vijayakrishnan, J., Sullivan, K., Matakidou, A., Wang, Y., Mills, G., Doheny, K., substances requires cannabinoid receptor activation. J. Neurosci. 27, 791–795.
Tsai, Y.Y., Chen, W.V., Shete, S., Spitz, M.R., Houlston, R.S., 2008. Genome-wide Childs, E., de Wit, H., 2009. Amphetamine-induced place preference in humans. Biol.
association scan of tag SNPs identifies a susceptibility locus for lung cancer at Psychiatry 65, 900–904.
15q25.1. Nat. Genet. 40, 616–622. Dajas-Bailador, F., Wonnacott, S., 2004. Nicotinic acetylcholine receptors and the
Balerio, G.N., Aso, E., Maldonado, R., 2005. Involvement of the opioid system in the regulation of neuronal signaling. Trends Pharmacol. Sci. 25, 317–324.
effects induced by nicotine on anxiety-like behavior in mice. Psychopharmacol- Damaj, M.I., Martin, B.R., 1996. Tolerance to the antinociceptive effect of epibatidine
ogy (Berl.) 181, 260–269. after acute and chronic administration in mice. Eur. J. Pharmacol. 300, 51–57.
Balerio, G.N., Aso, E., Berrendero, F., Murtra, P., Maldonado, R., 2004. Delta9- De Vries, T.J., Schoffelmeer, A.N., 2005. Cannabinoid CB1 receptors control condi-
tetrahydrocannabinol decreases somatic and motivational manifestations of tioned drug seeking. Trends Pharmacol. Sci. 26, 420–426.
nicotine withdrawal in mice. Eur. J. Neurosci. 20, 2737–2748. De Vries, T.J., de Vries, W., Janssen, M.C., Schoffelmeer, A.N., 2005. Suppression of con-
Benowitz, N.L., 2008. Clinical pharmacology of nicotine: implications for under- ditioned nicotine and sucrose seeking by the cannabinoid-1 receptor antagonist
standing, preventing, and treating tobacco addiction. Clin. Pharmacol. Ther. 83, SR141716A. Behav. Brain Res. 161, 164–168.
531–541. Decker, M.W., Brioni, I.D., Bannon, A.W., Arneric, S.P., 1995. Diversity of neuronal
Berrendero, F., Mendizabal, V., Robledo, P., Galeote, L., Bilkei-Gorzo, A., Zimmer, A., nicotinic acetylcholine receptors: lessons from behavior and implications for
Maldonado, R., 2005. Nicotine-induced antinociception, rewarding effects, and CNS therapeutics. Life Sci. 56, 545–570.
physical dependence are decreased in mice lacking the preproenkephalin gene. del Arbol, J.L., Muñoz, J.R., Ojeda, L., Cascales, A.L., Irles, J.R., Miranda, M.T., Ruiz
J. Neurosci. 25, 1103–1112. Requena, M.E., Aguirre, J.C., 2000. Plasma concentrations of beta-endorphin
Berrendero, F., Kieffer, B.L., Maldonado, R., 2002. Attenuation of nicotine-induced in smokers who consume different numbers of cigarettes per day. Pharmacol.
antinociception, rewarding effects, and dependence in mu-opioid receptor Biochem. Behav. 67, 25–28.
knock-out mice. J. Neurosci. 22, 10935–10940. Delfs, J.M., Kong, H., Mestek, A., Chen, Y., Yu, L., Reisine, T., Chesselet, M.F., 1994.
Berrettini, W., Yuan, X., Tozzi, F., Song, K., Francks, C., Chilcoat, H., Waterworth, D., Expression of mu opioid receptor mRNA in rat brain: an in situ hybridization
Muglia, P., Mooser, V., 2008. Alpha-5/alpha-3 nicotinic receptor subunit alleles study at the single cell level. J. Comp. Neurol. 345, 46–68.
increase risk for heavy smoking. Mol. Psychiatry 13, 368–373. Deroche-Gamonet, V., Belin, D., Piazza, P.V., 2004. Evidence for addiction-like behav-
Bespalov, A.Y., Dravolina, O.A., Sukhanov, I., Zakharova, E., Blokhina, E., Zvartau, E., ior in the rat. Science 305, 1014–1017.
Danysz, W., van Heeke, G., Markou, A., 2005. Metabotropic glutamate receptor Dhatt, R.K., Gudehithlu, K.P., Wemlinger, T.A., Tejwani, G.A., Neff, N.H., Hadji-
(mGluR5) antagonist MPEP attenuated cue- and schedule-induced reinstate- constantinou, M., 1995. Preproenkephalin mRNA and methionine-enkephalin
ment of nicotine self-administration behavior in rats. Neuropharmacology 49 content are increased in mouse striatum after treatment with nicotine. J. Neu-
(Suppl. 1), 167–178. rochem. 64, 1878–1883.
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 229

Di Chiara, G., Imperato, A., 1988. Drugs abused by humans preferentially increase Simonato, L., Lowry, R., Conway, D.I., Znaor, A., Healy, C., Zelenika, D., Boland,
synaptic dopamine concentrations in the mesolimbic system of freely moving A., Delepine, M., Foglio, M., Lechner, D., Matsuda, F., Blanche, H., Gut, I., Heath,
rats. Proc. Natl. Acad. Sci. U.S.A. 85, 5274–5278. S., Lathrop, M., Brennan, P., 2008. A susceptibility locus for lung cancer maps to
Di Matteo, V., Di Giovanni, G., Di Mascio, M., Esposito, E., 1999. SB 242084, a selective nicotinic acetylcholine receptor subunit genes on 15q25. Nature 452, 633–637.
serotonin2C receptor antagonist, increases dopaminergic transmission in the Hutcheson, D.M., Matthes, H.W., Valjent, E., Sánchez-Blázquez, P., Rodríguez-Díaz,
mesolimbic system. Neuropharmacology 38, 1195–1205. M., Garzón, J., Kieffer, B.L., Maldonado, R., 2001. Lack of dependence and reward-
Dravolina, O.A., Zakharova, E.S., Shekunova, E.V., Zvartau, E.E., Danysz, W., Bespalov, ing effects of deltorphin II in mu-opioid receptor-deficient mice. Eur. J. Neurosci.
A.Y., 2007. mGlu1 receptor blockade attenuates cue- and nicotine-induced 13, 153–161.
reinstatement of extinguished nicotine self-administration behavior in rats. Hutchison, K.E., Monti, P.M., Rohsenow, D.J., Swift, R.M., Colby, S.M., Gnys, M.,
Neuropharmacology 52, 263–269. Niaura, R.S., Sirota, A.D., 1999. Effects of naltrexone with nicotine replacement
Dwoskin, L.P., Smith, A.M., Wooters, T.E., Zhang, Z., Crooks, P.A., Bardo, M.T., 2009. on smoking cue reactivity: preliminary results. Psychopharmacology (Berl.) 142,
Nicotinic receptor-based therapeutics and candidates for smoking cessation. 139–143.
Biochem. Pharmacol. 78, 732–743. Ise, Y., Narita, M., Nagase, H., Suzuki, T., 2002. Modulation of kappa-opioidergic
Epstein, A.M., King, A.C., 2004. Naltrexone attenuates acute cigarette smoking behav- systems on mecamylamine-precipitated nicotine-withdrawal aversion in rats.
ior. Pharmacol. Biochem. Behav. 77, 29–37. Neurosci. Lett. 323, 164–166.
Fattore, L., Spano, M.S., Cossu, G., Scherma, M., Fratta, W., Fadda, P., 2009. Ise, Y., Narita, M., Nagase, H., Suzuki, T., 2000. Modulation of opioidergic system on
Baclofen prevents drug-induced reinstatement of extinguished nicotine- mecamylamine-precipitated nicotine-withdrawal aversion in rats. Psychophar-
seeking behaviour and nicotine place preference in rodents. Eur. Neuropsy- macology (Berl.) 151, 49–54.
chopharmacol. 19, 487–498. Isola, R., Zhang, H., Tejwani, G.A., Neff, N.H., Hadjiconstantinou, M., 2009. Acute
Fu, Y., Matta, S.G., Gao, W., Brower, V.G., Sharp, B.M., 2000. Systemic nicotine stim- nicotine changes dynorphin and prodynorphin mRNA in the striatum. Psy-
ulates dopamine release in nucleus accumbens: re-evaluation of the role of chopharmacology 201, 507–516.
N-methyl-D-aspartate receptors in the ventral tegmental area. J. Pharmacol. Exp. Isola, R., Zhang, H., Tejwani, G.A., Neff, N.H., Hadjiconstantinou, M., 2008. Dynorphin
Ther. 294, 458–465. and prodynorphin mRNA changes in the striatum during nicotine withdrawal.
Galeote, L., Berrendero, F., Bura, S.A., Zimmer, A., Maldonado, R., 2009. Prodynorphin Synapse 62, 448–455.
gene disruption increases the sensitivity to nicotine self-administration in mice. Isola, R., Zhang, H., Duchemin, A.M., Tejwani, G.A., Neff, N.H., Hadjiconstantinou, M.,
Int. J. Neuropsychopharmacol. 12, 615–625. 2002. Met-enkephalin and preproenkephalin mRNA changes in the striatum of
Galeote, L., Maldonado, R., Berrendero, F., 2008. Involvement of kappa/dynorphin the nicotine abstinence mouse. Neurosci. Lett. 325, 67–71.
system in the development of tolerance to nicotine-induced antinociception. J. Jackson, K.J., McIntosh, J.M., Brunzell, D.H., Sanjakdar, S.S., Damaj, M.I., 2009. The
Neurochem. 105, 1358–1368. role of alpha6-containing nicotinic acetylcholine receptors in nicotine reward
Galeote, L., Kieffer, B.L., Maldonado, R., Berrendero, F., 2006. Mu-opioid receptors are and withdrawal. J. Pharmacol. Exp. Ther. 331, 547–554.
involved in the tolerance to nicotine antinociception. J. Neurochem. 97, 416–423. Jackson, K.J., Martin, B.R., Changeux, J.P., Damaj, M.I., 2008. Differential role of
Göktalay, G., Cavun, S., Levendusky, M.C., Hamilton, J.R., Millington, W.R., 2006. nicotinic acetylcholine receptor subunits in physical and affective nicotine with-
Glycyl-glutamine inhibits nicotine conditioned place preference and with- drawal signs. J. Pharmacol. Exp. Ther. 325, 302–312.
drawal. Eur. J. Pharmacol. 530, 95–102. Johnson, P.M., Hollander, J.A., Kenny, P.J., 2008. Decreased brain reward function
Goldberg, S.R., Spealman, R.D., Goldberg, D.M., 1981. Persistent behavior at high during nicotine withdrawal in C57BL6 mice: evidence from intracranial self-
rates maintained by intravenous self-administration of nicotine. Science 214, stimulation (ICSS) studies. Pharmacol. Biochem. Behav. 90, 409–415.
573–575. Karras, A., Kane, J.M., 1980. Naloxone reduces cigarette smoking. Life Sci. 27,
Gorelick, D.A., Rose, J., Jarvik, M., 1989. Effect of naloxone on cigarette smoking. J. 1541–1545.
Subst. Abuse 1, 153–159. Kenny, P.J., Chartoff, E., Roberto, M., Carlezon Jr., W.A., Markou, A., 2009. NMDA
Grottick, A.J., Corrigall, W.A., Higgins, G.A., 2001. Activation of 5-HT(2C) receptors receptors regulate nicotine-enhanced brain reward function and intravenous
reduces the locomotor and rewarding effects of nicotine. Psychopharmacology nicotine self-administration: role of the ventral tegmental area and central
(Berl.) 157, 292–298. nucleus of the amygdala. Neuropsychopharmacology 34, 266–281.
Grottick, A.J., Trube, G., Corrigall, W.A., Huwyler, J., Malherbe, P., Wyler, R., Higgins, Kenny, P.J., Gasparini, F., Markou, A., 2003. Group II metabotropic and alpha-
G.A., 2000. Evidence that nicotinic alpha(7) receptors are not involved in the amino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA)/kainate glutamate
hyperlocomotor and rewarding effects of nicotine. J. Pharmacol. Exp. Ther. 294, receptors regulate the deficit in brain reward function associated with nicotine
1112–1119. withdrawal in rats. J. Pharmacol. Exp. Ther. 306, 1068–1076.
Hahn, B., Stolerman, I.P., Shoaib, M., 2000. Kappa-opioid receptor modulation of Kieffer, B.L., Gavériaux-Ruff, C., 2002. Exploring the opioid system by gene knockout.
nicotine-induced behaviour. Neuropharmacology 39, 2848–2855. Prog. Neurobiol. 66, 285–306.
Hasebe, K., Kawai, K., Suzuki, T., Kawamura, K., Tanaka, T., Narita, M., Nagase, H., Kieffer, B.L., Evans, C.J., 2009. Opioid receptors: from binding sites to visible
Suzuki, T., 2004. Possible pharmacotherapy of the opioid kappa receptor agonist molecules in vivo. Opioid receptors: from binding sites to visible molecules in
for drug dependence. Ann. N.Y. Acad. Sci. 1025, 404–413. vivo. Neuropharmacology 56 (Suppl. 1), 205–212.
Hatsukami, D.K., Stead, L.F., Gupta, P.C., 2008. Tobacco addiction. Lancet 371, Kilts, C.D., Schweitzer, J.B., Quinn, C.K., Gross, R.E., Faber, T.L., Muhammad, F., Ely,
2027–2038. T.D., Hoffman, J.M., Drexler, K.P., 2001. Neural activity related to drug craving in
Hayes, D.J., Mosher, T.M., Greenshaw, A.J., 2009. Differential effects of 5-HT2C recep- cocaine addiction. Arch. Gen. Psychiatry 58, 334–341.
tor activation by WAY 161503 on nicotine-induced place conditioning and King, A., de Wit, H., Riley, R.C., Cao, D., Niaura, R., Hatsukami, D., 2006. Efficacy of
locomotor activity in rats. Behav. Brain Res. 197, 323–330. naltrexone in smoking cessation: a preliminary study and an examination of sex
Heidbreder, C., Shoaib, M., Shippenberg, T.S., 1996. Differential role of delta-opioid differences. Nicotine Tob. Res. 8, 671–682.
receptors in the development and expression of behavioral sensitization to King, A.C., Meyer, P.J., 2000. Naltrexone alteration of acute smoking response in
cocaine. Eur. J. Pharmacol. 298, 207–216. nicotine-dependent subjects. Pharmacol. Biochem. Behav. 66, 563–572.
Hildebrand, B.E., Panagis, G., Svensson, T.H., Nomikos, G.G., 1999. Behavioral and bio- Knutson, B., Fong, G.W., Bennett, S.M., Adams, C.M., Hommer, D., 2003. A region of
chemical manifestations of mecamylamine-precipitated nicotine withdrawal in mesial prefrontal cortex tracks monetarily rewarding outcomes: characteriza-
the rat: role of nicotinic receptors in the ventral tegmental area. Neuropsy- tion with rapid event-related fMRI. Neuroimage 18, 263–272.
chopharmacology 21, 560–574. Koob, G.F., Le Moal, M., 2008. Review. Neurobiological mechanisms for opponent
Hildebrand, B.E., Nomikos, G.G., Bondjers, C., Nisell, M., Svensson, T.H., 1997. Behav- motivational processes in addiction. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363,
ioral manifestations of the nicotine abstinence syndrome in the rat: peripheral 3113–3123.
versus central mechanisms. Psychopharmacology (Berl.) 129, 348–356. Kosowski, A.R., Cebers, G., Cebere, A., Swanhagen, A.C., Liljequist, S., 2004. Nicotine-
Hollander, J.A., Lu, Q., Cameron, M.D., Kamenecka, T.M., Kenny, P.J., 2008. Insular induced dopamine release in the nucleus accumbens is inhibited by the novel
hypocretin transmission regulates nicotine reward. Proc. Natl. Acad. Sci. U.S.A. AMPA antagonist ZK200775 and the NMDA antagonist CGP39551. Psychophar-
105, 19480–19485. macology (Berl.) 175, 114–123.
Houdi, A.A., Dasgupta, R., Kindy, M.S., 1998. Effect of nicotine use and withdrawal Krishnan-Sarin, S., Meandzija, B., O’Malley, S., 2003. Naltrexone and nicotine patch
on brain preproenkephalin A mRNA. Brain Res. 799, 257–263. smoking cessation: a preliminary study. Nicotine Tob. Res. 5, 851–857.
Houdi, A.A., Pierzchala, K., Marson, L., Palkovits, M., Van Loon, G.R., 1991. Nicotine- Krishnan-Sarin, S., Rosen, M.I., O’Malley, S.S., 1999. Naloxone challenge in smokers.
induced alteration in Tyr-Gly-Gly and Met-enkephalin in discrete brain nuclei Preliminary evidence of an opioid component in nicotine dependence. Arch.
reflects altered enkephalin neuron activity. Peptides 12, 161–166. Gen. Psychiatry 56, 663–668.
Hughes, J.R., 2007. Effects of abstinence from tobacco: valid symptoms and time Lança, A.J., Adamson, K.L., Coen, K.M., Chow, B.L., Corrigall, W.A., 2000. The pedun-
course. Nicotine Tob. Res. 9, 315–327. culopontine tegmental nucleus and the role of cholinergic neurons in nicotine
Hughes, J.R., Hatsukami, D., 1986. Signs and symptoms of tobacco withdrawal. Arch. self-administration in the rat: a correlative neuroanatomical and behavioral
Gen. Psychiatry 43, 289–294. study. Neuroscience 96, 735–742.
Hung, R.J., McKay, J.D., Gaborieau, V., Boffetta, P., Hashibe, M., Zaridze, D., Mukeria, Laviolette, S.R., Alexson, T.O., van der Kooy, D., 2002. Lesions of the tegmental pedun-
A., Szeszenia-Dabrowska, N., Lissowska, J., Rudnai, P., Fabianova, E., Mates, D., culopontine nucleus block the rewarding effects and reveal the aversive effects
Bencko, V., Foretova, L., Janout, V., Chen, C., Goodman, G., Field, J.K., Liloglou, T., of nicotine in the ventral tegmental area. J. Neurosci. 22, 8653–8660.
Xinarianos, G., Cassidy, A., McLaughlin, J., Liu, G., Narod, S., Krokan, H.E., Skorpen, Le Foll, B., Wertheim, C.E., Goldberg, S.R., 2008. Effects of baclofen on conditioned
F., Elvestad, M.B., Hveem, K., Vatten, L., Linseisen, J., Clavel-Chapelon, F., Vineis, rewarding and discriminative stimulus effects of nicotine in rats. Neurosci. Lett.
P., Bueno-de-Mesquita, H.B., Lund, E., Martinez, C., Bingham, S., Rasmuson, T., 443, 236–240.
Hainaut, P., Riboli, E., Ahrens, W., Benhamou, S., Lagiou, P., Trichopoulos, D., Le Foll, B., Goldberg, S.R., 2004. Rimonabant, a CB1 antagonist, blocks nicotine-
Holcátová, I., Merletti, F., Kjaerheim, K., Agudo, A., Macfarlane, G., Talamini, R., conditioned place preferences. Neuroreport 15, 2139–2143.
230 F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231

Lerman, C., Wileyto, E.P., Patterson, F., Rukstalis, M., Audrain-McGovern, J., Res- Miller, D.K., Crooks, P.A., Dwoskin, L.P., 2000. Lobeline inhibits nicotine evoked
tine, S., Shields, P.G., Kaufmann, V., Redden, D., Benowitz, N., Berrettini, W.H., [3H]dopamine overflow from rat striatal slices and nicotine-evoked 86Rb+ efflux
2004. The functional mu opioid receptor (OPRM1) Asn40Asp variant predicts from thalamic synaptosomes. Neuropharmacology 39, 2654–2662.
short-term response to nicotine replacement therapy in a clinical trial. Pharma- Miner, L.L., Collins, A.C., 1988. The effect of chronic nicotine treatment on nicotine-
cogenomics J. 4, 184–192. induced seizures. Psychopharmacology (Berl.) 95, 52–55.
Liechti, M.E., Markou, A., 2008. Role of the glutamatergic system in nicotine depen- Mineur, Y.S., Picciotto, M.R., 2008. Genetics of nicotinic acetylcholine receptors:
dence: implications for the discovery and development of new pharmacological relevance to nicotine addiction. Biochem. Pharmacol. 75, 323–333.
smoking cessation therapies. CNS Drugs 22, 705–724. Mollereau, C., Parmentier, M., Mailleux, P., Butour, J.L., Moisand, C., Chalon, P., Caput,
Liechti, M.E., Lhuillier, L., Kaupmann, K., Markou, A., 2007. Metabotropic gluta- D., Vassart, G., Meunier, J.C., 1994. ORL1, a novel member of the opioid receptor
mate 2/3 receptors in the ventral tegmental area and the nucleus accumbens family. Cloning, functional expression and localization. FEBS Lett. 341, 33–38.
shell are involved in behaviors relating to nicotine dependence. J. Neurosci. 27, Nemeth-Coslett, R., Griffiths, R.R., 1986. Naloxone does not affect cigarette smoking.
9077–9085. Psychopharmacology 89, 261–264.
Maisonneuve, I.M., Glick, S.D., 1999. (+/−)Cyclazocine blocks the dopamine response Nestler, E.J., 2005. Is there a common molecular pathway for addiction? Nat. Neu-
to nicotine. Neuroreport 10, 693–696. rosci. 8, 1445–1449.
Maldonado, R., Valverde, O., Berrendero, F., 2006. Involvement of the endocannabi- Nisell, M., Nomikos, G.G., Svensson, T.H., 1994. Systemic nicotine-induced dopamine
noid system in drug addiction. Trends Neurosci. 29, 225–232. release in the rat nucleus accumbens is regulated by nicotinic receptors in the
Malin, D.H., Lake, J.R., Payne, M.C., Short, P.E., Carter, V.A., Cunningham, J.S., Wilson, ventral tegmental area. Synapse 16, 36–44.
O.B., 1996. Nicotine alleviation of nicotine abstinence syndrome is naloxone- Oakley, R.H., Laporte, S.A., Holt, J.A., Caron, M.G., Barak, L.S., 2000. Differential
reversible. Pharmacol. Biochem. Behav. 53, 81–85. affinities of visual arrestin, beta arrestin1, and beta arrestin2 for G protein-
Malin, D.H., Lake, J.R., Carter, V.A., Cunningham, J.S., Wilson, O.B., 1993. Naloxone pre- coupled receptors delineate two major classes of receptors. J. Biol. Chem. 275,
cipitates nicotine abstinence syndrome in the rat. Psychopharmacology (Berl.) 17201–17210.
112, 339–342. Olive, M.F., Koenig, H.N., Nannini, M.A., Hodge, C.W., 2001. Stimulation of endor-
Malin, D.H., Lake, J.R., Newlin-Maultsby, P., Roberts, L.K., Lanier, J.G., Carter, V.A., Cun- phin neurotransmission in the nucleus accumbens by ethanol, cocaine, and
ningham, J.S., Wilson, O.B., 1992. Rodent model of nicotine abstinence syndrome. amphetamine. J. Neurosci. 21, RC184.
Pharmacol. Biochem. Behav. 43, 779–784. O’Malley, S.S., Krishnan-Sarin, S., McKee, S.A., Leeman, R.F., Cooney, N.L., Meandzija,
Mansour, A., Fox, C.A., Akil, H., Watson, S.J., 1995. Opioid-receptor mRNA expression B., Wu, R., Makuch, R.W., 2008. Dose-dependent reduction of hazardous alco-
in the rat CNS: anatomical and functional implications. Trends Neurosci. 18, hol use in a placebo-controlled trial of naltrexone for smoking cessation. Int. J.
22–29. Neuropsychopharmacol. 17, 1–9.
Mansvelder, H.D., McGehee, D.S., 2003. Cellular and synaptic mechanisms of nicotine Panagis, G., Hildebrand, B.E., Svensson, T.H., Nomikos, G.G., 2000. Selective c-fos
addiction. J. Neurobiol. 53, 606–617. induction and decreased dopamine release in the central nucleus of amygdala
Mansvelder, H.D., Keath, J.R., McGehee, D.S., 2002. Synaptic mechanisms under- in rats displaying a mecamylamine-precipitated nicotine withdrawal syndrome.
lie nicotine-induced excitability of brain reward areas. Neuron 33, 905– Synapse 35, 15–25.
919. Paterson, N.E., Vlachou, S., Guery, S., Kaupmann, K., Froestl, W., Markou, A., 2008. Pos-
Marco, E.M., Granstrem, O., Moreno, E., Llorente, R., Adriani, W., Laviola, G., Viveros, itive modulation of GABA(B) receptors decreased nicotine self-administration
M.P., 2007. Subchronic nicotine exposure in adolescence induces long-term and counteracted nicotine-induced enhancement of brain reward function in
effects on hippocampal and striatal cannabinoid-CB1 and mu-opioid receptors rats. J. Pharmacol. Exp. Ther. 326, 306–314.
in rats. Eur. J. Pharmacol. 557, 37–43. Paterson, N.E., Markou, A., 2005. The metabotropic glutamate receptor 5 antagonist
Markou, A., Paterson, N.E., 2001. The nicotinic antagonist methyllycaconitine has MPEP decreased break points for nicotine, cocaine and food in rats. Psychophar-
differential effects on nicotine self-administration and nicotine withdrawal in macology (Berl.) 179, 255–261.
the rat. Nicotine Tob. Res. 3, 361–373. Paterson, N.E., Froestl, W., Markou, A., 2005. Repeated administration of the GABAB
Markou, A., Koob, G.F., 1993. Intracraneal self-stimulation thresholds as a measure receptor agonist CGP44532 decreased nicotine self-administration, and acute
of reward. In: Sahgal, A. (Ed.), Behavioural Neuroscience: A Practical Approach. administration decreased cue-induced reinstatement of nicotine-seeking in
University Press, Oxford, New York, pp. 93–115. rats. Neuropsychopharmacology 30, 119–128.
Marks, M.J., Romm, E., Gaffney, D.K., Collins, A.C., 1986. Nicotine-induced toler- Paterson, N.E., Froestl, W., Markou, A., 2004. The GABAB receptor agonists baclofen
ance and receptor changes in four mouse strains. J. Pharmacol. Exp. Ther. 237, and CGP44532 decreased nicotine self-administration in the rat. Psychophar-
809–819. macology (Berl.) 172, 179–186.
Martín-García, E., Barbano, M.F., Galeote, L., Maldonado, R., 2009. New operant Paterson, N.E., Semenova, S., Gasparini, F., Markou, A., 2003. The mGluR5 antagonist
model of nicotine-seeking behaviour in mice. Int. J. Neuropsychopharmacol. 12, MPEP decreased nicotine self-administration in rats and mice. Psychopharma-
343–356. cology (Berl.) 167, 257–264.
Marty, M.A., Erwin, V.G., Cornell, K., Zgombick, J.M., 1985. Effects of nicotine on beta- Perkins, K.A., Lerman, C., Grottenthaler, A., Ciccocioppo, M.M., Milanak, M., Conklin,
endorphin, alpha MSH, and ACTH secretion by isolated perfused mouse brains C.A., Bergen, A.W., Benowitz, N.L., 2008. Dopamine and opioid gene variants are
and pituitary glands, in vitro. Pharmacol. Biochem. Behav. 22, 317–325. associated with increased smoking reward and reinforcement owing to negative
Marubio, L.M., Gardier, A.M., Durier, S., David, D., Klink, R., Arroyo-Jimenez, M.M., mood. Behav. Pharmacol. 19, 641–649.
McIntosh, J.M., Rossi, F., Champtiaux, N., Zoli, M., Changeux, J.P., 2003. Effects Picciotto, M.R., Corrigall, W.A., 2002. Neuronal systems underlying behaviors related
of nicotine in the dopaminergic system of mice lacking the alpha4 sub- to nicotine addiction: neural circuits and molecular genetics. J. Neurosci. 22,
unit of neuronal nicotinic acetylcholine receptors. Eur. J. Neurosci. 17, 1329– 3338–3341.
1337. Picciotto, M.R., Zoli, M., Rimondini, R., Léna, C., Marubio, L.M., Pich, E.M., Fuxe, K.,
Maskos, U., 2007. Emerging concepts: novel integration of in vivo approaches to Changeux, J.P., 1998. Acetylcholine receptors containing the beta2 subunit are
localize the function of nicotinic receptors. J. Neurochem. 100, 596–602. involved in the reinforcing properties of nicotine. Nature 391, 173–177.
Maskos, U., Molles, B.E., Pons, S., Besson, M., Guiard, B.P., Guilloux, J.P., Evrard, A., Pomerleau, O.F., 1998. Endogenous opioids and smoking: a review of progress and
Cazala, P., Cormier, A., Mameli-Engvall, M., Dufour, N., Cloëz-Tayarani, I., Bemel- problems. Psychoneuroendocrinology 23, 115–130.
mans, A.P., Mallet, J., Gardier, A.M., David, V., Faure, P., Granon, S., Changeux, J.P., Pontieri, F.E., Tanda, G., Orzi, F., Di Chiara, G., 1996. Effects of nicotine on the nucleus
2005. Nicotine reinforcement and cognition restored by targeted expression of accumbens and similarity to those of addictive drugs. Nature 382, 255–257.
nicotinic receptors. Nature 436, 103–107. Pons, S., Fattore, L., Cossu, G., Tolu, S., Porcu, E., McIntosh, J.M., Changeaux,
McGehee, D.S., Heath, M.J., Gelber, S., Devay, P., Role, L.W., 1995. Nicotine enhance- J.P., Maskos, U., Fratta, W., 2008. Crucial role of ␣4 and ␣6 nicotinic acetyl-
ment of fast excitatory synaptic transmission in CNS by presynaptic receptors. choline receptor subunits from ventral tegmental area in systemic nicotine
Science 269, 1692–1696. self-administration. J. Neurosci. 28, 12318–12327.
McLellan, A.T., Lewis, D.C., O’Brien, C.P., Kleber, H.D., 2000. Drug dependence, a Quick, M.W., Lester, R.A., 2002. Desensitization of neuronal nicotinic receptors. J.
chronic medical illness: implications for treatment, insurance, and outcomes Neurobiol. 53, 457–478.
evaluation. JAMA 284, 1689–1695. Rasmussen, N.A., Farr, L.A., 2009. Beta-endorphin response to an acute pain stimulus.
Mendizábal, V., Zimmer, A., Maldonado, R., 2006. Involvement of kappa/dynorphin J. Neurosci. Methods 177, 285–288.
system in WIN 55,212-2 self-administration in mice. Neuropsychopharmacol- Ray, L.A., Miranda Jr., R., Kahler, C.W., Leventhal, A.M., Monti, P.M., Swift, R., Hutchi-
ogy 31, 1957–1966. son, K.E., 2007. Pharmacological effects of naltrexone and intravenous alcohol on
Merritt, L.L., Martin, B.R., Walters, C., Lichtman, A.H., Damaj, M.I., 2008. The craving for cigarettes among light smokers: a pilot study. Psychopharmacology
endogenous cannabinoid system modulates nicotine reward and dependence. (Berl.) 193, 449–456.
J. Pharmacol. Exp. Ther. 326, 483–492. Ray, R., Jepson, C., Patterson, F., Strasser, A., Rukstalis, M., Perkins, K., Lynch, K.G.,
Meunier, J.C., Mollereau, C., Toll, L., Suaudeau, C., Moisand, C., Alvinerie, P., Butour, O’Malley, S., Berrettini, W.H., Lerman, C., 2006. Association of OPRM1 A118G
J.L., Guillemot, J.C., Ferrara, P., Monsarrat, B., et al., 1995. Isolation and structure variant with the relative reinforcing value of nicotine. Psychopharmacology
of the endogenous agonist of opioid receptor-like ORL1 receptor. Nature 377, (Berl.) 188, 355–363.
532–535. Reinscheid, R.K., Nothacker, H.P., Bourson, A., Ardati, A., Henningsen, R.A., Bunzow,
Millar, N.S., Gotti, C., 2009. Diversity of vertebrate nicotinic acetylcholine receptors. J.R., Grandy, D.K., Langen, H., Monsma Jr., F.J., Civelli, O., 1995. Orphanin FQ: a
Neuropharmacology 56, 237–246. neuropeptide that activates an opioidlike G protein-coupled receptor. Science
Miller, D.K., Lever, J.R., Rodvelt, K.R., Baskett, J.A., Will, M.J., Kracke, G.R., 2007. 270, 792–794.
Lobeline, a potential pharmacotherapy for drug addiction, binds to mu opioid Risner, M.E., Goldberg, S.R., 1983. A comparison of nicotine and cocaine self-
receptors and diminishes the effects of opioid receptor agonists. Drug Alcohol administration in the dog: fixed-ratio and progressive-ratio schedules of
Depend. 89, 282–291. intravenous drug infusion. J. Pharmacol. Exp. Ther. 224, 319–326.
F. Berrendero et al. / Neuroscience and Biobehavioral Reviews 35 (2010) 220–231 231

Rohsenow, D.J., Monti, P.M., Hutchison, K.E., Swift, R.M., MacKinnon, S.V., Sirota, son, K., Skuladottir, H., Isaksson, H.J., Gudbjartsson, T., Jones, G.T., Mueller, T.,
A.D., Kaplan, G.B., 2007. High-dose transdermal nicotine and naltrexone: effects Gottsäter, A., Flex, A., Aben, K.K., de Vegt, F., Mulders, P.F., Isla, D., Vidal, M.J., Asin,
on nicotine withdrawal, urges, smoking, and effects of smoking. Exp. Clin. Psy- L., Saez, B., Murillo, L., Blondal, T., Kolbeinsson, H., Stefansson, J.G., Hansdottir, I.,
chopharmacol. 15, 81–92. Runarsdottir, V., Pola, R., Lindblad, B., van Rij, A.M., Dieplinger, B., Haltmayer, M.,
Roth-Deri, I., Green-Sadan, T., Yadid, G., 2008. Beta-endorphin and drug-induced Mayordomo, J.I., Kiemeney, L.A., Matthiasson, S.E., Oskarsson, H., Tyrfingsson, T.,
reward and reinforcement. Prog. Neurobiol. 86, 1–21. Gudbjartsson, D.F., Gulcher, J.R., Jonsson, S., Thorsteinsdottir, U., Kong, A., Ste-
Rukstalis, M., Jepson, C., Strasser, A., Lynch, K.G., Perkins, K., Patterson, F., Lerman, C., fansson, K., 2008. A variant associated with nicotine dependence, lung cancer
2005. Naltrexone reduces the relative reinforcing value of nicotine in a cigarette and peripheral arterial disease. Nature 452, 638–642.
smoking choice paradigm. Psychopharmacology 180, 41–48. Tome, A.R., Izaguirre, V., Rosario, L.M., Cena, V., Gonzalez-Garcia, C., 2001. Nalox-
Sakoori, K., Murphy, N.P., 2009. Enhanced nicotine sensitivity in nociceptin/orphanin one inhibits nicotine-induced receptor current and catecholamine secretion in
FQ receptor knockout mice. Neuropharmacology 56, 896–904. bovine chromaffin cells. Brain Res. 903, 62–65.
Salas, R., Sturm, R., Boulter, J., De Biasi, M., 2009. Nicotinic receptors in the Trigo, J.M., Zimmer, A., Maldonado, R., 2009. Nicotine anxiogenic and rewarding
habenulo-interpeduncular system are necessary for nicotine withdrawal in effects are decreased in mice lacking beta-endorphin. Neuropharmacology 56,
mice. J. Neurosci. 29, 3014–3018. 1147–1153.
Sanchis-Segura, C., Spanagel, R., 2006. Behavioural assessment of drug reinforce- Vanderschuren, L.J., Everitt, B.J., 2004. Drug seeking becomes compulsive after pro-
ment and addictive features in rodents: an overview. Addict. Biol. 11, 2–38. longed cocaine self-administration. Science 305, 1017–1019.
Scott, D.J., Domino, E.F., Heitzeg, M.M., Koeppe, R.A., Ni, L., Guthrie, S., Zubieta, J.K., Vihavainen, T., Piltonen, M., Tuominen, R.K., Korpi, E.R., Ahtee, L., 2008.
2007. Smoking modulation of mu-opioid and dopamine D2 receptor-mediated Morphine–nicotine interaction in conditioned place preference in mice after
neurotransmission in humans. Neuropsychopharmacology 32, 450–457. chronic nicotine exposure. Eur. J. Pharmacol. 587, 169–174.
Scherma, M., Fadda, P., Le Foll, B., Forget, B., Fratta, W., Goldberg, S.R., Tanda, G., Villégier, A.S., Lotfipour, S., McQuown, S.C., Belluzzi, J.D., Leslie, F.M., 2007. Tranyl-
2008. The endocannabinoid system: a new molecular target for the treatment cypromine enhancement of nicotine self-administration. Neuropharmacology
of tobacco addiction. CNS Neurol. Disord. Drug Targets 7, 468–481. 52, 1415–1425.
Schoepp, D.D., Wright, R.A., Levine, L.R., Gaydos, B., Potter, W.Z., 2003. LY354740, an Villégier, A.S., Salomon, L., Granon, S., Changeux, J.P., Belluzzi, J.D., Leslie, F.M.,
mGlu2/3 receptor agonist as a novel approach to treat anxiety/stress. Stress 6, Tassin, J.P., 2006. Monoamine oxidase inhibitors allow locomotor and rewarding
189–197. responses to nicotine. Neuropsychopharmacology 31, 1704–1713.
Shippenberg, T.S., LeFevour, A., Chefer, V.I., 2008. Targeting endogenous mu- and Volkow, N., Li, T.K., 2005. The neuroscience of addiction. Nat. Neurosci. 8, 1429–1430.
delta-opioid receptor systems for the treatment of drug addiction. CNS Neurol. Walsh, Z., Epstein, A., Munisamy, G., King, A., 2008. The impact of depressive symp-
Disord. Drug Targets 7, 442–453. toms on the efficacy of naltrexone in smoking cessation. J. Addict. Dis. 27, 65–72.
Shippenberg, T.S., Zapata, A., Chefer, V.I., 2007. Dynorphin and the pathophysiology Walters, C.L., Brown, S., Changeux, J.P., Martin, B., Damaj, M.I., 2006. The beta2
of drug addiction. Pharmacol. Ther. 116, 306–321. but not alpha7 subunit of the nicotinic acetylcholine receptor is required for
Shoaib, M., 2008. The cannabinoid antagonist AM251 attenuates nicotine self- nicotine-conditioned place preference in mice. Psychopharmacology (Berl.) 184,
administration and nicotine-seeking behaviour in rats. Neuropharmacology 54, 339–344.
438–444. Walters, C.L., Cleck, J.N., Kuo, Y.C., Blendy, J.A., 2005. Mu-opioid receptor and CREB
Shoaib, M., 2006. Effects of isoarecolone, a nicotinic receptor agonist in rodent mod- activation are required for nicotine reward. Neuron 46, 933–943.
els of nicotine dependence. Psychopharmacology (Berl.) 188, 252–257. Wewers, M.E., Dhatt, R.K., Snively, T.A., Tejwani, G.A., 1999. The effect of chronic
Shoaib, M., Schindler, C.W., Goldberg, S.R., 1997. Nicotine self-administration in rats: administration of nicotine on antinociception, opioid receptor binding and met-
strain and nicotine pre-exposure effects on acquisition. Psychopharmacology enkelphalin levels in rats. Brain Res. 822, 107–113.
(Berl.) 129, 35–43. Wewers, M.E., Dhatt, R., Tejwani, G.A., 1998. Naltrexone administration affects ad
Spaziante, R., Merola, B., Colao, A., Gargiulo, G., Cafiero, T., Irace, C., Rossi, E., Oliver, libitum smoking behavior. Psychopharmacology (Berl.) 140, 185–190.
C., Lombardi, G., Mazzarella, B., et al., 1990. Beta-endorphin concentrations both Wilkie, G.I., Hutson, P.H., Stephens, M.W., Whiting, P., Wonnacott, S., 1993. Hip-
in plasma and in cerebrospinal fluid in response to acute painful stimuli. J. pocampal nicotinic autoreceptors modulate acetylcholine release. Biochem. Soc.
Neurosurg. Sci. 34, 99–106. Trans. 21, 429–431.
Stolerman, I.P., Naylor, C., Elmer, G.I., Goldberg, S.R., 1999. Discrimination and self- Wong, G.Y., Wolter, T.D., Croghan, G.A., Croghan, I.T., Offord, K.P., Hurt, R.D., 1999. A
administration of nicotine by inbred strains of mice. Psychopharmacology (Berl.) randomized trial of naltrexone for smoking cessation. Addiction 94, 1227–1237.
141, 297–306. Xue, Y., Domino, E.F., 2008. Tobacco/nicotine and endogenous brain opioids. Prog.
Stolerman, I.P., Jarvis, M.J., 1995. The scientific case that nicotine is addictive. Psy- Neuropsychopharmacol. Biol. Psychiatry 32, 1131–1138.
chopharmacology (Berl.) 117, 2–10. Yang, X., Criswell, H.E., Breese, G.R., 1996. Nicotine-induced inhibition in medial
Sun, D., Ma, J.Z., Payne, T.J., Li, M.D., 2008. Beta-arrestins 1 and 2 are associated septum involves activation of presynaptic nicotinic cholinergic receptors on
with nicotine dependence in European American smokers. Mol. Psychiatry 13, gamma-aminobutyric acid-containing neurons. J. Pharmacol. Exp. Ther. 276,
398–406. 482–489.
Sutherland, G., Stapleton, J.A., Russell, M.A., Feyerabend, C., 1995. Naltrexone, Yoo, J.H., Lee, S.Y., Loh, H.H., Ho, I.K., Jang, C.G., 2004. Loss of nicotine-induced
smoking behaviour and cigarette withdrawal. Psychopharmacology (Berl.) 120, behavioral sensitization in micro-opioid receptor knockout mice. Synapse 51,
418–425. 219–223.
Suzuki, T., Tsuji, M., Mori, T., Ikeda, H., Misawa, M., Nagase, H., 1997. Involvement of Yoo, J.H., Cho, J.H., Lee, S.Y., Loh, H.H., Ho, I.K., Jang, C.G., 2005. Reduced nNOS expres-
dopamine-dependent and -independent mechanisms in the rewarding effects sion induced by repeated nicotine treatment in mu-opioid receptor knockout
mediated by delta opioid receptor subtypes in mice. Brain Res. 744, 327–334. mice. Neurosci. Lett. 380, 70–74.
Tanda, G., Di Chiara, G., 1998. A dopamine-mu1 opioid link in the rat ventral tegmen- Zarrindast, M.R., Faraji, N., Rostami, P., Sahraei, H., Ghoshouni, H., 2003.
tum shared by palatable food (Fonzies) and non-psychostimulant drugs of abuse. Cross-tolerance between morphine- and nicotine-induced conditioned place
Eur. J. Neurosci. 10, 1179–1187. preference in mice. Pharmacol. Biochem. Behav. 74, 363–369.
Tapper, A.R., McKinney, S.L., Nashmi, R., Schwarz, J., Deshpande, P., Labarca, C., Zadina, J.E., Hackler, L., Ge, L.J., Kastin, A.J., 1997. A potent and selective endogenous
Whiteaker, P., Marks, M.J., Collins, A.C., Lester, H.A., 2004. Nicotine activation agonist for the mu-opiate receptor. Nature 386, 499–502.
of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science Zhang, L., Kendler, K.S., Chen, X., 2006. The mu-opioid receptor gene and smoking
306, 1029–1032. initiation and nicotine dependence. Behav. Brain Funct. 2, 28.
Thorgeirsson, T.E., Geller, F., Sulem, P., Rafnar, T., Wiste, A., Magnusson, K.P., Zheng, H., Loh, H.H., Law, P.Y., 2008. Beta-arrestin-dependent mu-opioid receptor-
Manolescu, A., Thorleifsson, G., Stefansson, H., Ingason, A., Stacey, S.N., Bergth- activated extracellular signal-regulated kinases (ERKs) translocate to nucleus in
orsson, J.T., Thorlacius, S., Gudmundsson, J., Jonsson, T., Jakobsdottir, M., contrast to G protein-dependent ERK activation. Mol. Pharmacol. 73, 178–190.
Saemundsdottir, J., Olafsdottir, O., Gudmundsson, L.J., Bjornsdottir, G., Kristjans-

S-ar putea să vă placă și