Skip to main content
Cold Spring Harbor Perspectives in Medicine logoLink to Cold Spring Harbor Perspectives in Medicine
. 2013 Jan;3(1):a012112. doi: 10.1101/cshperspect.a012112

Molecular Mechanisms Underlying Behaviors Related to Nicotine Addiction

Marina R Picciotto 1, Paul J Kenny 2
PMCID: PMC3530035  NIHMSID: NIHMS445652  PMID: 23143843

Abstract

Tobacco smoking results in more than 5 million deaths each year and accounts for almost 90% of all deaths from lung cancer. Nicotine, the major reinforcing component of tobacco smoke, acts in the brain through the neuronal nicotinic acetylcholine receptors (nAChRs). The nAChRs are allosterically regulated, ligand-gated ion channels consisting of five membrane-spanning subunits. Twelve mammalian α subunits (α2–α10) and β subunits (β2–β4) have been cloned. The predominant nAChR subtypes in mammalian brain are those containing α4 and β2 subunits (denoted as α4β2* nAChRs). The α4β2* nAChRs mediate many behaviors related to nicotine addiction and are the primary targets for currently approved smoking cessation agents. Considering the large number of nAChR subunits in the brain, it is likely that nAChRs containing subunits in addition to α4 and β2 also play a role in tobacco smoking. Indeed, genetic variation in the CHRNA5-CHRNA3-CHRNB4 gene cluster, encoding the α5, α3, and β4 nAChR subunits, respectively, has been shown to increase vulnerability to tobacco dependence and smoking-associated diseases including lung cancer. Moreover, mice in which expression of α5 or β4 subunits has been genetically modified have profoundly altered patterns of nicotine consumption. In addition to the reinforcing properties of nicotine, the effects of nicotine on appetite, attention, and mood are also thought to contribute to establishment and maintenance of the tobacco smoking habit. Here we review recent insights into the behavioral actions of nicotine and the nAChRs subtypes involved, which likely contribute to the development of tobacco dependence in smokers.


The α4β2* nicotinic acetylcholine receptors (nAChRs) mediate behaviors related to nicotine addiction. Other α and β subunits can be incorporated into nAChRs and may modulate these behaviors.

NICOTINIC RECEPTOR SUBTYPES INVOLVED IN CONTROL OF THE MESOLIMBIC SYSTEM AND NICOTINE REINFORCEMENT

The mesolimbic dopamine (DA) system is a central mediator of drug reward and reinforcement (Koob 1992). Lesions of the ventral tegmental area (VTA) and its primary projection area, the nucleus accumbens (nAc), greatly attenuate nicotine self-administration and the psychostimulant properties of nicotine (its ability to increase locomotion [Clarke et al. 1988; Corrigall et al. 1992, 1994]). A great deal of progress has been made in identifying the nAChR subtypes expressed in both the dopaminergic and GABAergic neurons of the VTA and on neuronal terminals in the nAc (Klink et al. 2001; Zoli et al. 2002). DA neurons express heteromeric nAChRs containing the α4, α5, α6, β2, and β3 subunits in various combinations, with the predominant subtypes being α4/β2/α5 and α4/α6/β2/β3. The α6 subunit appears to be selectively expressed in DA neurons (Le Novère et al. 1996; Drenan et al. 2008b), although a recent report has suggested that there may be an effect of α6-containing receptors on GABA transmission in the VTA (Yang et al. 2011). In addition, α7 homomeric nAChRs are expressed in DA neurons (Klink et al. 2001), as well as on neuronal terminals on afferents to the VTA (Mansvelder et al. 2002; Wooltorton et al. 2003).

Electrophysiological studies have shown that nAChRs containing the β2 subunit are essential for the ability of nicotine to depolarize DA cell bodies in the VTA and to increase their firing rate (Picciotto et al. 1998; Zhou et al. 2001). While the predominant inward currents owing to nicotine in these neurons involve β2* nAChRs, nicotine can also modulate the presynaptic input to DA neurons from GABAergic and glutamatergic terminals impinging on them. In a slice preparation, nicotine can potentiate glutamate input to DA neurons through α7 nAChRs, resulting in long-term potentiation of those inputs (Mansvelder and McGehee 2000). In addition, nicotine can desensitize β2* nAChRs on GABAergic inputs to DA neurons, resulting in a shift from mixed excitation and inhibition of DA neurons by nicotine, to a more unmixed stimulation of nAChRs on presynaptic glutamatergic terminals (Mansvelder et al. 2002; Wooltorton et al. 2003).

Evidence from mouse genetic models with knockout or mutations of nAChR subunits suggests that the postsynaptic depolarization of DA neurons is essential for behaviors related to nicotine reward and reinforcement such as nicotine place preference and self-administration. Knockout of the β2 subunit abolishes nicotine-mediated DA release (Picciotto et al. 1998; Grady et al. 2001), nicotine-induced locomotor activation (King et al. 2004), nicotine self-administration (Picciotto et al. 1998; Maskos et al. 2005), and nicotine place preference (Walters et al. 2006; Brunzell et al. 2009; Mineur et al. 2009a). Similarly, knockout of the α4 subunit abolishes intracerebroventricular (i.c.v.) self-administration of nicotine, consistent with evidence that α4/β2* nAChRs are required for depolarization of DA neurons in the VTA (Exley et al. 2011; McGranahan et al. 2011). Conversely, knockin of a hypersensitive α4 subunit shifts the dose-response curve for nicotine-induced increases in DA neuron firing to the left, and results in nicotine place preference at very low doses of the drug (Tapper et al. 2004). Similarly, expression of a hypersensitive α6 subunit in a bacterial artificial chromosome (BAC) transgenic mouse line potentiates nicotine-induced burst firing in DA neurons, and potentiates nicotine place preference at low doses of nicotine (Drenan et al. 2008a, 2010).

A series of recent studies have provided further support for the involvement of α6* nAChRs in nicotine self-administration. Mice lacking the α6 subunit do not acquire intravenous nicotine self-administration (Pons et al. 2008). Similarly, conotoxins selective for α6/β2* nAChRs disrupt nicotine self-administration in the rat when infused into the VTA (Gotti et al. 2010) and following self-administration training, these conotoxins decrease the motivation to lever press for nicotine on a progressive ratio schedule (Brunzell et al. 2010). In contrast, mice with constitutive knockout of the α6 subunit, intra-VTA self-administration of nicotine is not disrupted, whereas α4* nAChRs are necessary and sufficient for both intra-VTA self-administration, as well as nicotine-induced increases in firing of DA neurons (Exley et al. 2011; McGranahan et al. 2011); however, α4 and α6 subunits are both required for the ability of nicotine to gate DA transmission in the nAc, suggesting that nAChRs in nAc may be more important in motivation to self-administer nicotine (Brunzell et al. 2010; Exley et al. 2011) and that this may affect acquisition of intravenous self-administration behavior (Pons et al. 2008; Gotti et al. 2010), as well as nicotine-dependent locomotor activation (Gotti et al. 2010).

Despite its contribution to nicotine-dependent plasticity in the VTA, knockout of the α7 subunit in mice does not affect nicotine place preference (Walters et al. 2006) or acquisition of nicotine self-administration (Pons et al. 2008); however, antagonizing α7-type nAChRs in the nAc or anterior cingulate cortex in the rat increases the motivation to self-administer nicotine, whereas infusion of a selective α7 agonist decreases motivation, as measured using a progressive ratio schedule (Brunzell and McIntosh 2012). α7-type nAChRs may modulate, rather than mediate, nicotine reinforcement and therefore the effect of α7 knockout may be more subtle than knockout of β2* nAChRs.

An important role for nAChRs in the VTA in nicotine reinforcement has been shown using both molecular genetic and pharmacological techniques. Selective viral reexpression of β2* (Maskos et al. 2005; Pons et al. 2008) or α4* nAChRs in the VTA is sufficient to support both intra-VTA (Maskos et al. 2005) and systemic nicotine self-administration (Pons et al. 2008), identifying the nAChR subtypes necessary for nicotine reinforcement, as well as demonstrating the importance of nAChRs within the VTA itself for this behavior. This is consistent with previous studies suggesting that nAChRs within the VTA are critical for nicotine reward, because local infusion of a nicotinic agonist into the VTA, but not the nAc, is sufficient for nicotine place preference in the rat (Museo and Wise 1994).

Thus, molecular genetic studies support the idea that α4/α6/β2* nAChRs on DA neurons in the VTA are essential for nicotine reinforcement. These experiments in mice are supported by pharmacological studies in rats, and provide a consistent molecular subtype and neuroanatomical locus for the rewarding and reinforcing effects of nicotine.

NICOTINIC RECEPTORS AND CIRCUITS INVOLVED IN AVERSION AND NICOTINE WITHDRAWAL: FOCUS ON THE HABENULA-INTERPEDUNCULAR PATHWAY

The habenula is a diencephalic structure located on the dorsomedial surface of the caudal thalamus that is segregated into medial (MHb) and lateral (LHb) domains (Lecourtier and Kelly 2007; Hikosaka 2010). The MHb and LHb are anatomically, chemically, and functionally distinct subnuclei, each with different complements of afferent and efferent connections. LHb receives afferent inputs from, and projects extensively to, midbrain and hindbrain sites. In particular, the LHb projects densely to the rostromedial tegmental nucleus (RMTg) (Jhou et al. 2009), and has a well-established inhibitory effect on the firing of midbrain dopamine neurons (Lecourtier and Kelly 2007; Matsumoto and Hikosaka 2009; Hikosaka 2010; Bromberg-Martin and Hikosaka 2011). LHb neurons are excited by omission of anticipated rewards or exposure to aversive stimuli (Lecourtier and Kelly 2007; Matsumoto and Hikosaka 2009; Hikosaka 2010; Bromberg-Martin and Hikosaka 2011). This has prompted considerable interest in the role for LHb neurons in encoding negative motivational states. Unlike the LHb, the MHb projects almost exclusively to the interpeduncular nucleus (IPN) via the fasciculus retroflexus (Fr) (Lecourtier and Kelly 2007; Hikosaka 2010). MHb is comprised of neurons that produce the neurotransmitters acetylcholine or substance P (Cuello et al. 1978; Eckenrode et al. 1987), and a small population that produce the cytokine interleukin-18 (IL-18) (Sugama et al. 2002). However, it is believed that most MHb neurons also produce and corelease glutamate, with this excitatory neurotransmitter considered the major functional transmitter at the MHb-IPN synapse (Mata et al. 1977; Vincent et al. 1980; Girod et al. 2000; Ren et al. 2011). The MHb contains some of the highest densities of nicotine-binding sites in brain (Mugnaini et al. 2002). In particular, the highest density of α5, α3, and β4 nAChR subunit expression in brain is detected in MHb and/or IPN (De Biasi and Salas 2008). Indeed, approximately 90%–100% of MHb neurons express α3, α4, α5, β2, and β4 nAChR subunits (Sheffield et al. 2000), and in mouse brain slices through the MHb, >85% of neurons respond to nicotine with an inward current, and these currents are not altered in mice lacking the β2 nAChR subunit (Picciotto et al. 1995). It is also hypothesized that ∼20% of functional nAChRs in rat MHb neurons that project to IPN contain α5 subunits (Grady et al. 2009).

The fact that the MHb-IPN pathway is enriched in α5, α3, and β4 nAChR subunits is of particular interest in the context of recent human genetics findings. It has been shown that allelic variation in the α5/α3/β4 nAChR subunit gene cluster located in chromosome region 15q25 significantly increases the risk of tobacco addiction (Saccone et al. 2007; Berrettini et al. 2008; Lips et al. 2010). For example, a single nucleotide polymorphism (SNP) in CHRNA5 (rs16969968) that is very common in those of European descent (minor allele frequency = 0.42) increases the risk of tobacco dependence by ∼30% in individuals carrying a single copy of the variant, and more than doubles the risk in those carrying two risk alleles (Bierut et al. 2008; Wang et al. 2009); a finding that has been consistently replicated (Berrettini et al. 2008; Bierut et al. 2008; Grucza et al. 2008; Stevens et al. 2008). The rs16969968 risk variant is associated with heavy smoking (Berrettini et al. 2008; Bierut et al. 2008; Grucza et al. 2008; Stevens et al. 2008), early onset of smoking behavior (Weiss et al. 2008), and with “pleasurable buzz” from tobacco (Sherva et al. 2008). In addition, the same genetic variability in CHRNA5 is also a major risk factor for lung cancer and chronic obstructive pulmonary disease (COPD) in smokers (Amos et al. 2008; Hung et al. 2008; Wang et al. 2010), likely reflecting higher levels of tobacco dependence in individuals carrying risk alleles and consequently greater exposure to carcinogens and toxins contained in tobacco smoke (Le Marchand et al. 2008; Thorgeirsson et al. 2008). In addition to the rs16969968 SNP in CHRNA5, there is also increased risk of tobacco dependence in individuals carrying the rs6495308, rs578776, or rs1051730 SNPs in CHRNA3 (Berrettini et al. 2008; Saccone et al. 2009), and rs1948 in CHRNB4 (Schlaepfer et al. 2008).

The above findings suggest that nAChRs containing α5, α3, and/or β4 nAChR subunits, densely expressed in the MHb-IPN pathway, regulate addiction-related actions of nicotine. Consistent with an important role for α5* nAChRs in regulating nicotine intake, it was recently shown that mice with a null mutation for this subunit intravenously self-administered far more nicotine than their wild-type littermates (Fowler et al. 2011). Interestingly, the knockout mice consumed more nicotine only when higher unit doses of the drug were available (Fowler et al. 2011). By using Fos immunoreactivity as a measure of neuronal activation, it was shown that the MHb-IPN pathway of the knockout mice was far less sensitive to nicotine-induced activation than wild-type mice (Fowler et al. 2011). Moreover, chemical inactivation of the MHb or the IPN using the local anesthetic lidocaine, or disruption of NMDA receptor-mediated glutamatergic transmission in these sites using the competitive antagonist LY2358959, increased nicotine self-administration behavior in rats in a manner similar to what was observed in α5 nAChR subunit knockout mice (Fowler et al. 2011). Virus-mediated reexpression of the α5 nAChR subunit in the MHb-IPN pathway of knockout mice abolished the increased nicotine intake seen at higher doses of nicotine (Fowler et al. 2011). Conversely, RNA interference-mediated knockdown of α5 nAChR subunits in the MHb-IPN pathway in rats resulted in increased nicotine intake at higher unit doses of the drug, very similar to the same behavioral profile detected in the knockout mice (Fowler et al. 2011). Finally, knockdown of α5 nAChR subunits in the MHb-IPN pathway in rats decreased their sensitivity to the reward-inhibiting (i.e., aversive) actions of higher nicotine doses compared with control rats, as measured by nicotine-induced elevations of intracranial self-stimulation (ICSS) reward thresholds (Fowler et al. 2011). Taken together, these findings suggest that nicotine activates the MHb-IPN pathway through stimulatory effects on α5* nAChRs. Nicotine-induced activation of the MHb-IPN pathway results in a negative motivational signal that serves to limit further nicotine intake. Hence, disruption of α5* nAChR signaling diminishes the stimulatory effects of nicotine on MHb-IPN activity, and thereby permits consumption of greater quantities of nicotine.

In addition to α5* nAChRs, evidence suggests that β4* nAChRs in the MHb-IPN pathway also play an important role in regulating nicotine consumption. Specifically, overexpression of β4* nAChRs in mice using BAC transgenic technology resulted in greatly diminished sensitivity to the reinforcing properties of orally consumed nicotine solutions and far less consumption of the drug than wild-type mice (Frahm et al. 2011). This finding suggests that, similar to α5* nAChRs, β4* nAChRs in the MHb-IPN pathway also regulate sensitivity to the aversive effects of nicotine that control the quantities of the drug consumed.

Dependence on tobacco smoking depends not only on the balance between the rewarding and aversive action of nicotine described above, but also on escape from the aversive consequences of nicotine withdrawal (Doherty et al. 1995; Kenny and Markou 2001). Indeed, withdrawal duration and severity predicts relapse in abstinent human smokers (Piasecki et al. 1998, 2000, 2003). The nicotine withdrawal syndrome in abstinent smokers is composed of “physical” or somatic components, and “affective” components. The most common somatic symptoms include bradycardia, gastrointestinal discomfort, and increased appetite. Affective symptoms primarily include depressed mood including anhedonia, dysphoria, anxiety, irritability, difficulty concentrating, and craving (Parrott 1993; Doherty et al. 1995; Kenny and Markou 2001). Similar to α5 subunits, α2 subunits are highly enriched in the IPN (Grady et al. 2009). Recently, it was shown that α5 and α2 subunit knockout mice that were dependent on nicotine (delivered through subcutaneously implanted osmotic minipumps) did not show somatic signs of nicotine withdrawal when withdrawal was precipitated with the nAChR antagonist mecamylamine (Salas et al. 2009). Moreover, direct infusion of mecamylamine into the IPN, but not the VTA, of nicotine-dependent wild-type mice precipitated the expression of somatic withdrawal signs (Salas et al. 2009). This suggests that α5* and α2* nAChRs in the MHb-IPN pathway, and perhaps other nAChR subtypes enriched in this pathway, regulate the expression of somatic signs of nicotine withdrawal. However, little is known concerning the role for nAChRs in the MHb-IPN tract in regulating affective aspects of nicotine withdrawal, and in particular, withdrawal-associated reward deficits that may motivate relapse during periods of abstinence in human smokers.

Taken together, the above findings support a key role for α5 and β4, and perhaps also α2 and α3 nAChRs, which are enriched in the MHb-IPN pathway in regulating nicotine reinforcement and the expression of the nicotine withdrawal syndrome in nicotine-dependent rodents. As such, nAChRs containing these subunits may be important targets for the development of novel therapeutics for smoking cessation.

NICOTINIC INVOLVEMENT IN BEHAVIORS RELATED TO ONGOING SMOKING: EFFECTS OF NICOTINE ON DEPRESSION, APPETITE, AND ATTENTION

Nicotine reinforcement and avoidance of the aversive effects of nicotine withdrawal are clearly fundamental for ongoing smoking, but a number of other factors are also likely to contribute to smoking behavior in humans. nAChRs are expressed throughout the brain on both excitatory and inhibitory neurons, with the ability to increase inhibition of circuits when excitation is high and to increase excitation when circuits are less active (Picciotto 2003). The result of this circuit-level integration is that nicotine can modulate behavioral function bidirectionally, acting as a stimulant and increasing anxiety under some conditions and decreasing activity and anxiety in others (Picciotto 2003).

Some individuals report that they smoke to improve attention (Rusted and Warburton 1992; Warburton et al. 1992), and the ability of smoking to improve attentional function in individuals with schizophrenia (George et al. 2002) is likely to contribute to their extremely high rates of smoking. Similarly, a large proportion of smokers report that they smoke to control symptoms of anxiety and depression (Picciotto et al. 2002), and the rate of smoking in individuals with affective disorders is more than double the rate in the general population (Kalman et al. 2005). The idea that some individuals smoke to self-medicate psychiatric symptoms is thought to underlie the high rate of smoking in individuals with psychiatric illness, and some estimates suggest that ∼44% of cigarettes are sold to individuals with a current psychiatric condition (Lasser et al. 2000).

Effects of nAChRs on Anxiety- and Depression-Like Behaviors

Studies in mouse genetic models have helped identify the nAChR subtypes involved in a number of behavioral effects of nicotine that may affect human smoking. Nicotine is known to have both anxiolytic and anxiogenic effects in rodents (File et al. 2000), and these effects are likely to depend on different nAChR subtypes. For example, chronic administration of nicotine increased anxiety-like behavior in female, but not male, mice (Caldarone et al. 2008), whereas knockout mice lacking the β4 subunit show less anxiety-like behaviors at baseline (Salas et al. 2003) and no difference in anxiety-like behaviors were seen at baseline or following nicotine administration in mice lacking the β2 subunit (Caldarone et al. 2008). Similarly, female, but not male, knockout mice lacking the α5 subunit showed reduced anxiety-like behavior, and this may be related to progesterone effects on α5 subunit expression (Gangitano et al. 2009). These data suggest that stimulation of α5β4* nAChRs is important for the anxiogenic effects of nicotine.

The effects of nicotine on depression-like behavior are also complex. Studies in the Flinders sensitive line of rats have shown that acute nicotine administration is antidepressant-like in the forced swim test and that this effect can be blocked by the nicotinic antagonist mecamylamine, suggesting that activation of nAChRs decreases depression-like behavior in this model (Tizabi et al. 2000). In contrast, the nicotinic antagonist mecamylamine has antidepressant-like effects in mice (Caldarone et al. 2004; Rabenstein et al. 2006; Andreasen et al. 2009), and can be effective as an add-on medication in depressed human subjects who are nonresponsive to an SSRI (George et al. 2008). Similarly, nicotinic partial agonists, that would be expected to decrease activity of acetylcholine at endogenous nAChRs when cholinergic tone is high but increase activity of nAChRs when cholinergic tone is low, are effective in mouse models of antidepressant efficacy (Mineur et al. 2007, 2009b, 2011b; Rollema et al. 2009; Caldarone et al. 2011) and in human smokers (Philip et al. 2009). These data suggest that inhibition of nAChRs in some neuronal subtypes or brain areas and activation in others may contribute to an antidepressant-like effect of nicotinic drugs, so the cycles of nAChR activation and desensitization experienced by smokers may result in fluctuations in depressive symptoms throughout the day. Both the antagonist mecamylamine (Rabenstein et al. 2006) and the partial agonist sazetidine (Caldarone et al. 2011), as well as the classical antidepressant amitriptyline (Caldarone et al. 2004), are ineffective in mice lacking the β2 subunit, and these knockout mice show decreased depression-like behavior at baseline, suggesting that β2* nAChRs are critical for the antidepressant-like effects of nicotinic drugs; however, mice lacking the α7 subunit are also resistant to the antidepressant-like effects of mecamylamine (Rabenstein et al. 2006), and the effects of the partial agonist sazetidine could be blocked with mecamylamine (Caldarone et al. 2011), suggesting that other nAChR subtypes may also contribute to the antidepressant-like effects of nicotinic drugs, and that activation as well as inhibition of nAChRs can result in antidepressant-like effects.

Effects of nAChRs on Behaviors Related to Attention

In addition to effects on anxiety and depression, nicotine and nicotinic drugs can improve attention in control subjects (Rusted and Warburton 1992) and individuals with schizophrenia (Sacco et al. 2004). Interestingly, after control subjects quit smoking and transition past the acute withdrawal period, their working memory function improves compared with when they were smoking (George et al. 2002). In contrast, individuals with schizophrenia show impaired attentional performance once they quit smoking (George et al. 2002). Genetic and functional studies have implicated α7 nAChRs in prepulse inhibition, a physiological marker associated with schizophrenia (Leonard et al. 2000; Freedman et al. 2003). Mice lacking the α7 subunit have been shown to have impaired trace eye-blink conditioning (Brown et al. 2010). These data suggest that optimal nAChR stimulation is achieved at baseline in control subjects or wild-type mice with normal α7 nAChR levels, whereas nicotine from tobacco smoke can further improve attention in individuals with schizophrenia.

Studies using knockout mice with lentiviral-mediated reexpression have shown that β2* nAChRs in the prelimbic medial prefrontal cortex (mPFC) are important for normal performance of the five-choice serial reaction time task measuring visual attention. Similarly, rapid acetylcholine transients in the mPFC are correlated with attention to brief cues, and mice lacking the β2, but not the α7, subunit show impaired performance in an attentional task (Parikh et al. 2007, 2008). Studies in rodents have also implicated nAChRs on glutamatergic thalamocortical neurons impinging on layer five pyramidal neurons in the prefrontal cortex as an important site for nAChR control of attention (Lambe et al. 2005; Bailey et al. 2010). Overall, it appears that nAChRs in thalamo-corticothalamic loops are important for regulating glutamate release in this circuit, and for mediating the effects of acetylcholine on attentional function (Heath and Picciotto 2009).

In addition to effects of nicotine on attentional function in adulthood, many studies have shown a role for nAChRs in maturation of circuits important for attention during development (reviewed in Heath and Picciotto 2009). Mice administered nicotine during the adolescent period show deficits in the five-choice serial reaction time task that are associated with decreased expression of mGluR2 receptors, and that are rescued by administration of mGluR2 agonists (Counotte et al. 2011). Similarly, α5/β2* nAChRs on layer six cortical glutamatergic projection neurons to the thalamus are essential in maturation of this circuit and for normal adult performance in passive avoidance, a somatosensory aversive learning task (King et al. 2003; Heath et al. 2010). Electrophysiological studies have shown that currents mediated through α5/β2* nAChRs are maximal in the early postnatal period (Kassam et al. 2008). Nicotine administration during this same period alters performance in the passive avoidance task in normal mice as well as in mice with expression of β2* nAChRs exclusively in corticothalamic neurons (Heath et al. 2010), suggesting that disrupting normal acetylcholine signaling through these nAChRs during a critical period has lasting effects on function of the corticothalamic circuit in passive avoidance behavior. Interestingly, modulation of nicotinic function through the lynx1 protein is also important for regulating the critical period for activity-dependent visual system development (Morishita et al. 2010).

Effects of nAChRs on Food Intake

The anorexic effects of smoking have been well-documented in human subjects, and the principal reason cited by female teenagers for why they smoke is weight control (Voorhees et al. 2002). On average, smokers weigh ∼5 kg less than nonsmokers and have significantly lower body mass index than nonsmokers (Albanes et al. 1987). Similarly, nicotine decreases feeding in animal models (Grunberg et al. 1987), suggesting that the nicotine in tobacco is important for the effects of smoking on appetite. Whereas β2α4α6* nAChRs are critical for nicotine reward and reinforcement, β4* nAChRs on proopiomelanocortin (POMC) neurons in the arcuate nucleus of the hypothalamus are necessary for the appetite-suppressing effects of nicotine (Mineur et al. 2011a). There are a number of nAChR subtypes expressed in the hypothalamus (Jo et al. 2002, 2005), and nicotine can stimulate the firing of both POMC neurons, which signal satiety, and neuropeptide Y (NPY) neurons, which stimulate food seeking (Huang et al. 2011). Interestingly, in a slice preparation, the effects of nicotine on firing of POMC neurons persist longer than firing of NPY neurons, showing that at the circuit level, stimulation of nAChRs shifts the balance toward neuronal patterns that signal satiety (Huang et al. 2011).

Although β4* nAChRs on POMC neurons can signal satiety, nAChRs in the mesolimbic dopamine system may be more important for the motivation to work for food. The DA system is important for the hedonic value of both drugs of abuse, like nicotine, and palatable foods (Kenny 2011). Food or sugar intake can increase acetylcholine release in the VTA (Hajnal et al. 1998; Rada et al. 2000), and withdrawal from binge eating increases acetylcholine release in the nAc (Avena et al. 2008). Interestingly, blocking α7 nicotinic AChRs in the VTA can decrease food seeking (Schilstrom et al. 1998). In contrast, previous nicotine exposure increases the motivation of mice to work for food, and this is attributable to non-β2* nAChRs (Brunzell et al. 2006). Taken together, these data show that, in addition to its effects on satiety mediated through POMC neuron signaling, acetylcholine in the mesolimbic system is also likely to affect motivation to seek palatable foods and to modulate their hedonic value through distinct nAChR subtypes.

CONCLUSIONS

The high-affinity α4β2* nAChRs play a key role in the behavioral actions of nicotine that contribute to the development of tobacco dependence, including its effects on brain circuitries involved in reinforcement, mood, attention, and food consumption. Recent evidence has shed important light on other nAChR subunits that may also be incorporated into the α4β2* nAChRs that regulate these processes. For example, incorporation of α6 and β3 nAChR subunits in α4β2* nAChRs in the mesoaccumbens pathway gives rise to a nAChR subtype (α4α6β2β3*) that appears to play a particularly important role in nicotine reinforcement. In addition, nAChR subtypes containing α5, α3, and/or β4 nAChR subunits have been implicated in regulating the aversive properties of nicotine that control the quantities of the drug consumed and in the development of tobacco dependence. In addition, β4* nAChRs also play an important role in appetite regulation, particularly the inhibitory effects of nicotine on appetite that underlie the anorectic effects of tobacco smoke. A more refined understanding of the precise contribution of discrete nAChR subtypes to these addiction-relevant properties of nicotine may reveal important new targets for the development of novel therapeutics for tobacco dependence. Moreover, such novel therapeutics could also have utility for the treatment of mood and attention disorders and the control of body weight.

ACKNOWLEDGMENTS

The authors are supported by grants DA14241 (M.R.P.) and DA020686 (P.J.K.) from the National Institute on Drug Abuse.

Footnotes

Editors: R. Christopher Pierce and Paul J. Kenny

Additional Perspectives on Addiction available at www.perspectivesinmedicine.org

REFERENCES

  1. Albanes D, Jones DY, Micozzi MS, Mattson ME 1987. Associations between smoking and body weight in the US population: Analysis of NHANES II. Am J Pub Health 77: 439–444 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Amos CI, Wu X, Broderick P, Gorlov IP, Gu J, Eisen T, Dong Q, Zhang Q, Gu X, Vijayakrishnan J, et al. 2008. Genome-wide association scan of tag SNPs identifies a susceptibility locus for lung cancer at 15q25.1. Nature Gen 40: 616–622 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Andreasen JT, Olsen GM, Wiborg O, Redrobe JP 2009. Antidepressant-like effects of nicotinic acetylcholine receptor antagonists, but not agonists, in the mouse forced swim and mouse tail suspension tests. J Psychopharmacology 23: 797–804 [DOI] [PubMed] [Google Scholar]
  4. Avena NM, Bocarsly ME, Rada P, Kim A, Hoebel BG 2008. After daily bingeing on a sucrose solution, food deprivation induces anxiety and accumbens dopamine/acetylcholine imbalance. Physiol Behav 94: 309–315 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bailey CDC, De Biasi M, Fletcher PJ, Lambe EK 2010. The nicotinic acetylcholine receptor α5 subunit plays a key role in attention circuitry and accuracy. J Neurosci 30: 9241–9252 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Berrettini W, Yuan X, Tozzi F, Song K, Francks C, Chilcoat H, Waterworth D, Muglia P, Mooser V 2008. α-5/α-3 Nicotinic receptor subunit alleles increase risk for heavy smoking. Mol Psychiatry 13: 368–373 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bierut LJ, Stitzel JA, Wang JC, Hinrichs AL, Grucza RA, Xuei X, Saccone NL, Saccone SF, Bertelsen S, Fox L, et al. 2008. Variants in nicotinic receptors and risk for nicotine dependence. Am J Psychiatry 165: 1163–1171 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bromberg-Martin ES, Hikosaka O 2011. Lateral habenula neurons signal errors in the prediction of reward information. Nat Neurosci 14: 1209–1216 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Brown K, Comalli D, De Biasi M, Woodruff-Pak D 2010. Trace eyeblink conditioning is impaired in α7 but not in β2 nicotinic acetylcholine receptor knockout mice. Front Behav Neurosci 4: 166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Brunzell D, McIntosh J 2012. α7 Nicotinic acetylcholine receptors modulate motivation to self-administer nicotine: Implications for smoking and schizophrenia. Neuropsychopharmacology 37: 1134–1143 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brunzell DH, Chang JR, Schneider B, Olausson P, Taylor JR, Picciotto MR 2006. β2-Subunit-containing nicotinic acetylcholine receptors are involved in nicotine-induced increases in conditioned reinforcement but not progressive ratio responding for food in C57BL/6 mice. Psychopharmacology 184: 328–338 [DOI] [PubMed] [Google Scholar]
  12. Brunzell D, Mineur Y, Neve R, Picciotto M 2009. Nucleus accumbens CREB activity is necessary for nicotine conditioned place preference. Neuropsychopharmacology 34: 1993–2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Brunzell D, Boschen K, Hendrick E, Beardsley P, McIntosh J 2010. α-Conotoxin MII-sensitive nicotinic acetylcholine receptors in the nucleus accumbens shell regulate progressive ratio responding maintained by nicotine. Neuropsychopharmacology 35: 665–673 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Caldarone BJ, Harrist A, Cleary MA, Beech RD, King SL, Picciotto MR 2004. High-affinity nicotinic acetylcholine receptors are required for antidepressant effects of amitriptyline on behavior and hippocampal cell proliferation. Biol Psychiatry 56: 657–664 [DOI] [PubMed] [Google Scholar]
  15. Caldarone BJ, King SL, Picciotto MR 2008. Sex differences in anxiety-like behavior and locomotor activity following chronic nicotine exposure in mice. Neurosci Lett 439: 187–191 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Caldarone B, Wang D, Paterson N, Manzano M, Fedolak A, Cavino K, Kwan M, Hanania T, Chellappan S, Kozikowski A, et al. 2011. Dissociation between duration of action in the forced swim test in mice and nicotinic acetylcholine receptor occupancy with sazetidine, varenicline, and 5-I-A85380. Psychopharmacology 217: 199–210 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Clarke PB, Fu DS, Jakubovic A, Fibiger HC 1988. Evidence that mesolimbic dopaminergic activation underlies the locomotor stimulant action of nicotine in rats. J Pharmacol Exp Ther 246: 701–708 [PubMed] [Google Scholar]
  18. Corrigall WA, Franklin KB, Coen KM, Clarke PB 1992. The mesolimbic dopaminergic system is implicated in the reinforcing effects of nicotine. Psychopharmacology 107: 285–289 [DOI] [PubMed] [Google Scholar]
  19. Corrigall WA, Coen KM, Adamson KL 1994. Self-administered nicotine activates the mesolimbic dopamine system through the ventral tegmental area. Brain Res 653: 278–284 [DOI] [PubMed] [Google Scholar]
  20. Counotte DS, Goriounova NA, Li KW, Loos M, van der Schors RC, Schetters D, Schoffelmeer ANM, Smit AB, Mansvelder HD, Pattij T, et al. 2011. Lasting synaptic changes underlie attention deficits caused by nicotine exposure during adolescence. Nat Neurosci 14: 417–419 [DOI] [PubMed] [Google Scholar]
  21. Cuello AC, Emson PC, Paxinos G, Jessell T 1978. Substance P containing and cholinergic projections from the habenula. Brain Res 149: 413–429 [DOI] [PubMed] [Google Scholar]
  22. De Biasi M, Salas R 2008. Influence of neuronal nicotinic receptors over nicotine addiction and withdrawal. Exp Biol Med 233: 917–929 [DOI] [PubMed] [Google Scholar]
  23. Doherty K, Kinnunen T, Militello FS, Garvey AJ 1995. Urges to smoke during the first month of abstinence: Relationship to relapse and predictors. Psychopharmacology 119: 171–178 [DOI] [PubMed] [Google Scholar]
  24. Drenan R, Grady S, Whiteaker P, McClure-Begley T, McKinney S, Miwa J, Bupp S, Heintz N, McIntosh J, Bencherif M, et al. 2008a. In vivo activation of midbrain dopamine neurons via sensitized, high-affinity α6* nicotinic acetylcholine receptors. Neuron 60: 123–136 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Drenan RM, Nashmi R, Imoukhuede P, Just H, McKinney S, Lester HA 2008b. Subcellular trafficking, pentameric assembly, and subunit stoichiometry of neuronal nicotinic acetylcholine receptors containing fluorescently labeled α6 and β3 subunits. Mol Pharmacol 73: 27–41 [DOI] [PubMed] [Google Scholar]
  26. Drenan RM, Grady SR, Steele AD, McKinney S, Patzlaff NE, McIntosh JM, Marks MJ, Miwa JM, Lester HA 2010. Cholinergic modulation of locomotion and striatal dopamine release is mediated by α6α4* nicotinic acetylcholine receptors. J Neurosci 30: 9877–9889 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Eckenrode TC, Barr GA, Battisti WP, Murray M 1987. Acetylcholine in the interpeduncular nucleus of the rat: Normal distribution and effects of deafferentation. Brain Res 418: 273–286 [DOI] [PubMed] [Google Scholar]
  28. Exley R, Maubourguet N, David V, Eddine R, Evrard A, Pons S, Marti F, Threlfell S, Cazala P, McIntosh JM, et al. 2011. Distinct contributions of nicotinic acetylcholine receptor subunit α4 and subunit α6 to the reinforcing effects of nicotine. Proc Natl Acad Sci 108: 7577–7582 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. File SE, Cheeta S, Kenny PJ 2000. Neurobiological mechanisms by which nicotine mediates different types of anxiety. Eur J Pharmacol 393: 231–236 [DOI] [PubMed] [Google Scholar]
  30. Fowler CD, Lu Q, Johnson PM, Marks MJ, Kenny PJ 2011. Habenular α5 nicotinic receptor subunit signalling controls nicotine intake. Nature 471: 597–601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Frahm S, Slimak MA, Ferrarese L, Santos-Torres J, Antolin-Fontes B, Auer S, Filkin S, Pons S, Fontaine J-F, Tsetlin V, et al. 2011. Aversion to nicotine is regulated by the balanced activity of β4 and α5 nicotinic receptor subunits in the medial habenula. Neuron 70: 522–535 [DOI] [PubMed] [Google Scholar]
  32. Freedman R, Olincy A, Ross RG, Waldo MC, Stevens KE, Adler LE, Leonard S 2003. The genetics of sensory gating deficits in schizophrenia. Curr Psychiatry Rep 5: 155–161 [DOI] [PubMed] [Google Scholar]
  33. Gangitano D, Salas R, Teng Y, Perez E, De Biasi M 2009. Progesterone modulation of α5 nAChR subunits influences anxiety-related behavior during estrus cycle. Genes Brain Behav 8: 398–406 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. George TP, Vessicchio JC, Termine A, Sahady DM, Head CA, Pepper WT, Kosten TR, Wexler BE 2002. Effects of smoking abstinence on visuospatial working memory function in schizophrenia. Neuropsychopharmacology 26: 75–85 [DOI] [PubMed] [Google Scholar]
  35. George TP, Sacco KA, Vessicchio JC, Weinberger AH, Shytle RD 2008. Nicotinic antagonist augmentation of selective serotonin reuptake inhibitor-refractory major depressive disorder: A preliminary study. J Clin Psychopharm 28: 340–344 [DOI] [PubMed] [Google Scholar]
  36. Girod R, Barazangi N, McGehee D, Role LW 2000. Facilitation of glutamatergic neurotransmission by presynaptic nicotinic acetylcholine receptors. Neuropharmacology 39: 2715–2725 [DOI] [PubMed] [Google Scholar]
  37. Gotti C, Guiducci S, Tedesco V, Corbioli S, Zanetti L, Moretti M, Zanardi A, Rimondini R, Mugnaini M, Clementi F, et al. 2010. Nicotinic acetylcholine receptors in the mesolimbic pathway: Primary role of ventral tegmental area α6β2* receptors in mediating systemic nicotine effects on dopamine release, locomotion, and reinforcement. J Neurosci 30: 5311–5325 [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Grady SR, Meinerz NM, Cao J, Reynolds A, Picciotto MR, Changeux J-P, McIntosh MJ, Marks MJ, Collins AC 2001. Nicotinic agonists stimulate acetylcholine release from mouse interpeduncular nucleus: A function mediated by a different nAChR than dopamine release from striatum. J Neurochem 76: 258–268 [DOI] [PubMed] [Google Scholar]
  39. Grady S, Moretti M, Zoli M, Marks M, Zanardi A, Pucci L, Clementi F, Gotti C 2009. Rodent habenulo-interpeduncular pathway expresses a large variety of uncommon nAChR subtypes, but only the α3β4* and α3β3β4* subtypes mediate acetylcholine release. J Neurosci 29: 2272–2282 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Grucza RA, Wang JC, Stitzel JA, Hinrichs AL, Saccone SF, Saccone NL, Bucholz KK, Cloninger CR, Neuman RJ, Budde JP, et al. 2008. A risk allele for nicotine dependence in CHRNA5 is a protective allele for cocaine dependence. Biol Psychiatry 64: 922–929 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Grunberg NE, Winders SE, Popp KA 1987. Sex differences in nicotine’s effects on consummatory behavior and body weight in rats. Psychopharmacology 91: 221–225 [DOI] [PubMed] [Google Scholar]
  42. Hajnal A, Pothos EN, Lenard L, Hoebel BG 1998. Effects of feeding and insulin on extracellular acetylcholine in the amygdala of freely moving rats. Brain Res 785: 41–48 [DOI] [PubMed] [Google Scholar]
  43. Heath CJ, Picciotto MR 2009. Nicotine-induced plasticity during development: Modulation of the cholinergic system and long-term consequences for circuits involved in attention and sensory processing. Neuropharmacology 56: 254–262 [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Heath CJ, King SL, Gotti C, Marks MJ, Picciotto MR 2010. Cortico-thalamic connectivity is vulnerable to nicotine exposure during early postnatal development through α4/β2/α5 nicotinic acetylcholine receptors. Neuropsychopharmacology 35: 2324–2338 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Hikosaka O 2010. The habenula: From stress evasion to value-based decision-making. Nature Rev Neurosci 11: 503–513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Huang H, Xu Y, van den Pol A 2011. Nicotine excites hypothalamic arcuate anorexigenic proopiomelanocortin neurons and orexigenic neuropeptide Y neurons: Similarities and differences. J Neurophysiol 106: 1191–1202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Hung RJ, McKay JD, Gaborieau V, Boffetta P, Hashibe M, Zaridze D, Mukeria A, Szeszenia-Dabrowska N, Lissowska J, Rudnai P, et al. 2008. A susceptibility locus for lung cancer maps to nicotinic acetylcholine receptor subunit genes on 15q25. Nature 452: 633–637 [DOI] [PubMed] [Google Scholar]
  48. Jhou TC, Fields HL, Baxter MG, Saper CB, Holland PC 2009. The rostromedial tegmental nucleus (RMTg), a GABAergic afferent to midbrain dopamine neurons, encodes aversive stimuli and inhibits motor responses. Neuron 61: 786–800 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Jo YH, Talmage DA, Role LW 2002. Nicotinic receptor-mediated effects on appetite and food intake. J Neurobiology 53: 618–632 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Jo Y-H, Wiedl D, Role LW 2005. Cholinergic modulation of appetite-related synapses in mouse lateral hypothalamic slice. J Neurosci 25: 11133–11144 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Kalman D, Morissette S, George T 2005. Co-morbidity of smoking in patients with psychiatric and substance use disorders. Am J Addictions 14: 106–123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Kassam S, Herman P, Goodfellow N, Alves N, Lambe E 2008. Developmental excitation of corticothalamic neurons by nicotinic acetylcholine receptors. J Neurosci 28: 8756–8764 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Kenny PJ 2011. Common cellular and molecular mechanisms in obesity and drug addiction. Nat Rev Neurosci 12: 638–651 [DOI] [PubMed] [Google Scholar]
  54. Kenny PJ, Markou A 2001. Neurobiology of the nicotine withdrawal syndrome. Pharmacol Biochem Behav 70: 531–549 [DOI] [PubMed] [Google Scholar]
  55. King SL, Marks MJ, Grady SR, Caldarone BJ, Koren AO, Mukhin AG, Collins AC, Picciotto MR 2003. Conditional expression in corticothalamic efferents reveals a developmental role for nicotinic acetylcholine receptors in modulation of passive avoidance behavior. J Neurosci 23: 3837–3843 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. King S, Caldarone B, Picciotto M 2004. β2-Subunit-containing nicotinic acetylcholine receptors are critical for dopamine-dependent locomotor activation following repeated nicotine administration. Neuropharmacology 47: 132–139 [DOI] [PubMed] [Google Scholar]
  57. Klink R, de Kerchove d’Exaerde A, Zoli M, Changeux JP 2001. Molecular and physiological diversity of nicotinic acetylcholine receptors in the midbrain dopaminergic nuclei. J Neurosci 21: 1452–1463 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Koob GF 1992. Drugs of abuse: Anatomy, pharmacology and function of reward pathways. Trends Pharmacol Sci 13: 177–184 [DOI] [PubMed] [Google Scholar]
  59. Lambe EK, Olausson P, Horst NK, Taylor JR, Aghajanian GK 2005. Hypocretin and nicotine excite the same thalamocortical synapses in prefrontal cortex: Correlation with improved attention in rat. J Neurosci 25: 5225–5229 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Lasser K, Boyd JW, Woolhandler S, Himmelstein DU, McCormick D, Bor DH 2000. Smoking and mental illness: A population-based prevalence study. JAMA 284: 2606–2610 [DOI] [PubMed] [Google Scholar]
  61. Lecourtier L, Kelly PH 2007. A conductor hidden in the orchestra? Role of the habenular complex in monoamine transmission and cognition. Neurosci Biobehav Rev 31: 658–672 [DOI] [PubMed] [Google Scholar]
  62. Le Marchand L, Derby KS, Murphy SE, Hecht SS, Hatsukami D, Carmella SG, Tiirikainen M, Wang H 2008. Smokers with the CHRNA lung cancer-associated variants are exposed to higher levels of nicotine equivalents and a carcinogenic tobacco-specific nitrosamine. Cancer Res 68: 9137–9140 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Le Novère N, Zoli M, Changeux J-P 1996. Neuronal nicotinic receptor α6 subunit mRNA is selectively concentrated in catecholaminergic nuclei of the rat brain. Eur J Neurosci 8: 2428–2439 [DOI] [PubMed] [Google Scholar]
  64. Leonard S, Breese C, Adams C, Benhammou K, Gault J, Stevens K, Lee M, Adler L, Olincy A, Ross R, et al. 2000. Smoking and schizophrenia: Abnormal nicotinic receptor expression. Eur J Pharmacol 393: 237–242 [DOI] [PubMed] [Google Scholar]
  65. Lips EH, Gaborieau V, McKay JD, Chabrier A, Hung RJ, Boffetta P, Hashibe M, Zaridze D, Szeszenia-Dabrowska N, Lissowska J, et al. 2010. Association between a 15q25 gene variant, smoking quantity and tobacco-related cancers among 17,000 individuals. Int J Epidemiol 39: 563–577 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Mansvelder HD, McGehee DS 2000. Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron 27: 349–357 [DOI] [PubMed] [Google Scholar]
  67. Mansvelder HD, Keath JR, McGehee DS 2002. Synaptic mechanisms underlie nicotine-induced excitability of brain reward areas. Neuron 33: 905–919 [DOI] [PubMed] [Google Scholar]
  68. Maskos U, Molles BE, Pons S, Besson M, Guiard BP, Guilloux JP, Evrard A, Cazala P, Cormier A, Mameli-Engvall M, et al. 2005. Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature 436: 103–107 [DOI] [PubMed] [Google Scholar]
  69. Mata MM, Schrier BK, Moore RY 1977. Interpeduncular nucleus: Differential effects of habenula lesions on choline acetyltransferase and glutamic acid decarboxylase. Exp Neurol 57: 913–921 [DOI] [PubMed] [Google Scholar]
  70. Matsumoto M, Hikosaka O 2009. Representation of negative motivational value in the primate lateral habenula. Nature Neurosci 12: 77–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. McGranahan TM, Patzlaff NE, Grady SR, Heinemann SF, Booker TK 2011. α4β2 nicotinic acetylcholine receptors on dopaminergic neurons mediate nicotine reward and anxiety relief. J Neurosci 31: 10891–10902 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Mineur YS, Somenzi O, Picciotto MR 2007. Cytisine, a partial agonist of high-affinity nicotinic acetylcholine receptors, has antidepressant-like properties in male C57BL/6J mice. Neuropharmacology 52: 1256–1262 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Mineur Y, Brunzell D, Grady S, Lindstrom J, McIntosh J, Marks M, King S, Picciotto M 2009a. Localized low level re-expression of high affinity mesolimbic nicotinic acetylcholine receptors restores nicotine-induced locomotion but not place conditioning. Genes Brain Behav 8: 257–266 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Mineur YS, Eibl C, Young G, Kochevar C, Papke RL, Gundisch D, Picciotto MR 2009b. Cytisine-based nicotinic partial agonists as novel antidepressant compounds. J Pharmacol Exp Therap 329: 377–386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Mineur YS, Abizaid A, Rao Y, Salas R, DiLeone RJ, Gundisch D, Diano S, De Biasi M, Horvath TL, Gao X-B, et al. 2011a. Nicotine decreases food intake through activation of POMC neurons. Science 332: 1330–1332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Mineur YS, Einstein EB, Seymour PA, Coe JW, O’Neill BT, Rollema H, Picciotto MR 2011b. α4β2 nicotinic acetylcholine receptor partial agonists with low intrinsic efficacy have antidepressant-like properties. Behav Pharmacol 22: 291–299 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Morishita H, Miwa JM, Heintz N, Hensch TK 2010. Lynx1, a cholinergic brake, limits plasticity in adult visual cortex. Science 330: 1238–1240 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Mugnaini M, Tessari M, Tarter G, Merlo Pich E, Chiamulera C, Bunnemann B 2002. Upregulation of [3H]methyllycaconitine binding sites following continuous infusion of nicotine, without changes of α7 or α6 subunit mRNA: An autoradiography and in situ hybridization study in rat brain. Euro J Neurosci 16: 1633–1646 [DOI] [PubMed] [Google Scholar]
  79. Museo E, Wise RA 1994. Place preference conditioning with ventral tegmental injections of cytisine. Life Sci 55: 1179–1186 [DOI] [PubMed] [Google Scholar]
  80. Parikh V, Kozak R, Martinez V, Sarter M, Parikh V, Kozak R, Martinez V, Sarter M 2007. Prefrontal acetylcholine release controls cue detection on multiple timescales. Neuron 56: 141–154 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Parikh V, Man K, Decker MW, Sarter M, Parikh V, Man K, Decker MW, Sarter M 2008. Glutamatergic contributions to nicotinic acetylcholine receptor agonist-evoked cholinergic transients in the prefrontal cortex. J Neurosci 28: 3769–3780 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Parrott AC 1993. Cigarette smoking: Effects upon self-rated stress and arousal over the day. Addictive Behav 18: 389–395 [DOI] [PubMed] [Google Scholar]
  83. Philip NS, Carpenter LL, Tyrka AR, Whiteley LB, Price LH 2009. Varenicline augmentation in depressed smokers: An 8-week, open-label study. J Clin Psychiatry 70: 1026–1031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Piasecki TM, Fiore MC, Baker TB 1998. Profiles in discouragement: Two studies of variability in the time course of smoking withdrawal symptoms. J Abnorm Psychol 107: 238–251 [DOI] [PubMed] [Google Scholar]
  85. Piasecki TM, Niaura R, Shadel WG, Abrams D, Goldstein M, Fiore MC, Baker TB 2000. Smoking withdrawal dynamics in unaided quitters. J Abnorm Psychol 109: 74–86 [DOI] [PubMed] [Google Scholar]
  86. Piasecki TM, Jorenby DE, Smith SS, Fiore MC, Baker TB 2003. Smoking withdrawal dynamics: II. Improved tests of withdrawal-relapse relations. J Abnorm Psychol 112: 14–27 [PubMed] [Google Scholar]
  87. Picciotto MR 2003. Nicotine as a modulator of behavior: Beyond the inverted U. Trends Pharmacol Sci 24: 493–499 [DOI] [PubMed] [Google Scholar]
  88. Picciotto MR, Zoli M, Léna C, Bessis A, Lallemand Y, Le Novère N, Vincent P, Merlo Pich E, Brulet P, Changeux J-P 1995. Abnormal avoidance learning in mice lacking functional high-affinity nicotine receptor in the brain. Nature 374: 65–67 [DOI] [PubMed] [Google Scholar]
  89. Picciotto MR, Zoli M, Rimondini R, Léna C, Marubio LM, Merlo Pich E, Fuxe K, Changeux JP 1998. Acetylcholine receptors containing the β-2 subunit are involved in the reinforcing properties of nicotine. Nature 391: 173–177 [DOI] [PubMed] [Google Scholar]
  90. Picciotto MR, Brunzell DH, Caldarone BJ 2002. Effect of nicotine and nicotinic receptors on anxiety and depression. Neuroreport 13: 1097–1106 [DOI] [PubMed] [Google Scholar]
  91. Pons S, Fattore L, Cossu G, Tolu S, Porcu E, McIntosh J, Changeux J, Maskos U, Fratta W 2008. Crucial role of α4 and α6 nicotinic acetylcholine receptor subunits from ventral tegmental area in systemic nicotine self-administration. J Neurosci 28: 12318–12327 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Rabenstein RL, Caldarone BJ, Picciotto MR 2006. The nicotinic antagonist mecamylamine has antidepressant-like effects in wild type but not β2 or α7 nicotinic acetylcholine receptor knockout mice. Psychopharmacology 189: 395–401 [DOI] [PubMed] [Google Scholar]
  93. Rada PV, Mark GP, Yeomans JJ, Hoebel BG 2000. Acetylcholine release in ventral tegmental area by hypothalamic self-stimulation, eating, and drinking. Pharmacol Biochem Behav 65: 375–379 [DOI] [PubMed] [Google Scholar]
  94. Ren J, Qin C, Hu F, Tan J, Qiu L, Zhao S, Feng G, Luo M 2011. Habenula “cholinergic” neurons co-release glutamate and acetylcholine and activate postsynaptic neurons via distinct transmission modes. Neuron 69: 445–452 [DOI] [PubMed] [Google Scholar]
  95. Rollema H, Guanowsky V, Mineur Y, Shrikhande A, Coe J, Seymour P, Picciotto M 2009. Varenicline has antidepressant-like activity in the forced swim test and augments sertraline’s effect. Eur J Pharmacol 605: 114–116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Rusted JM, Warburton DM 1992. Facilitation of memory by post-trial administration of nicotine: Evidence for an attentional explanation. Psychopharmacology 108: 452–455 [DOI] [PubMed] [Google Scholar]
  97. Sacco KA, Bannon KL, George TP 2004. Nicotinic receptor mechanisms and cognition in normal states and neuropsychiatric disorders. J Psychopharmacology 18: 457–474 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Saccone SF, Hinrichs AL, Saccone NL, Chase GA, Konvicka K, Madden PAF, Breslau N, Johnson EO, Hatsukami D, Pomerleau O, et al. 2007. Cholinergic nicotinic receptor genes implicated in a nicotine dependence association study targeting 348 candidate genes with 3713 SNPs. Hum Mol Genetics 16: 36–49 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Saccone NL, Saccone SF, Hinrichs AL, Stitzel JA, Duan W, Pergadia ML, Agrawal A, Breslau N, Grucza RA, Hatsukami D, et al. 2009. Multiple distinct risk loci for nicotine dependence identified by dense coverage of the complete family of nicotinic receptor subunit (CHRN) genes. Am J Med Genet B 150B: 453–466 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Salas R, Pieri F, Fung B, Dani JA, De Biasi M 2003. Altered anxiety-related responses in mutant mice lacking the β4 subunit of the nicotinic receptor. J Neurosci 23: 6255–6263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Salas R, Sturm R, Boulter J, De Biasi M 2009. Nicotinic receptors in the habenulo-interpeduncular system are necessary for nicotine withdrawal in mice. J Neurosci 29: 3014–3018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Schilstrom B, Svensson HM, Svensson TH, Nomikos GG 1998. Nicotine and food induced dopamine release in the nucleus accumbens of the rat: Putative role of α7 nicotinic receptors in the ventral tegmental area. Neurosci 85: 1005–1009 [DOI] [PubMed] [Google Scholar]
  103. Schlaepfer IR, Hoft NR, Collins AC, Corley RP, Hewitt JK, Hopfer CJ, Lessem JM, McQueen MB, Rhee SH, Ehringer MA 2008. The CHRNA5/A3/B4 gene cluster variability as an important determinant of early alcohol and tobacco initiation in young adults. Biol Psychiatry 63: 1039–1046 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Sheffield EB, Quick MW, Lester RA 2000. Nicotinic acetylcholine receptor subunit mRNA expression and channel function in medial habenula neurons. Neuropharmacology 39: 2591–2603 [DOI] [PubMed] [Google Scholar]
  105. Sherva R, Wilhelmsen K, Pomerleau CS, Chasse SA, Rice JP, Snedecor SM, Bierut LJ, Neuman RJ, Pomerleau OF 2008. Association of a single nucleotide polymorphism in neuronal acetylcholine receptor subunit α5 (CHRNA5) with smoking status and with “pleasurable buzz” during early experimentation with smoking. Addiction 103: 1544–1552 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Stevens VL, Bierut LJ, Talbot JT, Wang JC, Sun J, Hinrichs AL, Thun MJ, Goate A, Calle EE 2008. Nicotinic receptor gene variants influence susceptibility to heavy smoking. Cancer Epidemiol Biomarkers Prev 17: 3517–3525 [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Sugama S, Cho BP, Baker H, Joh TH, Lucero J, Conti B 2002. Neurons of the superior nucleus of the medial habenula and ependymal cells express IL-18 in rat CNS. Brain Res 958: 1–9 [DOI] [PubMed] [Google Scholar]
  108. Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C, Whiteaker P, Marks MJ, Collins AC, Lester HA 2004. Nicotine activation of α4* receptors: Sufficient for reward, tolerance, and sensitization. Science 306: 1029–1032 [DOI] [PubMed] [Google Scholar]
  109. Thorgeirsson TE, Geller F, Sulem P, Rafnar T, Wiste A, Magnusson KP, Manolescu A, Thorleifsson G, Stefansson H, Ingason A, et al. 2008. A variant associated with nicotine dependence, lung cancer and peripheral arterial disease. Nature 452: 638–642 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Tizabi Y, Rezvani AH, Russell LT, Tyler KY, Overstreet DH 2000. Depressive characteristics of FSL rats: Involvement of central nicotinic receptors. Pharmacol Biochem Behav 66: 73–77 [DOI] [PubMed] [Google Scholar]
  111. Vincent SR, Staines WA, McGeer EG, Fibiger HC 1980. Transmitters contained in the efferents of the habenula. Brain Res 195: 479–484 [DOI] [PubMed] [Google Scholar]
  112. Voorhees CC, Schreiber GB, Schumann BC, Biro F, Crawford PB 2002. Early predictors of daily smoking in young women: The national heart, lung, and blood institute growth and health study. Prevent Med 34: 616–624 [DOI] [PubMed] [Google Scholar]
  113. Walters C, Brown S, Changeux J-P, Martin B, Damaj M 2006. The β2 but not α7 subunit of the nicotinic acetylcholine receptor is required for nicotine-conditioned place preference in mice. Psychopharmacol 184: 339–344 [DOI] [PubMed] [Google Scholar]
  114. Wang JC, Cruchaga C, Saccone NL, Bertelsen S, Liu P, Budde JP, Duan W, Fox L, Grucza RA, Kern J, et al. 2009. Risk for nicotine dependence and lung cancer is conferred by mRNA expression levels and amino acid change in CHRNA5. Hum Mol Genetics 18: 3125–3135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Wang Y, Broderick P, Matakidou A, Eisen T, Houlston RS 2010. Role of 5p15.33 (TERT-CLPTM1L), 6p21.33 and 15q25.1 (CHRNA5-CHRNA3) variation and lung cancer risk in never-smokers. Carcinogenesis 31: 234–238 [DOI] [PubMed] [Google Scholar]
  116. Warburton DM, Rusted JM, Fowler J 1992. A comparison of the attentional and consolidation hypotheses for the facilitation of memory by nicotine. Psychopharmacology 108: 443–447 [DOI] [PubMed] [Google Scholar]
  117. Weiss RB, Baker TB, Cannon DS, von Niederhausern A, Dunn DM, Matsunami N, Singh NA, Baird L, Coon H, McMahon WM, et al. 2008. A candidate gene approach identifies the CHRNA5-A3-B4 region as a risk factor for age-dependent nicotine addiction. PLoS Genet 4: e1000125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Wooltorton JRA, Pidoplichko VI, Broide RS, Dani JA 2003. Differential desensitization and distribution of nicotinic acetylcholine receptor subtypes in midbrain dopamine areas. J Neurosci 23: 3176–3185 [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Yang K, Buhlman L, Khan GM, Nichols RA, Jin G, McIntosh JM, Whiteaker P, Lukas RJ, Wu J 2011. Functional nicotinic acetylcholine receptors containing α6 subunits are on GABAergic neuronal boutons adherent to ventral tegmental area dopamine neurons. J Neurosci 31: 2537–2548 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Zhou FM, Liang Y, Dani JA 2001. Endogenous nicotinic cholinergic activity regulates dopamine release in the striatum. Nat. Neurosci 4: 1224–1229 [DOI] [PubMed] [Google Scholar]
  121. Zoli M, Moretti M, Zanardi A, McIntosh JM, Clementi F, Gotti C 2002. Identification of the nicotinic receptor subtypes expressed on dopaminergic terminals in the rat striatum. J Neurosci 22: 8785–8789 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Medicine are provided here courtesy of Cold Spring Harbor Laboratory Press

RESOURCES