Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jan 1.
Published in final edited form as: Neuropharmacology. 2013 Apr 28;76(0 0):10.1016/j.neuropharm.2013.04.028. doi: 10.1016/j.neuropharm.2013.04.028

Molecules and circuits involved in nicotine addiction: the many faces of smoking

M R Picciotto 1,*, Y S Mineur 1
PMCID: PMC3772953  NIHMSID: NIHMS473741  PMID: 23632083

Abstract

Tobacco smoking in humans is one of the most persistent and widespread addictions and is driven by nicotine in tobacco smoke. Over the last several decades, understanding of the molecular and cellular basis for nicotine addiction has increased tremendously as a result of pharmacological, molecular genetic, electrophysiological and behavioral studies of nicotine reinforcement. Studies of the biological basis for nicotine reinforcement has helped in the design of new treatments for smoking cessation such as varenicline; however, smokers report that they smoke for many reasons, including the ability to control symptoms of anxiety and depression or the desire to control appetite. Further, developmental exposure to tobacco smoke increases the likelihood of adult smoking. Here we review what is known about the molecular and circuit basis for a number of behaviors related to tobacco smoking. Leveraging the knowledge from studies of different behaviors mediated by nicotine receptors in multiple brain circuits could provide points of convergence that will inform future therapeutic development for smoking cessation.

Keywords: tobacco, smoking, molecular basis, animal models

Introduction

There is an essential paradox at the heart of smoking behavior. Among humans, tobacco is a highly addictive drug, however, in animal models, the primary addictive component of tobacco, nicotine, is less reinforcing than some other drugs of abuse, such as cocaine (Risner and Goldberg, 1983). There are likely to be many reasons for this discrepancy. First, tobacco is legal and access is almost universal, so there is more chance of exposure to tobacco than to illicit drugs, and the stigma of use may be lower. Second, there are more than 4,000 constituents of tobacco smoke, and other constituents in addition to nicotine may contribute to tobacco use. Third, and perhaps most important, nicotine has many effects on brain circuits and behavior in addition to its ability to stimulate neuronal systems involved in primary reinforcement, and the complex actions on these systems may contribute to smoking behavior and relapse in human tobacco users. In addition to the rewarding and reinforcing effects of nicotine, several other factors contribute to the initiation and maintenance of tobacco intake. Smokers report that they use tobacco because is it pleasurable, but they also report that that they use it to control appetite, to help with mood symptoms and to focus their attention. For instance, weight control is cited as the primary reason for initiation of cigarette use in teenage girls in the United States (Fulkerson and French, 2003). Given the current high rates of obesity and the focus on limiting food intake, this connection is alarming. Further, the rate of smoking in depressed patients is twice that of the average population (Kalman et al., 2005). While the connection between mood disorders and tobacco addiction is not fully understood, it is clear that depression, which affects up to 10% of the population, is a significant risk factor for tobacco use. Finally, maternal smoking is associated with an increased risk of attention hyperactivity disorder (ADHD) in children exposed to tobacco smoke in utero, and this points to one potential link between smoke exposure and circuits involved in attention (Heath and Picciotto, 2009; Picciotto et al., 2012). The deleterious effects of developmental smoke exposure also highlight the fact that tobacco use can have long lasting, transgenerational effects due to changes in the developing brain that can last many years after exposure.

Therefore, while there are many behavioral consequences of smoking, this review will focus on the multivariate effects of nicotine on behaviors related to depression and food intake, as well as the more classical brain systems and behaviors related to addiction. In addition, since one of the risk factors for tobacco smoking is parental tobacco use, we will review the effects of nicotine on brain development. It is clear that there are social and societal factors that support smoking and contribute to relapse, but in addition, there are molecular and neurobiological mechanisms that make it difficult to quit smoking. It is these diverse mechanisms that we will review here.

Molecular and anatomical basis for nicotine reinforcement

The primary reason that people smoke is that the nicotine in tobacco is addictive. Like other drugs of abuse, nicotine stimulates dopamine (DA) release from neurons in the mesolimbic system originating in the ventral tegmental area (VTA) and terminating in the nucleus accumbens (NAc). Nicotine can stimulate the firing rate of VTA neurons (Grenhoff et al., 1986; Picciotto et al., 1998), induces DA release from isolated nerve terminals (synaptosomes) (Grady et al., 1992) and increases the excitatory glutamatergic drive onto DA cell bodies in the VTA (Mansvelder et al., 2002; McGehee et al., 1995). Consistent with the ability of nicotine to potentiate DA signaling, peripheral nicotine administration can increase extracellular DA levels in the NAc of rodents for more than an hour (Benwell and Balfour, 1992; Picciotto et al., 1998). The ability of nicotine to potentiate glutamatergic signaling onto DA neurons in the VTA has been proposed as a mechanism underlying this prolonged nicotine-induced DA release, that outlasts the acute effects of nicotine on firing rate of DA neurons (Mansvelder and McGehee, 2000; Tang and Dani, 2009). Thus, nicotine is highly effective at stimulating the DA system, a circuit necessary for drug reinforcement (Koob, 1992).

If nicotine is so effective at stimulating the DA system, why isn’t it as effective as other psychostimulants in supporting drug self-administration in rodents? One reason may be that nicotine stimulates both glutamatergic and GABAergic inputs onto VTA DA neurons, resulting in a mixed excitation and inhibition of this circuit (Mansvelder et al., 2002; Wooltorton et al., 2003). Thus, nicotine efficiently activates the mesolimbic system, but the action of nicotine on many neuronal subtypes can mitigate the ability of nicotine to drive DA signaling (Picciotto, 2003).

Another reason may be that nicotine can stimulate its molecular targets in the brain, the nicotinic acetylcholine receptors (nAChRs), but also rapidly desensitizes these receptors (Grady et al., 1994; Picciotto et al., 2008; Pidoplichko et al., 1997). Therefore, nicotine can limit its own actions in the mesolimbic system. Repeated nicotine exposure results in decreased stimulation of DA neuron firing rate (Pidoplichko et al., 1997) and nicotine-elicited DA release from synaptosomes (Grady et al., 1994). Despite the limiting consequences of repeated nicotine administration, a number of studies have now shown that nAChR desensitization can promote DAergic signaling in the VTA by decreasing nicotine-stimulated GABA release (Mansvelder et al., 2002; Wooltorton et al., 2003), and by increasing signal to noise in the NAc through decreased tonic DA release but maintained phasic DA signaling (Rice and Cragg, 2004; Zhang and Sulzer, 2004). The complex interaction between nAChR stimulation and desensitization varies across nAChR subtypes, and therefore across brain regions and neuronal subtype (Picciotto et al., 2008), potentially contributing to individual differences in susceptibility to nicotine addiction. For example, even though nAChRs containing the α7 subunit desensitize very rapidly in vitro in response to high concentrations of nicotine (reviewed in (Picciotto et al., 2008)), this nAChR subtype mediates the ability of low concentrations of nicotine to stimulate glutamate release from terminals in VTA slices, even after several minutes of exposure. Thus, nAChR desensitization rates can vary depending on the cellular context in which they are measured, and the dose of nicotine used.

The nAChR subtypes that contribute to the ability of nicotine to stimulate the DA system and to support behaviors related to nicotine addiction have been characterized using transgenic mice with mutations in specific nAChR subunits (Table 1). Mouse VTA DA neurons express mRNAs encoding many different nAChR subunits, including α3, α4, α5, α6, α7, β2, β3 and β4 (Klink et al., 2001). In the rat NAc, the primary subtypes of nAChRs on DA terminals include α4/α5/β2, α4/α6/β2/β3 and α6/β2/β3 nAChRs (Zoli et al., 2002). Studies of transgenic mice expressing hypersensitive nAChRs subunits have shown that activation of receptors containing the α4 or α6 subunits is sufficient for nicotine-mediated stimulation of DA neuron firing and for nicotine preference in conditioned place preference paradigm (Drenan et al., 2008; Tapper et al., 2004). Similarly, knockout studies have shown that β2 subunit-containing (β2*) nAChRs are necessary for nicotine-elicited DA neuron firing, DA release from synaptosomes, locomotor activation, conditioned place preference, conditioned reinforcement and self-administration (Brunzell et al., 2009; Brunzell et al., 2006; Grady et al., 2001; King et al., 2004; Mineur et al., 2009a; Picciotto et al., 1998). The brain areas and neuronal subtypes involved in the ability of α4/β2* nAChRs to support nicotine-induced DA neuron firing and self-administration have been elucidated using viral-mediated rescue throughout the VTA (Maskos et al., 2005; Pons et al., 2008) or in specific VTA cell types (Tolu et al., 2012). Interestingly, these studies show that β2* nAChR expression in VTA is sufficient for neuronal firing and nicotine reinforcement (Maskos et al., 2005), but that these nAChRs must be expressed in both DA and GABA neurons of the VTA for nicotine to induce burst firing of DA neurons and to support nicotine self-administration (Tolu et al., 2012). Similarly, knockout of the α4 nAChR subunit only in DA neurons shows that α4/β2* nAChRs in DA neurons are required for nicotine reinforcement (McGranahan et al., 2011). Finally, recent studies using constitutive knockout mice or highly selective conotoxins have shown that α6* nAChRs are critical for nicotine-elicited dopamine release and self-administration in both mice and rats (Brunzell et al., 2010; Gotti et al., 2010; Pons et al., 2008). Taken together, a combination of pharmacological, molecular genetic, electrophysiological, neurochemical and behavioral approaches has identified α4/α6/β2* nAChRs in DA neurons of the VTA as essential for nicotine reinforcement, and suggests an important contribution of β2* nAChRs in GABA neurons of the VTA to this circuit (Table 1). Thus, medications targeted to these nAChR subtypes are likely to be the most effective for smoking cessation. Varenicline, a partial agonist at α4/β2* nAChRs currently in use for smoking cessation (Rollema et al., 2007), is therefore an example of molecular-based drug design that has helped in smoking cessation

Table 1.

role of nAChR subunits in nicotine-related behaviors.

β2* activation β2* blockade β4* α4* α5* α6* α7
VTA (DA neurons) ↑ CPP, self-admin (Tolu et al., 2013), DA drive (Mansvelder et al., 2002; Wooltorton et al., 2003) and release (Grady et al., 1992; Tolu et al., 2013) ↓ DA release (Grady et al., 2001; Maskos et al., 2005; Picciotto et al., 1998) ↓ CPP and locomotion (Brunzell et al., 2009; Maskos et al., 2005; Mineur et al., 2009a; Walters et al., 2006) ↑ self-admin (Pons et al., 2008), CPP and DA release (McGranahan et al., 2011; Tapper et al., 2004) ↑ self-administration (Pons et al., 2008), CPP and DA release (Drenan et al., 2008)
VTA (DA neurons) ↑ CPP, self-admin (Tolu et al., 2013), DA drive (Mansvelder et al., 2002; Wooltorton et al., 2003) and release (Grady et al., 1992; Tolu et al., 2013) ↓ DA release (Grady et al., 2001; Maskos et al., 2005; Picciotto et al., 1998) ↓ CPP and locomotion (Brunzell et al., 2009; Maskos et al., 2005; Mineur et al., 2009a; Walters et al., 2006)
VTA (GABA neurons) ↑ Self-admin (when â2 also in DA neurons) (Tolu et al., 2013) ↓ GABA release (Lu et al., 1998)
Striatum ↓ nAChR binding in model of depression (Tizabi et al., 2009) ↑self-admin (Brunzell et al., 2010)
PPtg ↑ VTA neuron stimulation (Mansvelder and McGehee, 2000)
Habenula Nicotine aversion (Fowler et al., 2011)
Amygdala ↓ depression-like behavior (Mineur et al., 2007)
Hippocampus ↑ depression-like behavior and anxiety (ACh; Mineur et al, 2013) ↓ anxiety (5HT terminals) (Andreasen et al., 2012; File et al., 2000; Tucci et al., 2003)
OFC ↓ binding in depressed patients (Saricicek et al., 2012)
Dorsal raphe ↑anxiety (w/low level nic) (Tucci et al., 2003)
Arcuate Nucleus ↓ feeding (Mineur et al., 2011)

Molecular and anatomical basis for effects of nicotine on mood and depression

One critical risk factor that greatly decreases the success of smoking cessation is a history of major depression (Glassman et al., 1990). Smokers have twice the rate of depressive disorders than non-smokers (Diwan et al., 1998; Glassman et al., 1990) and common etiological factors, including shared genetic susceptibility (Cinciripini et al., 2004; Lerman et al., 1998), are likely to underlie this connection. It has been suggested that some smokers use tobacco to control their mood symptoms, and that smoking cessation can exacerbate symptoms of depression (Glassman et al., 1990); however, recent epidemiological studies suggest that the use of nicotine may lead to development of depressive symptoms.

In a computational analysis investigating the behavior of young adults between the ages of 18 and 25, the best fitting causal model suggested that depression symptoms are precipitated by regular use of nicotine (Boden et al., 2010). There is also strong evidence for a bidirectional relationship between smoking and depression. In a longitudinal study of adolescents, those with major depression at the first measurement had an increased risk for daily smoking at follow up, while daily smoking in early adolescence increased the risk for major depression at follow up (Breslau et al., 1998). Whereas acute withdrawal increases stress and the likelihood of depression, measures of stress have been reported to be lower following prolonged smoking cessation (Parrott, 1995, 2000), supporting the idea that smoking can actually increase negative affect (Munafo and Araya, 2010). The timing of smoking initiation, duration of smoking behavior and duration of abstinence are critical elements that determine self-rated mood status. Smokers report increased feelings of anxiety and stress between cigarettes (Hughes et al., 1990). Nicotine delivery can alleviate these negative effects of short term withdrawal, resulting in self-report that smoking is perceived as a positive experience (Parrott, 1999). These mood changes are positively correlated with the quantity of tobacco used, even in people with no history of depression (Parrott and Murphy, 2012). Thus, smokers may perceive a positive effect of smoking on mood because nicotine delivery counteracts the negative effects of withdrawal, but over the course of the day, the cyclical mood changes reported by smokers may mask a longer term increase in depression symptoms. But, despite these clinical and epidemiological studies, the direction of causality between nicotine exposure and depression remains an open question (Morrell et al., 2010). Although there is no conclusive evidence that depression drives initiation of smoking, a constellation of studies suggests that, as for most drugs of abuse, smokers continue to smoke to avoid negative affect (Baker et al., 2004a; Baker et al., 2004b)

The possibility that smoking could induce signs of depression is in line with studies of acetylcholine (ACh), the endogenous ligand for nAChRs, which demonstrate that a balance exists between cholinergic and noradrenergic systems, such that hyperactivation of cholinergic signaling, as could be induced by nicotine in tobacco, could lead to, or worsen, depressive symptoms (Janowsky et al., 1972; Oppenheim et al., 1979). Clinical observations demonstrate that administration of physostigmine, a cholinesterase antagonist that increases ACh levels in brain, to patients or volunteer healthy subjects resulted in negative effects on mood (Janowsky et al., 1974). From these observations, it was suggested that cholinergic hypersensitivity could represent a risk factor for the development of depression (Janowsky et al., 1994). Based on this hypothesis, a rat model of depression was developed (the Flinders sensitive line) by selecting for vulnerability to a depressive-like response to a cholinesterase antagonist (Overstreet, 1993). Human imaging studies have also shown that depressed patients have increased central concentrations of choline, the precursor of ACh (Charles et al., 1994), and patients with depression have decreased brain nAChR availability with no change in receptor number, suggestive of increased endogeneous ACh levels (Saricicek et al., 2012).

Initial studies used the muscarinic cholinergic antagonist scopolamine to reverse the effects of physostigmine on depressive symptoms (Janowsky et al., 1983), but rodent experiments demonstrated that nicotinic receptors are critical in the ability of scopolamine to reverse the depression-like effects of physostigmine (Rhodes et al., 2001). These data suggest that regulation of mood by ACh is complex, could be altered by nicotine in tobacco, and involves multiple receptor subtypes. Likely, multiple AChR subtypes and brain circuits are involved, and the net homeostatic balance of activity in these pathways influences mood. For instance, an increase in cholinergic tone in amygdala by stimulation with nicotine from tobacco could be sufficient to induce depression-like symptoms (Mineur and Picciotto, 2010), but a study in mouse has suggested that decreasing cholinergic tone in striatum can also lead to depression-like symptoms (Warner-Schmidt et al., 2012). These results highlight the likelihood that nicotine in tobacco smoke induces heterogeneous effects in different brain areas, with potentially opposing behavioral consequences. In addition, nAChRs are largely neuromodulatory, so effects of these receptors on ACh signaling in the brain depend on the baseline activity of individual circuits.

The integration of nicotinic signaling across brain regions with different levels of activity under conditions of stress could also explain the apparent discrepancy between the idea that smoking induces depressive symptoms and that smoking ameliorates depressive symptoms. Depressed smokers are thought to use cigarettes as a method of nicotine delivery in an effort to self-medicate their mood symptoms. Smokers report a decrease in symptoms immediately after using tobacco products, however, these positive effects are only observed after abstinence, when withdrawal-induced depression symptoms are high (Perkins et al., 2010).

At the molecular level, nicotine enters the brain rapidly and acts as an agonist at nAChRs, but it subsequently desensitizes most subtypes, particularly those containing the β2 subunit (Picciotto et al., 2008). Significant nAChR desensitization would mimic the effects of a nicotinic antagonist or low-efficacy partial agonist, and may therefore underlie at least some of the antidepressant-like effects of nicotinic drugs. The potent antidepressant-like effects observed with nicotinic agonists (Ferguson et al., 2000; Gatto et al., 2004), partial agonists (Mineur et al., 2009b; Mineur et al., 2007; Rollema et al., 2009), and antagonists (Ferguson et al., 2000; Shytle et al., 2002) is puzzling unless one recognizes that all these agents can decrease ACh activity by desensitizing, limiting activity of, or blocking nAChRs, respectively. This would also be consistent with the state-dependent effects of smoking on mood. Chronic nicotine delivery, by smoking or nicotine patches, leads to long term desensitization of nAChRs (Reitstetter et al., 1999), and functional antagonism of most nAChRs over time (Quick and Lester, 2002). However, chronic smoking also leads to long-term upregulation of high affinity nAChRs containing the β2 subunit (Staley et al., 2006), potentially increasing nAChR activity once nicotine-induced desensitization is relieved following abstinence. This increase in ACh-mediated signaling could contribute to depressive symptoms associated with smoking withdrawal.

While a broad nicotinic antagonist has antidepressant-like effects in animal models (Caldarone et al., 2004; Rabenstein et al., 2006), it is likely that multiple nAChR subtypes have effects on depression-like behaviors. For example, mice lacking either the β2 or the α7 subunit do not exhibit the antidepressant-like effects of mecamylamine (Rabenstein et al., 2006). Both α7- (Pichat et al., 2007) and β2- (Mineur et al., 2009b; Mineur et al., 2007; Rollema et al., 2009) selective partial agonists have antidepressant-like properties in rodents, suggesting that decreasing ACh signaling through either subtype is effective; however, in rats, the ability of nicotine to act like an antidepressant can be antagonized by mecamylamine (Tizabi et al., 2000) and nicotine agonism can potentiate the effects of the antidepressant citalopram (Andreasen and Redrobe, 2009), supporting the idea that nAChRs can have complex effects on circuits involved in depression-like behavior.

The co-morbidity between depression and smoking has also influenced antidepressant drug development. Despite the failure of a recent clinical trial, nAChR antagonists or partial agonists can show antidepressant effects in combination with ineffective regimens of selective serotonin reuptake inhibitors in human depressed patients (George et al., 2008; Philip et al., 2009). Both α7 and β2 nAChRs are located throughout the brain, so the neuroanatomical basis for the antidepressant effects of nicotinic drugs is not yet known. Recent studies in subjects with a history of depression indicated that nicotine restored functional asymmetry in frontal cortical areas following tryptophan depletion, a model of induced depressive symptoms (Knott et al., 2012). Nicotine acting through presynaptic α4β2 nAChRs (Garduno et al., 2012) also increases the firing rate of serotonergic neurons projecting to the nucleus accumbens, which is suspected to improve mood and affect (Chang et al., 2011). Doses of nicotinic antagonists and partial agonists that have antidepressant properties also decrease neuronal activity (as measured by cfos activity) in the amygdala (Mineur et al., 2007), a brain region that is often hyperactive in depressed patients (Drevets et al., 2008). Finally, increased ACh activity decreases resilience to stress and results in depressant-like effects (Mineur et al., 2013). Taken together, these studies identify a potential molecular basis for the co-morbidity between depression and smoking, and provides potential avenues for medication development to help depressed smokers quit smoking.

Molecular and anatomical basis for effects of nicotine on appetite

A frequent reason cited by smokers for ongoing cigarette use is appetite control, and among smokers, female teenagers reported that their motivation to smoke was primarily to control body weight (Voorhees et al., 2002). Following the intake of nicotinic drugs, immediate effects are observed on food intake, appetite, hunger and fullness (Grebenstein et al., 2013), likely due to the activation of melanocortin system (Huang et al., 2011; Mineur et al., 2011). These effects seem to be maintained over time, suggesting that long-term smoking does not lead to tolerance to the appetite suppressing effects of nicotine. This is in line with data demonstrating that, in the aggregate, smokers are leaner than non-smokers, and why nicotine decreases weight gain over long periods of time in animals models (Mineur et al., 2011; Grebenstein et al., 2013). Conversely, smoking cessation often leads to weight gain. Nicotine administration limits weight gain and food intake in former smokers and in laboratory rodents (Bellinger et al., 2005; Blaha et al., 1998; Filozof et al., 2004; Hughes and Hatsukami, 1997; Perkins, 1993). As in human ex-smokers, rodents gain weight after cessation of a chronic nicotine administration regimen (Bishop et al., 2004; Filozof et al., 2004; Fornari et al., 2007). Nicotine contained in tobacco products can act at many different levels throughout the body and brain, therefore it is likely to include alteration of metabolism, caloric intake and feeding behaviors (Zoli and Picciotto, 2012). Other factors including oral stimulation (smoking instead of snacking, for instance) likely also contribute to the effects of cigarettes on weight and feeding, but we will focus on the physiological and neuronal basis of nicotine’s impact on food intake.

Calorie intake is only one part of the complex homeostatic regulation of the balance between energy expenditure and food intake, and the nicotinic system can act on each of these levels. Humans and animals tightly regulate nicotine intake (Gritz et al., 1976; Vieyra-Reyes et al., 2008), and this likely allows for careful titration of the effects of nicotine on different nAChR subtypes and circuits. The arcuate nucleus (ARC), located at the bottom of the ventral hypothalamus, is the main brain structure that controls energy expenditure, food intake and feeding patterns (Sainsbury and Zhang, 2010). Neuropeptide Y (NPY) and proopiomelanocortin (POMC) expressing neurons are the two main neuronal populations in the ARC and activation of these neuronal pathways stimulate or inhibit feeding, respectively. Equilibrium binding experiments have identified multiple nAChR subclasses in ARC, including α7 and β2* nAChRs, the most widely expressed subtypes (Han et al., 2000; Han et al., 2003). In situ and rtPCR experiments have indicated that other, less widely expressed, subunits are also expressed in these neurons (Gotti et al., 2006; Han et al., 2000; Mineur et al., 2011), likely in discrete cell types, providing a potential molecular target for nicotine in the ARC in regulation of food intake.

Rat studies have shown that nicotine administration can decrease food intake and body weight, with greater effects in female animals (Grunberg et al., 1987). A similar nicotine regimen also decreases body weight and fat mass in mice as a result of β4* nAChR-mediated activation of POMC neurons and subsequent activation of MC4 receptors on second order neurons in the paraventricular nucleus of the hypothalamus (Mineur et al., 2011). NPY neurons are also activated by nicotine acutely, however, consistent with the continued ability of nicotine to decrease food intake after chronic administration, repeated stimulation by nicotine potentiates signaling in POMC, but not NPY, neurons (Huang et al., 2011). The fact that nicotine-induced firing of NPY neurons was not as sustained as the effects on POMC neurons may also explain why nicotinic-induced cfos activation was observed primarily in POMC neurons (Mineur et al., 2011). More indirect effects of nicotine are also likely to modulate feeding by altering ARC neuron activity. For instance, activation of nAChRs located on orexin-positive neurons in the lateral hypothalamus can trigger the release of glutamate and ACh, two neurotransmitters that could then induce a feed-forward stimulation of the LH-ARC pathway once primed by nicotine (Pasumarthi and Fadel, 2010). These data suggest that nicotine effects on hypothalamic nAChRs may decrease food intake in a receptor subtype- and cell type-specific manner and raises the possibility that medications targeting specific nAChR subtypes could help moderate weight gain following smoking cessation.

Effects of nicotine on neuronal development

The contribution of parental smoking to initiation of tobacco use in children is likely to involve greater access to cigarettes, effects of socio-economic status and other social factors. In animal models, nicotine exposure during early development and adolescence also promotes later nicotine reinforcement. For example, nicotine administration to pregnant dams or to adolescent rodents increases later nicotine preference and potentiates the acquisition of nicotine self-administration (Adriani et al., 2003; Klein et al., 2003), suggesting that biological effects of nicotine during development in children exposed to tobacco smoke also contribute to susceptibility to tobacco addiction. It is clear that nAChRs contribute to normal neuronal development and that nicotine can perturb the maturation of neurons (Role and Berg, 1996), likely leading to these long-term changes in behavior. While the specific molecular and cellular basis of the effects of developmental nicotine exposure on susceptibility to nicotine reinforcement is not yet established, it is clear that nAChR stimulation affects the maturation of a number of neuronal subtypes, and that nicotine can alter neuronal function at several stages of development. Thus, parental smoking can increase risk of smoking by their children, and the effects of nicotine in tobacco on neuronal development is likely to be one important factor underlying that risk.

At early stages of neuronal development, nAChRs are critical for the timing of maturation of GABA signaling in the CNS (Liu et al., 2006). Whereas GABA signaling is inhibitory in adult neurons, GABA depolarizes and stimulates firing of immature neurons. The switch from GABA-mediated excitation to inhibition results from the expression of an adult isoform of the chloride transporter which is induced by β2* nAChR signaling (Liu et al., 2006). This finding suggests that fetal nicotine exposure could accelerate the transition to inhibitory GABA signaling, potentially altering the timing of neuronal maturation events dependent on GABA-induced depolarization. With respect to addiction, alterations in the balance of GABA signaling in brain areas involved in nicotine reinforcement, such as the VTA, could result in greater susceptibility to nicotine reward later in life.

At the synaptic level, recent studies have identified nAChR-dependent alterations in development of glutamatergic synapses (Lozada et al., 2012a, b). Activation of postynaptic α7 nAChRs in cultured neurons or organotypic slices promotes glutamatergic synapse formation, whereas knockout of the α7 subunit in mice decreases the number of dendritic spines (Lozada et al., 2012b) and glutamatergic synapses, although the number of GABAergic synapses was unchanged, suggesting that cholinergic signaling through this nAChR subtype is normally important in modulating the number of excitatory synapses and can modulate the ratio of excitatory to inhibitory inputs (Lozada et al., 2012a). As above, alterations in this ratio could alter susceptibility to the rewarding and reinforcing effects of nicotine later in life.

The effects of nicotine on neuronal development can also be seen at the circuit level. One well-studied site for developmental actions of nicotine is the cortico-thalamo-cortical connections that are critical for the gating of sensory information (Heath and Picciotto, 2009). Prospective studies of children exposed to tobacco smoke in utero through maternal smoking have identified early changes in auditory processing (Fried and Watkinson, 1988) and long lasting changes in performance of auditory and other sensory attention tasks (Fried et al., 2003; Jacobsen et al., 2007b). Some smokers report that they use tobacco to help focus their attention; therefore the ability of developmental nicotine exposure to perturb these circuits could contribute to later smoking to attempt to self-medicate attentional deficits.

The circuits important for the developmental effects of tobacco exposure on sensory attention have been investigated with in vivo imaging techniques in human subjects (Jacobsen et al., 2007a) and in animal models (Aramakis et al., 2000; Heath et al., 2010b; King et al., 2003; Liang et al., 2006; Liang et al., 2008), and appear to involve altered signaling in the connections between cortical and thalamic neurons. A number of molecular and electrophysiological studies have implicated α7 nAChR signaling in thalamocortical neurons (Aramakis and Metherate, 1998; Aramakis et al., 2000) and β2* nAChR signaling in layer 6 cortical pyramidal neurons (Heath et al., 2010b; King et al., 2003) as important for development of cortico-thalamo-cortical loops (Heath and Picciotto, 2009). Layer 6 pyramidal neurons projecting to the thalamus express high levels of α5* nAChRs during early postnatal development that are critical for maturation of cortico-thalamic communication (Bailey et al., 2010; Kassam et al., 2008). In mice, pharmacological disruption of nAChR signaling during development by administration of nicotine during the postnatal period, or genetically by knocking out the β2 nAChR subunit results in impaired performance in an auditory discrimination task (Horst et al., 2012) and hypersensitive passive avoidance learning that is dependent on expression of β2* nAChRs in layer 6 cortical neurons that project to the thalamus (Heath et al., 2010b; King et al., 2003; Picciotto et al., 1995). Nicotine administered through the dam does not result in changes in maternal behavior, suggesting that altered neurodevelopment is likely to explain the behavioral consequences of early nicotine exposure (Heath et al., 2010a). Thus, specific nAChR subtypes and a potential neuronal circuit have been implicated in the effects of developmental tobacco and nicotine exposure on sensory attention.

In addition to effects on sensory processing, early tobacco exposure also alters later likelihood of smoking in humans (Ernst et al., 2001), whereas developmental nicotine exposure increases later nicotine self-administration in rodents (Adriani et al., 2006). While a number of molecular consequences of developmental nicotine exposure have been identified (Heath and Picciotto, 2009), there are not studies identifying a causal link between these molecular changes and altered nicotine reinforcement. One hint may come from a recent study showing that nicotine can result in epigenetic alterations that lead to increased drug reinforcement (Levine et al., 2011). This study identified a nicotine-dependent decrease in histone deacetylase activity and a concomitant increase in ΔfosB expression in the striatum that potentiated the effects of another drug of abuse (cocaine) on plasticity in medium spiny neurons. This was associated with an increase in drug place preference. Similar molecular mechanisms acting as a result of nicotine exposure during critical periods of development could potentially be involved in alterations in nicotine and tobacco preference.

Taken together, the human epidemiological data and the rodent behavioral data suggest that early exposure to nicotine in tobacco smoke can increase the likelihood of later nicotine addiction. The molecular and cellular events mediated through nAChR signaling alter neural development and result in plasticity in a number of brain areas that could contribute to increased susceptibility to smoking in children exposed to tobacco during development.

Integration of systems level effects of nicotine: common and disparate molecular mechanisms

It is now very clear that the nicotine in tobacco interacts with nAChRs in the brain, and alters their function. It is therefore important to understand the role of these receptors in the brain if we are to understand how nicotine supports tobacco addiction. We have briefly reviewed the effects of nicotine on reward circuits, appetite, stress-related behaviors and neuronal development. While the subtypes of nAChR and the cell type on which they exert their effects differ across the behaviors and developmental time points discussed here, a number of fundamental mechanisms have been identified that likely operate at many sites throughout the brain to mediate the effects of nicotine. For example, one common mechanism that contributes to all the behavioral effects of nicotine is the the ability of nicotine to potentiate neurotransmitter release from presynaptic sites as a key electrophysiological function of nAChRs (McGehee and Role, 1996; Wonnacott, 1997). The ability of nAChRs to increase DA, ACh, GABA and glutamate release (Dickinson et al., 2008; Grady et al., 1992; Grady et al., 2001; Lu et al., 1998; Radcliffe and Dani, 1998) is likely to underlie at least some of its effects on behavior. Another common mechanism is the ability of nicotine to desensitize, as well as activate, its receptors. The differential desensitization of different presynaptic nAChR subtypes is likely to shape the postsynaptic consequences of nicotine application in many brain areas, as has been shown in both the VTA, where desensitization results in maintained glutamate release but less GABA release following extended nicotine exposure (Mansvelder et al., 2002), and the striatum, where desensitization decreases tonic DA release while leaving phasic DA signaling intact (Rice and Cragg, 2004; Zhang and Sulzer, 2004). These mechanisms have been studied in specific brain areas, but are likely to be common across many brain areas, and to mediate the effects of nicotine in other circuits in which nAChRs are expressed, such as those involved in appetite (Huang et al., 2011) or affective behaviors.

In addition to the broad themes of presynaptic action on neurotransmitter release and the relevance of desensitization to nicotine’s actions, it is important to emphasize that activation of nAChRs coordinates the firing of ensembles of neurons to modulate the output of circuits involved in reinforcement, appetite and stress response. Nicotine depolarizes and increases the firing of neurons in the VTA, thalamus, cortex and hippocampus (Grenhoff et al., 1986; Gu and Yakel, 2011; Kassam et al., 2008; Picciotto et al., 1995; Picciotto et al., 1998). Nicotine can coordinate the firing of neuronal ensembles by decreasing the threshold for action potential firing (Kawai et al., 2007). This ability of nicotine to coordinate the firing of neurons in brain areas relevant for behaviors related to reinforcement, attention and stress involves a combination of pre- and post-synaptic effects of nAChRs shape the output of brain circuits that mediate nicotine-dependent behaviors. In the mesolimbic system, the multiple levels of nicotine effects on nAChR activity result in increased burst firing of DA neurons (Grenhoff et al., 1986), potentiation of excitatory input onto VTA DA neurons (Mansvelder and McGehee, 2000) but decreased GABA input (Mansvelder et al., 2002; Wooltorton et al., 2003), increased DA release from NAc DA terminals (Grady et al., 1992), and decreased salience of tonic DA release (Exley et al., 2008; Grady et al., 2012), all of which combine to result in long-term increases in nicotine-induced DA signaliing (Balfour et al., 1998), likely strengthening association between nicotine administration and salient environmental cues that drive ongoing nicotine intake (Palmatier et al., 2007). Thus, actions of nicotine at each of these points in the mesolimbic circuit come together to mediate the ability of nicotine in tobacco to drive ongoing smoking.

While extensive research at the molecular, electrophysiological and behavioral levels has begun to provide a coherent picture of how effects of nicotine at multiple points in the mesolimbic system coordinates DA dynamics, there is still more work to do to understand how the nicotine in tobacco promotes behaviors that contribute to ongoing smoking. For example, ACh in hippocampus may be important for the association between smoking and depression (Mineur et al, 2013) and the effects of nicotine on corticothalamic loops may explain why smokers report that smoking focuses their attention (Heath and Picciotto, 2009), whereas the ability of nicotine to stimulate neurons in the ventral hypothalamus may contribute to the perception that smoking can be used to regulate appetite (Mineur et al, 2011). Nicotine-induced changes in hippocampal GABA and glutamate neuron activity and plasticity are mediated through a number of different nAChR subtypes acting at different cell types and subcellular locations in the hippocampus (Fujii and Sumikawa, 2001; Ge and Dani, 2005; Ji et al., 2001; Jones and Yakel, 1997; Leao et al., 2012; Nakauchi et al., 2007; Radcliffe and Dani, 1998; Shao and Yakel, 2000). Similarly, α7 and β2 nAChRs can modulate cortico-thalamic signaling at multiple different points in the circuit (Heath and Picciotto, 2009; Kawai et al., 2007; Metherate and Hsieh, 2004). Finally, postsynaptic effects of nicotine in the ventral hypothalamus (Mineur et al., 2011), and differential desensitization of nAChRs on NPY- and POMC-expressing neurons (Huang et al., 2011) contribute to effects of nicotine on this brain center critical for food intake. In each of these brain areas, some of the mechanisms that have been identified in the mesolimbic system have not yet been investigated. As a result, studies of the effects of nicotine on circuits involved in addiction could be extended to these brain areas, particularly studies of how differential desensitization of nAChR subtypes contributes to the ability of nicotine to alter the output of hippocampal, corticothalamic or hypothalamic circuits. Overall, while there have been significant advances in the understanding of nicotine’s effects on circuits related to nicotine reinforcement, a more complete understanding of the mechanisms underlying effects of nicotine on circuits involved in effects of nicotine on other behaviors that also drive tobacco smoking will allow more targeted interventions for smoking cessation.

Conclusions

Over the last several decades there has been a great deal of progress understanding the molecular and cellular basis of behaviors related to nicotine addiction (Changeux, 2010), and this understanding has led to targeted drug discovery leading to new therapeutics for smoking cessation such as varenicline (Coe et al., 2005). These advances show that fundamental studies of the neurobiological basis of drug abuse can increase our knowledge of why individuals become addicted and what drives ongoing smoking, but can also lead to novel methods of intervention to help people quit and stay abstinent. The knowledge that has been gained about the mechanisms underlying nicotine reinforcement has been applied to understanding other behaviors that drive ongoing smoking. Targeting the multimodal basis for nicotine intake may therefore result in more effective treatments for smoking cessation going forward.

Highlights.

  • Many behaviors contribute to tobacco smoking.

  • Nicotine is addictive, but can affect mood and appetite as well.

  • Developmental effects of nicotine can promote later smoking.

  • We review advances on molecular and cellular basis of nicotine-related behaviors.

  • Molecular knowledge could help design novel smoking cessation therapies.

Acknowledgments

This work was supported by grants DA14241, DA033945 and MH077681 from the National Institutes of Health.

Footnotes

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

Literature Cited

  1. Adriani W, Deroche-Gamonet V, Le Moal M, Laviola G, Piazza PV. Preexposure during or following adolescence differently affects nicotine-rewarding properties in adult rats. Psychopharmacology. 2006;184:382–390. doi: 10.1007/s00213-005-0125-1. [DOI] [PubMed] [Google Scholar]
  2. Adriani W, Spijker S, Deroche-Gamonet V, Laviola G, Le Moal M, Smit AB, Piazza PV. Evidence for enhanced neurobehavioral vulnerability to nicotine during periadolescence in rats. Journal of Neuroscience. 2003;23:4712–4716. doi: 10.1523/JNEUROSCI.23-11-04712.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Andreasen JT, Redrobe JP. Nicotine, but not mecamylamine, enhances antidepressant-like effects of citalopram and reboxetine in the mouse forced swim and tail suspension tests. Behav Brain Res. 2009;197:150–156. doi: 10.1016/j.bbr.2008.08.016. [DOI] [PubMed] [Google Scholar]
  4. Andreasen JT, Redrobe JP, Nielsen EO. Combined alpha7 nicotinic acetylcholine receptor agonism and partial serotonin transporter inhibition produce antidepressant-like effects in the mouse forced swim and tail suspension tests: a comparison of SSR180711 and PNU-282987. Pharmacol Biochem Behav. 2012;100:624–629. doi: 10.1016/j.pbb.2011.11.004. [DOI] [PubMed] [Google Scholar]
  5. Aramakis V, Metherate R. Nicotine selectively enhances NMDA receptor-mediated synaptic transmission during postnatal development in sensory neocortex. J Neurosci. 1998;18:8485–8495. doi: 10.1523/JNEUROSCI.18-20-08485.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Aramakis VB, Hsieh CY, Leslie FM, Metherate R. A critical period for nicotine-induced disruption of synaptic development in rat auditory cortex. J Neurosci. 2000;20:6106–6116. doi: 10.1523/JNEUROSCI.20-16-06106.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bailey CDC, De Biasi M, Fletcher PJ, Lambe EK. The nicotinic acetylcholine receptor alpha5 subunit plays a key role in attention circuitry and accuracy. J Neurosci. 2010;30:9241–9252. doi: 10.1523/JNEUROSCI.2258-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Baker TB, Brandon TH, Chassin L. Motivational influences on cigarette smoking. Annu Rev Psychol. 2004a;55:463–491. doi: 10.1146/annurev.psych.55.090902.142054. [DOI] [PubMed] [Google Scholar]
  9. Baker TB, Piper ME, McCarthy DE, Majeskie MR, Fiore MC. Addiction motivation reformulated: an affective processing model of negative reinforcement. Psychol Rev. 2004b;111:33–51. doi: 10.1037/0033-295X.111.1.33. [DOI] [PubMed] [Google Scholar]
  10. Balfour DJK, Benwell MEM, Birrell CE, Kelly RJ, Alaloul M. Sensitization of the mesoaccumbens dopamine response to nicotine. Pharmacol Biochem Behav. 1998;59:1021–1030. doi: 10.1016/s0091-3057(97)00537-6. [DOI] [PubMed] [Google Scholar]
  11. Bellinger LL, Wellman PJ, Cepeda-Benito A, Kramer PR, Guan G, Tillberg CM, Gillaspie PR, Hill EG. Meal patterns in female rats during and after intermittent nicotine administration. Pharmacol Biochem Behav. 2005;80:437–444. doi: 10.1016/j.pbb.2004.12.010. [DOI] [PubMed] [Google Scholar]
  12. Benwell ME, Balfour DJ. The effects of acute and repeated nicotine treatment on nucleus accumbens dopamine and locomotor activity. Br J Pharmacol. 1992;105:849–856. doi: 10.1111/j.1476-5381.1992.tb09067.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bishop C, Parker GC, Coscina DV. Systemic nicotine alters whole-body fat utilization in female rats. Physiol Behav. 2004;80:563–567. doi: 10.1016/j.physbeh.2003.10.011. [DOI] [PubMed] [Google Scholar]
  14. Blaha V, Yang ZJ, Meguid M, Chai JK, Zadak Z. Systemic nicotine administration suppresses food intake via reduced meal sizes in both male and female rats. Acta Medica (Hradec Kralove) 1998;41:167–173. [PubMed] [Google Scholar]
  15. Breslau N, Peterson EL, Schultz LR, Chilcoat HD, Andreski P. Major depression and stages of smoking. A longitudinal investigation. Arch Gen Psychiatry. 1998;55:161–166. doi: 10.1001/archpsyc.55.2.161. [DOI] [PubMed] [Google Scholar]
  16. Brunzell DH, Boschen KE, Hendrick ES, Beardsley PM, McIntosh JM. Alpha-conotoxin MII-sensitive nicotinic acetylcholine receptors in the nucleus accumbens shell regulate progressive ratio responding maintained by nicotine. Neuropsychopharmacology. 2010;35:665–673. doi: 10.1038/npp.2009.171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Brunzell DH, Chang JR, Schneider B, Olausson P, Taylor JR, Picciotto MR. beta2-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. 2006;184:328–338. doi: 10.1007/s00213-005-0099-z. [DOI] [PubMed] [Google Scholar]
  18. Brunzell DH, Mineur YS, Neve RL, Picciotto MR. Nucleus accumbens CREB activity is necessary for nicotine conditioned place preference. Neuropsychopharmacology. 2009;34:1993–2001. doi: 10.1038/npp.2009.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Caldarone BJ, Harrist A, Cleary MA, Beech RD, King SL, Picciotto MR. High-affinity nicotinic acetylcholine receptors are required for antidepressant effects of amitriptyline on behavior and hippocampal cell proliferation. Biol Psychiatry. 2004;56:657–664. doi: 10.1016/j.biopsych.2004.08.010. [DOI] [PubMed] [Google Scholar]
  20. Chang B, Daniele CA, Gallagher K, Madonia M, Mitchum RD, Barrett L, Vezina P, McGehee DS. Nicotinic excitation of serotonergic projections from dorsal raphe to the nucleus accumbens. Journal of neurophysiology. 2011;106:801–808. doi: 10.1152/jn.00575.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Changeux J-P. Nicotine addiction and nicotinic receptors: lessons from genetically modified mice. Nature Revi Neurosci. 2010;11:389–401. doi: 10.1038/nrn2849. [DOI] [PubMed] [Google Scholar]
  22. Charles HC, Lazeyras F, Krishnan KR, Boyko OB, Payne M, Moore D. Brain choline in depression: in vivo detection of potential pharmacodynamic effects of antidepressant therapy using hydrogen localized spectroscopy. Prog Neuro-Psychopharm Biol Psych. 1994;18:1121–1127. doi: 10.1016/0278-5846(94)90115-5. [DOI] [PubMed] [Google Scholar]
  23. Cinciripini P, Wetter D, Tomlinson G, Tsoh J, De Moor C, Cinciripini L, Minna J. The effects of the DRD2 polymorphism on smoking cessation and negative affect: evidence for a pharmacogenetic effect on mood. Nic Tobacco Res. 2004;6:229–239. doi: 10.1080/14622200410001676396. [DOI] [PubMed] [Google Scholar]
  24. Coe JW, Brooks PR, Vetelino MG, Wirtz MC, Arnold EP, Huang J, Sands SB, Davis TI, Lebel LA, Fox CB, Shrikhande A, Heym JH, Schaeffer E, Rollema H, Lu Y, Mansbach RS, Chambers LK, Rovetti CC, Schulz DW, Tingley FD, 3rd, O’Neill BT. Varenicline: an alpha4beta2 nicotinic receptor partial agonist for smoking cessation. J Med Chem. 2005;48:3474–3477. doi: 10.1021/jm050069n. [DOI] [PubMed] [Google Scholar]
  25. Dickinson JA, Kew JNC, Wonnacott S. Presynaptic alpha 7- and beta 2-containing nicotinic acetylcholine receptors modulate excitatory amino acid release from rat prefrontal cortex nerve terminals via distinct cellular mechanisms. Mol Pharmacol. 2008;74:348–359. doi: 10.1124/mol.108.046623. [DOI] [PubMed] [Google Scholar]
  26. Diwan A, Castine M, Pomerleau CS, Meador-Woodruff JH, Dalack GW. Differential prevalence of cigarette smoking in patients with schizophrenic vs mood disorders. Schizophrenia Research. 1998;33:113–118. doi: 10.1016/s0920-9964(98)00045-0. [DOI] [PubMed] [Google Scholar]
  27. Drenan RM, Grady SR, Whiteaker P, McClure-Begley T, McKinney S, Miwa JM, Bupp S, Heintz N, McIntosh JM, Bencherif M, Marks MJ, Lester HA. In vivo activation of midbrain dopamine neurons via sensitized, high-affinity alpha 6 nicotinic acetylcholine receptors. Neuron. 2008;60:123–136. doi: 10.1016/j.neuron.2008.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Drevets WC, Price JL, Furey ML. Brain structural and functional abnormalities in mood disorders: implications for neurocircuitry models of depression. Brain Struct Funct. 2008;213:93–118. doi: 10.1007/s00429-008-0189-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Ernst M, Moolchan ET, Robinson ML. Behavioral and neural consequences of prenatal exposure to nicotine. J Am Acad Child Adol Psychiatry. 2001;40:630–641. doi: 10.1097/00004583-200106000-00007. [DOI] [PubMed] [Google Scholar]
  30. Exley R, Clements MA, Hartung H, McIntosh JM, Cragg SJ. Alpha6-containing nicotinic acetylcholine receptors dominate the nicotine control of dopamine neurotransmission in nucleus accumbens. Neuropsychopharmacology. 2008;33:2158–2166. doi: 10.1038/sj.npp.1301617. [DOI] [PubMed] [Google Scholar]
  31. Ferguson SM, Brodkin JD, Lloyd GK, Menzaghi F. Antidepressant-like effects of the subtype-selective nicotinic acetylcholine receptor agonist, SIB-1508Y, in the learned helplessness rat model of depression. Psychopharmacology. 2000;152:295–303. doi: 10.1007/s002130000531. [DOI] [PubMed] [Google Scholar]
  32. File SE, Kenny PJ, Cheeta S. The role of the dorsal hippocampal serotonergic and cholinergic systems in the modulation of anxiety. Pharmacol Biochem Behav. 2000;66:65–72. doi: 10.1016/s0091-3057(00)00198-2. [DOI] [PubMed] [Google Scholar]
  33. Filozof C, Fernandez Pinilla MC, Fernandez-Cruz A. Smoking cessation and weight gain. Obes Rev. 2004;5:95–103. doi: 10.1111/j.1467-789X.2004.00131.x. [DOI] [PubMed] [Google Scholar]
  34. Fornari A, Pedrazzi P, Lippi G, Picciotto MR, Zoli M, Zini I. Nicotine withdrawal increases body weight, neuropeptide Y and Agouti-related protein expression in the hypothalamus and decreases uncoupling protein-3 expression in the brown adipose tissue in high-fat fed mice. Neurosci Lett. 2007;411:72–76. doi: 10.1016/j.neulet.2006.10.014. [DOI] [PubMed] [Google Scholar]
  35. Fowler CD, Lu Q, Johnson PM, Marks MJ, Kenny PJ. Habenular alpha5 nicotinic receptor subunit signalling controls nicotine intake. Nature. 2011;471:597–601. doi: 10.1038/nature09797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Fried PA, Watkinson B. 12- and 24-month neurobehavioural follow-up of children prenatally exposed to marihuana, cigarettes and alcohol. Neurotoxicology & Teratology. 1988;10:305–313. doi: 10.1016/0892-0362(88)90032-3. [DOI] [PubMed] [Google Scholar]
  37. Fried PA, Watkinson B, Gray R. Differential effects on cognitive functioning in 13- to 16-year-olds prenatally exposed to cigarettes and marihuana. Neurotoxicology & Teratology. 2003;25:427–436. doi: 10.1016/s0892-0362(03)00029-1. [DOI] [PubMed] [Google Scholar]
  38. Fujii S, Sumikawa K. Nicotine accelerates reversal of long-term potentiation and enhances long-term depression in the rat hippocampal CA1 region. Brain Res. 2001;894:340–346. doi: 10.1016/s0006-8993(01)02058-3. [DOI] [PubMed] [Google Scholar]
  39. Fulkerson JA, French SA. Cigarette smoking for weight loss or control among adolescents: gender and racial/ethnic differences. J Adolesc Health. 2003;32:306–313. doi: 10.1016/s1054-139x(02)00566-9. [DOI] [PubMed] [Google Scholar]
  40. Garduno J, Galindo-Charles L, Jimenez-Rodriguez J, Galarraga E, Tapia D, Mihailescu S, Hernandez-Lopez S. Presynaptic alpha4beta2 nicotinic acetylcholine receptors increase glutamate release and serotonin neuron excitability in the dorsal raphe nucleus. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2012;32:15148–15157. doi: 10.1523/JNEUROSCI.0941-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Gatto GJ, Bohme GA, Caldwell WS, Letchworth SR, Traina VM, Obinu MC, Laville M, Reibaud M, Pradier L, Dunbar G, Bencherif M. TC-1734: an orally active neuronal nicotinic acetylcholine receptor modulator with antidepressant, neuroprotective and long-lasting cognitive effects. CNS Drug Reviews. 2004;10:147–166. doi: 10.1111/j.1527-3458.2004.tb00010.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Ge S, Dani J. Nicotinic acetylcholine receptors at glutamate synapses facilitate long-term depression or potentiation. J Neurosci. 2005;25:6084–6091. doi: 10.1523/JNEUROSCI.0542-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. George TP, Sacco KA, Vessicchio JC, Weinberger AH, Shytle RD. Nicotinic antagonist augmentation of selective serotonin reuptake inhibitor-refractory major depressive disorder: A preliminary study. J Clin Psychopharm. 2008;28:340–344. doi: 10.1097/JCP.0b013e318172b49e. [DOI] [PubMed] [Google Scholar]
  44. Glassman AH, Helzer JE, Covey LS, Cottler LB, Stetner F, Tipp JE, Johnson J. Smoking, smoking cessation, and major depression. Jama. 1990;264:1546–1549. [PubMed] [Google Scholar]
  45. Gotti C, Guiducci S, Tedesco V, Corbioli S, Zanetti L, Moretti M, Zanardi A, Rimondini R, Mugnaini M, Clementi F, Chiamulera C, Zoli M. Nicotinic acetylcholine receptors in the mesolimbic pathway: primary role of ventral tegmental area alpha6beta2* receptors in mediating systemic nicotine effects on dopamine release, locomotion, and reinforcement. J Neurosci. 2010;30:5311–5325. doi: 10.1523/JNEUROSCI.5095-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Gotti C, Zoli M, Clementi F. Brain nicotinic acetylcholine receptors: native subtypes and their relevance. Trends in pharmacological sciences. 2006;27:482–491. doi: 10.1016/j.tips.2006.07.004. [DOI] [PubMed] [Google Scholar]
  47. Grady S, Marks MJ, Wonnacott S, Collins AC. Characterization of nicotinic receptor-mediated [3H]dopamine release from synaptosomes prepared from mouse striatum. J Neurochem. 1992;59:848–856. doi: 10.1111/j.1471-4159.1992.tb08322.x. [DOI] [PubMed] [Google Scholar]
  48. Grady SR, Marks MJ, Collins AC. Desensitization of nicotine-stimulated [3H]dopamine release from mouse striatal synaptosomes. J Neurochem. 1994;62:1390–1398. doi: 10.1046/j.1471-4159.1994.62041390.x. [DOI] [PubMed] [Google Scholar]
  49. Grady SR, Meinerz NM, Cao J, Reynolds A, Picciotto MR, Changeux JP, McIntosh MJ, Marks MJ, Collins AC. Nicotinic agonists stimulate acetylcholine release from mouse interpeduncular nucleus: a function mediated by a different nAChR than dopamine release from striatum. J Neurochem. 2001;76:258–268. doi: 10.1046/j.1471-4159.2001.00019.x. [DOI] [PubMed] [Google Scholar]
  50. Grady SR, Wageman CR, Patzlaff NE, Marks MJ. Low concentrations of nicotine differentially desensitize nicotinic acetylcholine receptors that include alpha5 or alpha6 subunits and that mediate synaptosomal neurotransmitter release. Neuropharmacology. 2012;62:1935–1943. doi: 10.1016/j.neuropharm.2011.12.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Grebenstein PE, Thompson IE, Rowland NE. The effects of extended intravenous nicotine administration on body weight and meal patterns in male Sprague-Dawley Rats. Psychopharmacology (Berl) 2013 doi: 10.1007/s00213-013-3043-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Grenhoff J, Aston-Jones G, Svensson TH. Nicotinic effects on the firing pattern of midbrain dopamine neurons. Acta Physiol Scand. 1986;128:351–358. doi: 10.1111/j.1748-1716.1986.tb07988.x. [DOI] [PubMed] [Google Scholar]
  53. Gritz ER, Baer-Weiss V, Jarvik ME. Titration of nicotine intake with full-length and half-length cigarettes. Clinical pharmacology and therapeutics. 1976;20:552–556. doi: 10.1002/cpt1976205552. [DOI] [PubMed] [Google Scholar]
  54. Grunberg NE, Winders SE, Popp KA. Sex differences in nicotine’s effects on consummatory behavior and body weight in rats. Psychopharmacology. 1987;91:221–225. doi: 10.1007/BF00217067. [DOI] [PubMed] [Google Scholar]
  55. Gu Z, Yakel JL. Timing-dependent septal cholinergic induction of dynamic hippocampal synaptic plasticity. Neuron. 2011;71:155–165. doi: 10.1016/j.neuron.2011.04.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Han ZY, Le Novere N, Zoli M, Hill JA, Jr, Champtiaux N, Changeux JP. Localization of nAChR subunit mRNAs in the brain of Macaca mulatta. Eur J Neurosci. 2000;12:3664–3674. doi: 10.1046/j.1460-9568.2000.00262.x. [DOI] [PubMed] [Google Scholar]
  57. Han ZY, Zoli M, Cardona A, Bourgeois JP, Changeux JP, Le Novere N. Localization of [3H]nicotine, [3H]cytisine, [3H]epibatidine, and [125I]alpha-bungarotoxin binding sites in the brain of Macaca mulatta. J Comp Neurol. 2003;461:49–60. doi: 10.1002/cne.10659. [DOI] [PubMed] [Google Scholar]
  58. Heath CJ, Horst NK, Picciotto MR. Oral nicotine consumption does not affect maternal care or early development in mice but results in modest hyperactivity in adolescence. Physiology & Behavior. 2010a;101:764–769. doi: 10.1016/j.physbeh.2010.08.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Heath CJ, King SL, Gotti C, Marks MJ, Picciotto MR. Cortico-thalamic connectivity is vulnerable to nicotine exposure during early postnatal development through alpha4/beta2/alpha5 nicotinic acetylcholine receptors. Neuropsychopharmacology. 2010b;35:2324–2338. doi: 10.1038/npp.2010.130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Heath CJ, Picciotto MR. Nicotine-induced plasticity during development: modulation of the cholinergic system and long-term consequences for circuits involved in attention and sensory processing. Neuropharmacology. 2009;56:254–262. doi: 10.1016/j.neuropharm.2008.07.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Horst NK, Heath CJ, Neugebauer NM, Kimchi EY, Laubach M, Picciotto MR. Impaired auditory discrimination learning following perinatal nicotine exposure or beta2 nicotinic acetylcholine receptor subunit deletion. Behav Brain Res. 2012;231:170–180. doi: 10.1016/j.bbr.2012.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Huang H, Xu Y, van den Pol AN. Nicotine excites hypothalamic arcuate anorexigenic proopiomelanocortin neurons and orexigenic neuropeptide Y neurons: similarities and differences. Journal of neurophysiology. 2011;106:1191–1202. doi: 10.1152/jn.00740.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Hughes JR, Gust SW, Keenan RM, Fenwick JW. Effect of dose on nicotine’s reinforcing, withdrawal-suppression and self-reported effects. J Pharmacol Exp Ther. 1990;252:1175–1183. [PubMed] [Google Scholar]
  64. Hughes JR, Hatsukami DK. Effects of three doses of transdermal nicotine on post-cessation eating, hunger and weight. Journal of substance abuse. 1997;9:151–159. doi: 10.1016/s0899-3289(97)90013-4. [DOI] [PubMed] [Google Scholar]
  65. Jacobsen LK, Picciotto MR, Heath CJ, Frost SJ, Tsou KA, Dwan RA, Jackowski MP, Constable RT, Mencl WE. Prenatal and adolescent exposure to tobacco smoke modulates the development of white matter microstructure. J Neurosci. 2007a;27:13491–13498. doi: 10.1523/JNEUROSCI.2402-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Jacobsen LK, Slotkin TA, Mencl WE, Frost SJ, Pugh KR. Gender-specific effects of prenatal and adolescent exposure to tobacco smoke on auditory and visual attention. Neuropsychopharm. 2007b;32:2453–2464. doi: 10.1038/sj.npp.1301398. [DOI] [PubMed] [Google Scholar]
  67. Janowsky DS, el-Yousef MK, Davis JM. Acetylcholine and depression. Psychosomatic medicine. 1974;36:248–257. doi: 10.1097/00006842-197405000-00008. [DOI] [PubMed] [Google Scholar]
  68. Janowsky DS, el-Yousef MK, Davis JM, Sekerke HJ. A cholinergic-adrenergic hypothesis of mania and depression. Lancet. 1972;2:632–635. doi: 10.1016/s0140-6736(72)93021-8. [DOI] [PubMed] [Google Scholar]
  69. Janowsky DS, Overstreet DH, Nurnberger JI., Jr Is cholinergic sensitivity a genetic marker for the affective disorders? Am J Med Genetics. 1994;54:335–344. doi: 10.1002/ajmg.1320540412. [DOI] [PubMed] [Google Scholar]
  70. Janowsky DS, Risch SC, Gillin JC. Adrenergic-cholinergic balance and the treatment of affective disorders. Progress in Neuro-Psychopharmacology & Biological Psychiatry. 1983;7:297–307. doi: 10.1016/0278-5846(83)90119-7. [DOI] [PubMed] [Google Scholar]
  71. Ji D, Lape R, Dani JA. Timing and location of nicotinic activity enhances or depresses hippocampal synaptic plasticity. Neuron. 2001;31:131–141. doi: 10.1016/s0896-6273(01)00332-4. [DOI] [PubMed] [Google Scholar]
  72. Jones S, Yakel JL. Functional nicotinic ACh receptors on interneurones in the rat hippocampus. J Physiol. 1997;504:603–610. doi: 10.1111/j.1469-7793.1997.603bd.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Kalman D, Morissette SB, George TP. Co-morbidity of smoking in patients with psychiatric and substance use disorders. Am J Addict. 2005;14:106–123. doi: 10.1080/10550490590924728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Kassam S, Herman P, Goodfellow N, Alves N, Lambe E. Developmental excitation of corticothalamic neurons by nicotinic acetylcholine receptors. J Neurosci. 2008;28:8756–8764. doi: 10.1523/JNEUROSCI.2645-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Kawai H, Lazar R, Metherate R. Nicotinic control of axon excitability regulates thalamocortical transmission. Nature Neurosci. 2007;10:1168–1175. doi: 10.1038/nn1956. [DOI] [PubMed] [Google Scholar]
  76. King S, Caldarone B, Picciotto M. beta2-subunit-containing nicotinic acetylcholine receptors are critical for dopamine-dependent locomotor activation following repeated nicotine administration. Neuropharmacology. 2004;47:132–139. doi: 10.1016/j.neuropharm.2004.06.024. [DOI] [PubMed] [Google Scholar]
  77. King SL, Marks MJ, Grady SR, Caldarone BJ, Koren AO, Mukhin AG, Collins AC, Picciotto MR. Conditional expression in corticothalamic efferents reveals a developmental role for nicotinic acetylcholine receptors in modulation of passive avoidance behavior. J Neurosci. 2003;23:3837–3843. doi: 10.1523/JNEUROSCI.23-09-03837.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Klein LC, Stine MM, Pfaff DW, Vandenbergh DJ. Maternal nicotine exposure increases nicotine preference in periadolescent male but not female C57B1/6J mice. Nic Tobacco Res. 2003;5:117–124. doi: 10.1080/14622200307257. [DOI] [PubMed] [Google Scholar]
  79. Klink R, de Kerchove d’Exaerd A, Zol M, Changeu JP. Molecular and physiological diversity of nicotinic acetylcholine receptors in the midbrain dopaminergic nuclei. J Neurosci. 2001;21:1452–1463. doi: 10.1523/JNEUROSCI.21-05-01452.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Knott V, Thompson A, Shah D, Ilivitsky V. Neural expression of nicotine’s antidepressant properties during tryptophan depletion: an EEG study in healthy volunteers at risk for depression. Biological psychology. 2012;91:190–200. doi: 10.1016/j.biopsycho.2012.06.002. [DOI] [PubMed] [Google Scholar]
  81. Koob GF. Drugs of abuse: anatomy, pharmacology and function of reward pathways. Trends Pharmacol Sci. 1992;13:177–184. doi: 10.1016/0165-6147(92)90060-j. [DOI] [PubMed] [Google Scholar]
  82. Leao RN, Mikulovic S, Leao KE, Munguba H, Gezelius H, Enjin A, Patra K, Eriksson A, Loew LM, Tort AB, Kullander K. OLM interneurons differentially modulate CA3 and entorhinal inputs to hippocampal CA1 neurons. Nat Neurosci. 2012;15:1524–1530. doi: 10.1038/nn.3235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Lerman C, Caporaso N, Main D, Audrain J, Boyd NR, Bowman ED, Shields PG. Depression and self-medication with nicotine: the modifying influence of the dopamine D4 receptor gene. Health Psychology. 1998;17:56–62. doi: 10.1037//0278-6133.17.1.56. comment. [DOI] [PubMed] [Google Scholar]
  84. Levine A, Huang Y, Drisaldi B, Griffin EJ, Pollak D, Xu S, Yin D, Schaffran C, Kandel D, Kandel E. Molecular mechanism for a gateway drug: epigenetic changes initiated by nicotine prime gene expression by cocaine. Sci Transl Med. 2011;3:107ra109. doi: 10.1126/scitranslmed.3003062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Liang K, Poytress BS, Chen Y, Leslie FM, Weinberger NM, Metherate R. Neonatal nicotine exposure impairs nicotinic enhancement of central auditory processing and auditory learning in adult rats. Eur J Neurosci. 2006;24:857–866. doi: 10.1111/j.1460-9568.2006.04945.x. [DOI] [PubMed] [Google Scholar]
  86. Liang K, Poytress BS, Weinberger NM, Metherate R. Nicotinic modulation of tone-evoked responses in auditory cortex reflects the strength of prior auditory learning. Neurobiol Learn Mem. 2008;90:138–146. doi: 10.1016/j.nlm.2008.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Liu Z, Neff RA, Berg DK. Sequential interplay of nicotinic and GABAergic signaling guides neuronal development. Science. 2006;314:1610–1613. doi: 10.1126/science.1134246. [DOI] [PubMed] [Google Scholar]
  88. Lozada AF, Wang X, Gounko NV, Massey KA, Duan J, Liu Z, Berg DK. Glutamatergic synapse formation is promoted by alpha7-containing nicotinic acetylcholine receptors. J Neurosci. 2012a;32:7651–7661. doi: 10.1523/JNEUROSCI.6246-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Lozada AF, Wang X, Gounko NV, Massey KA, Duan J, Liu Z, Berg DK. Induction of Dendritic Spines by beta2-Containing Nicotinic Receptors. J Neurosci. 2012b;32:8391–8400. doi: 10.1523/JNEUROSCI.6247-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Lu Y, Grady S, Marks MJ, Picciotto M, Changeux JP, Collins AC. Pharmacological characterization of nicotinic receptor-stimulated GABA release from mouse brain synaptosomes. J Pharmacol Exp Ther. 1998;287:648–657. [PubMed] [Google Scholar]
  91. Mansvelder HD, Keath JR, McGehee DS. Synaptic mechanisms underlie nicotine-induced excitability of brain reward areas. Neuron. 2002;33:905–919. doi: 10.1016/s0896-6273(02)00625-6. [DOI] [PubMed] [Google Scholar]
  92. Mansvelder HD, McGehee DS. Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron. 2000;27:349–357. doi: 10.1016/s0896-6273(00)00042-8. [DOI] [PubMed] [Google Scholar]
  93. Maskos U, Molles BE, Pons S, Besson M, Guiard BP, Guilloux JP, Evrard A, Cazala P, Cormier A, Mameli-Engvall M, Dufour N, Cloez-Tayarani I, Bemelmans AP, Mallet J, Gardier AM, David V, Faure P, Granon S, Changeux JP. Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature. 2005;436:103–107. doi: 10.1038/nature03694. [DOI] [PubMed] [Google Scholar]
  94. McGehee DS, Heath MJ, Gelber S, Devay P, Role LW. Nicotine enhancement of fast excitatory synaptic transmission in CNS by presynaptic receptors. Science. 1995;269:1692–1696. doi: 10.1126/science.7569895. [DOI] [PubMed] [Google Scholar]
  95. McGehee DS, Role LW. Presynaptic ionotropic receptors. Curr Opin Neurobiol. 1996;6:342–349. doi: 10.1016/s0959-4388(96)80118-8. [DOI] [PubMed] [Google Scholar]
  96. McGranahan TM, Patzlaff NE, Grady SR, Heinemann SF, Booker TK. alpha4beta2 nicotinic acetylcholine receptors on dopaminergic neurons mediate nicotine reward and anxiety relief. J Neurosci. 2011;31:10891–10902. doi: 10.1523/JNEUROSCI.0937-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Metherate R, Hsieh CY. Synaptic mechanisms and cholinergic regulation in auditory cortex. Prog Brain Res. 2004;145:143–156. doi: 10.1016/s0079-6123(03)45010-3. [DOI] [PubMed] [Google Scholar]
  98. Mineur YS, Abizaid A, Rao Y, Salas R, DiLeone RJ, Gundisch D, Diano S, De Biasi M, Horvath TL, Gao X-B, Picciotto MR. Nicotine decreases food intake through activation of POMC neurons. Science. 2011;332:1330–1332. doi: 10.1126/science.1201889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Mineur YS, Brunzell DH, Grady SR, Lindstrom JM, McIntosh JM, Marks MJ, King SL, Picciotto MR. Localized low-level re-expression of high-affinity mesolimbic nicotinic acetylcholine receptors restores nicotine-induced locomotion but not place conditioning. Genes Brain Behav. 2009a;8:257–266. doi: 10.1111/j.1601-183X.2008.00468.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Mineur YS, Eibl C, Young G, Kochevar C, Papke RL, Gundisch D, Picciotto MR. Cytisine-based nicotinic partial agonists as novel antidepressant compounds. J Pharmacol Exp Therap. 2009b;329:377–386. doi: 10.1124/jpet.108.149609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Mineur YS, Obayemi A, Wigestrand MB, Fote G, Calarco CA, Li AM, Picciotto MR. Cholinergic signaling in the hippocampus regulates social stress resilience, anxiety- and depression-like behavior. Proc Natl Acad Sci U S A. 2013 doi: 10.1073/pnas.1219731110. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Mineur YS, Picciotto MR. Nicotine receptors and depression: revisiting and revising the cholinergic hypothesis. Tr Pharmacol Sci. 2010;31:580–586. doi: 10.1016/j.tips.2010.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Mineur YS, Somenzi O, Picciotto MR. Cytisine, a partial agonist of high-affinity nicotinic acetylcholine receptors, has antidepressant-like properties in male C57BL/6J mice. Neuropharmacology. 2007;52:1256–1262. doi: 10.1016/j.neuropharm.2007.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Morrell HE, Cohen LM, McChargue DE. Depression vulnerability predicts cigarette smoking among college students: Gender and negative reinforcement expectancies as contributing factors. Addict Behav. 2010;35:607–611. doi: 10.1016/j.addbeh.2010.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Munafo MR, Araya R. Cigarette smoking and depression: a question of causation. Br J Psychiatry. 2010;196:425–426. doi: 10.1192/bjp.bp.109.074880. [DOI] [PubMed] [Google Scholar]
  106. Nakauchi S, Brennan RJ, Boulter J, Sumikawa K. Nicotine gates long-term potentiation in the hippocampal CA1 region via the activation of α2* nicotinic ACh receptors. Eur J Neurosci. 2007;25:2666–2681. doi: 10.1111/j.1460-9568.2007.05513.x. [DOI] [PubMed] [Google Scholar]
  107. Oppenheim G, Ebstein RP, Belmaker RH. Effect of lithium on the physostigmine-induced behavioral syndrome and plasma cyclic GMP. J Psychiatric Res. 1979;15:133–138. doi: 10.1016/0022-3956(79)90024-4. [DOI] [PubMed] [Google Scholar]
  108. Overstreet DH. The Flinders sensitive line rats: a genetic animal model of depression. Neurosci Biobehav Rev. 1993;17:51–68. doi: 10.1016/s0149-7634(05)80230-1. [DOI] [PubMed] [Google Scholar]
  109. Palmatier MI, Liu X, Caggiula AR, Donny EC, Sved AF. The role of nicotinic acetylcholine receptors in the primary reinforcing and reinforcement-enhancing effects of nicotine. Neuropsychopharmacology. 2007;32:1098–1108. doi: 10.1038/sj.npp.1301228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Parrott AC. Smoking cessation leads to reduced stress, but why? Int J Addict. 1995;30:1509–1516. doi: 10.3109/10826089509055846. [DOI] [PubMed] [Google Scholar]
  111. Parrott AC. Does cigarette smoking cause stress? Am Psychol. 1999;54:817–820. doi: 10.1037//0003-066x.54.10.817. [DOI] [PubMed] [Google Scholar]
  112. Parrott AC. Cigarette smoking does cause stress. Am Psychologist. 2000;55:1159–1160. doi: 10.1037//0003-066x.55.10.1159. [DOI] [PubMed] [Google Scholar]
  113. Parrott AC, Murphy RS. Explaining the stress-inducing effects of nicotine to cigarette smokers. Human psychopharmacology. 2012;27:150–155. doi: 10.1002/hup.1247. [DOI] [PubMed] [Google Scholar]
  114. Pasumarthi RK, Fadel J. Stimulation of lateral hypothalamic glutamate and acetylcholine efflux by nicotine: implications for mechanisms of nicotine-induced activation of orexin neurons. J Neurochem. 2010;113:1023–1035. doi: 10.1111/j.1471-4159.2010.06666.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Perkins KA. Weight gain following smoking cessation. Journal of consulting and clinical psychology. 1993;61:768–777. doi: 10.1037//0022-006x.61.5.768. [DOI] [PubMed] [Google Scholar]
  116. Perkins KA, Karelitz JL, Conklin CA, Sayette MA, Giedgowd GE. Acute negative affect relief from smoking depends on the affect situation and measure but not on nicotine. Biol Psychiatry. 2010;67:707–714. doi: 10.1016/j.biopsych.2009.12.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Philip NS, Carpenter LL, Tyrka AR, Whiteley LB, Price LH. Varenicline Augmentation in Depressed Smokers: An 8-Week, Open-Label Study. J Clin Psychiatry. 2009;70:1026–1031. doi: 10.4088/jcp.08m04441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Picciotto M, Addy N, Mineur Y, Brunzell D. It is not “either/or”: Activation and desensitization of nicotinic acetylcholine receptors both contribute to behaviors related to nicotine addiction and mood. Prog Neurobiol. 2008;84:329–342. doi: 10.1016/j.pneurobio.2007.12.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Picciotto MR. Nicotine as a modulator of behavior: beyond the inverted U. Tr Pharmacol Sci. 2003;24:493–499. doi: 10.1016/S0165-6147(03)00230-X. [DOI] [PubMed] [Google Scholar]
  120. Picciotto MR, Higley MJ, Mineur YS. Acetylcholine as a neuromodulator: cholinergic signaling shapes nervous system function and behavior. Neuron. 2012;76:116–129. doi: 10.1016/j.neuron.2012.08.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Picciotto MR, Zoli M, Léna C, Bessis A, Lallemand Y, Le Novère N, Vincent P, Merlo Pich E, Brulet P, Changeux J-P. Abnormal avoidance learning in mice lacking functional high-affinity nicotine receptor in the brain. Nature. 1995;374:65–67. doi: 10.1038/374065a0. [DOI] [PubMed] [Google Scholar]
  122. Picciotto MR, Zoli M, Rimondini R, Léna C, Marubio LM, Merlo Pich E, Fuxe K, Changeux JP. Acetylcholine receptors containing the beta-2 subunit are involved in the reinforcing properties of nicotine. Nature. 1998;391:173–177. doi: 10.1038/34413. [DOI] [PubMed] [Google Scholar]
  123. Pichat P, Bergis OE, Terranova JP, Urani A, Duarte C, Santucci V, Gueudet C, Voltz C, Steinberg R, Stemmelin J, Oury-Donat F, Avenet P, Griebel G, Scatton B. SSR180711, a novel selective alpha7 nicotinic receptor partial agonist: (II) efficacy in experimental models predictive of activity against cognitive symptoms of schizophrenia. Neuropsychopharmacology. 2007;32:17–34. doi: 10.1038/sj.npp.1301188. [DOI] [PubMed] [Google Scholar]
  124. Pidoplichko VI, Debiasi M, Williams JT, Dani JA. Nicotine activates and desensitizes midbrain dopamine neurons. Nature. 1997;390:401–404. doi: 10.1038/37120. [DOI] [PubMed] [Google Scholar]
  125. Pons S, Fattore L, Cossu G, Tolu S, Porcu E, McIntosh JM, Changeux JP, Maskos U, Fratta W. Crucial role of alpha4 and alpha6 nicotinic acetylcholine receptor subunits from ventral tegmental area in systemic nicotine self-administration. J Neurosci. 2008;28:12318–12327. doi: 10.1523/JNEUROSCI.3918-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Quick MW, Lester RA. Desensitization of neuronal nicotinic receptors. J Neurobiol. 2002;53:457–478. doi: 10.1002/neu.10109. [DOI] [PubMed] [Google Scholar]
  127. Rabenstein RL, Caldarone BJ, Picciotto MR. The nicotinic antagonist mecamylamine has antidepressant-like effects in wild type but not β2 or α7 nicotinic acetylcholine receptor knockout mice. Psychopharmacology. 2006;189:395–401. doi: 10.1007/s00213-006-0568-z. [DOI] [PubMed] [Google Scholar]
  128. Radcliffe KA, Dani JA. Nicotinic stimulation produces multiple forms of increased glutamatergic synaptic transmission. J Neurosci. 1998;18:7075–7083. doi: 10.1523/JNEUROSCI.18-18-07075.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Reitstetter R, Lukas RJ, Gruener R. Dependence of nicotinic acetylcholine receptor recovery from desensitization on the duration of agonist exposure. J Pharmacol Exp Ther. 1999;289:656–660. [PubMed] [Google Scholar]
  130. Rhodes ME, O’Toole SM, Czambel RK, Rubin RT. Male-female differences in rat hypothalamic-pituitary-adrenal axis responses to nicotine stimulation. Brain Res Bull. 2001;54:681–688. doi: 10.1016/s0361-9230(01)00488-9. [DOI] [PubMed] [Google Scholar]
  131. Rice ME, Cragg SJ. Nicotine amplifies reward-related dopamine signals in striatum. Nature Neurosci. 2004;7:583–584. doi: 10.1038/nn1244. [DOI] [PubMed] [Google Scholar]
  132. Risner ME, Goldberg SR. A comparison of nicotine and cocaine self-administration in the dog: fixed-ratio and progressive-ratio schedules of intravenous drug infusion. J Pharmacol Exp Ther. 1983;224:319–326. [PubMed] [Google Scholar]
  133. Role LW, Berg DK. Nicotinic receptors in the development and modulation of CNS synapses. Neuron. 1996;16:1077–1085. doi: 10.1016/s0896-6273(00)80134-8. [DOI] [PubMed] [Google Scholar]
  134. Rollema H, Coe JW, Chambers LK, Hurst RS, Stahl SM, Williams KE. Rationale, pharmacology and clinical efficacy of partial agonists of alpha4beta2 nACh receptors for smoking cessation. Tr Pharmacol Sci. 2007;28:316–325. doi: 10.1016/j.tips.2007.05.003. [DOI] [PubMed] [Google Scholar]
  135. Rollema H, Guanowsky V, Mineur Y, Shrikhande A, Coe J, Seymour P, Picciotto M. Varenicline has antidepressant-like activity in the forced swim test and augments sertraline’s effect. Eur J Pharmacol. 2009;605:114–116. doi: 10.1016/j.ejphar.2009.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Sainsbury A, Zhang L. Role of the arcuate nucleus of the hypothalamus in regulation of body weight during energy deficit. Molecular and cellular endocrinology. 2010;316:109–119. doi: 10.1016/j.mce.2009.09.025. [DOI] [PubMed] [Google Scholar]
  137. Saricicek A, Esterlis I, Maloney KH, Mineur YS, Ruf BM, Muralidharan A, Chen JI, Cosgrove KP, Kerestes R, Ghose S, Tamminga CA, Pittman B, Bois F, Tamagnan G, Seibyl J, Picciotto MR, Staley JK, Bhagwagar Z. Persistent beta2*-Nicotinic Acetylcholinergic Receptor Dysfunction in Major Depressive Disorder. Am J Psychiatry. 2012;169:851–859. doi: 10.1176/appi.ajp.2012.11101546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Shao Z, Yakel JL. Single channel properties of neuronal nicotinic ACh receptors in stratum radiatum interneurons of rat hippocampal slices. J Physiol. 2000;527:507–513. doi: 10.1111/j.1469-7793.2000.00507.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Shytle RD, Silver AA, Lukas RJ, Newman MB, Sheehan DV, Sanberg PR. Nicotinic acetylcholine receptors as targets for antidepressants. Mol Psychiatry. 2002;7:525–535. doi: 10.1038/sj.mp.4001035. [DOI] [PubMed] [Google Scholar]
  140. Staley JK, Krishnan-Sarin S, Cosgrove KP, Krantzler E, Frohlich E, Perry E, Dubin JA, Estok K, Brenner E, Baldwin RM, Tamagnan GD, Seibyl JP, Jatlow P, Picciotto MR, London ED, O’Malley S, van Dyck CH. Human tobacco smokers in early abstinence have higher levels of beta2* nicotinic acetylcholine receptors than nonsmokers. J Neurosci. 2006;26:8707–8714. doi: 10.1523/JNEUROSCI.0546-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Tang J, Dani JA. Dopamine enables in vivo synaptic plasticity associated with the addictive drug nicotine. Neuron. 2009;63:673–682. doi: 10.1016/j.neuron.2009.07.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C, Whiteaker P, Marks MJ, Collins AC, Lester HA. Nicotine activation of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science. 2004;306:1029–1032. doi: 10.1126/science.1099420. [DOI] [PubMed] [Google Scholar]
  143. Tizabi Y, Getachew B, Rezvani AH, Hauser SR, Overstreet DH. Antidepressant-like effects of nicotine and reduced nicotinic receptor binding in the Fawn-Hooded rat, an animal model of co-morbid depression and alcoholism. Progress in neuro-psychopharmacology & biological psychiatry. 2009;33:398–402. doi: 10.1016/j.pnpbp.2008.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Tizabi Y, Rezvani AH, Russell LT, Tyler KY, Overstreet DH. Depressive characteristics of FSL rats: involvement of central nicotinic receptors. Pharmacol Biochem Behav. 2000;66:73–77. doi: 10.1016/s0091-3057(00)00236-7. [DOI] [PubMed] [Google Scholar]
  145. Tolu S, Eddine R, Marti F, David V, Graupner M, Pons S, Baudonnat M, Husson M, Besson M, Reperant C, Zemdegs J, Pages C, Hay YA, Lambolez B, Caboche J, Gutkin B, Gardier AM, Changeux JP, Faure P, Maskos U. Co-activation of VTA DA and GABA neurons mediates nicotine reinforcement. Mol Psychiatry. 2012 doi: 10.1038/mp.2012.83. [DOI] [PubMed] [Google Scholar]
  146. Tolu S, Eddine R, Marti F, David V, Graupner M, Pons S, Baudonnat M, Husson M, Besson M, Reperant C, Zemdegs J, Pages C, Hay YA, Lambolez B, Caboche J, Gutkin B, Gardier AM, Changeux JP, Faure P, Maskos U. Co-activation of VTA DA and GABA neurons mediates nicotine reinforcement. Mol Psychiatry. 2013;18:382–393. doi: 10.1038/mp.2012.83. [DOI] [PubMed] [Google Scholar]
  147. Tucci SA, Genn RF, File SE. Methyllycaconitine (MLA) blocks the nicotine evoked anxiogenic effect and 5-HT release in the dorsal hippocampus: possible role of alpha7 receptors. Neuropharmacology. 2003;44:367–373. doi: 10.1016/s0028-3908(02)00391-x. [DOI] [PubMed] [Google Scholar]
  148. Vieyra-Reyes P, Picciotto MR, Mineur YS. Voluntary oral nicotine intake in mice down-regulates GluR2 but does not modulate depression-like behaviors. Neurosci Lett. 2008;434:18–22. doi: 10.1016/j.neulet.2008.01.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Voorhees CC, Schreiber GB, Schumann BC, Biro F, Crawford PB. Early predictors of daily smoking in young women: the national heart, lung, and blood institute growth and health study. Preventive medicine. 2002;34:616–624. doi: 10.1006/pmed.2002.1026. [DOI] [PubMed] [Google Scholar]
  150. Walters C, Brown S, Changeux J-P, Martin B, Damaj M. The beta2 but not alpha7 subunit of the nicotinic acetylcholine receptor is required for nicotine-conditioned place preference in mice. Psychopharmacol. 2006;184:339–344. doi: 10.1007/s00213-005-0295-x. [DOI] [PubMed] [Google Scholar]
  151. Warner-Schmidt JL, Schmidt EF, Marshall JJ, Rubin AJ, Arango-Lievano M, Kaplitt MG, Ibanez-Tallon I, Heintz N, Greengard P. Cholinergic interneurons in the nucleus accumbens regulate depression-like behavior. Proc Natl Acad Sci U S A. 2012;109:11360–11365. doi: 10.1073/pnas.1209293109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Wonnacott S. Presynaptic nicotinic ACh receptors. Tr Neurosci. 1997;20:92–98. doi: 10.1016/s0166-2236(96)10073-4. [DOI] [PubMed] [Google Scholar]
  153. Wooltorton JRA, Pidoplichko VI, Broide RS, Dani JA. Differential desensitization and distribution of nicotinic acetylcholine receptor subtypes in midbrain dopamine areas. J Neurosci. 2003;23:3176–3185. doi: 10.1523/JNEUROSCI.23-08-03176.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Zhang H, Sulzer D. Frequency-dependent modulation of dopamine release by nicotine. Nature Neurosci. 2004;7:581–582. doi: 10.1038/nn1243. [DOI] [PubMed] [Google Scholar]
  155. Zoli M, Moretti M, Zanardi A, McIntosh JM, Clementi F, Gotti C. Identification of the nicotinic receptor subtypes expressed on dopaminergic terminals in the rat striatum. J Neurosci. 2002;22:8785–8789. doi: 10.1523/JNEUROSCI.22-20-08785.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Zoli M, Picciotto MR. Nicotinic regulation of energy homeostasis. Nicotine Tob Res. 2012;14:1270–1290. doi: 10.1093/ntr/nts159. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES