Sei sulla pagina 1di 13

REVIEWS

Nicotine addiction and nicotinic receptors: lessons from genetically modified mice
Jean-Pierre Changeux

Abstract | The past decades have seen a revolution in our understanding of brain diseases and in particular of drug addiction. This has been largely due to the identification of neurotransmitter receptors and the development of animal models, which together have enabled the investigation of brain functions from the molecular to the cognitive level. Tobacco smoking, the principal yet avoidable cause of lung cancer is associated with nicotine addiction. Recent studies in mice involving deletion and replacement of nicotinic acetylcholine receptor subunits have begun to identify the molecular mechanisms underlying nicotine addiction and might offer new therapeutic strategies to treat this addiction.
Every year, more than five million people worldwide die from the consequences of smoking. These deaths, principally from lung cancer, are avoidable. A formidable obstacle to the prevention of these deaths is that tobacco contains nicotine the major, if not sole, compound responsible for driving the strong addiction to smoking 1. Of those who try to quit smoking, only 35% are successful without the use of nicotine replacement therapies, and no more than one-third are successful with them2. How can a simple chemical substance such as nicotine have so strong an effect on human behaviour? Answering this question is becoming possible now that the molecular and cellular targets and the physiological and behavioural effects of nicotine are being unravelled. The actions of nicotine are mediated by nicotinic acetylcholine (ACh) receptors (nAChRs) (FIG. 1). To increase our understanding of the contribution of different nAChR oligomers3,4 to nicotine addiction, new strategies have been developed. These include, first, deletions in mice5 of nearly all known AChR subunit genes6, and targeted knock-in gene mutations yielding gain-of-function receptor subunits 7,8; second, the re-expression of a deleted gene either using inducible transgenic expression systems9 or by stereotaxic injection of a lentiviral vector carrying the missing gene10,11 or the relevant small interfering RNA11,12; third, the quantitative analysis of the neuronal firing patterns13 and behaviours14 elicited by nicotine in these mice. This Review aims to bridge the gap from genes to cognition in the understanding of nicotine addiction, on the basis of the recent advances in the molecular biology of nAChRs and of animal models with modified nAChR gene expression.

Collge de France and the Institut Pasteur CNRS URA 2182, 25 rue du Dr Roux, 75015 Paris, France. e-mail: changeux@pasteur.fr doi:10.1038/nrn2849

Nicotinic receptors Nicotine interacts with a broad population of nAChR homopentamers and heteropentamers that are distributed throughout the PNS and CNS15. nAChRs are allosteric membrane proteins that respond to ACh and nicotinic agonists by the fast opening (s to ms range) of a cationic channel that is permeable to Na+, K+ and, in some cases, Ca2+ ions4. In the PNS, nAChRs mediate the rapid, phasic effects of local, short-lasting, high ACh levels. In the brain, few examples of fast transmission are documented; but here, nAChRs are also the target of tonically released ACh in lower, modulatory concentrations. Of key importance for the understanding of nicotine addiction, chronic exposure to ACh or nicotinic drugs causes a gradual decrease in the rate of this ionic response (100 ms to minutes), leading to a high-affinity, desensitized, closed state of the receptor 3,4,16 and to additional long-term changes in receptor properties. It also causes an upregulation of the number of high-affinity receptors in the brain17. nAChRs are transmembrane oligomers consisting of five subunits (FIG. 1) that result, in mammals, from the combinatorial assembly of different nAChR subunits encoded by 17 genes. Of these, nine genes encoding -subunits and three encoding -subunits are expressed in the brain. The various nAChR subunit combinations differ in their pharmacological and kinetic properties4
vOlumE 11 | juNE 2010 | 389

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
and also in their cellular and subcellular localization in the brain15. The two main types of nAChRs in the brain are the 7 homo-oligomer, characterized by a fast activation, a low affinity and a high Ca2+ permeability; and the 42 hetero-oligomer, which is typified by a high affinity and slow desensitization3,4. 2 subunitcontaining nAChRs (2*nAChRs) and 7*nAChRs are widely expressed in the brain15, whereas other nAChR subunits have a more restricted expression (FIG. 2). The existence of several functional oligomers with many subunit types, different active site interface profiles and brain distributions15 has made it difficult to elucidate the molecular mechanism of nicotine addiction4. midbrain, which project to the prefrontal cortex (PFC) and to limbic and striatal structures, in particular the nucleus accumbens6,17; and the extended amygdala, the brain stress system 18, the hippocampus 19, the habenulo-interpeduncular system20 and the prefrontal (including insular) cortex 21. Functional mRI studies have shown that acute nicotine injections in anaesthetized mice increase activation in prelimbic, anterior frontal, motor and somatosensory cortices, the nucleus accumbens, vTA, substantia nigra and the thalamus, but not in the motor cortex and amygdala22. Addiction is viewed as the end point of a series of transitions23,24 that are characterized by a compulsion to seek and voluntarily take the drug owing to its reinforcing effects; this leads to a loss of control of intake such that it becomes habitual and ultimately compulsive (BOX 1). when access to the drug is prevented, a negative state emerges that manifests in somatic and affective signs, reflecting a motivational withdrawal syndrome24. Addiction thus emerges as a chronic relapsing disease of the brain. Nicotine addiction differs from other drug addictions in that it has fewer obvious signs of the bingeintoxication stage and a less dramatic withdrawal effect than, for example, opioid or alcohol addiction24. Nicotine selfadministration does not result in a progressive escalation manifested by a compulsively increased drug intake24,25 but by a moderate change in a nicotine-rewarded learning process26. moreover, there is little evidence that nicotine is abused in its pure form27.

CNS circuits in nicotine addiction many brain areas are involved in nicotine addiction (FIG. 2) . These include the dopaminergic neurons in the ventral tegmental area (vTA; nucleus A10) of the

a
Na+, K+, Ca2+

Homomeric 7 7

Heteromeric 42 2

ACh

ACh Extracellular

7 7 7

2 4 2

Intracellular

C M1 M2 M3 M4

Behavioural effects of nAChR subunits To understand the effects of nicotine exposure that lead to addiction, which are mediated by nAChRs, we must examine the role of nAChR subunits on behaviour, learning and reward. The endogenous ligand of nAChRs is ACh; thus, deleting a particular nAChR subunit may affect the behaviour of mice even in the absence of nicotine.
Contribution of nAChRs to cognition and locomotion. Nicotine and nicotinic agonists enhance cognitive and psychomotor behaviours and, conversely, nicotinic antagonists and loss of nAChRs impair cognitive performance2831. However, these findings have been debated and the behavioural consequences of knocking out specific nAChR subunits have therefore been investigated. mice lacking the 2 nAChR subunit (2/ mice) (FIG. 3) show normal spatial orientation learning in the water maze5. By contrast, they do not show the nicotine-induced memory enhancement of an avoidance response to a mild foot shock that is observed in wild-type animals5,9,32, indicating a contribution of the 2 nAChR subunit to the retention of aversive memory. In experiments in which two objects are placed in an open field, wild-type mice progressively increase the number of trajectories between the two objects, whereas 2/ mice do not, indicating that cognitive learning strategies involving spatial memory may also require 2*nAChRs33. moreover, in a
www.nature.com/reviews/neuro

Figure 1 | structure of nAchrs. a | Nicotinic acetylcholine receptors (nAChRs) are transmembrane oligomers consisting of five subunits. b | Each subunit consists of a large Nature Reviews | Neuroscience aminoterminal extracellular domain with an immunoglobulinlike sandwich, a transmembrane domain and a variable cytoplasmic domain. The extracellular domain carries the acetylcholine (ACh)nicotine binding sites at the boundary between subunits. The number of binding sites per pentamer ranges, depending on its composition, from two (in muscle nAChRs or brain 42 nAChRs) to five (in the 7 homopentamer). Sites for allosteric modulators are located in the transmembrane domain. c | The two main types of brain nAChRs are the 7 homooligomer, characterized by a fast activation, a low affinity and a high Ca2+ permeability, and the 42 heterooligomer, typified by a high affinity and slow desensitization3,4,146. In addition to the principal subunits (2, 3, 4 and 6) and complementary subunits (2 and 4) of the nAChR, other subunits (5 and 3), the role of which has been recently reevaluated6,37,113, form part of the organization of various heterooligomers. d | Side view of an 7 nAChR pentamer model, showing five nicotine molecules (dark grey) in the binding sites and the volume of the ion channel (dark blue). Parts a, c and d are reproduced, with permission, from ReF. 4 (2009) Macmillan Publishers Ltd. All rights reserved.
390 | juNE 2010 | vOlumE 11

2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
Prefrontal cortex: orbital, medial, cingulate and insula (executive control)

42 7

Amygdala (emotions)

Hippocampus (contextual information)

2 42 452 34 7

22 (IPN) 42 HBIPN 34 (HB) (withdrawal signs) 334 7

NAc shellcore (reinforcement)

42 Dorsal striatum 452 (habits) 623 6423

Hypothalamus Brain stem (stress symptoms)

VTA

SNpc

4(5)2 4(5)62 623 34 7

LDTg 42

PPTg 42

Residentintruder test
A test for social interaction and aggressive behaviour in rodents. An unfamiliar mouse (the intruder) is introduced into the cage of a mouse that has been kept isolated in its resident cage for several months.

Navigation
Spontaneous locomotor behaviour characterized by large movements at fast speed, aimed at acquiring general information about the environment.

Figure 2 | Neuronal mechanisms involved in nicotine addiction: a model. Many brain areas contain nicotinic acetylcholine receptor (nAChR) subunits and are involved in nicotine addiction. First, the somata of the dopaminergic Nature Reviews | Neuroscience neurons that contribute to nicotine intake and reinforcement are in the ventral tegmental area (VTA) of the midbrain: they project to the prefrontal cortex and to limbic areas, in particular the hippocampus and nucleus accumbens (NAc) in the striatum10,147149. These VTA neurons receive cholinergic innervation from the pedunculopontine tegmental nucleus (PPTg) and the adjacent laterodorsal tegmental nucleus (LDTg)150,151. Second, the emergence of a negative emotional state and withdrawal syndrome following smoking cessation or nicotine deprivation mobilizes distinct neural circuits that can include the extended amygdala and brain stress systems18, the hypothalamus, hippocampus19, substantia nigra pars compacta (SNpc), and/or the habenulainterpeduncular (HBIPN) system20. Third, the switch from voluntary nicotine use to compulsive drug use may represent a global topdown gating transition from control by a prefrontal (cortical and insular) global neuronal workspace (BOX 1) to subcortical (striatal) control21,82,130.

residentintruder test,

Exploration
Spontaneous locomotor behaviour characterized by small and slow movements, enabling more precise investigation of the environment.

Nigrostriatal pathway
The dopaminergic pathway from the substantia nigra to the striatum, which is associated with motor control.

Mesolimbic pathway
The dopaminergic pathway from the ventral tegmental area to the nucleus accumbens and limbic areas, which is associated with reward processing.

2/ resident mice show more approach behaviour and attempt fewer escapes to stop the interaction, suggesting that 2 nAChR subunits have a role in social interaction33. mouse locomotor behaviour as assessed in an open field can be divided into navigation and exploration14,33. In 2/ mice10,33 the balance between these behaviours is shifted in favour of navigation and their exploratory behaviour is different from that of wild-type mice14,33. This indicates a role of 2*nAChRs in exploration strategies and decision making 14. Re-expression of the 2 subunit in the vTA by stereotaxic gene expression10,11 restored exploratory behaviour in an open field without modifying navigation. Distinct neural mechanisms thus underlie exploration and navigation. This dissociation was further investigated by evaluating the contribution of endogenous ACh acting on 2*nAChRs in the ascending nigrostriatal pathway and mesolimbic pathway11. Re-expression of 2 subunits in the substantia nigra pars compacta (SNpc) of 2/ mice restored navigation but not exploration to wild-type

levels; the opposite effect occurred when 2 subunits were re-expressed in the vTA. Results from RNA silencing experiments were consistent with these findings11. These data indicate a dissociation between the upstream, cholinergic, nAChR-mediated control of the nigrostriatal pathway and that of the mesolimbic pathway in regulating exploration and navigation11. mice lacking the gene encoding the 4 subunit, the principal partner of the 2 subunit, have normal baseline locomotor activity. However, compared with wild-type mice they show a sustained increase in cocaine-elicited locomotor activity and recover more quickly from nicotine-elicited locomotor suppression34. nAChR gain-of-function mutations 3,4 that result in higher receptor affinity and loss of desensitization of 4* and 6*nAChRs cause (in the absence of chronic nicotine exposure) increased locomotion and deficits in motor learning 8,35. Thus, both the 4 subunit and the 6 subunit contribute to spontaneous locomotor behaviour 12, which was anticipated from their association with the 2 subunit in high-affinity oligomers3,4,15.
vOlumE 11 | juNE 2010 | 391

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
Fear conditioning
A form of learning in which an aversive stimulus (for example, an electric shock) is associated with a particular neutral context (for example, a room) or neutral stimulus (for example, a tone), resulting in the expression of fear responses to the originally neutral stimulus or context.

Fractional anisotropy
A parameter in diffusion tensor imaging, which images brain structures by measuring the diffusion properties of water molecules. It provides information about the microstructural integrity of the white matter.

The 3 subunit can assemble with 6, 4 and 2 subunits in nigrostriatal dopaminergic pathways. mice lacking the gene encoding the 3 subunit show altered locomotor activity and prepulse inhibition of the acoustic startle response (both of which are partially regulated by dopaminergic transmission), supporting a role for 3*nAChRs in modulating these behaviours36. However, behavioural changes elicited by nicotine have not been tested in these mice. 3/ mice suffer from high perinatal mortality, which impairs research on the effects of nicotine in these mice. Compared with wild-type mice, heterozygous 3+/ mice were partially resistant to nicotine-elicited seizures but had normal sensitivity to the hypo-locomotor effects of 0.5 mg per kg nicotine (possibly owing to partial expression of this subunit in the heterozygous mice)37. mice lacking 7 nAChRs38 show normal contextual and fear conditioning, spatial memory and anxiety 39. However, in an attentional task that relies on the integrity of the prefrontal cortex (at least in rats40), 7/ mice were slower and less efficient than their wild-type littermates41. moreover, mice lacking both the 7 and 2 subunits show greater impairment in passive avoidance learning than 2/ mice42. In conclusion, the available data from mouse studies are consistent with a role for activation of 4*, 6*, 2* and 7*nAChRs by endogenous ACh in locomotor

behaviour and cognitive function. The data underline the contribution of these subunits in ascending dopaminergic pathways and also in more elaborate, top-down gating strategies involving attentional control (BOX 1; FIG. 2). The contribution of 3 and 3 subunits to cognition and locomotion has been noted but needs further investigation, together with that of the still poorly investigated 5 subunit. Role of nAChR subunits in the rewarding effects of nicotine. The reinforcing effects of nicotine can be demonstrated in tests of volitional nicotine self-administration, which can be systemic (by intravenous infusion or introduction into the drinking water 43,44) or local (by nicotine injection into the vTA10,45). when mice were primed with systemic cocaine self-administration46 and cocaine was subsequently substituted with nicotine, wild-type mice continued to self-administer nicotine, whereas 2/ mice did not46,47. Drug-naive mice lacking the 2, 4 and 6 subunits also did not self-administer nicotine systemically, but this response was normalized when these subunits were re-expressed in the vTA44. By contrast, deletion of the 7 subunit did not affect systemic self-administration of nicotine. when local, intra-vTA self-administration was tested, mice lacking the 2 subunit again did not show nicotine self-administration10. Re-expressing the gene in the vTA normalized this behaviour. In 4/ mice,

Box 1 | Nicotine addiction and the global neuronal workspace


Addiction is viewed as the end point of a series of transitions from initial voluntary drug use to the loss of conscious control over this behaviour. The neural correlates of consciousness have been the subject of abundant discussion and model building. In particular, the global neuronal workspace (GNW) model130 proposes that conscious information processing recruits networks of neurons with long-range axons that reciprocally connect distinct cortical areas, including the prefrontal, parietal, temporal and cingulate cortices130133. In the GNW model, the striatum, hippocampus and amygdala are under the top-down control of the prefrontal cortex (FIG. 2). According to this proposal, the GNW is engaged in decision-making and might therefore have a key role in the loss of control that characterizes the ultimate stage of nicotine addiction3. Patients with damage to the ventromedial prefrontal cortex and people who are addicted to drugs of abuse show similar deficits in cognitive behaviour21. In addition, smokers with brain damage involving the insula undergo a disruption of smoking addiction and are more likely to quit smoking than smokers with brain damage not involving the insula134, an observation corroborated by a study in rats135. The insula sends excitatory projections to the nucleus accumbens and acts in concert with regions that are involved in maintaining specific goals, such as the dorsolateral prefrontal cortex136. Activation by nicotine might therefore influence the GNW, subverting decision-making to seek and procure drugs134 and thereby hijacking the cognitive functions exerting inhibitory control to resist drug use21. Consistent with this possibility, repeated high-frequency transcranial magnetic stimulation of the left dorsolateral prefrontal cortex combined with either smoking-related or neutral cues reduced cigarette consumption and nicotine dependence. Furthermore, this treatment blocked the craving induced by daily presentation of smoking-related pictures137. Relevant to a possible effect of chronic drug use on the GNW, diffusion tensor imaging studies have shown reduced frontal white matter integrity in people who show cocaine dependence138 or abuse heroin139. The same method has revealed that prenatal and adolescent exposure to tobacco smoke alters the development of the microstructure of the white matter, with increased fractional anisotropy in right and left frontal regions and in the genu of the corpus callosum140. These observations suggested that nicotine might directly act on white matter. There is electrophysiological evidence supporting a direct action of nicotine on axon conduction, possibly at the level of the node of Ranvier80,141,142. Moreover, low levels of nicotine applied to thalamocortical axons in slice preparations enhance prefrontal cortical activity80,143. These observations support a direct control of the GNW by nicotine at the white matter level and are consistent with the expression of functional nicotinic acetylcholine receptors on prefrontal cortical neurons79. The observations also support the proposal that a gating circuit modulates, in a top-down manner, the nicotine-elicited activation of the reward system involving the ventral tegmental area101 (FIG. 2), which is engaged in the volitional consumption of nicotine. According to the GNW model, a functional relationship becomes established between nicotine self-administration and the noradrenaline- and hypocretin-regulated state of awareness144,145, which is a prerequisite for the self-discipline of conscious behaviour.

392 | juNE 2010 | vOlumE 11 2010 Macmillan Publishers Limited. All rights reserved

www.nature.com/reviews/neuro

REVIEWS
Genomic RNA packaging system 5-LTR 1 kb Rev response element Gene of interest Central polypurine tract Central termination sequence (cPPT) FLAP PGK XhoI 2 NheI IRES2 EGFP SalI BstBI WPRE 3-LTR U3

Promoter

Figure 3 | Method of 2 subunit gene re-expression using a lentiviral vector. The lentiviral expression vector contains a bicistronic cassette that simultaneously expresses the 2 subunit and enhanced green fluorescent protein (EGFP) to enable Nature Reviews | Neuroscience detection of transduced cells. Vectors were based on the construct termed pTRIPU3 from which the elongation factor 1 (EF1) promoter was removed. The mouse phosphoglycerate kinase (PGK) promoter was amplified by PCR and inserted into the plasmid, with the restriction sites EcoRI and BamHI upstream and downstream of the PGK promoter, respectively. To create the 2internal ribosome entry site 2 (IRES2)EGFP construct, a new SalI restriction site was created 3 to the EGFP stop codon in the pIRES2EGFP expression plasmid by mutagenesis, and a new XhoI restriction site was created downstream of the NheI restriction site. The wildtype mouse 2 subunit was then ligated between the XhoINheI sites of pIRES2EGFP. The 2IRES2EGFP construct was then ligated into the pTRIPU3PGK vector using XhoISalI sites. Finally, the WPRE (Woodchuck hepatitis posttranscriptional regulatory element) sequence was added using SalIBstBI sites. LTR, long terminal repeat. Figure is modified, with permission, from ReF. 10 (2005) Macmillan Publishers Ltd. All rights reserved.

Gene cluster
A group of neighbouring genes on a chromosome.

Conditioned place preference task


A test for nicotine addiction potential in which the mouse receives a dose of nicotine in a distinct environment, after which the amount of time the mouse spends in that environment is assessed.

Nicotine discrimination
The ability of mice to discriminate nicotine from saline using a two-bar operant procedure with a tandem schedule of food reinforcement.

intra-vTA self-administration initially increased and then decreased, and re-expression of the 4 subunit in the vTA normalized self-administration. Deletion of the 6 subunit had no significant effect on intracranial nicotine self-administration14. Thus, 42* and 62*nAChRs, but not 7*nAChRs, are necessary and sufficient for systemic nicotine reinforcement in drug-naive mice44; in intra-vTA self-administration, the 6 subunit has no major role but the 4 subunit is required, together with the 2 subunit, for long-term self-administration behaviour. By contrast, deletion of 5 subunits or overexpression of 4 subunits enhanced nicotine self-administration48,49, suggesting a modulatory action of these subunits on nicotine reward pathways. Interestingly, the 5 subunit is abundantly expressed in dopaminergic neurons50. moreover, the gene encoding this and the 3 and 4 subunits form a gene cluster that has been associated in genome-wide association studies with lung cancer. The genes encoding the 3 and 6 subunits also form a gene cluster that has been associated with nicotine dependence (see Supplementary information S1 (box)). The rewarding effects of nicotine can also be tested in the conditioned place preference task, in which a dose of nicotine is paired with a distinct environmental cue. mice lacking the 2 subunit did not show nicotineconditioned place preference or nicotine discrimination, whereas 7/ mice did51,52. Knock-in mice expressing hypersensitive, high-affinity 4-containing receptors displayed conditioned place preference with nicotine concentrations 50-fold lower than wild-type mice8. This test also showed that the 6 subunit has a role in mediating the rewarding effects of nicotine53. Overall, these observations provide evidence that nAChRs containing 4, 2 and to some extent 5, 6 and 4 subunits mediate the rewarding effects of nicotine8,46, whereas the role of the 7 subunit is less clear (however, see ReF. 54).

Physiological effects of nAChR subunits Differential effects of nAChR subunits on neural activity in the VTA. Because the dopaminergic system of the vTA mediates the reinforcing effects of nicotine55,56, nicotine is expected to modulate firing of these dopaminergic neurons. various nAChR oligomers are indeed expressed on the GABA (-aminobutyric acid)-containing interneurons in the vTA, on excitatory or inhibitory inputs to the vTA57 and on dopaminergic vTA neurons. Dopaminergic vTA neurons express nAChR oligomers with various subunit compositions (FIG. 4), the distribution of which may differ in the somatodendritic and axonal compartments15. Single-cell recordings of vTA dopaminergic neurons reveal two neuronal firing rhythms in vivo, which occur both spontaneously and in the presence of nicotine: a slow, regular single-spike firing and a bursting mode58,59 (FIG. 5). The burst-firing mode causes a substantially larger dopamine release than regular spiking and greater activation of immediate early genes in the target areas of the dopaminergic neurons60,61. The transition from regular (tonic) firing to bursting (phasic) activity has been associated with receiving reward-predicting stimuli and unpredicted rewards62. Burst firing is absent in midbrain slice preparations (which lack afferent fibres)63, indicating that it depends on the input to the dopaminergic neurons, including glutamatergic afferents originating in part from the PFC, and cholinergic and glutamatergic input from the tegmental pedunculopontine nucleus and the laterodorsal tegmental nucleus, respectively 58,64,65 (FIG. 4). A closer analysis of the firing patterns of vTA dopaminergic neurons13 reveals four modes of spontaneous firing in wild-type anaesthetized mice: first, a low-frequency, regular firing mode without bursting (low frequency, low bursting); second, a low-frequency firing mode with abundant bursts (low frequency, high bursting); third, a high-frequency firing mode with few bursting events (high frequency, low bursting); and
vOlumE 11 | juNE 2010 | 393

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
PFC PPTg VTA Glu 7 GABA NAc 462(3) 4(5)2 62 7 42 GABA 42 42 42 LDTg PPTg

DA

ACh

42 ACh

4(5)2 4(5)62 62(3) (34) 7

Figure 4 | nAchr subunits control the afferent and efferent connectivity of dopaminergic neurons from the VTA. Dopaminergic neurons (DA; shown |in red) in the Nature Reviews Neuroscience ventral tegmental area (VTA) receive two main types of excitatory inputs. First, from cholinergic neurons (shown in blue) of the laterodorsal tegmental nucleus (LDTg) and pedunculopontine nucleus (PPTg); and second, from glutamatergic neurons (shown in green) of various sources, including the prefrontal cortex (PFC) and the PPTg. Dopaminergic neurons also receive inhibitory input from GABA (aminobutyric acid)ergic neurons (shown in yellow) and send axonal projections to the nucleus accumbens (NAc), where they receive inputs from intrinsic cholinergic interneurons (shown in blue). Behavioural and physiological studies with genetically modified mice are revealing the role of the indicated nicotinic acetylcholine receptor (nAChR) subunits in nicotine reward.

fourth, a mode with high-frequency firing and many bursts (high frequency, high bursting) (FIG. 5). The contribution of the various nAChR subunits to these patterns have been investigated in mutant mice. In vTA dopaminergic neurons of 2/ mice, only the low-frequency, low-bursting mode persisted (FIG. 5), whereas those of 7/ mice showed only the highfrequency, high-bursting mode. This suggests that high-frequency, high-burst firing require the activation of 2* but not 7*nAChRs. In wild-type mice, an intravenous injection of nicotine caused a fast and sustained (~10 min) increase in firing frequency in vTA dopaminergic neurons. This response was abolished in 2/ mice46 and restored by lentiviral re-expression of 2 subunits in the vTA10; however, this effect did not persist for more than 2 min10. This suggests that the sustained firing of dopaminergic neurons in the vTA also depends on the presence of 2*nAChRs in excitatory structures projecting to the vTA, such as glutamate-containing neurons from the prefrontal cortex or ponto-tegmental afferents (FIG. 4). In 7/ mice, intravenous nicotine injections caused a large and rapid increase in the firing rate of dopaminergic neurons13 that was short-lasting and followed by a plateau at a slightly lower firing rate (compared with the initial increase). This suggests that the absence of 7*nAChR-mediated glutamatergic excitation unmasks an inhibitory effect of GABAergic neurons (which do not contain 7*nAChRs) (FIG. 4). The data imply that a concomitant activation of 7* and 2*nAChRs may be necessary for the full expression of events leading to nicotine reinforcement, with 2*nAChRs switching neurons from a resting to an excited state (global, tonic regulation), and 7*nAChRs finely tuning this state after 2*nAChRs have been activated13.
394 | juNE 2010 | vOlumE 11

Intravenous nicotine injections in 4/ mice66 caused an increase in the firing rate in the vTA that was lower than in wild-type mice, without affecting the mean bursting activity. By contrast, intravenous nicotine injections caused a similar response in vTA neurons of 6/ mice and wild-type mice. As the spontaneous activity of vTA neurons of 4/ and 6/ mice was similar to that of wild-type mice, these findings indicate that the 4 subunit, but not the 6 subunit, is required for the transition from tonic to phasic firing that is crucial for reinforcement14,66 (in particular, in response to reward-predicting stimuli and unpredicted rewards62). As discussed above, the bursting activity of vTA dopaminergic neurons enhances dopamine release in the striatum (FIG. 4). Tonic ACh release from striatal cholinergic (excitatory) interneurons is thought to control dopamine release in the striatum by activating presynaptic nAChRs on the dopaminergic terminals6770 (FIG. 4). According to one hypothesis71, continued exposure to nicotine desensitizes these nAChRs and so enhances striatal dopamine release from bursts6770, causing facilitation of dopamine-dependent reinforcement. In brain slices of mice lacking the 6 or the 4 subunit, dopamine release in the nucleus accumbens is not modified by nicotine application, in contrast to wild-type brain slices, showing that both 6* and 4*nAChRs have a key role in the control of striatal dopamine release by ACh or nicotine in the nucleus accumbens70. By favouring dopamine release from bursts while depressing dopamine release from tonic signals, nicotine acting on 46*nAChRs effectively increases the signal-to-noise ratio of reward-related burst activity along dopamine afferents72. The precise roles of 4 and 6 subunits in nicotine reinforcement might depend on the differential cellular distribution of the functional 4* and 6*nAChRs in vTA dopaminergic neurons (the 2 subunit being necessary for both 4* and 6*nAChR activation). Thus, 462*nAChRs might be preferentially expressed on the axon terminating in the nucleus accumbens, with 42*nAChRs remaining in the soma. However, this has still to be confirmed experimentally. Although nicotine targets nAChRs on both dopaminergic and GABAergic neurons in the vTA (FIG. 4) , it is unclear which of the two neuron types is more important for mediating the effects of nicotine on dopamine release in the nucleus accumbens. In a study on brain slices73, nicotine application caused a robust increase in inhibitory postsynaptic current (IPSC) frequency in dopaminergic neurons of the vTA, which was followed by a decrease to below baseline after the removal of nicotine. This effect was reduced by blocking 7*nAChRs, and completely abolished by blocking non- 7*nAChRs. The authors73 proposed that the nicotine-elicited increase in IPSC frequency was due to activation of 42*nAChRs on GABAergic neurons (FIG. 4) rather than activation of nAChRs on dopaminergic neurons. As nicotine is known to have an overall stimulatory action on dopaminergic neurons, the authors further proposed that the transient nicotine-elicited increase in GABA transmission to dopaminergic neurons is followed by a desensitization
www.nature.com/reviews/neuro

2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
a b
Spike in burst (%)
100 80 60 40 20 0 0 2 4 6 8 10

WT

HFHB

1 mV

LFHB LFLB HFLB

1 sec

Mean frequency (Hz)

c
Spike in burst (%)

100 80 60 40 20 0 0

2/

d
Spike in burst (%)

100 80 60 40 20 0

VTA Vec

LFHB LFLB
0 2 4 6

LFLB

HFLB
8 10

10

Mean frequency (Hz)

Mean frequency (Hz)

Figure 5 | role of nAchr subunits in spontaneous spike firing and burst firing in Nature Reviews | Neuroscience VTA dopaminergic neurons. a | Sample traces showing two types of neuronal firing by dopaminergic neurons in the ventral tegmental area (VTA): singlespike, regular firing (top) and burst firing (bottom). A typical burst starts with a short interval (< 80 ms) and ends with a long interval (> 160 ms). bd | Mean firing frequency is plotted against the percentage of spikes that are within a burst for individual cells of wildtype (WT) mice (b), 2/ mice (c), and in 2/ mice after reexpression of 2 subunits in the VTA using a lentiviral vector (Vec) (d). Four main groups of cells can be distinguished: highfiring, highbursting (HFHB) cells; lowfiring, highbursting (LFHB) cells; highfiring, lowbursting (HFLB) cells; and lowfiring, lowbursting (LFLB) cells. VTA neurons from 2/ mice show only LFLB activity, whereas reexpressing the subunit in the VTA restores HFLB and LFHB activity. Figure is reproduced, with permission, from ReF. 13 (2006) Cell Press.

Spike timing-dependent plasticity


A form of plasticity that results from functional changes in neurons and/or synapses and that depends on the precise timing of action potentials in connected neurons.

Giant depolarizing potentials


An early type of electrical activity that arises in the course of development and is probably due to excitatory actions of GABA.

of the 42*nAChRs, removing the tonic cholinergic drive to GABAergic neurons and thereby disinhibiting dopaminergic neurons73. Conversely, in vivo studies have emphasized the importance of 2*nAChRs on vTA dopaminergic neurons in mediating the effect of nicotine. Intravenous injection of nicotine caused a fast, temporary but robust increase in the firing rate and bursting of these neurons13. This effect was abolished in 2/ mice10,46 and reduced in amplitude and duration in 7/ mice. It was proposed that nicotine acting on 2*nAChRs on dopaminergic neurons causes a switch from a resting to an excited state of these cells, with a less important contribution from 7*nAChRs. The differences of interpretation of the in vitro and in vivo experiments regarding the role in reinforcement of nAChRs on GABAergic versus dopaminergic neurons, respectively, can be addressed using a lentiviral system to control nAChR subunit expression specifically in dopaminergic or GABAergic neurons74. Preliminary in vivo results show that 2 / mice in which the 2 subunit is exclusively re-expressed in dopaminergic neurons of the vTA actively, though transiently, self-administer nicotine 75; this finding is consistent with the proposal of the in vivo study described above13.

Physiological effects in other brain areas. Nicotine affects attention and cognitive performance in rodents, primates and humans76,77, suggesting that the medial prefrontal cortex (mPFC) might be a target of nicotine. Indeed, 2* and 7*nAChRs are expressed in the mouse PFC5,15,78, and iontophoretic application of nicotinic agonists to pyramidal cells in layers IIIII of the rat prelimbic area modulates excitatory synaptic transmission79. Also, glutamate release from layer v pyramidal neurons in response to nicotine applied to thalamocortical input neurons is abolished in 2/ mice, demonstrating a contribution of this subunit to information processing in the PFC80. In slice preparations of mouse PFC, spike timing-dependent plasticity (STDP) can be elicited by pairing stimulation of the excitatory inputs to pyramidal neurons of PFC layer v with postsynaptic spikes81. Nicotine increases the threshold for STDP, eliminating long-term potentiation (lTP) and causing synaptic depression of the excitatory inputs to these cells. This inhibitory effect of nicotine seems to contradict the known cognitive-enhancing effect of nicotine (see ReF. 81 for a discussion), but could be explained by the absence of certain features of in situ information processing in cortical slices13,63. The mPFC, amygdala and hippocampus provide glutamatergic input to neurons in the nucleus accumbens, which are also regulated by dopaminergic neurons in the vTA (FIG. 2). The convergence of these inputs in the nucleus accumbens is thought to be the neural basis of a top-down control, or gating function, of attention on neural processes82 (BOX 1). Here, I briefly review the role of nAChRs in several components of this circuit. Simultaneous extracellular recordings of neuronal activity in the PFC, ventral hippocampus and nucleus accumbens in anaesthetized rats reveal that the mPFC is required for the activation of the nucleus accumbens by the ventral hippocampus83. Several nAChR oligomers are expressed in GABAergic and glutamatergic hippocampal neurons. Nicotine enhances several types of hippocampus-dependent learning and modulates the induction of lTP84, the putative cellular substrate of learning and memory. Specifically, nicotine elicits lTP at hippocampal synapses through 7* and/or 42*nAChRs85 and upregulates 4*nAChR levels in glutamate afferents and GABAergic interneurons86. At birth, most hippocampal glutamatergic synapses are immature and functionally silent, but they can be converted presynaptically into conductive synapses by a brief application of nicotine at immature Schaffer collateralCA1 connections87. Studies with 7/ and 2/ mice showed that 7* and 2*nAChRs are sufficient to modulate nicotine-elicited increases in the frequency of spontaneously occurring giant depolarizing potentials88. Recordings from freely moving mice further showed that nicotine-elicited dopamine signalling in the hippocampus was necessary for lTP at entorhinaldentate gyrus synapses89. GABAergic interneurons in the hippocampal stratum oriensalveus express 2 mRNA90,91. Interestingly, 2*nAChRs do not desensitize in response to nicotine exposure, and continuous exposure to nicotine results in sustained activation of 2*nAChRs and continuous discharge of these interneurons92. This raises the
vOlumE 11 | juNE 2010 | 395

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
Intracranial electrical self-stimulation
A procedure in which rats work (for example, pressing a lever) to obtain rewarding electrical self-stimulation through an electrode implanted in the brains reward system. It is used to measure the sensitivity of the brain reward system in vivo.

possibility that modulation of 2*nAChRs through inhibitory interneurons in the hippocampus contributes to the overall circuit. In conclusion, 2, 4, 7 and 2 nAChR subunits contribute to the effect of nicotine on hippocampal synaptic plasticity. Nicotine affects the ventral hippocampus-toaccumbens pathway. Preliminary findings in mice suggest that the pattern of phasic bursting activity in the nucleus accumbens and the degree of hippocampus accumbens coherence are affected by nicotine and that the altered patterns of activity can persist long after nicotine has been removed from the system 93. This suggests a possible mechanism by which nicotine affects the concerted action of the nucleus accumbens and the hippocampus through the top-down, gating control of the so-called global neuronal workspace (BOX 1) . Co-cultures of ventral hippocampus neurons from 7/ mice and nucleus accumbens neurons from wild-type mice have recently demonstrated the contribution of the 7*nAChR to the hippocampusto- accumbens pathway. These ongoing experiments reveal that 7*nAChRs are required to convert the transient effects of nicotine that are observed in this preparation into long-lasting changes in synaptic transmission94.
VTA Increased dopaminergic neuron bursting activity

Both 7* and 2*nAChRs are expressed in the basal and lateral amygdala nuclei 95,96. A single, low-dose administration of nicotine to mouse basolateral amygdala slices elicits long-lasting facilitation of glutamatergic transmission at cortico-amygdala connections97. Pharmacological studies in wild-type and 7/ mice further reveal that, in the basolateral amygdala, activation of presynaptic nAChRs containing the 7 subunit (together with non-7*nAChRs) facilitates glutamatergic transmission in an activity-dependent manner 97. By controlling neuronal activity in the amygdala, which is a brain region known to have a role in contextual learning (for example, of smoking-related cues), nicotine could modulate an important aspect of nicotine addiction76. In summary, gene deletion and re-expression studies have revealed distinct contributions of 4, 6, 7 and 2 subunits (and possibly the still largely unexplored 3, 5 and 4 subunits) to the short-term effects of nicotine on the physiology of dopaminergic neurons in the vTA versus the SNpc, as well as on dopaminergic versus GABAergic neurons in the vTA. moreover, the studies have distinguished the roles of these nAChR subunits in the acute behavioural effects of nicotine, as manifested by cognitive behaviours and nicotine self-administration, for example. There is also evidence that acute nicotine exposure might influence (through activation of 42* and 7*nAChRs) a global gating circuit that includes the striatum, hippocampus and amygdala, under the top-down control of the PFC (BOX 1).

IFBNST Increased glutamate output 42*nAChR upregulation NAc Increased dopamine release

a b

Increased dopamine neuron activity

Chronic nicotine exposure

Presynaptic 2*nAChR downregulation

Increased presynaptic 7*nAChR transmission

Increased dopaminergic neuron activity

Effects of chronic nicotine exposure Chronic nicotine exposure alters brain reward systems. It increases dopamine release in the nucleus accumbens55,98100, and studies in rats revealed that volitionally consumed (but not passively administered) nicotine lowers the threshold of intracranial electrical self-stimulation for several weeks25. At the cellular level, nicotine self-administration for 2 months causes hyperactivity of dopaminergic neurons in the vTA101. Together, these observations indicate that self-administered nicotine resets dopamine reward circuits and increases their sensitivity to subsequent exposures to nicotine (sensitization). This seems to be unique to nicotine, as other drugs of abuse cause decreased sensitivity of these circuits25,26. Sensitization is thought to be a key step in the process leading to addiction, and several lines of research have explored the molecular mechanisms underlying this effect (FIG. 6).
Upregulation of VTA nAChRs. A possible mechanism for sensitization involves an upregulation of high-affinity brain nAChRs (mostly 42-containing oligomers) induced by long-term exposure to nicotine17,102,103. This upregulation is due to post-transcriptional mechanisms that include an increase in receptor subunit assembly and conformational maturation of the oligomer104,105 accompanied by structural changes17. There is some evidence that the upregulation has a role in the increased locomotor activity and brain reward responses associated with the early stages of chronic nicotine exposure98100. Blocking nAChRs in the vTA (but not in the nucleus accumbens)
www.nature.com/reviews/neuro

Figure 6 | Possible mechanisms for short- to long-term changes caused by chronic nicotine exposure. Chronic nicotine exposure alters the sensitivity of dopaminergic Nature Reviews | Neuroscience reward circuits. Several molecular mechanisms have been proposed to underlie this sensitization. a | Upregulation of highaffinity nicotinic acetylcholine receptors (nAChRs), mostly 42containing (42*) oligomers, in the ventral tegmental area (VTA) or on midbrain dopaminergic and GABA (aminobutyric acid)ergic neurons. b | Changes in cholinergic transmission (mediated by 2* and 7*nAChRs) in brain circuits at the presynaptic level. It is possible that chronic nicotine exposure causes a downregulation of 2*nAChRs (possibly following an initial upregulation), followed by an enhanced 7*nAChRmediated cholinergic transmission. The brain locus of this process is still uncertain. c | Topdown enhancement of bursting of VTA neurons. Potentiation of activity in anterior regions of the brain (infralimbic cortex (IF) and bed nucleus of the stria terminalis (BNST)) induced by chronic nicotine exposure would increase the bursting activity of dopaminergic neurons by enhancing glutamatergic input to the VTA.
396 | juNE 2010 | vOlumE 11

2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
inhibits long-term locomotor hyperactivity in mice, and moderate exposure to the same nAChR antagonists produces a transient nAChR upregulation in the vTA but not in the nucleus accumbens17,106. However, there is a complex relationship (graphically represented as an inverted u shape) between systemic nicotine concentrations and effects on lTP and behaviour 107. This relationship can be explained by the initial enhanced activity (that is, upregulation) of nAChRs by nicotine, followed by desensitization, particularly at high nicotine concentrations107. upregulation of nAChRs may therefore contribute to an early step in a sequence of molecular and cellular events that ultimately result in enhanced reward responses to nicotine following chronic nicotine exposure17,108. Chronic nicotine exposure may also affect nAChRs on both midbrain dopaminergic and GABAergic neurons, although possibly in a different manner 86. Direct whole-cell patch-clamp recordings show that passive chronic nicotine administration increases the baseline firing rate and the excitatory effect of acutely administered nicotine on GABAergic neurons in the substantia nigra pars reticulata, but decreases these measures in dopaminergic neurons86. It is possible that the upregulation of nAChRs on GABAergic neurons results in an enhanced inhibition of midbrain dopaminergic neurons102. Chronic exposure to nicotine would then result in an enhanced inhibition rather than a sensitization of midbrain dopaminergic neurons, mediated by nAChRs on GABAergic neurons (FIG. 4). The potential contribution of such a process to the effects of chronic nicotine can now be evaluated using differential expression of genes encoding nAChR subunits in dopaminergic versus GABAergic neurons74. Opposing contribution of 7* and 2*nAChR-mediated processes. A second possible mechanism of sensitization involves changes in cholinergic transmission in brain circuits at the presynaptic level. In 2/ mice, passive long-term nicotine treatment normalized the altered firing patterns of vTA dopaminergic neurons and the exploratory deficit observed in nicotine-naive 2/ mice. However, it did not cause an upregulation of heteromeric non-2*nAChRs nor of homomeric 7*nAChRs in the brain, suggesting a presynaptic effect109. moreover, chronic nicotine treatment of 2/ mice normalized the increased density of presynaptic, high-affinity choline uptake sites that is seen in treatment-naive 2/ mice, particularly in the caudate putamen109. most of these restorative effects were blocked by the 7*nAChR inhibitor methyllycaconitine (mlA), consistent with a presynaptic alteration at the level of tonic, 7-mediated cholinergic transmission. Indeed, the phenotype of wild-type mice exposed to chronic nicotine and with 7*nAChRs pharmacologically blocked was similar to that of 2/ mice. Thus, 2* and 7*nAChRs differentially contribute to the long-term changes that accompany chronic nicotine exposure109: chronic nicotine probably causes a downregulation of 2*nAChRs (possibly following an initial upregulation), resulting in an enhanced 7-mediated cholinergic transmission, the brain locus of which is still uncertain.
NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

Recent studies on the effects of long-term active nicotine self-administration (through the drinking water) in 2/ and 7/ mice confirmed this proposed mechanism54. Initially, 2/ mice, but not 7/ mice, showed decreased nicotine consumption relative to wild-type mice. However, after approximately 3 weeks of nicotine access, the 2/ mice returned to wild-type levels of nicotine consumption, whereas the 7/ mice continued to show decreased nicotine consumption54. These data suggest that, following chronic nicotine exposure in wild-type mice, there is a functional balance between opposing processes mediated by different nAChR subtypes, with the decreased contribution of 2*nAChR-mediated processes being compensated for, in the long-term, by 7-mediated synaptic processes at the level of ACh release109,110. Top-down enhanced bursting of VTA neurons. A third possible mechanism for sensitization has recently been suggested on the basis of in vivo recordings in the rat during long-term exposure to nicotine in an active self-administration paradigm101. Exposure to self-administered nicotine potentiated the activity of dopaminergic neurons in the vTA, manifested by an increase in bursting activity 101. This potentiation did not, according to the authors, result from a direct action of nicotine on the nAChRs expressed on dopaminergic neurons, but from the potentiation of anterior regions of the brain (infralimbic cortex and bed nucleus of the stria terminalis), which are involved in the volitional aspect of nicotine consumption. This potentiation would then increase, in a top-down manner (BOX 1), the bursting activity of dopaminergic neurons by enhancing glutamatergic input to the vTA111 (FIG. 4). According to the authors, this mechanism, which involves enhanced bursting activity (discussed above), can account for the effects of long-term self administration of nicotine in freely behaving rats101. In summary, the key step in nicotine addiction that results in the altered response to nicotine associated with repeated administration remains insufficiently explored, and a single well-defined mechanism is therefore lacking. methods and models26 are available to distinguish or specify the relative contributions of the longterm, direct effects of nicotine on nAChRs in the reward systems from the top-down regulation of nicotine consumption by higher brain centres.

nAChRs in nicotine withdrawal symptoms The withdrawal syndrome that accompanies smoking cessation after chronic tobacco use is one of the factors that precludes success in quitting 112. In humans, this includes somatic symptoms such as bradycardia, gastrointestinal discomfort and increased appetite accompanied by weight gain; and affective symptoms such as irritability, anxiety, depressed mood, difficulty concentrating, disrupted cognition and nicotine craving 112,113. In rodent models, rearing, jumping, shaking, abdominal constrictions, chewing, scratching, facial tremor, hyperalgesia and a change in locomotor activity are considered somatic withdrawal symptoms, whereas affective signs
vOlumE 11 | juNE 2010 | 397

REVIEWS
Medial habenula 2, (4), 5, 4 Limbic system 7 2 MHb LHb Affective withdrawal symptoms Somatic withdrawal symptoms 6, 2 2, 5, 7, 4

LC (noradrenaline) 6

Basal ganglia 7 2 VTA (dopamine) Thalamus IPN 2, 5 2

Raphe nucleus (serotonin)

Figure 7 | contribution of nAchr subunits to nicotine withdrawal symptoms at the level of the medial habenulainterpeduncular system. Subunits shown in green are implicated in somatic withdrawal symptoms; subunits shown in red are implicated in affective withdrawal symptoms. IPN, interpeduncular nucleus; MHb, medial habenula; LC, locus coeruleus; LHb, lateral habenula; VTA, ventral tegmental area.
Nature Reviews | Neuroscience

Contextual fear conditioning


A behavourial test in which an aversive stimulus is given to an animal in a conditioning chamber, such that the fear response can subsequently be elicited in the conditioning chamber in the absence of the aversive stimulus.

Conditioned place aversion


A form of classical Pavlovian conditioning in which an animal learns to avoid a compartment that was previously paired with an aversive stimulus; for example, the aversive stimulus can be the negative affective state caused by nicotine withdrawal.

include anxiety-like behaviours, elevated reward thresholds, withdrawal-induced contextual fear conditioning and withdrawal-induced conditioned place aversion113,114 (FIG. 7). The receptor-mediated mechanisms underlying nicotine withdrawal can be studied in animals in which nicotine withdrawal symptoms are elicited by nAChR antagonists115, and in mice lacking specific subunits of the nAChR113,116118. Importantly, 3, 5, 6 and 4 subunits are predominantly expressed in the habenulainterpeduncular pathways, whereas 2* and 7*nAChRs are widely expressed15 (FIG. 2). Somatic withdrawal signs caused by the nAChR antagonist mecamylamine following chronic nicotine exposure were similar in wild-type and 2/ mice113,118 but were less evident in 2 /, 5 /, 7/ and 4 / mice113,116,117,119. Affective withdrawal signs are mediated by different nAChR subunits, as wild-type mice and 5/ and 7/ mice showed withdrawal-elicited anxietylike behaviour and conditioned place aversion, but 2/ mice did not 113. Furthermore, 7/ mice showed similar changes in fear conditioning to wild-type mice undergoing nicotine withdrawal; these changes were reduced in 2/ mice120. 6*nAChRs are involved in the affectiveanxietylike withdrawal signs, but not in physical withdrawal signs53. The 6*nAChRs are located in the vTA and the locus coeruleus, where they modulate noradrenaline release121,122. The contribution of 6*nAChRs in the vTA to nicotine reward is well established, but it is possible that, in parallel, 6*nAChRs in the locus coeruleus might contribute to the regulation of nicotine withdrawal53. There is abundant expression of the 4 subunit in the medial habenula and of the 2 and 5 subunits in the interpeduncular nucleus20,119, which is connected to the habenula through the fasciculus retroflexus. In mice chronically treated with nicotine, mecamylamine caused somatic withdrawal symptoms when microinjected into the habenula or the interpeduncular nucleus, an effect that was abolished in 2/ and 5/ mice20. Also, in knock-in mice expressing the hypersensitive 4l9A nAChR, but

not wild-type mice, injection of a low dose of nicotine increased Fos expression (indicating neuronal activation) in the ventrolateral region of the medial habenula, suggesting a possible contribution to withdrawal symptoms by the 4 subunit in the habenula123. The habenula system has also been implicated in withdrawal from other drugs of addiction, including opiates, cocaine and alcohol124. Animal studies have provided evidence that the habenula has inhibitory projections to the vTA and the substantia nigra and that habenula lesions increase dopamine turnover in the nucleus accumbens and the PFC125. Activity in the habenula and interpeduncular nucleus increases during nicotine withdrawal possibly through reduced activation of nAChRs thereby controlling the activity of dopaminergic neurons in the vTA and serotonergic neurons in the dorsal raphe nucleus126, to which they project. In conclusion, the 2, 5, 7 and 4 subunits regulate the expression of somatic symptoms of withdrawal119, whereas 2 and 6 subunits contribute to affective components of the nicotine withdrawal syndrome113 (FIG. 7). The pattern and distribution of nAChR subunits involved in the nicotine withdrawal syndrome might differ from those engaged in the short-term, phasic effects of nicotine, which would offer novel targets for long-term smoking cessation therapies.

Conclusions and future directions There are substantial differences between the response of wild-type mice and mice lacking specific nAChR subunits to acute or chronic administration of nicotine. The mutant mice share features with human genetic variants, which legitimizes their use as models to investigate the molecular biology of nicotine addiction (see ReF. 127 and Supplementary information S1 (box)). Inactivation with or without re-expression of nAChR genes in the mouse5,10 has identified sets of nAChR subunits that characterize the successive steps of nicotine addiction, from the acute effects of nicotine consumption to long-term nicotine addiction. The subset of nAChR subunits engaged in the acute effects of nicotine on dopaminergic reward neurons in the vTA are being elucidated, as are those involved in nicotine withdrawal symptoms. However, the nAChR subunits engaged in the transition from the short-term to the long-term effects of nicotine exposure are still far from being understood. This is a path that requires further exploration to help develop drugs that are more effective than those currently available, which are primarily designed to target the acute effects of nicotine. Human genetic studies on nicotine dependence (see Supplementary information S1 (box)) have so far only revealed weak associations with the 2 and 7 subunit genes, which play an important part in short-term nicotine reward. By contrast, these studies consistently show an association of a 3, 5 and 4 subunit polymorphisms and 3 and 6 subunit polymorphisms with nicotine dependence128. In animal models, some of these subunits seem to be involved in nicotine withdrawal rather than in nicotine reward129; however, ongoing studies suggest a
www.nature.com/reviews/neuro

398 | juNE 2010 | vOlumE 11 2010 Macmillan Publishers Limited. All rights reserved

REVIEWS
possible contribution to reward processes as well. mouse models and the human genetics data of nicotine addiction thus point to largely unexplored targets in the design of smoking-cessation drugs. An interesting feature of the present discussion is that, in addition to the much-studied effect of nicotine on dopaminergic reward systems, there is evidence that a global gating brain circuit 82 is involved in nicotine addiction, which includes the striatum, hippocampus and amygdala under the top-down control of the PFC101,130. This is consistent with neurological and brain imaging data in humans that support the notion that the loss of control caused by nicotine addiction may correspond to a disruption of prefrontal cortex and insula
1. 2. Dome, P. et al. Smoking, nicotine and neuropsychiatric disorders. Neurosci. Biobehav. Rev. 34, 295342 (2009). Stead, L. F., Perera, R., Bullen, C., Mant, D. & Lancaster, T. Nicotine replacement therapy for smoking cessation. Cochrane Database Syst. Rev. 1, CD000146 (2008). Changeux, J. P. & Edelstein, S. J. Nicotinic Acetylcholine Receptors: From Molecular Biology to Cognition (Odile Jacob, New York, 2005). Taly, A., Corringer, P. J., Guedin, D., Lestage, P. & Changeux, J. P. Nicotinic receptors: allosteric transitions and therapeutic targets in the nervous system. Nature Rev. Drug Discov. 8, 733750 (2009). A recent review on nicotinic drugs that are in development for the treatment of brain diseases. Picciotto, M. R. et al. Abnormal avoidance learning in mice lacking functional high-affinity nicotine receptor in the brain. Nature 374, 6567 (1995). Greenbaum, L. & Lerer, B. Differential contribution of genetic variation in multiple brain nicotinic cholinergic receptors to nicotine dependence: recent progress and emerging open questions. Mol. Psychiatry 14, 912945 (2009). Orr-Urtreger, A. et al. W. Mice homozygous for the L250T mutation in the 7 nicotinic acetylcholine receptor show increased neuronal apoptosis and die within 1 day of birth. J. Neurochem. 74, 21542166 (2000). Tapper, A. R. et al. Nicotine activation of 4* receptors: sufficient for reward, tolerance, and sensitization. Science 306, 10291032 (2004). King, S. L. et al. Conditional expression in corticothalamic efferents reveals a developmental role for nicotinic acetylcholine receptors in modulation of passive avoidance behavior. J. Neurosci. 23, 38373843 (2003). Maskos, U. et al. Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature 436, 103107 (2005). An efficient method of nAChR subunit gene re-expression using lentiviral vectors. Avale, M. E. et al. Interplay of 2* nicotinic receptors and dopamine pathways in the control of spontaneous locomotion. Proc. Natl Acad. Sci. USA 105, 1599115996 (2008). Le Novere, N. et al. Involvement of 6 nicotinic receptor subunit in nicotine-elicited locomotion, demonstrated by in vivo antisense oligonucleotide infusion. Neuroreport 10, 24972501 (1999). Mameli-Engvall, M. et al. Hierarchical control of dopamine neuron-firing patterns by nicotinic receptors. Neuron 50, 911921 (2006). Maubourguet, N., Lesne, A., Changeux, J. P., Maskos, U. & Faure, P. Behavioral sequence analysis reveals a novel role for 2* nicotinic receptors in exploration. PLoS Comput. Biol. 4, e1000229 (2008). Gotti, C. et al. Structural and functional diversity of native brain neuronal nicotinic receptors. Biochem. Pharmacol. 78, 703711 (2009). Katz, B. & Thesleff, S. A study of the desensitization produced by acetylcholine at the motor end-plate. J. Physiol. 138, 6380 (1957). Govind, A. P., Vezina, P. & Green, W. N. Nicotineinduced upregulation of nicotinic receptors: underlying mechanisms and relevance to nicotine addiction. Biochem. Pharmacol. 78, 756765 (2009).

functions in a global neuronal workspace (BOX 1). This disruption would subvert attention, reasoning, planning and decision-making processes that exert inhibitory control to resist drug use21. Nicotine dependence might then be tentatively viewed as a global neuronal workspace disconnection syndrome130, which would point to novel targets that mediate nicotine action. The consequences of these studies for the understanding and treatment of diseases such as depression, Alzheimers disease and Parkinsons disease remain to be explored. Nevertheless, these mouse model experiments open a whole field of investigations, in particular in primates and humans, to understand the neuro-cognitive processes that are crucial to nicotine addiction.
35. Labarca, C. et al. Point mutant mice with hypersensitive 4 nicotinic receptors show dopaminergic deficits and increased anxiety. Proc. Natl Acad. Sci. USA 98, 27862791 (2001). 36. Cui, C. et al. The 3 nicotinic receptor subunit: a component of l-conotoxin MII-binding nicotinic acetylcholine receptors that modulate dopamine release and related behaviors. J. Neurosci. 23, 1104511053 (2003). 37. Salas, R., Cook, K. D., Bassetto, L. & De Biasi, M. The 3 and 4 nicotinic acetylcholine receptor subunits are necessary for nicotine-induced seizures and hypolocomotion in mice. Neuropharmacology 47, 401407 (2004). 38. Fernandes, C., Hoyle, E., Dempster, E., Schalkwyk, L. C. & Collier, D. A. Performance deficit of 7 nicotinic receptor knockout mice in a delayed matching-to-place task suggests a mild impairment of working/episodiclike memory. Genes Brain Behav. 5, 433440 (2006). An elegant demonstration of the contribution of the 7 nAChR subunit to working memory. 39. Paylor, R. et al. 7 nicotinic receptor subunits are not necessary for hippocampal-dependent learning or sensorimotor gating: a behavioral characterization of Acra7-deficient mice. Learn. Mem. 5, 302316 (1998). 40. McGaugh, J. L., McIntyre, C. K. & Power, A. E. Amygdala modulation of memory consolidation: interaction with other brain systems. Neurobiol. Learn. Mem. 78, 539552 (2002). 41. Young, J. W. et al. Nicotine improves sustained attention in mice: evidence for involvement of the 7 nicotinic acetylcholine receptor. Neuropsychopharmacology 29, 891900 (2004). 42. Marubio, L. M. & Paylor, R. Impaired passive avoidance learning in mice lacking central neuronal nicotinic acetylcholine receptors. Neuroscience 129, 575582 (2004). 43. Corrigall, W. Nicotine self-administration in animals as a dependence model. Nicotine Tob. Res. 1, 1120 (1999). 44. Pons, S. et al. Crucial role of 4 and 6 nicotinic acetylcholine receptor subunits from ventral tegmental area in systemic nicotine self-administration. J. Neurosci. 28, 1231812327 (2008). 45. David, V., Besson, M., Changeux, J. P., Granon, S. & Cazala, P. Reinforcing effects of nicotine microinjections into the ventral tegmental area of mice: dependence on cholinergic nicotinic and dopaminergic D1 receptors. Neuropharmacology 50, 10301040 (2006). 46. Picciotto, M. R. et al. Acetylcholine receptors containing the 2 subunit are involved in the reinforcing properties of nicotine. Nature 391, 173177 (1998). 47. Epping-Jordan, M. P., Picciotto, M. R., Changeux, J. P. & Pich, E. M. Assessment of nicotinic acetylcholine receptor subunit contributions to nicotine selfadministration in mutant mice. Psychopharmacology (Berl.) 147, 2526 (1999). 48. Frahm, S. et al. The acetylcholine receptor 4-subunit is rate limiting for nicotinic function in vivo. Society for Neuroscience Annual Meeting (Chicago, Illinois) Abstract 228.13 (2009). 49. Fowler, C. D. & Kenny, P. J. Intravenous nicotine selfadministration in wild type and 5 nicotinic acetylcholine receptor subunit knock-out mice. Society for Neuroscience Annual Meeting (Chicago, Illinois) Abstract 447.15 (2009).

3. 4.

5. 6.

7.

8. 9.

10.

11.

12.

13. 14.

15. 16. 17.

18. Koob, G. F. A role for brain stress systems in addiction. Neuron 59, 1134 (2008). 19. Davis, J. A. & Gould, T. J. Hippocampal nAChRs mediate nicotine withdrawal-related learning deficits. Eur. Neuropsychopharmacol. 19, 551561 (2009). 20. Salas, R., Sturm, R., Boulter, J. & De Biasi, M. Nicotinic receptors in the habenula-interpeduncular system are necessary for nicotine withdrawal in mice. J. Neurosci. 29, 30143018 (2009). This study provides evidence for a role of the habenulointerpeduncular system in nicotine withdrawal. 21. Naqvi, N. H. & Bechara, A. The hidden island of addiction: the insula. Trends Neurosci. 32, 5667 (2009). 22. Suarez, S. V. et al. Brain activation by short-term nicotine exposure in anesthetized wild-type and 2-nicotinic receptors knockout mice: a BOLD fMRI study. Psychopharmacology (Berl.) 202, 599610 (2009). 23. Everitt, B. J. & Robbins, T. W. Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nature Neurosci. 8, 14811489 (2005). 24. Koob, G. F. & Le Moal, M. Neurobiological mechanisms for opponent motivational processes in addiction. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, 31133123 (2008). 25. Kenny, P. J. & Markou, A. Nicotine self-administration acutely activates brain reward systems and induces a long-lasting increase in reward sensitivity. Neuropsychopharmacology 31, 12031211 (2006). 26. Gutkin, B. S., Dehaene, S. & Changeux, J. P. A neurocomputational hypothesis for nicotine addiction. Proc. Natl Acad. Sci. USA 103, 11061111 (2006). 27. Balfour, D. J. The neurobiology of tobacco dependence: a preclinical perspective on the role of the dopamine projections to the nucleus accumbens. Nicotine Tob. Res. 6, 899912 (2004). 28. Newhouse, P. A., Potter, A. & Singh, A. Effects of nicotinic stimulation on cognitive performance. Curr. Opin. Pharmacol. 4, 3646 (2004). 29. Levin, E. D. Nicotinic receptor subtypes and cognitive function. J. Neurobiol. 53, 633640 (2002). 30. Stolerman, I. P., Garcha, H. S. & Mirza, N. R. Dissociations between the locomotor stimulant and depressant effects of nicotinic agonists in rats. Psychopharmacology (Berl.) 117, 430437 (1995). 31. Poorthuis, R. B., Goriounova, N. A., Couey, J. J. & Mansvelder, H. D. Nicotinic actions on neuronal networks for cognition: general principles and longterm consequences. Biochem. Pharmacol. 78, 668676 (2009). 32. Caldarone, B. J., Duman, C. H. & Picciotto, M. R. Fear conditioning and latent inhibition in mice lacking the high affinity subclass of nicotinic acetylcholine receptors in the brain. Neuropharmacology 39, 27792784 (2000). 33. Granon, S., Faure, P. & Changeux, J. P. Executive and social behaviors under nicotinic receptor regulation. Proc. Natl Acad. Sci. USA 100, 95969601 (2003). 34. Marubio, L. M. et al. Effects of nicotine in the dopaminergic system of mice lacking the alpha4 subunit of neuronal nicotinic acetylcholine receptors. Eur. J. Neurosci. 17, 13291337 (2003).

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

vOlumE 11 | juNE 2010 | 399

REVIEWS
50. Klink, R., de Kerchove dExaerde, A., Zoli, M. & Changeux, J. P. Molecular and physiological diversity of nicotinic acetylcholine receptors in the midbrain dopaminergic nuclei. J. Neurosci. 21, 14521463 (2001). 51. Walters, C. L., Brown, S., Changeux, J. P., Martin, B. & Damaj, M. I. The 2 but not 7 subunit of the nicotinic acetylcholine receptor is required for nicotine conditioned place preference in mice. Psychopharmacology (Berl.) 184, 339344 (2006). 52. Stolerman, I. P., Chamberlain, S., Bizarro, L., Fernandes, C. & Schalkwyk, L. The role of nicotinic receptor 7 subunits in nicotine discrimination. Neuropharmacology 46, 363371 (2004). 53. Jackson, K. J., McIntosh, J. M., Brunzell, D. H., Sanjakdar, S. S. & Damaj, M. I. The role of 6-containing nicotinic acetylcholine receptors in nicotine reward and withdrawal. J. Pharmacol. Exp. Ther. 331, 547554 (2009). 54. Levin, E. D. et al. Nicotinic 7- or 2-containing receptor knockout: effects on radial-arm maze learning and long-term nicotine consumption in mice. Behav. Brain Res. 196, 207213 (2009). This study demonstrates the long-term consequences of nAChR subunit gene deletion on mouse behaviour. 55. Balfour, D. J., Benwell, M. E., Birrell, C. E., Kelly, R. J. & Al-Aloul, M. Sensitization of the mesoaccumbens dopamine response to nicotine. Pharmacol. Biochem. Behav. 59, 10211030 (1998). 56. Watkins, S. S., Koob, G. F. & Markou, A. Neural mechanisms underlying nicotine addiction: acute positive reinforcement and withdrawal. Nicotine Tob. Res. 2, 1937 (2000). 57. Overton, P. G. & Clark, D. Burst firing in midbrain dopaminergic neurons. Brain Res. Brain Res. Rev. 25, 312334 (1997). 58. Kitai, S. T., Shepard, P. D., Callaway, J. C. & Scroggs, R. Afferent modulation of dopamine neuron firing patterns. Curr. Opin. Neurobiol. 9, 690697 (1999). 59. Grace, A. A. & Bunney, B. S. The control of firing pattern in nigral dopamine neurons J. Neurosci. 4, 28662876 (1984). 60. Chergui, K., Nomikos, G. G., Mathe, J. M., Gonon, F. & Svensson, T. H. Burst stimulation of the medial forebrain bundle selectively increase Fos-like immunoreactivity in the limbic forebrain of the rat. Neuroscience 72, 141156 (1996). 61. Chergui, K. et al. Increased expression of NGFI-A mRNA in the rat striatum following burst stimulation of the medial forebrain bundle. Eur. J. Neurosci. 9, 23702382 (1997). 62. Schultz, W. Getting formal with dopamine and reward. Neuron 36, 241263, (2002). 63. Mansvelder, H. D., De Rover, M., McGehee, D. S. & Brussaard, A. B. Cholinergic modulation of dopaminergic reward areas: upstream and downstream targets of nicotine addiction. Eur. J. Pharmacol. 480, 117123 (2003). 64. Floresco, S. B., West, A. R., Ash, B., Moore, H. & Grace, A. A. Afferent modulation of dopamine neuron firing differentially regulates tonic and phasic dopamine transmission. Nature Neurosci. 6, 968973 (2003). 65. Lodge, D. J. & Grace, A. A. The hippocampus modulates dopamine neuron responsivity by regulating the intensity of phasic neuron activation. Neuropsychopharmacology 31, 13561361 (2006). 66. Faure, P. et al. Differential role for 4 and 6 nicotinic subunit in burst response to nicotine in dopamine neurons. FENS Forum of European Neuroscience (Amsterdam, the Netherlands) Abstract P14006 (2010). 67. Rice, M. E. & Cragg, S. J. Nicotine amplifies rewardrelated dopamine signals in striatum. Nature Neurosci. 7, 583584 (2004). 68. Zhang, H. & Sulzer, D. Frequency-dependent modulation of dopamine release by nicotine. Nature Neurosci. 7, 581582 (2004). 69. Zhou, F. M., Liang, Y. & Dani, J. A. Endogenous nicotinic cholinergic activity regulates dopamine release in the striatum. Nature Neurosci. 4, 12241229 (2001). 70. Exley, R., Clements, M. A., Hartung, H., McIntosh, J. M. & Cragg, S. J. 6- containing nicotinic acetylcholine receptors dominate the nicotine control of dopamine neurotransmission in nucleus accumbens. Neuropsychopharmacology 33, 21582166 (2008). 71. Cragg, S. J. Meaningful silences: how dopamine listens to the ACh pause. Trends Neurosci. 29, 125131 (2006). 72. Zhang, T. et al. Dopamine signaling differences in the nucleus accumbens and dorsal striatum exploited by nicotine. J. Neurosci. 29, 40354043 (2009). 73. Mansvelder, H. D., Keath, J. R. & McGehee, D. S. Synaptic mechanisms underlie nicotine-induced excitability of brain reward areas. Neuron 33, 905919 (2002). 74. Tolu, S. et al. A versatile system for the neuronal subtype specific expression of lentiviral vectors. FASEB J. 24, 723730 (2010). An in vivo method to distinguish the contribution of nAChRs expressed in dopaminergic versus GABAergic neurons. 75. Tolu, S. et al. A versatile system for the neuronal subtype specific expression of lentiviral vectors. FENS Forum of European Neuroscience (Amsterdam, the Netherlands). Abstract 07549 (2010). 76. Levin, E. D., McClernon, F. J. & Rezvani, A. H. Nicotinic effects on cognitive function: behavioral characterization, pharmacological specification, and anatomic localization. Psychopharmacology (Berl.) 184, 523539 (2006). 77. Parikh, V. & Sarter, M. Cholinergic mediation of attention: contributions of phasic and tonic increases in prefrontal cholinergic activity. Ann. NY Acad. Sci. 1129, 225235 (2008). 78. Orr-Urtreger, A. et al. Mice deficient in the 7 neuronal nicotinic acetylcholine receptor lack -bungarotoxin binding sites and hippocampal fast nicotinic currents. J. Neurosci. 17, 91659171 (1997). 79. Vidal, C. & Changeux, J. P. Nicotinic and muscarinic modulations of excitatory synaptic transmission in the rat prefrontal cortex in vitro. Neuroscience 56, 2332 (1993). 80. Lambe, E. K., Picciotto, M. R. & Aghajanian, G. K. Nicotine induces glutamate release from thalamocortical terminals in prefrontal cortex. Neuropsychopharmacology 28, 216225 (2003). 81. Couey, J. J. et al. Distributed network actions by nicotine increase the threshold for spike-timingdependent plasticity in prefrontal cortex. Neuron 54, 7387 (2007). 82. Grace, A. A. Gating of information flow within the limbic system and the pathophysiology of schizophrenia. Brain Res. Brain Res. Rev. 31, 330341 (2000). 83. Belujon, P. & Grace, A. A. Critical role of the prefrontal cortex in the regulation of hippocampus-accumbens information flow. J. Neurosci. 28, 97979805 (2008). 84. Placzek, A. N., Zhang, T. A. & Dani, J. A. Nicotinic mechanisms influencing synaptic plasticity in the hippocampus. Acta Pharmacol. Sin. 30, 752760 (2009). 85. Matsuyama, S., Matsumoto, A., Enomoto, T. & Nishizaki, T. Activation of nicotinic acetylcholine receptors induces long-term potentiation in vivo in the intact mouse dentate gyrus. Eur. J. Neurosci. 12, 37413747 (2000). 86. Nashmi, R. et al. Chronic nicotine cell specifically upregulates functional 4* nicotinic receptors: basis for both tolerance in midbrain and enhanced longterm potentiation in perforant path. J. Neurosci. 27, 82028218 (2007). 87. Maggi, L., Le Magueresse, C., Changeux, J. P. & Cherubini, E. Nicotine activates immature silent connections in the developing hippocampus. Proc. Natl Acad. Sci. USA 100, 20592064 (2003). 88. Le Magueresse, C., Safiulina, V., Changeux, J. P. & Cherubini, E. Nicotinic modulation of network and synaptic transmission in the immature hippocampus investigated with genetically modified mice. J. Physiol. 576, 533546 (2006). 89. Tang, J. & Dani, J. A. Dopamine enables in vivo synaptic plasticity associated with the addictive drug nicotine. Neuron 63, 673682 (2009). An in vivo electrophysiological exploration of nicotine-induced reward responses. 90. Wada, E. et al. Distribution of alpha 2, alpha 3, alpha 4, and beta 2 neuronal nicotinic receptor subunit mRNAs in the central nervous system: a hybridization histochemical study in the rat. J. Comp. Neurol. 284, 314335 (1989). 91. Ishii, K., Wong, J. K. & Sumikawa, K. Comparison of 2 nicotinic acetylcholine receptor subunit mRNA expression in the central nervous system of rats and mice. J. Comp. Neurol. 493, 241260 (2005). 92. Jia, Y., Yamazaki, Y., Nakauchi, S. & Sumikawa, K. 2 nicotine receptors function as a molecular switch to continuously excite a subset of interneurons in rat hippocampal circuits. Eur. J. Neurosci. 29, 15881603 (2009). 93. Marshall, S. P., Adhikari, A., Nason, M. W., Role, L. & Gordon, J. A. Altered theta and gamma activity profiles in the ventral hippocampal-accumbal pathway of Type III neuregulin 1 heterozygous mice. Society for Neuroscience Annual Meeting (Chicago, Illinois) Abstract 102.9 (2009). 94. Zhong, C. et al. Presynaptic type III neuregulin 1 is required for sustained enhancement of hippocampal transmission by nicotine and for axonal targeting of 7 nicotinic acetylcholine receptors. J. Neurosci. 28, 91119116 (2008). 95. Hill, J. A. Jr, Zoli, M., Bourgeois, J. & Changeux, J. P. Immunocytochemical localization of a neuronal nicotinic receptor: the 2-subunit. J. Neurosci. 13, 1551 1568 (1993). 96. Seguela, P., Wadiche, J., Dineley-Miller, K., Dani, J. A. & Patrick, J. W. Molecular cloning, functional properties, and distribution of rat brain 7: a nicotinic cation channel highly permeable to calcium. J. Neurosci. 13, 596604 (1993). 97. Jiang, L. & Role, L. W. Facilitation of cortico-amygdala synapses by nicotine: activity-dependent modulation of glutamatergic transmission. J. Neurophysiol. 99, 19881999 (2008). 98. Benwell, M. E., Balfour, D. J. & Birrell, C. E. Desensitization of the nicotine-induced mesolimbic dopamine responses during constant infusion with nicotine. Br. J. Pharmacol. 114, 454460 (1995). 99. Schoffelmeer, A. N., De Vries, T. J., Wardeh, G., van de Ven, H. W. & Vanderschuren, L. J. Psychostimulantinduced behavioral sensitization depends on nicotinic receptor activation. J. Neurosci. 22, 32693276 (2002). 100. Vezina, P. Sensitization, drug addiction and psychopathology in animals and humans. Prog. Neuropsychopharmacol. Biol. Psychiatry 31, 15531555 (2007). 101. Caille, S., Guillem, K., Cador, M., Manzoni, O. & Georges, F. Voluntary nicotine consumption triggers in vivo potentiation of cortical excitatory drives to midbrain dopaminergic neurons. J. Neurosci. 29, 1041010415 (2009). An in vivo analysis of the top-down regulation of short-term to long-term nicotine consumption in the rat. 102. Nashmi, R. & Lester, H. Cell autonomy, receptor autonomy, and thermodynamics in nicotine receptor up-regulation. Biochem. Pharmacol. 74, 11451154 (2007). 103. Marks, M. J., Burch, J. B. & Collins, A. C. Effects of chronic nicotine infusion on tolerance development and nicotinic receptors. J. Pharmacol. Exp. Ther. 226, 817825 (1983). 104. Sallette, J. et al. An extracellular protein microdomain controls up-regulation of neuronal nicotinic acetylcholine receptors by nicotine. J. Biol. Chem. 279, 1876718775 (2004). 105. Sallette, J. et al. Nicotine upregulates its own receptors through enhanced intracellular maturation. Neuron 46, 595607 (2005). 106. Baker, L. K. et al. Nicotine upregulates VTA nAChRs and requires these receptors to induce locomotor sensitization. Society for Neuroscience Annual Meeting (Washington, DC) Abstract 1027.17 (2005). 107. Picciotto, M. R. Nicotine as a modulator of behavior: beyond the inverted U. Trends Pharmacol. Sci. 24, 493499 (2003). 108. Vezina, P., McGehee, D. S. & Green, W. N. Exposure to nicotine and sensitization of nicotine-induced behaviors. Prog. Neuropsychopharmacol. Biol. Psychiatry 31, 16251638 (2007). 109. Besson, M. et al. Long-term effects of chronic nicotine exposure on brain nicotinic receptors. Proc. Natl Acad. Sci. USA 104, 81558160 (2007). 110. Koob, G. & Bloom, F. Cellular and molecular mechanisms of drug dependence. Science 242, 715723 (1988). 111. Zweifel, L. S. et al. Disruption of NMDAR-dependent burst firing by dopamine neurons provides selective assessment of phasic dopamine-dependent behavior. Proc. Natl Acad. Sci. USA 106, 72817288 (2009). 112. West, R. J., Hajek, P. & Belcher, M. Severity of withdrawal symptoms as a predictor of outcome of an attempt to quit smoking. Psychol. Med. 19, 981985 (1989). 113. Jackson, K. J., Martin, B. R., Changeux, J. P. & Damaj, M. I. Differential role of nicotinic acetylcholine receptor subunits in physical and affective nicotine withdrawal signs. J. Pharmacol. Exp. Ther. 325, 302312 (2008). This study shows the role of non-classical nAChR subunits in withdrawal symptoms.

400 | juNE 2010 | vOlumE 11 2010 Macmillan Publishers Limited. All rights reserved

www.nature.com/reviews/neuro

REVIEWS
114. Johnson, P. M., Hollander, J. A. & Kenny, P. J. Decreased brain reward function during nicotine withdrawal in C57BL6 mice: evidence from intracranial selfstimulation (ICSS) studies. Pharmacol. Biochem. Behav. 90, 409415 (2008). 115. Damaj, M. I., Meyer, E. M. & Martin, B. R. The antinociceptive effects of 7 nicotinic agonists in an acute pain model. Neuropharmacology 39, 27852791 (2000). 116. Salas, R., Pieri, F. & De Biasi, M. Decreased signs of nicotine withdrawal in mice null for the 4 nicotinic acetylcholine receptor subunit. J. Neurosci. 24, 1003510039 (2004). 117. Salas, R., Main, A., Gangitano, D. & De Biasi, M. Decreased withdrawal symptoms but normal tolerance to nicotine in mice null for the 7 nicotinic acetylcholine receptor subunit. Neuropharmacology 53, 863869 (2007). 118. Besson, M. et al. Genetic dissociation of two behaviors associated with nicotine addiction: beta-2 containing nicotinic receptors are involved in nicotine reinforcement but not in withdrawal syndrome. Psychopharmacology (Berl.) 187, 189199 (2006). 119. Salas, R. et al. Nicotine relieves anxiogenic-like behavior in mice that overexpress the read-through variant of acetylcholinesterase but not in wild-type mice. Mol. Pharmacol. 74, 16411648 (2008). 120. Portugal, G. S., Kenney, J. W. & Gould, T. J. 2 subunit containing acetylcholine receptors mediate nicotine withdrawal deficits in the acquisition of contextual fear conditioning. Neurobiol. Learn. Mem. 89, 106113 (2008). 121. Lena, C. et al. Diversity and distribution of nicotinic acetylcholine receptors in the locus ceruleus neurons. Proc. Natl Acad. Sci. USA 96, 1212612131 (1999). 122. Champtiaux, N. et al. Distribution and pharmacology of 6-containing nicotinic acetylcholine receptors analyzed with mutant mice. J. Neurosci. 22, 12081217, (2002). 123. Fonck, C. et al. Demonstration of functional 4-containing nicotinic receptors in the medial habenula. Neuropharmacology 56, 247253 (2009). 124. Maisonneuve, I. M. & Glick, S. D. Anti-addictive actions of an iboga alkaloid congener: a novel mechanism for a novel treatment. Pharmacol. Biochem. Behav. 75, 607618 (2003). 125. Matsumoto, M. & Hikosaka, O. Lateral habenula as a source of negative reward signals in dopamine neurons. Nature 447, 11111115 (2007). 126. Hikosaka, O., Sesack, S. R., Lecourtier, L. & Shepard, P. D. Habenula: crossroad between the basal ganglia and the limbic system. J. Neurosci. 28, 1182511829 (2008). 127. National Cancer Institute. Phenotypes and endophenotypes: foundations for genetic studies of nicotine use and dependence. NCI Tobacco Control Monograph 20 (eds Swan, G. E. et al.) (2009). 128. Saccone, S. F. et al. Supplementing high-density SNP microarrays for additional coverage of disease-related genes: addiction as a paradigm. PLoS ONE 4, e5225 (2009). 129. Li, M. D. & Burmeister, M. New insights into the genetics of addiction. Nature Rev. Genet. 10, 225231 (2009). 130. Changeux, J. & Dehaene, S. in Learning and Memory: A Comprehensive Reference (ed. Byrne, J.) Vol. 1, 729758 (Elsevier, Oxford, 2008). A general review about the global neuronal workspace hypothesis of consciousness. 131. Sergent, C. & Dehaene, S. Neural processes underlying conscious perception: experimental findings and a global neuronal workspace framework. J. Physiol. Paris 98, 374384 (2004). 132. Sergent, C., Baillet, S. & Dehaene, S. Timing of the brain events underlying access to consciousness during the attentional blink. Nature Neurosci. 8, 13911400 (2005). 133. Del Cul, A., Baillet, S. & Dehaene, S. Brain dynamics underlying the nonlinear threshold for access to consciousness. PLoS Biol. 5, e260 (2007). 134. Naqvi, N. H., Rudrauf, D., Damasio, H. & Bechara, A. Damage to the insula disrupts addiction to cigarette smoking. Science 315, 531534 (2007). 135. Contreras, M., Ceric, F. & Torrealba, F. Inactivation of the interoceptive insula disrupts drug craving and malaise induced by lithium. Science 318, 655658 (2007). 136. Miller, E. K. The prefrontal cortex and cognitive control. Nature Rev. Neurosci. 1, 5965 (2000). 137. Amiaz, R., Levy, D., Vainiger, D., Grunhaus, L. & Zangen, A. Repeated high-frequency transcranial magnetic stimulation over the dorsolateral prefrontal cortex reduces cigarette craving and consumption. Addiction 104, 653660 (2009). 138. Lim, K. O., Choi, S. J., Pomara, N., Wolkin, A. & Rotrosen, J. P. Reduced frontal white matter integrity in cocaine dependence: a controlled diffusion tensor imaging study. Biol. Psychiatry 51, 890895 (2002). 139. Liu, H. et al. Disrupted white matter integrity in heroin dependence: a controlled study utilizing diffusion tensor imaging. Am. J. Drug Alcohol Abuse 34, 562575 (2008). 140. Jacobsen, L. K. et al. Prenatal and adolescent exposure to tobacco smoke modulates the development of white matter microstructure. J. Neurosci. 27, 1349113498 (2007). This study demonstrates the long-term consequences of tobacco smoking on white matter organization. 141. Gil, Z., Connors, B. W. & Amitai, Y. Differential regulation of neocortical synapses by neuromodulators and activity. Neuron 19, 679686 (1997). 142. Lena, C., Changeux, J. P. & Mulle, C. Evidence for preterminal nicotinic receptors on GABAergic axons in the rat interpeduncular nucleus. J. Neurosci. 13, 26802688 (1993). 143. Kawai, H., Lazar, R. & Metherate, R. Nicotinic control of axon excitability regulates thalamocortical transmission. Nature Neurosci. 10, 11681175 (2007). 144. Hollander, J. A., Lu, Q., Cameron, M. D., Kamenecka, T. M. & Kenny, P. J. Insular hypocretin transmission regulates nicotine reward. Proc. Natl Acad. Sci. USA 105, 1948019485 (2008). 145. Paterson, N. E., Semenova, S. & Markou, A. The effects of chronic versus acute desipramine on nicotine withdrawal and nicotine self-administration in the rat. Psychopharmacology (Berl.) 198, 351362 (2008). 146. Changeux, J. P. & Taly, A. Nicotinic receptors, allosteric proteins and medicine. Trends Mol. Med. 14, 93102 (2008). 147. Balfour, D. J., Wright, A. E., Benwell, M. E. & Birrell, C. E. The putative role of extra-synaptic mesolimbic dopamine in the neurobiology of nicotine dependence. Behav. Brain. Res. 113, 7383 (2000). 148. Balfour, D. J. The neuronal pathways mediating the behavioral and addictive properties of nicotine. Handb. Exp. Pharmacol.192, 209233 (2009). 149. Di Chiara, G. Role of dopamine in the behavioural actions of nicotine related to addiction. Eur. J. Pharmacol. 393, 295314 (2000). 150. Picciotto, M. R. & Corrigall, W. A. Neuronal systems underlying behaviors related to nicotine addiction: neural circuits and molecular genetics. J. Neurosci. 22, 33383341 (2002). 151. Maskos, U. The cholinergic mesopontine tegmentum is a relatively neglected nicotinic master modulator of the dopaminergic system: relevance to drugs of abuse and pathology. Br. J. Pharmacol. 153, S438S445 (2008).

Acknowledgements

I thank I. Damaj, P. Faure, G. Koob, U. Maskos and S. Tolu for helpful comments. The work was supported by CNRS URA 2182 and by the IPSEN foundation (neuronal plasticity award).

Competing interests statement

The author declares competing financial interests: see web version for details.

FURTHER INFORMATION
Jean-Pierre Changeuxs Wikipedia page: http://en.wikipedia.org/wiki/JeanPierre_Changeux

SUPPLEMENTARY INFORMATION
See online article: S1 (box)
All liNks Are AcTiVe iN The oNliNe Pdf

NATuRE REvIEwS | NeuroscieNce 2010 Macmillan Publishers Limited. All rights reserved

vOlumE 11 | juNE 2010 | 401

Potrebbero piacerti anche