Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Oct 15.
Published in final edited form as: Biochem Pharmacol. 2011 Jul 20;82(8):800–807. doi: 10.1016/j.bcp.2011.07.067

Naturally-expressed nicotinic acetylcholine receptor subtypes

Jie Wu 1, Ronald J Lukas 2
PMCID: PMC3163308  NIHMSID: NIHMS316579  PMID: 21787755

Abstract

Nicotinic acetylcholine receptors (nAChRs) warrant attention, as they play many critical roles in brain and body function and have been implicated in a number of neurological and psychiatric disorders, including nicotine dependence. nAChRs are composed as diverse subtypes containing specific combinations of genetically-distinct subunits and that have different functional properties, distributions, and pharmacological profiles. There had been confidence that the rules that define ranges of assembly partners for specific subunits were well-established, especially for the more prominent nAChR subtypes. However, we review here some newer findings indicating that nAChRs having largely the same, major subunits exist as isoforms with unexpectedly different properties. Moreover, we also summarize our own studies indicating that novel nAChR subtypes exist and/or have distributions not heretofore described. Importantly, the nAChRs that exist as new isoforms or subtypes or have interesting distributions require alteration in thinking about their roles in health and disease.

Keywords: nicotine, nicotinic receptor, acetylcholine, Alzheimer’s disease, drug dependence

1. INTRODUCTION

nAChRs are prototypical members of the ligand-gated ion channel superfamily of neurotransmitter receptors. nAChRs represent both classic and contemporary models for the establishment of concepts pertaining to mechanisms of drug action, synaptic transmission, and structural/functional diversity of transmembrane signaling molecules (see reviews [17]). nAChRs are found throughout the nervous system (e.g., in muscle, autonomic and sensory ganglia, and the CNS). They are very important, because they play many critical roles in brain and body function, making it logical that nAChRs also are implicated in a number of neurological and psychiatric disorders, as they are in nicotine dependence. nAChRs exist as multiple, diverse subtypes composed as pentamers of unique combinations from a family of at least seventeen (α1–α10, β1–β4, γ, δ,ε) similar, but genetically-distinct, subunits. nAChR subtypes are named according to their known subunit composition (using an “*” to indicate possible additional assembly partners; [6]). Each subunit gene has a unique promoter, even though some are collected in a cluster, suggesting a means for cell-specific expression. There also are unique protein sequence elements for each, especially in the large, cytoplasmic loop, suggesting means for differential post-translational control of subunit trafficking. There is evidence for specificity of targeting of nAChR subunit proteins and the relevant nAChR assemblies to sub- or peri-synaptic destinations in somatodendritic domains, but also down axons to pre-terminal or synaptic terminal locations. Although many nAChR subtypes are possible in theory, there seem to be some rules that define and limit the number of viable subunit combinations. Most of these nAChR subtypes appear to exist as heteropentamers containing two or more different kinds of subunit. For example, heterologous expression studies suggest that α2, α3, α4, or α6 subunits can combine in binary fashion with β2 or β4 subunits to form ligand-binding and/or functional nAChRs (e.g., α4β2-nAChRs). β3 and α5 subunits are “wild-cards” not able to form nAChRs alone or with any other single type of subunit. However, they are capable of integrating into complexes with two other subunit types found in binary complexes to form distinctive, trinary complexes (such as α4β2α5-or α3β4α5-nAChRs (found naturally expressed). They also can contribute to formation of quaternary complexes that contain more than one of the α2-4 or α6 subunits or that contain both β2 or β4 subunits (for example, α4α6β2β3- or α3β2β4α5-nAChRs). In addition, mammalian muscle-type nAChRs are quaternary complexes composed of α1, β1, δ and either γ (fetal) or ε (adult) subunits. By contrast, phylogentically ancient nAChR α7 subunits are able to form functional homopentamers, the simplest possible prototype for a ligand-gated ion channel. Although nAChR α9 subunits also are able to form functional homomers with modest channel activity, function is markedly enhanced when they and α10 subunits co-assemble to form a novel binary complex [8] (note that these subunits and the unusual nAChRs they constitute are not substantially expressed in the brain). nAChRs containing α7 subunits (α7-nAChRs) are the most abundant curaremimetic neurotoxin-binding nAChRs in the brain. nAChRs containing α4 and β2 subunits (α4β2*-nAChRs) are the most abundant high affinity nicotine-binding nAChRs in the brain. However, other, less abundant nAChRs (e.g., α3*-nAChRs, α6*-nAChRs) must exist and may play important physiological roles.

Nevertheless, the field has been somewhat altered by realization that the lack of two-fold symmetry in pentameric assemblies allows for more diversity across nAChR subtypes than heretofore realized. More recent work has indicated that even for α4β2*-nAChR, having two α4β2 subunit cassettes thought to provide an α4:β2 subunit interface where nicotinic agonists bind to gate channel opening, there exist unique isoforms that have different subunits occupying the “fifth” or “accessory” position in the pentamer [912]. Remarkably, the pharmacological properties of these isoforms can be quite different. For example, α4β2*-nAChR having 2 α4 subunits and 3 β2 subunits [(α4)2(β2)3-nAChR; i.e., having a β2 subunit in the “fifth” or “accessory” position] have higher sensitivity for many nicotinic agonists than “low sensitivity” (α4)3(β2)2-nAChR having an α4 subunit in the accessory position. Moreover, wild-card subunits α5 or β3 can occupy the fifth position, creating (α4)2(β2)2α5- or (α4)2(β2)2β3-nAChR having yet again distinctive pharmacological character. It is likely that further diversity exists in other complexes that contain, for example, α4 and α6 subunits. The characterization of these isoforms presents new and larger challenges than before. Although the physiological implications of this unexpectedly broader diversity are currently poorly understood, they are bound to influence our understanding of phenomena such as nicotine dependence as well as strategies for translation of nAChR drug discovery to treatment of neuropsychiatric disorders.

Functionally, nAChRs in the brain play roles not only in the mediation of classical, excitatory, cholinergic neurotransmission at selected loci, but also and perhaps more globally in the modulation of neurotransmission by other chemical messengers, including glutamate, GABA, the monoamines dopamine, norepinephrine and serotonin, and acetylcholine (ACh) itself [25, 1317]. This means that some nAChR subtypes have postsynaptic (or peri-synaptic), somatodendritic localizations, whereas others have pre-synaptic dispositions (i.e., on neuronal terminals). However, care should be exercised in calling some nAChRs according to their disposition in synaptic space. Indeed, so called “pre-synaptic” nAChRs that reside on nerve terminals and that perhaps locally modulate neurotransmitter release might actually be called “post-synaptic” if they lie under cholinergic nerve endings. It probably is wise to speak of nAChRs with respect to their location on soma, dendrites, nerve terminals, or even processes slightly distal to nerve terminals. Moreover, some nAChRs have been implicated in processes such as the structuring and maintenance of neurites and synapses [1820] and even in modulation of neuronal viability/death [2124]. Thus, nAChR subtypes in the brain play complex and interesting roles in modulation of the chemical milieu of the brain, in completion of neuronal circuits, and perhaps in development and architecture of synapses. In this review, we summarize some of the recent progress in studies of naturally-expressed nAChR subtypes in the brain and their function, and we highlight just some of the possible roles for nAChRs in diseases.

2. DISCUSSION

2.1 α4*-nAChRs in the brain

nAChRs that bind radiolabeled nicotine with the highest affinity contain α4 subunits (α4*-nAChR; see reviews and/or tables by [17]. Immunoassays have shown that the predominant, naturally expressed form of α4*-nAChRs in the vertebrate brain contains α4 and β2 subunits (α4β2-nAChRs) [25, 26]. α4β2-nAChRs have been implicated in nicotine self-administration, reward, and dependence, and in diseases such as Alzheimer's and epilepsy [15, 2733]. α4 subunits can also assemble with β4 subunits to form α4β4-nAChRs that have comparably high nicotine affinity [3436]. β4 subunit mRNA colocalizes with α4 subunit mRNA in many brain regions [3738] that could be involved in complex behaviors including nicotine dependence. Moreover, in addition to existence in "binary" nAChR complexes containing two types of subunits, α4 subunits can have more than one type of assembly partner [3956]. For example, nAChRs containing α4, β2, and α5 subunits are expressed naturally, and heterologously-expressed α4β2α5-nAChRs have interesting properties distinct from those of simpler α4β2-nAChR. To the first approximation, properties of heterologously expressed α4*-nAChRs and naturally-expressed α4*-nAChRs of the same composition (as best can be ascertained) are very similar (op. cit.). However, given their evident physiological importance and their potential to form as diverse combinations of subunits, including α4β2α5- or α4β2β3-nAChR isoforms containing the indicated subunits in the accessory position, more work on heterologously expressed α4*-nAChRs is warranted, as are careful comparisons of properties of these α4*-nAChRs of defined subunit composition(s) with properties of naturally expressed α4*-nAChRs in vertebrate brain neurons.

Considerable attention has been given to effects of chronic nicotine exposure on nAChRs because of relevance to habitual use of tobacco products [57] but information about these effects remains incomplete. Chronic nicotine exposure for periods of days induces increases in numbers of nAChR-like radioligand binding sites (or, in some studies, subunit polypetptides) in human or non-human animal brain or in cells from a variety of primary or continuous culture systems [25, 26, 5862]. Although these binding sites can be in intracellular pools of possible assembly intermediates and not always on functional, cell surface nAChRs, the magnitude of their upregulation varies considerably across experiments and experimental systems, and concentration-response studies show that nAChR subtypes most sensitive to nicotine-induced upregulation include α4 subunits [26, 61, 6364]. Upregulation seems to reflect nicotine-mediated, “chaperone-like” facilitation of α4 and β2 subunit assembly that is manifest as stabilization of α4β2-nAChR assembly intermediates in a state or states detectable by radioagonist binding assays [64]. In terms of functional effects, nicotine acts acutely much in the way that ACh does, causing opening of nAChR channels. However, many studies indicate that nicotine exposure for seconds-minutes-hours leads to losses in nAChR function. Again, there is great diversity in experimental systems used and in magnitudes, kinetics, and dose-dependence of effects, partly reflecting variations in techniques and diversity in experimental systems used [26, 61, 6568]. There seems to be a consensus that there is at least one stage of reversible functional loss on shorter-term (seconds-minutes) nicotine exposure called "desensitization," and that a more slowly reversible (or irreversible?) loss of nAChR function occurs on longer-term (tens of minutes-hours) nicotine exposure via a process called "persistent inactivation" by some [61, 6869]. Desensitization or persistent inactivation of a particular nAChR subtype occur at much lower concentrations of nicotine and for shorter exposure times than necessary to induce upregulation of binding sites, and acute function or persistent inactivation can be blocked under conditions that allow upregulation to occur, suggesting distinction between these processes [61]. There seemed to be a consensus that persistent inactivation of α4β2-nAChR function also occurs on chronic nicotine exposure not only in the brain but also in a variety of heterologous expression systems and for α4β2-nAChRs from various species including humans [7073]. Perhaps importantly, most studies examining effects of chronic nicotine exposure in vivo or in vitro typically use 10 days or shorter exposure. Because human smokers are exposed to nicotine for months-years, laboratory work to date might be providing a misleading view of effects on nAChR expression related to smoking behavior.

2.2 α7*-nAChRs in the brain

Another prominent nAChR subtype found in vertebrate central and autonomic nervous systems contains α7 subunits (α7-nAChRs). These sites have been known to exist for many years based on their ability to bind the curaremimetic neurotoxin, α-bungarotoxin (Bgt) [7478]. They long have been known to exhibit many of the biochemical and pharmacological features of true nAChRs, to have brain distributions sub- or peri-synaptic to cholinergic terminals, to have levels of expression sensitive to chronic nicotine exposure and/or modification of cholinergic inputs, and to reveal hints of functional significance in electrophysiological studies. However, their physiological relevance was elusive, and their functional study was confounded until heterologous expression studies of α7-nAChR composed as homomers revealed unusually rapid, agonist-induced, calcium ion-permeable channel opening and inactivation [7983]. Subsequently, renewed searches for functions of natural Bgt-binding nAChRs uncovered short-lived, nicotine-gated, toxin-sensitive, inward currents and/or elevations of intracellular Ca2+ in chick autonomic neurons [84], in human ganglionic neuron-like clonal cells [85], or in rat CNS neurons [16, 4044, 8691]. Many of the cell types naturally expressing Bgt-sensitive, functional nAChRs have been shown to express α7 genes as well as some native form of Bgt-binding nAChRs [68, 9293]. Knock out of the α7 gene leads to absence of Bgt-binding nAChRs in cell lines or in mice [85, 94]. Thus, correlations have been drawn between the expression of a major form of Bgt-binding nAChRs and α7 gene products.

By virtue of their unique subcellular localizations, channel kinetics and Ca2+ permeability, α7-nAChRs appear to have novel functional roles in addition to (i.e., distinct from) the mediation of classical excitatory neurotransmission. For example, Bgt-sensitive α7-nAChRs have been implicated in processes such as vicinal control of neurotransmitter release [7, 14], development and maintenance of neurites and synapses [1820], long-term potentiation [9596], seizures [97], and neuronal viability/death [2124]. These intriguing findings underscore the need for further characterization of functional α7-nAChRs.

Because of their homomeric nature in heterologous expression systems, α7-nAChRs and related chimeric receptors have proven to be very useful in mutagenesis-based studies that have provisionally identified structures and residues contributing to ligand binding, subunit interactions, and/or lining the ligand-gated ion channel [98104]. These α7-nAChR homomers, perhaps reflecting the phylogentically ancient status of α7 subunits, represent one of the simplest possible prototypes for elucidation of nAChR structure and function, and they are ideally suited for studies of nAChR structure-function relationships. Bgt-sensitive, functional or ligand binding activities of heterologously-expressed, homomeric α7-nAChRs are similar to those of native Bgt-binding nAChRs in cells or tissues that express α7 subunits [85, 105107], suggesting the possibility that many if not all nAChRs containing α7 subunits exist as homomers. Some studies of native α7-nAChRs also suggest that they exist as homomers [25, 108109].

However, some nAChRs in neurons have been postulated to contain α7 plus other kinds of subunit [5, 13, 110112]. For instance, heterologous expression work has indicated that nAChR α7 and β2 subunits can assemble together to form heteromeric, functional channels [113], and histological studies have shown that there is co-expression of nAChR α7 and β2 subunits in most forebrain cholinergic neurons [110]. We discovered a pharmacologically-unusual nAChR subtype that is naturally expressed in the rat basal forebrain, and we obtained evidence that this subtype is a heteropentameric α7β2-nAChR that is absent in nAChR β2 subunit knock-out mice [91]. Interestingly, these α7β2-nAChRs exhibit high sensitivity to pathological levels of β-amyloid [91]. These data suggest that whereas most α7-nAChRs are formed as homomeric pentamers, others may exist as heteromers, including possible α7β2-nAChR pentamers (Fig. 1). The fact that oligomeric amyloid-β seems to have high affinity for α7β2-nAChRs expressed in the brain region that is damaged early in Alzheimer’s disease may have implications for disease etiopathogenesis.

Fig. 1.

Fig. 1

Schematic illustrations are shown as viewed from synaptic space of selected nicotinic acetylcholine receptor (nAChR) subtypes, all of which are assembled as pentamers of subunits forming a rosette with a central pore, which is the ion channel. In the upper part of the figure are shown the presumed structures of: left - a homomeric α7-nAChR composed only of α7 subunits (blue) as found, for example, in the ventral tegmental area; right - a heteromeric α7β2-nAChR composed of α7 subunits (blue) and β2 subunits (gray) as found, for example, on basal forebrain cholinergic neurons except in mice lacking β2 subunits [91]. The heteromeric, presumably α7β2-nAChR subtype displays elevated sensitivity to functional blockade by β-amyloid or dihydro-β-erythrodine than seen for homomeric α7-nAChRs [91]. In the lower part of the figure are shown isoforms of α4β2-nAChRs having a 3:2 (left) or a 2:3 (right) ratio of α4:β2 subunits (α4 subunits, lavender). (α4)3(β2)2-nAChRs have comparatively low sensitivity to nicotinic agonists when compared to “high sensitivity,” (α4)2(β2)3-nAChRs [912]. Studies based on heterozygotic mice with lower gene doses for nAChR α4 subunits indicate that proportions of functional expression of high sensitivity (α4)2(β2)3-nAChRs are decreased and that the opposite occurs in nAChR subunit β2+/− heterozygotic mice, suggesting that natural expression of these two isoforms occurs, could be regulated, and has physiological relevance [143].

2.3 α6*-nAChRs in the brain

α6 subunits are not widely expressed in the brain, but they are prevalent in midbrain dopaminergic (DAergic) regions associated with pleasure, reward, and mood control (Fig. 2) [114115], suggesting that α6*-nAChRs play critical roles in nicotine dependence and in the ability to modulate mood and emotion attributed to nicotine [116]. Until a report from Lindstrom’s laboratory [117] of heterologous expression in oocytes of nAChRs containing α6 subunits, these subunits were of “orphan” status. Kuryatov et al. [118] found functional expression of human α6 plus β4 plus β3 subunits in an oocyte system, but functional expression in that system was poor for other combinations and was poor for responses to nicotine relative to those for ACh. More recent results suggest success in using alternative strategies to express α6*-like nAChRs [119120]. Using transgenic mice expressing gain-of-function nAChR α6 subunits, the presence of functional α6*-nAChRs on ventral tegmental area (VTA) DAergic neurons has been reported [121]. Recently, we reported a novel discovery that functional α6*-nAChRs are located on GABAergic presynaptic boutons associated with VTA DAergic neurons, where these α6*-nAChRs mediate nicotinic modulation of GABA release onto those DAergic neurons [122]. This adds to literature indicative of many roles for nAChR in direct and indirect modulation of DAergic neuronal function.

Fig. 2.

Fig. 2

Diagram indicating some of the complexity in nicotinic acetylcholine receptor (nAChR) expression across cell types and at different locations relative to the synapse, using the ventral tegmental area (VTA) as an example. It has been known for some time that α7-nAChRs (turquoise) are present on glutamatergic nerve terminals (Glu, red) synapsing on dopaminergic (DA) cell bodies (and perhaps dendrites; DA, orange triangle) in the VTA [25]. Moreover, different nAChRs, presumably responding to acetylcholine released from cholinergic nerve terminals (purple) have a variety of distributions on DA neurons in the VTA. In fact, it recently became clear that functional nAChR phenotypes allow “binning” of VTA DA type ID, type IID and type IIID neuronal subsets predominantly expressing, respectively, α4β2-nAChRs (lavender), α7-nAChRs (turquoise), or β4*-nAChRs (blue) on soma and/or dendrites [90]. This means that there isn’t just one kind of VTA DA neuron, at least with respect to nAChR expression profiles. Whether these VTA DA functional nAChR phenotypes change with nicotine exposure and/or are on VTA VDA neurons that also receive distinctive inputs or have different targets remains to be determined. Moreover, newly appreciated is the presence of α6*-nAChRs (green) not only on VTA DA nerve terminals ending in the NAc and perhaps on VTA soma (not shown in this illustration) [121, 130131], but also on GABAergic boutons (black) associated with VTA DA neuronal soma and/or proximal dendrites [122]. In addition, on GABAergic terminals, pre-terminals and/or soma, there are α4β2-nAChRs [144]. The balance between activation and desensitization or persistent inactivation of different nAChR subtypes on VTA nerve terminals ending on DA neurons and on those neurons and their own terminals likely dictates VTA DA neuronal activity relevant to nicotine dependence and perhaps to cholinergic control of dependence on other agents as well as more generally in mood and reward.

The development of α-conotoxins (α-Ctx) as tools for investigating neuronal nAChRs [123] has provided much-needed assistance to α6*-nAChR investigations. Initially, α-CtxMII was thought to be highly selective for α3β2*-nAChRs [124] and became a useful tool for investigating receptor structure [125]. Based on sensitivity to α-CtxMII, pre-synaptic nAChRs on striatal DAergic terminals were divided into two classes: "α-CtxMII-sensitive" and "α-CtxMII-resistant" [126127]. Because α-CtxMII was expected to block any nAChRs containing the α3β2 subunit interface, the α-CtxMII-sensitive receptors were identified as α3β2*-nAChRs, but in α3 knockout mice, the majority of 125I-α-CtxMII binding (including in terminal regions of VTA/substantia nigra (SN) DAergic projections) is not eliminated [128], suggesting that non-α3*-nAChRs account for most α-CtxMII-sensitive nAChRs. The α6 subunit represented a likely alternate component of 125I-α-CtxMII binding sites since it is closely related to the α3 subunit, and residues in α3 subunits found to be critical for α-CtxMII sensitivity [125] are conserved in α6 subunits.

Other "α3β2-specific" antagonists, such as neuronal- or kappa-bungarotoxin and α-CtxPnIA, have also turned out to block α6*-nAChRs [129]. The sensitivity of α6*-nAChRs to α-CtxMII and the colocalization of α6 and β3 with β2 subunits in VTA DAergic neurons further suggested that the α-CtxMII-sensitive component of nicotine-mediated DA release might contain these subunits. Indeed, striatal 125I-α-CtxMII binding was eliminated upon α6 subunit gene deletion [130], and a β3 subunit-null mutation drastically reduced (although it did not abolish) expression of these receptors [131]. Thus, α-CtxMII appears to be a useful tool for recognizing naturally-expressed α6*-nAChRs, but one limitation is that this compound poorly distinguishes between α6β2*- and α3β2*-nAChR subtypes. The more selective antagonist of α6*-nAChR, α-CtxPIA [132134] exhibits 75-fold higher affinity for rat α6β2*-nAChRs than for rat α3β2*-nAChRs [134].

There is an emerging consensus that nAChR α6 subunit mRNA and proteins are distributed in brain regions thought to be involved in reward and drug reinforcement, in theory being involved in DA release [135]. This makes the hypothesis particularly intriguing that α6*-nAChRs play roles in nicotine dependence. These same brain regions in which α6 messages and proteins have been found also are implicated in attention, mood control, and neurological or psychiatric disorders, including anxiety, depression, and Parkinson’s disease [116]. Other studies from our laboratory [136137] and from others [138139] suggest interactions between antidepressants and some nAChR subtypes, but specific studies examining α6*-nAChRs have not yet been performed. The high incidence of tobacco use by mentally ill individuals, coupled with reports of the mood-stabilizing effects of nicotine in tobacco users, implicate nAChRs in the control of mood and emotion, and α6*-nAChRs are candidates for mediation of these effects [116].

There is good evidence that α6*-nAChRs, in particular, modulate neurotransmitter release in multiple brain regions. Based on studies using α-CtxMII in combination with nAChR subunit-null mutant mice, Salminen et al. [140] determined that striatal pre-synaptic nAChRs mediating DA release contain one (α4α6β2β3 subtype) or two (α6β2β3 subtype) α-CtxMII-sensitive ACh binding sites indicating that pre-synaptic α6*-nAChRs contribute to nicotinic modulation of DA release in the striatum. These subunit assignments were confirmed using immunochemical approaches by Gotti et al. [131]. Also, Endo et al. [141] found naturally-expressed α3β2- and α6β2-nAChRs on superior colliculus neurons, and these receptors are likely located on pre-synaptic terminals of GABAergic neurons where they modulate GABA release. Interestingly, Champtiaux et al. [130] demonstrated that the majority of DAergic cell-body nAChR function in the VTA is mediated by non-α6*-nAChRs. Most of the α6*-nAChRs synthesized in VTA DAergic cell bodies appear to be transported to cell terminal regions (such as striatum and nucleus accumbens) [131]. However, our recent findings indicate that there also are functional nAChRs located on GABAergic, pre-synaptic boutons associated with these VTA DAergic cell bodies (Fig. 2) [122]. Agonist activation of these nAChRs results in increased inhibitory post synaptic currents (IPSCs) measured at DAergic cell bodies. Our data indicate that these receptors are predominantly of the α6β2* subtype, because treatment with α6β2*-selective nAChR antagonists (either α-CtxMII or α-CtxPIA; 1 nM) or deletion of the nAChR β2 subunit abolishes nicotine-induced increases in inhibitory post-synaptic currents [122].

2.4 Studies of heterologously expressed nAChRs

Given the complexity of subunit composition (types and stoichiometries) in nAChR subtypes, it might be expected that heterologous expression in oocytes or cell lines might be a boon. However, although there have been some successes (see above and [2, 5, 142], it has not been easy to develop cell lines that stably express specific nAChR subtypes. Moreover, when expression is driven by introduction of loose subunits, each encoded by its own vector, investigators are at the mercy of cell with regard to the mixture of closed assemblies. There is some success and significant promise in work using concatenated subunits to allow investigators to control and specify subunit stoichiometries and arrangements, but we will likely benefit from review of that work in the near future rather than here, allowing the field to further develop.

2.5 Conclusions

It is clear that naturally-expressed nAChR diversity is broader than heretofore suspected, bringing new challenges to receptor characterization. However, great opportunities and new ideas with profound translational implications come from this. For example, the realization that distinctive pharmacological properties in native nAChRs in regions where α4 and β2 are expressed could be due to expression as α4 or β2 or other isoforms of α4β2-nAChRs (Fig. 1) [143] could explain puzzling, earlier findings. Perhaps more importantly, consideration needs to be given to how high sensitivity and low sensitivity α4β2-nAChRs in humans would respond to nicotine at pharmacological concentrations attained in smokers, interstitial fluid acetylcholine that is present at concentrations that rival those of nicotine in smokers, or synaptic acetylcholine that is one hundred times more concentrated or greater. The finding that the brain region affected early in Alzheimer’s disease contains a unique, heteropentameric α7β2-nAChRs highly sensitive to β-amyloid [91] might alter thinking about roles for nAChRs in that disease. Discovery of α6*-nAChRs on GABA terminals associated with VTA DAergic neuronal soma [122] brings a new dimension to our understanding of nicotine’s effects and nAChR’s roles in drug dependence. It is likely that additional findings will also emerge, contributing to further evolution in our understanding about nAChRs and their physiological, pathological and pharmacological roles.

TABLE 1.

Different properties and locations of homomeric and heteromeric α7*-nAChR. Studies of acutely dissociated, dopaminergic neurons from the ventral tegmental area (VTA) indicate their expression of homomeric α7-nAChR, whereas studies using similar techniques and cholinergic neurons from the ventral aspect of the nucleus of the diagonal band (VDB) are consistent with expression there of heteromeric α7β2-nAChR that are absent in nAChR β2 subunit knock-out mice [9091]. Some of the properties distinguishing these α7- and α7β2-nAChR are summarized here.

α7β2 nAChRs in VDB α7 nAChRs in VTA
Choline response
Rising time (ms) 72.1 ± 9.1 29.1 ± 2.9 (n=12)***
Decay time (ms) 28.6 ± 2.8 10.2 ± 1.5 (n=12)***
IC50 (nM)
Methyllycaconitine 0.7 0.4
Dihydro-β-erythroidine 200 >100,000
1–42 inhibition (%)
1 nM 35 ± 8 7 ± 3**
10 nM 70 ± 9 10 ± 10**

Note:

**

p<0.01,

***

p<0.001 for comparisons between α7β2- and α7-nAChRs.

ACKNOWLEDGEMENTS

The authors acknowledge support over many years from several sources, including National Institutes of Health grants R21 DA030045, U19 DA019377, R21 DA027070, U19 MH085193 UIC sazetidine, R21 DA026627 and R01 DA015389. Other effort was supported by the Barrow Neurological Foundation, by the Arizona Biomedical Research Commission (9730 and 9615), and by Philip Morris USA Inc. and Philip Morris International. The contents of this report are solely the responsibility of the authors and do not necessarily represent the views of the National Institutes of Health or any of the aforementioned awarding agencies.

Non-standard abbreviations

nAChR(s)

nicotinic acetylcholine receptor(s)

VTA

ventral tegmental area

DAergic

dopaminergic

DA

dopamine

SN

substantia nigra

ACh

acetylcholine

IPSC(s)

inhibitory post synaptic current(s)

Bgt

α-bungarotoxin

αCtx

α-conotoxin

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.

REFERENCES

  • 1.Clementi F, Fornasari D, Gotti C. Neuronal nicotinic receptors, important new players in brain function. European Journal of Pharmacology. 2000;393:3–10. doi: 10.1016/s0014-2999(00)00066-2. [DOI] [PubMed] [Google Scholar]
  • 2.Jensen AA, Frolund B, Lijefors T, Krogsgaard-Larsen P. Neuronal nicotinic acetylcholine receptors: Structural revelations, target identifications, and therapeutic inspirations. Journal of Medicinal Chemistry. 2005;48:4705–4745. doi: 10.1021/jm040219e. [DOI] [PubMed] [Google Scholar]
  • 3.Lindstrom J. Neuronal nicotinic acetylcholine receptors. In: Narahashi T, editor. Ion Channels. New York: Plenum; 1996. pp. 377–450. [DOI] [PubMed] [Google Scholar]
  • 4.Lindstrom JM. Nicotinic acetylcholine receptors of muscles and nerves - Comparison of their structures, functional roles, and vulnerability to pathology. Myasthenia Gravis and Related Disorders. 2003:41–52. doi: 10.1196/annals.1254.007. [DOI] [PubMed] [Google Scholar]
  • 5.Lukas RJ, Bencherif M. Recent developments in nicotinic acetylcholine receptor biology. In: Arias HR, editor. Biological and Biophysical Aspects of Ligand-Gated Ion Channel Receptor Superfamilies. Kerala, India: Research Signpost; 2006. pp. 27–59. [Google Scholar]
  • 6.Lukas RJ, Changeux JP, Le Novere N, Albuquerque EX, Balfour DJ, Berg DK, et al. International Union of Pharmacology. XX. Current status of the nomenclature for nicotinic acetylcholine receptors and their subunits. Pharmacological Reviews. 1999;51:397–401. [PubMed] [Google Scholar]
  • 7.Albuquerque EX, Pereira EF, Mike A, Eisenberg HM, Maelicke A, Alkondon M. Neuronal nicotinic receptors in synaptic functions in humans and rats: physiological and clinical relevance. Behav Brain Res. 2000;113:131–141. doi: 10.1016/s0166-4328(00)00208-4. [DOI] [PubMed] [Google Scholar]
  • 8.Elgoyhen AB, Vetter DE, Katz E, Rothlin CV, Heinemann SF, Boulter J. alpha10: a determinant of nicotinic cholinergic receptor function in mammalian vestibular and cochlear mechanosensory hair cells. Proc Natl Acad Sci U S A. 2001;98:3501–3506. doi: 10.1073/pnas.051622798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Carbone AL, Moroni M, Groot-Kormelink PJ, Bermudez I. Pentameric concatenated (alpha 4)(2)(beta 2)(3) and (alpha 4)(3)(beta 2)(2) nicotinic acetylcholine receptors: subunit arrangement determines functional expression. British Journal of Pharmacology. 2009;156:970–981. doi: 10.1111/j.1476-5381.2008.00104.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Moroni M, Vijayan R, Carbone A, Zwart R, Biggin PC, Bermudez I. Non-agonist-binding subunit interfaces confer distinct functional signatures to the alternate stoichiometries of the alpha 4 beta 2 nicotinic receptor: An alpha 4-alpha 4 interface is required for Zn2+ potentiation. Journal of Neuroscience. 2008;28:6884–6894. doi: 10.1523/JNEUROSCI.1228-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Nelson ME, Kuryatov A, Choi CH, Zhou Y, Lindstrom J. Alternate stoichiometries of alpha 4 beta 2 nicotinic acetylcholine receptors. Molecular Pharmacology. 2003;63:332–341. doi: 10.1124/mol.63.2.332. [DOI] [PubMed] [Google Scholar]
  • 12.Zwart R, Carbone AL, Moroni M, Bermudez I, Mogg AJ, Folly EA, et al. Sazetidine-A is a potent and selective agonist at native and recombinant alpha 4 beta 2 nicotinic acetylcholine receptors. Molecular Pharmacology. 2008;73:1838–1843. doi: 10.1124/mol.108.045104. [DOI] [PubMed] [Google Scholar]
  • 13.McGehee DS, Role LW. Physiological diversity of nicotinic acetylcholine receptors expressed by vertebrate neurons. Annu Rev Physiol. 1995;57:521–546. doi: 10.1146/annurev.ph.57.030195.002513. [DOI] [PubMed] [Google Scholar]
  • 14.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]
  • 15.McGehee DS. Molecular diversity of neuronal nicotinic acetylcholine receptors. Ann N Y Acad Sci. 1999;868:565–577. doi: 10.1111/j.1749-6632.1999.tb11330.x. [DOI] [PubMed] [Google Scholar]
  • 16.Alkondon M, Pereira EF, Cortes WS, Maelicke A, Albuquerque EX. Choline is a selective agonist of alpha7 nicotinic acetylcholine receptors in the rat brain neurons. Eur J Neurosci. 1997;9:2734–2742. doi: 10.1111/j.1460-9568.1997.tb01702.x. [DOI] [PubMed] [Google Scholar]
  • 17.Schicker KW, Dorostkar MM, Boehm S. Modulation of transmitter release via presynaptic ligand-gated ion channels. Curr Mol Pharmacol. 2008;1:106–129. doi: 10.2174/1874467210801020106. [DOI] [PubMed] [Google Scholar]
  • 18.Freeman JA. Possible regulatory function of acetylcholine receptor in maintenance of retinotectal synapses. Nature. 1977;269:218–222. doi: 10.1038/269218a0. [DOI] [PubMed] [Google Scholar]
  • 19.Chan J, Quik M. A role for the nicotinic alpha-bungarotoxin receptor in neurite outgrowth in PC12 cells. Neuroscience. 1993;56:441–451. doi: 10.1016/0306-4522(93)90344-f. [DOI] [PubMed] [Google Scholar]
  • 20.Pugh PC, Berg DK. Neuronal acetylcholine receptors that bind alpha-bungarotoxin mediate neurite retraction in a calcium-dependent manner. J Neurosci. 1994;14:889–896. doi: 10.1523/JNEUROSCI.14-02-00889.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Renshaw GM. [125I]-alpha-bungarotoxin binding co-varies with motoneurone density during apoptosis. Neuroreport. 1994;5:1949–1952. doi: 10.1097/00001756-199410000-00027. [DOI] [PubMed] [Google Scholar]
  • 22.Renshaw GM, Dyson SE. Alpha-BTX lowers neuronal metabolism during the arrest of motoneurone apoptosis. Neuroreport. 1995;6:284–288. doi: 10.1097/00001756-199501000-00015. [DOI] [PubMed] [Google Scholar]
  • 23.Hory-Lee F, Frank E. The nicotinic blocking agents d-tubocurare and alpha-bungarotoxin save motoneurons from naturally occurring death in the absence of neuromuscular blockade. J Neurosci. 1995;15:6453–6460. doi: 10.1523/JNEUROSCI.15-10-06453.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Treinin M, Chalfie M. A mutated acetylcholine receptor subunit causes neuronal degeneration in C. elegans. Neuron. 1995;14:871–877. doi: 10.1016/0896-6273(95)90231-7. [DOI] [PubMed] [Google Scholar]
  • 25.Whiting P, Lindstrom J. Affinity labelling of neuronal acetylcholine receptors localizes acetylcholine-binding sites to their beta-subunits. FEBS Lett. 1987;213:55–60. doi: 10.1016/0014-5793(87)81464-3. [DOI] [PubMed] [Google Scholar]
  • 26.Flores CM, Rogers SW, Pabreza LA, Wolfe BB, Kellar KJ. A subtype of nicotinic cholinergic receptor in rat brain is composed of alpha 4 and beta 2 subunits and is up-regulated by chronic nicotine treatment. Mol Pharmacol. 1992;41:31–37. [PubMed] [Google Scholar]
  • 27.Picciotto MR, Zoli M, Lena C, Bessis A, Lallemand Y, Le Novere N, et al. 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]
  • 28.Picciotto MR, Zoli M, Rimondini R, Lena C, Marubio LM, Pich EM, et al. Acetylcholine receptors containing the beta2 subunit are involved in the reinforcing properties of nicotine. Nature. 1998;391:173–177. doi: 10.1038/34413. [DOI] [PubMed] [Google Scholar]
  • 29.Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C, et al. 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]
  • 30.Cordero-Erausquin M, Marubio LM, Klink R, Changeux JP. Nicotinic receptor function: new perspectives from knockout mice. Trends Pharmacol Sci. 2000;21:211–217. doi: 10.1016/s0165-6147(00)01489-9. [DOI] [PubMed] [Google Scholar]
  • 31.Jurgensen S, Ferreira ST. Nicotinic receptors, amyloid-beta, and synaptic failure in Alzheimer's disease. J Mol Neurosci. 40:221–229. doi: 10.1007/s12031-009-9237-0. [DOI] [PubMed] [Google Scholar]
  • 32.Buckingham SD, Jones AK, Brown LA, Sattelle DB. Nicotinic acetylcholine receptor signalling: roles in Alzheimer's disease and amyloid neuroprotection. Pharmacol Rev. 2009;61:39–61. doi: 10.1124/pr.108.000562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Steinlein OK, Bertrand D. Nicotinic receptor channelopathies and epilepsy. Pflugers Arch. 460:495–503. doi: 10.1007/s00424-009-0766-8. [DOI] [PubMed] [Google Scholar]
  • 34.Barabino B, Vailati S, Moretti M, McIntosh JM, Longhi R, Clementi F, et al. An alpha4beta4 nicotinic receptor subtype is present in chick retina: identification, characterization and pharmacological comparison with the transfected alpha4beta4 and alpha6beta4 subtypes. Mol Pharmacol. 2001;59:1410–1417. doi: 10.1124/mol.59.6.1410. [DOI] [PubMed] [Google Scholar]
  • 35.Moaddel R, Jozwiak K, Whittington K, Wainer IW. Conformational mobility of immobilized alpha3beta2, alpha3beta4, alpha4beta2, and alpha4beta4 nicotinic acetylcholine receptors. Anal Chem. 2005;77:895–901. doi: 10.1021/ac048826x. [DOI] [PubMed] [Google Scholar]
  • 36.Hamouda AK, Jin X, Sanghvi M, Srivastava S, Pandhare A, Duddempudi PK, et al. Photoaffinity labeling the agonist binding domain of alpha4beta4 and alpha4beta2 neuronal nicotinic acetylcholine receptors with [(125)I]epibatidine and 5[(125)I]A-85380. Biochim Biophys Acta. 2009;1788:1987–1995. doi: 10.1016/j.bbamem.2009.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Winzer-Serhan UH, Leslie FM. Codistribution of nicotinic acetylcholine receptor subunit alpha3 and beta4 mRNAs during rat brain development. The Journal of comparative neurology. 1997;386:540–554. doi: 10.1002/(sici)1096-9861(19971006)386:4<540::aid-cne2>3.0.co;2-2. [DOI] [PubMed] [Google Scholar]
  • 38.Quik M, Polonskaya Y, Gillespie A, Jakowec M, Lloyd GK, Langston JW. Localization of nicotinic receptor subunit mRNAs in monkey brain by in situ hybridization. The Journal of comparative neurology. 2000;425:58–69. doi: 10.1002/1096-9861(20000911)425:1<58::aid-cne6>3.0.co;2-x. [DOI] [PubMed] [Google Scholar]
  • 39.Conroy WG, Berg DK. Neurons can maintain multiple classes of nicotinic acetylcholine receptors distinguished by different subunit compositions. J Biol Chem. 1995;270:4424–4431. doi: 10.1074/jbc.270.9.4424. [DOI] [PubMed] [Google Scholar]
  • 40.Alkondon M, Albuquerque EX. Initial characterization of the nicotinic acetylcholine receptors in rat hippocampal neurons. J Recept Res. 1991;11:1001–1021. doi: 10.3109/10799899109064693. [DOI] [PubMed] [Google Scholar]
  • 41.Alkondon M, Albuquerque EX. Diversity of nicotinic acetylcholine receptors in rat hippocampal neurons. I. Pharmacological and functional evidence for distinct structural subtypes. J Pharmacol Exp Ther. 1993;265:1455–1473. [PubMed] [Google Scholar]
  • 42.Albuquerque EX, Costa AC, Alkondon M, Shaw KP, Ramoa AS, Aracava Y. Functional properties of the nicotinic and glutamatergic receptors. J Recept Res. 1991;11:603–625. doi: 10.3109/10799899109066430. [DOI] [PubMed] [Google Scholar]
  • 43.Albuquerque EX, Pereira EF, Castro NG, Alkondon M, Reinhardt S, Schroder H, et al. Nicotinic receptor function in the mammalian central nervous system. Ann N Y Acad Sci. 1995;757:48–72. doi: 10.1111/j.1749-6632.1995.tb17464.x. [DOI] [PubMed] [Google Scholar]
  • 44.Luetje CW, Patrick J. Both alpha- and beta-subunits contribute to the agonist sensitivity of neuronal nicotinic acetylcholine receptors. J Neurosci. 1991;11:837–845. doi: 10.1523/JNEUROSCI.11-03-00837.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Vernino S, Rogers M, Radcliffe KA, Dani JA. Quantitative measurement of calcium flux through muscle and neuronal nicotinic acetylcholine receptors. J Neurosci. 1994;14:5514–5524. doi: 10.1523/JNEUROSCI.14-09-05514.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Buisson B, Gopalakrishnan M, Arneric SP, Sullivan JP, Bertrand D. Human alpha4beta2 neuronal nicotinic acetylcholine receptor in HEK 293 cells: A patch-clamp study. J Neurosci. 1996;16:7880–7891. doi: 10.1523/JNEUROSCI.16-24-07880.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Buisson B, Vallejo YF, Green WN, Bertrand D. The unusual nature of epibatidine responses at the alpha4beta2 nicotinic acetylcholine receptor. Neuropharmacology. 2000;39:2561–2569. doi: 10.1016/s0028-3908(00)00158-1. [DOI] [PubMed] [Google Scholar]
  • 48.Elliott KJ, Ellis SB, Berckhan KJ, Urrutia A, ChavezNoriega LE, Johnson EC, et al. Comparative structure of human neuronal alpha 2-alpha 7 and beta 2-beta 4 nicotinic acetylcholine receptor subunits and functional expression of the alpha 2, alpha 3, alpha 4, alpha 7, beta 2, and beta 4 subunits. Journal of Molecular Neuroscience. 1996;7:217–228. doi: 10.1007/BF02736842. [DOI] [PubMed] [Google Scholar]
  • 49.Gopalakrishnan M, Molinari EJ, Sullivan JP. Regulation of human alpha4beta2 neuronal nicotinic acetylcholine receptors by cholinergic channel ligands and second messenger pathways. Molecular Pharmacology. 1997;52:524–534. [PubMed] [Google Scholar]
  • 50.Gopalakrishnan M, Monteggia LM, Anderson DJ, Molinari EJ, Piattoni-Kaplan M, Donnelly-Roberts D, et al. Stable expression, pharmacologic properties and regulation of the human neuronal nicotinic acetylcholine alpha 4 beta 2 receptor. J Pharmacol Exp Ther. 1996;276:289–297. [PubMed] [Google Scholar]
  • 51.Harvey SC, Maddox FN, Luetje CW. Multiple determinants of dihydro-beta-erythroidine sensitivity on rat neuronal nicotinic receptor alpha subunits. J Neurochem. 1996;67:1953–1959. doi: 10.1046/j.1471-4159.1996.67051953.x. [DOI] [PubMed] [Google Scholar]
  • 52.Wang F, Gerzanich V, Wells GB, Anand R, Peng X, Keyser K, et al. Assembly of human neuronal nicotinic receptor alpha 5 subunits with alpha 3, beta 2, and beta 4 subunits. Journal of Biological Chemistry. 1996;271:17656–17665. doi: 10.1074/jbc.271.30.17656. [DOI] [PubMed] [Google Scholar]
  • 53.Chavez-Noriega LE, Crona JH, Washburn MS, Urrutia A, Elliott KJ, Johnson EC. Pharmacological characterization of recombinant human neuronal nicotinic acetylcholine receptors h alpha 2 beta 2, h alpha 2 beta 4, h alpha 3 beta 2, h alpha 3 beta 4, h alpha 4 beta 2, h alpha 4 beta 4 and h alpha 7 expressed in Xenopus oocytes. J Pharmacol Exp Ther. 1997;280:346–356. [PubMed] [Google Scholar]
  • 54.Forsayeth JR, Kobrin E. Formation of oligomers containing the beta3 and beta4 subunits of the rat nicotinic receptor. J Neurosci. 1997;17:1531–1538. doi: 10.1523/JNEUROSCI.17-05-01531.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Stauderman KA, Mahaffy LS, Akong M, Velicelebi G, Chavez-Noriega LE, Crona JH, et al. Characterization of human recombinant neuronal nicotinic acetylcholine receptor subunit combinations alpha 2 beta 4, alpha 3 beta 4 and alpha 4 beta 4 stably expressed in HEK293 cells. Journal of Pharmacology and Experimental Therapeutics. 1998;284:777–789. [PubMed] [Google Scholar]
  • 56.Buisson B, Bertrand D. Chronic exposure to nicotine upregulates the human (alpha)4((beta)2 nicotinic acetylcholine receptor function. J Neurosci. 2001;21:1819–1829. doi: 10.1523/JNEUROSCI.21-06-01819.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Henningfield JE, Heishman SJ. The addictive role of nicotine in tobacco use. Psychopharmacology (Berl) 1995;117:11–13. doi: 10.1007/BF02245089. [DOI] [PubMed] [Google Scholar]
  • 58.Benwell ME, Balfour DJ, Anderson JM. Evidence that tobacco smoking increases the density of (−)-[3H]nicotine binding sites in human brain. J Neurochem. 1988;50:1243–1247. doi: 10.1111/j.1471-4159.1988.tb10600.x. [DOI] [PubMed] [Google Scholar]
  • 59.Wonnacott S. The paradox of nicotinic acetylcholine receptor upregulation by nicotine. Trends Pharmacol Sci. 1990;11:216–219. doi: 10.1016/0165-6147(90)90242-z. [DOI] [PubMed] [Google Scholar]
  • 60.Dani JA, Heinemann S. Molecular and cellular aspects of nicotine abuse. Neuron. 1996;16:905–908. doi: 10.1016/s0896-6273(00)80112-9. [DOI] [PubMed] [Google Scholar]
  • 61.Ke L, Eisenhour CM, Bencherif M, Lukas RJ. Effects of chronic nicotine treatment on expression of diverse nicotinic acetylcholine receptor subtypes. I. Dose- and time-dependent effects of nicotine treatment. J Pharmacol Exp Ther. 1998;286:825–840. [PubMed] [Google Scholar]
  • 62.Perry DC, Davila-Garcia MI, Stockmeier CA, Kellar KJ. Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography studies. J Pharmacol Exp Ther. 1999;289:1545–1552. [PubMed] [Google Scholar]
  • 63.Peng X, Gerzanich V, Anand R, Whiting PJ, Lindstrom J. Nicotine-induced increase in neuronal nicotinic receptors results from a decrease in the rate of receptor turnover. Mol Pharmacol. 1994;46:523–530. [PubMed] [Google Scholar]
  • 64.Bencherif M, Fowler K, Lukas RJ, Lippiello PM. Mechanisms of up-regulation of neuronal nicotinic acetylcholine receptors in clonal cell lines and primary cultures of fetal rat brain. J Pharmacol Exp Ther. 1995;275:987–994. [PubMed] [Google Scholar]
  • 65.Sharp BM, Beyer HS, Levine AS, Morley JE, McAllen KM. Attenuation of the plasma prolactin response to restraint stress after acute and chronic administration of nicotine to rats. J Pharmacol Exp Ther. 1987;241:438–442. [PubMed] [Google Scholar]
  • 66.Hulihan-Giblin BA, Lumpkin MD, Kellar KJ. Effects of chronic administration of nicotine on prolactin release in the rat: inactivation of prolactin response by repeated injections of nicotine. J Pharmacol Exp Ther. 1990;252:21–25. [PubMed] [Google Scholar]
  • 67.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]
  • 68.Lukas RJ. Diversity and patterns of regulation of nicotinic receptor subtypes. Ann N Y Acad Sci. 1995;757:153–168. doi: 10.1111/j.1749-6632.1995.tb17471.x. [DOI] [PubMed] [Google Scholar]
  • 69.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]
  • 70.Hsu YN, Amin J, Weiss DS, Wecker L. Sustained nicotine exposure differentially affects alpha 3 beta 2 and alpha 4 beta 2 neuronal nicotinic receptors expressed in Xenopus oocytes. J Neurochem. 1996;66:667–675. doi: 10.1046/j.1471-4159.1996.66020667.x. [DOI] [PubMed] [Google Scholar]
  • 71.Fenster CP, Rains MF, Noerager B, Quick MW, Lester RA. Influence of subunit composition on desensitization of neuronal acetylcholine receptors at low concentrations of nicotine. J Neurosci. 1997;17:5747–5759. doi: 10.1523/JNEUROSCI.17-15-05747.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Olale F, Gerzanich V, Kuryatov A, Wang F, Lindstrom J. Chronic nicotine exposure differentially affects the function of human alpha3, alpha4, and alpha7 neuronal nicotinic receptor subtypes. J Pharmacol Exp Ther. 1997;283:675–683. [PubMed] [Google Scholar]
  • 73.Gentry CL, Lukas RJ. Regulaton of nicotinic acetylcholine receptor numbers and function by chronic nicotine exposure. Curr Dru Targets CNS Neurol Disord. 2002;1:359–385. doi: 10.2174/1568007023339184. [DOI] [PubMed] [Google Scholar]
  • 74.Morley BJ, Kemp GE, Salvaterra P. alpha-Bungarotoxin binding sites in the CNS. Life Sci. 1979;24:859–872. doi: 10.1016/0024-3205(79)90335-7. [DOI] [PubMed] [Google Scholar]
  • 75.Schmidt JT, Freeman JA. Electrophysiologic evidence that retinotectal synaptic transmission in the goldfish is nicotinic cholinergic. Brain research. 1980;187:129–142. doi: 10.1016/0006-8993(80)90499-0. [DOI] [PubMed] [Google Scholar]
  • 76.Clarke PB. The fall and rise of neuronal alpha-bungarotoxin binding proteins. Trends Pharmacol Sci. 1992;13:407–413. doi: 10.1016/0165-6147(92)90125-p. [DOI] [PubMed] [Google Scholar]
  • 77.Lukas RJ, Bencherif M. Heterogeneity and regulation of nicotinic acetylcholine receptors. Int Rev Neurobiol. 1992;34:25–131. doi: 10.1016/s0074-7742(08)60097-5. [DOI] [PubMed] [Google Scholar]
  • 78.Horch HL, Sargent PB. Perisynaptic surface distribution of multiple classes of nicotinic acetylcholine receptors on neurons in the chicken ciliary ganglion. J Neurosci. 1995;15:7778–7795. doi: 10.1523/JNEUROSCI.15-12-07778.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Couturier S, Bertrand D, Matter JM, Hernandez MC, Bertrand S, Millar N, et al. A neuronal nicotinic acetylcholine receptor subunit (alpha 7) is developmentally regulated and forms a homo-oligomeric channel blocked by alpha-BTX. Neuron. 1990;5:847–856. doi: 10.1016/0896-6273(90)90344-f. [DOI] [PubMed] [Google Scholar]
  • 80.Peng JH, Fryer JD, Hurst RS, Schroeder KM, George AA, Morrissy S, et al. High-affinity epibatidine binding of functional, human alpha7-nicotinic acetylcholine receptors stably and heterologously expressed de novo in human SH-EP1 cells. J Pharmacol Exp Ther. 2005;313:24–35. doi: 10.1124/jpet.104.079004. [DOI] [PubMed] [Google Scholar]
  • 81.Peng X, Katz M, Gerzanich V, Anand R, Lindstrom J. Human alpha 7 acetylcholine receptor: cloning of the alpha 7 subunit from the SH-SY5Y cell line and determination of pharmacological properties of native receptors and functional alpha 7 homomers expressed in Xenopus oocytes. Mol Pharmacol. 1994;45:546–554. [PubMed] [Google Scholar]
  • 82.Schoepfer R, Conroy WG, Whiting P, Gore M, Lindstrom J. Brain Alpha-Bungarotoxin Binding-Protein Cdnas and Mabs Reveal Subtypes of This Branch of the Ligand-Gated Ion Channel Gene Superfamily. Neuron. 1990;5:35–48. doi: 10.1016/0896-6273(90)90031-a. [DOI] [PubMed] [Google Scholar]
  • 83.Seguela P, Wadiche J, Dineley-Miller K, Dani JA, Patrick JW. Molecular cloning, functional properties, and distribution of rat brain alpha 7: a nicotinic cation channel highly permeable to calcium. J Neurosci. 1993;13:596–604. doi: 10.1523/JNEUROSCI.13-02-00596.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Vijayaraghavan S, Pugh PC, Zhang ZW, Rathouz MM, Berg DK. Nicotinic receptors that bind alpha-bungarotoxin on neurons raise intracellular free Ca2+ Neuron. 1992;8:353–362. doi: 10.1016/0896-6273(92)90301-s. [DOI] [PubMed] [Google Scholar]
  • 85.Puchacz E, Buisson B, Bertrand D, Lukas RJ. Functional expression of nicotinic acetylcholine receptors containing rat alpha 7 subunits in human SH-SY5Y neuroblastoma cells. FEBS Lett. 1994;354:155–159. doi: 10.1016/0014-5793(94)01108-7. [DOI] [PubMed] [Google Scholar]
  • 86.Zorumski CF, Thio LL, Isenberg KE, Clifford DB. Nicotinic acetylcholine currents in cultured postnatal rat hippocampal neurons. Mol Pharmacol. 1992;41:931–936. [PubMed] [Google Scholar]
  • 87.Alkondon M, Pereira EF, Albuquerque EX. Mapping the location of functional nicotinic and gamma-aminobutyric acidA receptors on hippocampal neurons. J Pharmacol Exp Ther. 1996;279:1491–1506. [PubMed] [Google Scholar]
  • 88.Wu J, George AA, Schroeder KM, Xu L, Marxer-Miller S, Lucero L, et al. Electrophysiological, pharmacological, and molecular evidence for alpha7-nicotinic acetylcholine receptors in rat midbrain dopamine neurons. The Journal of pharmacology and experimental therapeutics. 2004;311:80–91. doi: 10.1124/jpet.104.070417. [DOI] [PubMed] [Google Scholar]
  • 89.Bonfante-Cabarcas R, Swanson KL, Alkondon M, Albuquerque EX. Diversity of nicotinic acetylcholine receptors in rat hippocampal neurons. IV. Regulation by external Ca++ of alpha-bungarotoxin-sensitive receptor function and of rectification induced by internal Mg++ J Pharmacol Exp Ther. 1996;277:432–444. [PubMed] [Google Scholar]
  • 90.Yang K, Hu J, Lucero L, Liu Q, Zheng C, Zhen X, et al. Distinctive nicotinic acetylcholine receptor functional phenotypes of rat ventral tegmental area dopaminergic neurons. The Journal of physiology. 2009;587:345–361. doi: 10.1113/jphysiol.2008.162743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Liu Q, Huang Y, Xue F, Simard A, DeChon J, Li G, et al. A novel nicotinic acetylcholine receptor subtype in basal forebrain cholinergic neurons with high sensitivity to amyloid peptides. J Neurosci. 2009;29:918–929. doi: 10.1523/JNEUROSCI.3952-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Lukas RJ. Expression of ganglia-type nicotinic acetylcholine receptors and nicotinic ligand binding sites by cells of the IMR-32 human neuroblastoma clonal line. J Pharmacol Exp Ther. 1993;265:294–302. [PubMed] [Google Scholar]
  • 93.Vernallis AB, Conroy WG, Berg DK. Neurons assemble acetylcholine receptors with as many as three kinds of subunits while maintaining subunit segregation among receptor subtypes. Neuron. 1993;10:451–464. doi: 10.1016/0896-6273(93)90333-m. [DOI] [PubMed] [Google Scholar]
  • 94.Orr-Urtreger A, Goldner FM, Saeki M, Lorenzo I, Goldberg L, De Biasi M, et al. Mice deficient in the alpha7 neuronal nicotinic acetylcholine receptor lack alpha-bungarotoxin binding sites and hippocampal fast nicotinic currents. J Neurosci. 1997;17:9165–9171. doi: 10.1523/JNEUROSCI.17-23-09165.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Hunter BE, de Fiebre CM, Papke RL, Kem WR, Meyer EM. A novel nicotinic agonist facilitates induction of long-term potentiation in the rat hippocampus. Neurosci Lett. 1994;168:130–134. doi: 10.1016/0304-3940(94)90433-2. [DOI] [PubMed] [Google Scholar]
  • 96.Morales MA, Bachoo M, Collier B, Polosa C. Pre- and postsynaptic components of nicotinic long-term potentiation in the superior cervical ganglion of the cat. J Neurophysiol. 1994;72:819–824. doi: 10.1152/jn.1994.72.2.819. [DOI] [PubMed] [Google Scholar]
  • 97.Miner LL, Marks MJ, Collins AC. Classical genetic analysis of nicotine-induced seizures and nicotinic receptors. J Pharmacol Exp Ther. 1984;231:545–554. [PubMed] [Google Scholar]
  • 98.Bertrand D, Galzi JL, Devillers-Thiery A, Bertrand S, Changeux JP. Mutations at two distinct sites within the channel domain M2 alter calcium permeability of neuronal alpha 7 nicotinic receptor. Proc Natl Acad Sci U S A. 1993;90:6971–6975. doi: 10.1073/pnas.90.15.6971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Galzi JL, Changeux JP. Neuronal nicotinic receptors: molecular organization and regulations. Neuropharmacology. 1995;34:563–582. doi: 10.1016/0028-3908(95)00034-4. [DOI] [PubMed] [Google Scholar]
  • 100.Garcia-Guzman M, Sala F, Sala S, Campos-Caro A, Criado M. Role of two acetylcholine receptor subunit domains in homomer formation and intersubunit recognition, as revealed by alpha 3 and alpha 7 subunit chimeras. Biochemistry. 1994;33:15198–15203. doi: 10.1021/bi00254a031. [DOI] [PubMed] [Google Scholar]
  • 101.Corringer PJ, Galzi JL, Eisele JL, Bertrand S, Changeux JP, Bertrand D. Identification of a new component of the agonist binding site of the nicotinic alpha 7 homooligomeric receptor. J Biol Chem. 1995;270:11749–11752. doi: 10.1074/jbc.270.20.11749. [DOI] [PubMed] [Google Scholar]
  • 102.Corringer PJ, Bertrand S, Bohler S, Edelstein SJ, Changeux JP, Bertrand D. Critical elements determining diversity in agonist binding and desensitization of neuronal nicotinic acetylcholine receptors. J Neurosci. 1998;18:648–657. doi: 10.1523/JNEUROSCI.18-02-00648.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Campos-Caro A, Sala S, Ballesta JJ, Vicente-Agullo F, Criado M, Sala F. A single residue in the M2–M3 loop is a major determinant of coupling between binding and gating in neuronal nicotinic receptors. Proc Natl Acad Sci U S A. 1996;93:6118–6123. doi: 10.1073/pnas.93.12.6118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Lukas RJ, Lucero L, Buisson B, Galzi JL, Puchacz E, Fryer JD, et al. Neurotoxicity of channel mutations in heterologously expressed alpha7-nicotinic acetylcholine receptors. Eur J Neurosci. 2001;13:1849–1860. doi: 10.1046/j.0953-816x.2001.01560.x. [DOI] [PubMed] [Google Scholar]
  • 105.Gopalakrishnan M, Buisson B, Touma E, Giordano T, Campbell JE, Hu IC, et al. Stable expression and pharmacological properties of the human alpha 7 nicotinic acetylcholine receptor. European journal of pharmacology. 1995;290:237–246. doi: 10.1016/0922-4106(95)00083-6. [DOI] [PubMed] [Google Scholar]
  • 106.Quik M, Choremis J, Komourian J, Lukas RJ, Puchacz E. Similarity between rat brain nicotinic alpha-bungarotoxin receptors and stably expressed alpha-bungarotoxin binding sites. J Neurochem. 1996;67:145–154. doi: 10.1046/j.1471-4159.1996.67010145.x. [DOI] [PubMed] [Google Scholar]
  • 107.Peng M, Conforti L, Millhorn DE. Expression of somatostatin receptor genes and acetylcholine receptor development in rat skeletal muscle during postnatal development. Int J Mol Med. 1998;1:841–848. doi: 10.3892/ijmm.1.5.841. [DOI] [PubMed] [Google Scholar]
  • 108.Chen D, Patrick JW. The alpha-bungarotoxin-binding nicotinic acetylcholine receptor from rat brain contains only the alpha7 subunit. J Biol Chem. 1997;272:24024–24029. doi: 10.1074/jbc.272.38.24024. [DOI] [PubMed] [Google Scholar]
  • 109.Drisdel RC, Green WN. Neuronal alpha-bungarotoxin receptors are alpha7 subunit homomers. J Neurosci. 2000;20:133–139. doi: 10.1523/JNEUROSCI.20-01-00133.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Azam L, Winzer-Serhan U, Leslie FM. Co-expression of alpha7 and beta2 nicotinic acetylcholine receptor subunit mRNAs within rat brain cholinergic neurons. Neuroscience. 2003;119:965–977. doi: 10.1016/s0306-4522(03)00220-3. [DOI] [PubMed] [Google Scholar]
  • 111.Yu CR, Role LW. Functional contribution of the alpha7 subunit to multiple subtypes of nicotinic receptors in embryonic chick sympathetic neurones. J Physiol. 1998;509(Pt 3):651–665. doi: 10.1111/j.1469-7793.1998.651bm.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Sudweeks SN, Yakel JL. Functional and molecular characterization of neuronal nicotinic ACh receptors in rat CA1 hippocampal neurons. J Physiol. 2000;527(Pt 3):515–528. doi: 10.1111/j.1469-7793.2000.00515.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Khiroug SS, Harkness PC, Lamb PW, Sudweeks SN, Khiroug L, Millar NS, et al. Rat nicotinic ACh receptor alpha7 and beta2 subunits co-assemble to form functional heteromeric nicotinic receptor channels. J Physiol. 2002;540:425–434. doi: 10.1113/jphysiol.2001.013847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Klink R, de Kerchove d'Exaerde A, Zoli M, Changeux 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]
  • 115.Azam L, Winzer-Serhan UH, Chen Y, Leslie FM. Expression of neuronal nicotinic acetylcholine receptor subunit mRNAs within midbrain dopamine neurons. J Comp Neurol. 2002;444:260–274. doi: 10.1002/cne.10138. [DOI] [PubMed] [Google Scholar]
  • 116.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]
  • 117.Gerzanich V, Kuryatov A, Anand R, Lindstrom J. "Orphan" alpha6 nicotinic AChR subunit can form a functional heteromeric acetylcholine receptor. Mol Pharmacol. 1997;51:320–327. [PubMed] [Google Scholar]
  • 118.Kuryatov A, Olale F, Cooper J, Choi C, Lindstrom J. Human alpha6 AChR subtypes: subunit composition, assembly, and pharmacological responses. Neuropharmacology. 2000;39:2570–2590. doi: 10.1016/s0028-3908(00)00144-1. [DOI] [PubMed] [Google Scholar]
  • 119.Capelli AM, Castelletti L, Chen YH, Vab der Keyl H, Pucci L, Oliosi B, et al. Stable expression and functional characterisation of a human nicotinic acetylcholine receptor with a6b2 properties: Discovery of selective antagonists. British Journal of Pharmacology. 2011 doi: 10.1111/j.1476-5381.2011.01213.x. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Kuryatov A, Lindstrom J. Expression of functional human a6b2b3* acetylcholine receptors in Xenopus laevis oocytes achieved through subunit chimeras and concatamers. Journal of Molecular Pharmacology. 2011;79:126–140. doi: 10.1124/mol.110.066159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Drenan RM, Grady SR, Whiteaker P, McClure-Begley T, McKinney S, Miwa JM, et al. 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]
  • 122.Yang K, Buhlman L, Khan GM, Nichols RA, Jin G, McIntosh JM, et al. Functional nicotinic acetylcholine receptors containing alpha6 subunits are on GABAergic neuronal boutons adherent to ventral tegmental area dopamine neurons. J Neurosci. 2011;31:2537–2548. doi: 10.1523/JNEUROSCI.3003-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.McIntosh JM, Olivera BM, Cruz LJ. Conus peptides as probes for ion channels. Methods Enzymol. 1999;294:605–624. doi: 10.1016/s0076-6879(99)94034-x. [DOI] [PubMed] [Google Scholar]
  • 124.Cartier GE, Yoshikami D, Gray WR, Luo S, Olivera BM, McIntosh JM. A new alpha-conotoxin which targets alpha3beta2 nicotinic acetylcholine receptors. J Biol Chem. 1996;271:7522–7528. doi: 10.1074/jbc.271.13.7522. [DOI] [PubMed] [Google Scholar]
  • 125.Harvey SC, McIntosh JM, Cartier GE, Maddox FN, Luetje CW. Determinants of specificity for alpha-conotoxin MII on alpha3beta2 neuronal nicotinic receptors. Mol Pharmacol. 1997;51:336–342. doi: 10.1124/mol.51.2.336. [DOI] [PubMed] [Google Scholar]
  • 126.Kulak JM, Nguyen TA, Olivera BM, McIntosh JM. Alpha-conotoxin MII blocks nicotine-stimulated dopamine release in rat striatal synaptosomes. J Neurosci. 1997;17:5263–5270. doi: 10.1523/JNEUROSCI.17-14-05263.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kaiser SA, Soliakov L, Harvey SC, Luetje CW, Wonnacott S. Differential inhibition by alpha-conotoxin-MII of the nicotinic stimulation of [3H]dopamine release from rat striatal synaptosomes and slices. J Neurochem. 1998;70:1069–1076. doi: 10.1046/j.1471-4159.1998.70031069.x. [DOI] [PubMed] [Google Scholar]
  • 128.Whiteaker P, Peterson CG, Xu W, McIntosh JM, Paylor R, Beaudet AL, et al. Involvement of the alpha3 subunit in central nicotinic binding populations. J Neurosci. 2002;22:2522–2529. doi: 10.1523/JNEUROSCI.22-07-02522.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Everhart D, Reiller E, Mirzoian A, McIntosh JM, Malhotra A, Luetje CW. Identification of residues that confer alpha-conotoxin-PnIA sensitivity on the alpha 3 subunit of neuronal nicotinic acetylcholine receptors. J Pharmacol Exp Ther. 2003;306:664–670. doi: 10.1124/jpet.103.051656. [DOI] [PubMed] [Google Scholar]
  • 130.Champtiaux N, Gotti C, Cordero-Erausquin M, David DJ, Przybylski C, Lena C, Clementi F, Moretti M, Rossi FM, Le Novere N, McIntosh JM, Gardier AM, Changeux JP. Subunit composition of functional nicotinic receptors in dopaminergic neurons investigated with knock-out mice. J Neurosci. 2003;23:7820–7829. doi: 10.1523/JNEUROSCI.23-21-07820.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Gotti C, Moretti M, Clementi F, Riganti L, McIntosh JM, Collins AC, et al. Expression of nigrostriatal alpha 6-containing nicotinic acetylcholine receptors is selectively reduced, but not eliminated, by beta 3 subunit gene deletion. Mol Pharmacol. 2005;67:2007–2015. doi: 10.1124/mol.105.011940. [DOI] [PubMed] [Google Scholar]
  • 132.Dowell C, Olivera BM, Garrett JE, Staheli ST, Watkins M, Kuryatov A, et al. Alpha-conotoxin PIA is selective for alpha6 subunit-containing nicotinic acetylcholine receptors. J Neurosci. 2003;23:8445–8452. doi: 10.1523/JNEUROSCI.23-24-08445.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Azam L, McIntosh JM. Effect of novel alpha-conotoxins on nicotine-stimulated [3H]dopamine release from rat striatal synaptosomes. J Pharmacol Exp Ther. 2005;312:231–237. doi: 10.1124/jpet.104.071456. [DOI] [PubMed] [Google Scholar]
  • 134.Chi SW, Lee SH, Kim DH, Kim JS, Olivera BM, McIntosh JM, et al. Solution structure of alpha-conotoxin PIA, a novel antagonist of alpha6 subunit containing nicotinic acetylcholine receptors. Biochem Biophys Res Commun. 2005;338:1990–1997. doi: 10.1016/j.bbrc.2005.10.176. [DOI] [PubMed] [Google Scholar]
  • 135.Di Chiara G, Imperato A. Drugs abused by humans preferentially increase synaptic dopamine concentrations in the mesolimbic system of freely moving rats. Proc Natl Acad Sci U S A. 1988;85:5274–5278. doi: 10.1073/pnas.85.14.5274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Fryer JD, Lukas RJ. Antidepressants noncompetitively inhibit nicotinic acetylcholine receptor function. J Neurochem. 1999;72:1117–1124. doi: 10.1046/j.1471-4159.1999.0721117.x. [DOI] [PubMed] [Google Scholar]
  • 137.Fryer JD, Lukas RJ. Noncompetitive functional inhibition at diverse, human nicotinic acetylcholine receptor subtypes by bupropion, phencyclidine, and ibogaine. J Pharmacol Exp Ther. 1999;288:88–92. [PubMed] [Google Scholar]
  • 138.Tizabi Y, Overstreet DH, Rezvani AH, Louis VA, Clark E, Jr, Janowsky DS, et al. Antidepressant effects of nicotine in an animal model of depression. Psychopharmacology (Berl) 1999;142:193–199. doi: 10.1007/s002130050879. [DOI] [PubMed] [Google Scholar]
  • 139.Slemmer JE, Martin BR, Damaj MI. Bupropion is a nicotinic antagonist. J Pharmacol Exp Ther. 2000;295:321–327. [PubMed] [Google Scholar]
  • 140.Salminen O, Murphy KL, McIntosh JM, Drago J, Marks MJ, Collins AC, et al. Subunit composition and pharmacology of two classes of striatal presynaptic nicotinic acetylcholine receptors mediating dopamine release in mice. Mol Pharmacol. 2004;65:1526–1535. doi: 10.1124/mol.65.6.1526. [DOI] [PubMed] [Google Scholar]
  • 141.Endo T, Yanagawa Y, Obata K, Isa T. Nicotinic acetylcholine receptor subtypes involved in facilitation of GABAergic inhibition in mouse superficial superior colliculus. J Neurophysiol. 2005;94:3893–3902. doi: 10.1152/jn.00211.2005. [DOI] [PubMed] [Google Scholar]
  • 142.Lukas RJ, Fryer JD, Eaton JB, Gentry CL. Some methods for studies of nicotinic acetylcholine receptor pharmacology. In: Levin ED, editor. Nicotinic Receptors and Nervous System. Boca Raton: CRC Press; 2002. pp. 3–27. [Google Scholar]
  • 143.Marks MJ, Whiteaker P, Calcaterra J, Stitzel JA, Bullock AE, Grady SR, et al. Two pharmacologically distinct components of nicotinic receptor-mediated rubidium efflux in mouse brain require the beta 2 subunit. J Pharm Exper Thera. 1999;289:1090–1103. [PubMed] [Google Scholar]
  • 144.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]

RESOURCES