GET THE APP

Nicotinic Acetylcholine Receptors in the Ventral Segmental Area a
Biochemistry & Pharmacology: Open Access

Biochemistry & Pharmacology: Open Access
Open Access

ISSN: 2167-0501

+44-20-4587-4809

Review Article - (2013) Volume 0, Issue 0

Nicotinic Acetylcholine Receptors in the Ventral Segmental Area are Important Targets for Nicotine and Ethanol Co-dependence

Devin H. Taylor1,2, Scott C. Steffensen3 and Jie Wu1,2*
1Interdisciplinary Program in Neuroscience, Graduate College, Arizona State University, USA
2Division of Neurology, Barrow Neurological Institute, St. Joseph’s Hospital and Medical Center, USA
3Departments of Psychology, and Physiology & Developmental Biology, Brigham Young University, USA
*Corresponding Author: Jie Wu, MD, Ph.D., Professor, Division of Neurology, Barrow Neurological Institute, St. Joseph’s Hospital and Medical Center, USA, Tel: (602) 406-6376 Email:

Abstract

Tobacco and alcohol are the most commonly abused drugs. The nicotine (NIC) in tobacco and the ethanol (EtOH) in alcoholic drinks are responsible for their dependence respectively. The magnitude of tobacco smoking is drastically higher among alcoholics, suggesting a NIC-EtOH co-dependence. However, the mechanisms of NIC-EtOH interaction are not fully known, and the clarification of this action is clinically relevant. The majority of the NIC-EtOH interaction utilizes the ventral tegmental area (VTA) through both dopamine (DA) and non-DA systems. EtOH has been shown to bind directly to some nicotinic acetylcholine receptors (nAChRs) as, of course, does NIC. The non-selective/noncompetitive
nAChR antagonist mecamylamine (MEC) has been shown to partially block the DA releasing action of EtOH in the nucleus accumbens (NAc), while both the α4β2 nAChR antagonist dihydro-β-erythroidine (DHβE) and the α7 nAChR antagonist methyllycaconitine (MLA) do not. However, the α6-containing nAChRs (α6*-nAChRs) are responsible for both NIC-induced effects on DA release in the NAc and EtOH-induced GABA release in the VTA, suggesting that the α6*-nAChRs likely play a significant role in NIC-EtOH interactions. In this review, we have summarized current studies that reveal how EtOH reward through VTA nAChRs and what nAChR subtypes play roles in mediation of EtOH’s effects
in mesolimbic circuits. The accumulating lines of evidence suggest that the nAChRs, especially α6*-nAChRs in the VTA are likely important targets for NIC-EtOH interactions and NIC-EtOH co-dependence.

<

Keywords: Nicotinic acetylcholine receptors; Nicotine; Ethanol; VTA; Dopamine; GABA

Introduction

Tobacco and alcohol are the most commonly abused drugs by humans. Nicotine (NIC) is the major contributor in the continuance of tobacco use [1], while ethanol (EtOH) is the intoxicating agent in alcoholic drinks that can lead to abuse and dependence [2]. Alcohol use has been ascribed both positive and negative effects. While alcohol in low doses has been shown to provide cardiovascular protection [3], binge drinking is associated with higher incidents of cardiovascular disease and associated mortality [4,5]. As with alcohol, tobacco smoking has also been associated with cardiovascular problems. It has also been linked to coronary heart disease [6,7] and strokes [8,9]. Tobacco and alcohol use are leading causes of preventable death in the United States [10]. Smoking tobacco, the leading cause of preventable death, is accountable for approximately 467,000 deaths per year, while alcohol contributes to another 90,000 [10]. The most common type of polydrug use is alcohol and tobacco taken in concert [11,12]. The magnitude of tobacco smoking is extremely high among alcoholics [13] and is drastically higher than the rate in the general population [14,15]. Those who smoke are ten times more likely to be alcoholics than those who do not [16]. Those who are not alcoholics have been more successful than their alcoholic counterparts in quitting smoking, 49% to 7% respectively [16]. Although we know that the co-use of tobacco and alcohol is prevalent, little is known about the mechanisms of action when the two are used collectively. Clarification of these actions would be clinically useful in the treatment for the abuse of both tobacco and alcohol, as many requiring treatment for one also use the other.

Dopamine Dependent Mechanisms in the Mesolimbic System

Projections from the ventral tegmental area (VTA) to the nucleus accumbens (NAc), by way of the medial forebrain bundle, make up a vital component of the mesolimbic pathway [17-20]. The rewarding effects of both NIC and EtOH have been linked to the mesolimbic dopamine (DA) system [21-23], wherein an increase in DA in the NAc is thought to be vital for reward signaling. This system has been connected to the rewarding effects of many abused drugs [22,24-27]. The VTA consists of three major types of neurons: DA, γ-aminobutyric acid (GABA), and glutamate neurons. The most numerous are DA neurons that project to the NAc. The second, GABA neurons, inhibit DA neurons in local circuitry and project to other brain nuclei. Finally, there is a small population of glutamatergic neurons [28] which can innervate both DA and GABA neurons. The NAc is part of the ventral forebrain and is segregated into two regions: the shell and the core. Of the two regions, the shell has been shown to be important for the rewarding effects [29]. The medial VTA seems to consist of the highest number of DA neurons innervating the NAc shell [30].

Dopamine Independent Mechanisms in the Mesolimbic System

Although the mesolimbic DA system’s involvement has been known to be critical for most drugs of reward, drugs such as morphine, phencyclidine and NIC also manifest DA independent mechanisms. The necessity of this DA system in the rewarding properties of benzodiazepines, barbiturates and caffeine is also questioned [29,31,32]. A rising hypothesis asserts that DA is not requisite for all rewarding effects of opiates, cannabis, cocaineand NIC. The idea that DA in the mesolimbic system is the only way by which reward occurs is perhaps too limiting. Evidence supporting a lack of DA involvement in drug reinforcement has been demonstrated in cocaine self-administration [33], and conditioned place preference [34,35]. Additionally, there has been confirmation that GABA neurons in the mesolimbic pathway are involved in the rewarding properties of opiates [36-38]. It has been reported that GABAA receptors in the VTA could be a gating mechanism wherein opiate naïve animals utilize a DA independent system, while opiate dependant, and opiate withdrawn, animals utilize a DA dependent system [39]. Following opiate exposure and withdrawal, VTA GABAA receptors change from acting in an inhibitory manner to an excitatory one. Moreover, high doses of the DA antagonist haloperidol neglected to stop the reinstatement of heroin seeking behavior, giving credence to a notion of a DA independent system, at least in the case of opiates [40,41]. Some assert that DA neurons are not exactly reward neurons, but instead may be pivotal for the initiation and reinstatement of drug use [42]. These conflicting reports on the necessity of DA for the reinstatement of drug use shows that the role of DA is not fully understood. Therefore, DA transmission in the mesolimbic pathway may be important for the motivational effects of abused drugs in dependent animals, while other systems could be exploited when animals are naïve [43,44]. These studies supply verification for the existence of DA independent mechanisms that also contribute to the reinforcing properties of drugs of abuse.

Nicotine and Ethanol Reward Associated with the Ventral Tegmental Area

The effects of both NIC and EtOH involve the VTA [45,46]. The direct excitatory effects of EtOH on neurons in the VTA have been observed [47,48]; both in vivo and in vitro recordings have shown this effect [49,50]. Rodents will self-administer NIC and EtOH individually into this region [51-55] and synaptic plasticity has been demonstrated in the VTA in response to both substances [56], giving further support to this theory. It is also well known that NIC binds to nAChRs throughout the brain. Rodents with NIC infusions into the VTA demonstrate conditioned place preference. However, similar infusions into areas dorsal or caudal to the VTA do not produce this preference, even if heavily populated by nicotinic receptors [57]. This demonstrates that the medial VTA is essential for the rewarding behaviors of NIC [58]. For the above-mentioned reasons, it is likely that the neural substrates underlying the co-use of NIC and EtOH depend on VTA neuronal activity.

Other midbrain tegmental regions are involved in the reinforcing characteristics of drugs such as opiates and NIC [51,59]. Cholinergic receptors work together with neurons in the NAc, VTA, and pedunculopontine tegmental nucleus (PPTg) to produce the rewarding effects of NIC [20,59,60]. As many drugs of abuse have demonstrated both rewarding and aversive properties [29,61], it has been proposed that the VTA is involved in mediating both of these qualities in actions of NIC [57]. The aversive properties of NIC are reported as being mediated by the mesolimbic DA system, while its rewarding effects are mediated by non-DA neurons projecting from the VTA to the PPTg [57]. Although these two separate effects are thought to be mediated by the same region, two different systems are involved. Blocking the mesolimbic DA pathway with the DA antagonist α-flupenthixol greatly increases the sensitivity to NIC reward in rodents [57]. In fact, a reduction in the amount of DA D1 and D2 receptors is positively correlated with NIC addiction, additionally supporting this finding [62]. Therefore, the VTA mediates the rewarding effects of EtOH and the aversive effects of NIC via DAergic projections to the NAc, while the rewarding properties of NIC are mediated via non-DAergic projections from the VTA to the PPTg [57]. This mediation of both NIC reward and aversion in the VTA could aid in explaining the cross-tolerance observed with NIC and EtOH interactions.

It is currently understood that the mesolimbic system, especially in the VTA, is involved in NIC-EtOH reward. There is, however, a question as to which type or subtype of receptor is most important and on which category of neuron they are found. The origin of long-term potentiation (LTP) induction in NAc DA neurons has been reported to be from presynaptic neurons [63]. GABA neurons play an integral role in the rewarding effects of drugs of abuse [48]. In fact, stimulation of GABAA receptors is reinforcing [51] and inhibition of GABA neurons in the VTA could lead to increased DA release in the NAc [64]. nAChRs can be found on postsynaptic, preterminal, and presynaptic segments of GABA neurons [65-67], and the reinforcing properties of EtOH is influenced by these receptors. These studies suggest GABA neurons in the VTA serve as an important locus for the modulation of the EtOH effects, possibly by nAChRs.

Nicotine and Ethanol Interactions

Interactions between NIC and EtOH have been demonstrated in an assortment of experiments. Alteration of nAChRs in response to EtOH has been verified [68]. Mouse and rat studies have displayed cross-tolerance between EtOH and NIC [69-72]. Additional testing has elucidated aspects of the interaction between NIC and EtOH on nAChRs. For example, locomotor stimulation in mice by EtOH was partially impeded by the non-selective/non-competitive nAChR antagonist mecamylamine (MEC) [46]. Systemic EtOH induces DA release in the rat NAc and can be blocked by MEC in the VTA but not the NAc [46]. The EtOH ingestion and preference in high EtOHpreferring rats was also decreased by MEC [73,74]. Together, this research confirms that EtOH’s effects are partially facilitated through nAChRs and suggest these receptors as likely candidates for NIC-EtOH interaction.

On the other hand, it is known that the DAergic portion of the mesolimbic pathway is not the only contributor to the reinforcing effects of NIC and EtOH. Many neuron systems and receptor types have been implicated in the interaction involving NIC and EtOH. The serotonin [75], endogenous opioid [76], glutamatergic [77], and cholinergic [60,78] systems have been associated with NIC and EtOH interactions. Cholinergic receptors, especially nAChRs, have been implicated in this association for some time, but they are not the only mediators of the NIC-EtOH interaction. Aside from nAChRs, endocannabinoid CB1 receptors have been implicated in EtOH and NIC seeking [79], NICEtOH cross-sensitization [71], and interactive effects of NIC and EtOH involved in passive avoidance learning [80]. Although nAChRs are not the sole agents involved in the NIC-EtOH interaction, they seem to have greater effects on this relationship than CB1 receptors in both number and impact.

Nicotinic Acetylcholine Receptors and Ethanol

In the VTA, nAChRs are involved in mediating some reinforcing properties of EtOH [81]. nAChRs are ligand-gated ion channels expressed in a variety of compositions with two subtypes, α and β. Nine types of α subunits (α2-α10) are known to be expressed in vertebrates, as well as three β subunit types (β2-β4) [82]. The pentameric structure of each individual nAChR determines the variety of ion that is able to pass through the receptor’s channel [82]. For example, the α4β2 receptor mostly permits the passage of sodium through its pore while the α7 receptor has relevantly high Ca2+ permeability [82]. The known subunits found in the human brain are thought to be α3-α7 and β2-β4, although not all are presently known [82,83]. Many nicotinic receptors, composed of diverse combinations of subunits, are present in the human brain. The most common nicotinic pentamers consist of heteromeric α4 and β2 subunits or homomeric α7 subunits. The heteromeric pentamers could be joined as α4(2)β2(3) or α4(3)β2(2). Upregulation of some nAChRs in the mouse midbrain has been shown in the presence of NIC and EtOH together [84]. Some have argued that EtOH is simply a co-agonist and requires NIC to elicit a cholinergic response [85]. However, EtOH is not only a co-agonist in the presence of a ligand binding to cholinergic receptors, but also operates directly on some types of nAChRs in vitro [83,86] and in vivo [87-89]. The sensitivity and effects elicited by EtOH binding to nAChRs are dependent upon subunit composition [90].

Because α4β2 and α7 nAChRs are the most numerous of the subtypes in the human brain [86,91], they have been investigated for their relevance in the NIC and EtOH relationship. Human nAChRs expressed in Xenopus oocytes have demonstrated that α4β2 and α2β4 nAChRs have the highest affinity to EtOH, while α4β4, α2β2 and α7 nAChRs also respond to EtOH [90]. All combinations of α2, α4, β2 and β4 subunits enhanced receptor function in response to EtOH, while EtOH inhibited the functional α7 homomeric nAChRs expressed in these Xenopus oocytes [92,93]. The results seen in α4β2 and α7 nAChRs have also been confirmed in cultured rat neurons [83,94]. A microdialysis study has shown that DA release because of systemic EtOH involved nAChRs in the VTA [95]. It has also been proposed that α4 containing nAChRs enable modulation of the withdrawal effects of EtOH in mice [78]. Together, these data illustrate the crucial role of nAChRs in the interaction of these two substances.

As previously stated, EtOH acts as an antagonist on α7 nAChRs [93,96,97]. However, the intraperitoneal administration of the selective α7 nAChR antagonist methyllycaconitine (MLA) did not obstruct either the locomotor activity or DA overflow induced by systemic EtOH [81,98]. As α7 nAChRs are located on glutamatergic terminals in the VTA [21] which innervate both GABA and DA neurons [99], the effects of this blockade could cause changes in neuronal firing in the VTA local circuitry as well as projections to the NAc and PPTg. Interestingly, the α4β2 nAChR antagonist, DHβE also failed to block changes in DA levels recorded from the NAc when it was microinfused into the VTA [85]. Since a change of DA levels is normally found in response to EtOH [81,85], this failure is puzzling, as there is evidence of EtOH binding to α4β2 nAChRs in oocytes. Pretreatment with MEC significantly attenuated alcohol drinking in a rat limited access paradigm, but pretreatment with DHβE had no effect [100]. Thus, nAChRs are partially responsible for the reinforcing effects of EtOH, but the roles of both α4β2 and α7 nAChRs in the association between NIC and EtOH are unclear.

The α6 Subunit in Nicotinic Acetylcholine Receptors

Almost two decades ago, the α-conotoxin MII (α-CtxMII), derived from the Conus magnus marine snail, was identified and was shown to antagonize α3β2 containing nAChRs [101]. Soon thereafter, it was discovered that β2 knockout mice did not self-administer NIC, nor were they sensitive to NIC induced DA release in the NAc much unlike their wild-type opposites [102]. These data display that the β2 subunit in the VTA is necessary for the rewarding properties of NIC [103,104]. It is evident that the β2 subunit is critical for NIC reinforcement, but not when paired with the α4 subunit alone because DHβE does not block NIC induced DA effects in the NAc [85]. More recently, α6 knockout mice revealed that α-CtxMII binds to α6 containing nAChRs (α6*-nAChRs). The α6β2 containing pentamer rather than α3β2 pentamer was found to modulate NIC induced changes in DA systems [105]. In addition, the α4 subunit could play a role in NIC reward when paired with the α6 subunit [106-108]. Studies using immunoprecipitation discovered that not only were α6 and β2 subunits expressed in the same receptors, but the β3 subunit was also found in most α6*-nAChR pentamers in mesolimbic and nigrostriatal DA pathways [109,110]. β3 knockout mouse studies confirmed that this subunit plays a role in α-CtxMII binding [109,111-114], and this subunit may be involved in control of ion permeability and receptor location [82]. α-CtxMII administered in the VTA was able to reduce EtOH induced DA release in the NAc [115] and locomotor activity [116]. In addition, α6 knockout mice failed to self-administer NIC, while self-administration of the drug was restored with the reintroduction of the α6 subunit [117]. Fast-scan cyclic voltammetry studies have shown that α6β2 subunit containing nAChRs are responsible for the majority of NIC induced affects on DA release in the NAc [118]. In further behavioral studies, α-CtxMII perfusion into the VTA blocked recognition of EtOH associated cues [119] and voluntary EtOH drinking in rodents [115]. Genetic, electrophysiological, and pharmacological techniques have been employed to demonstrate functional α6*-nAChRs situated on GABA terminals innervating DA neurons in the VTA [120]. The combined data robustly propose α6 and β2 containing nAChRs are located on these terminals, however α4 subunits are not [120]. Therefore, the majority of α6*-nAChRs in the mesolimbic pathway are part of either an α6(1)α4(1)β2(2)β3(1) or an α6(2) β2(2)β3(1) heteromeric pentamer [104,109,113,121-123] with the later located on VTA GABAergic boutons [120]; both these receptors may have a significant role in the actions of NIC and EtOH.

Conclusion

The neural network underlying the interaction between NIC and EtOH is complex. Their interaction utilizes the mesolimbic DA system and the majority of its mediation takes place within the VTA. The VTA mediates rewarding effects of EtOH and aversive effects of NIC through the NAc, while the rewarding properties of NIC are meditated through the PPTg. The nAChR antagonist MEC has been shown to attenuate EtOH induced DA release in the NAc. However, both α7 (MLA) and α4β2 (DHβE) antagonists could not block this effect. The mixed results involving α7 and α4β2 nAChRs suggest that more research is needed in order to uncover their involvement in the mediation of EtOH reward. However, the α6*-nAChR antagonist α-CtxMII was helpful in the identification of the critical role α6 subunits have in the rewarding effects of both NIC and EtOH. Many types of nAChRs affect NIC-EtOH co-use, however, the α6(2)β2(2)β3(1) nAChR pentamers in the meslolimbic DA pathway situated on VTA GABA terminals are a likely site for NICEtOH interactions. Future research could target α6*-nAChRs in order to combat NIC-EtOH co-dependence.

References

  1. Benowitz NL (1996) Pharmacology of nicotine: addiction and therapeutics. Annu Rev Pharmacol Toxicol 36: 597-613.
  2. Gilpin NW, Koob GF (2008) Neurobiology of Alcohol Dependence: Focus on Motivational Mechanisms. Alcohol Res Health 31: 185-195.
  3. Mukamal KJ, Conigrave KM, Mittleman MA, Camargo CA Jr, Stampfer MJ, et al. (2003) Roles of drinking pattern and type of alcohol consumed in coronary heart disease in men. N Engl J Med 348: 109-118.
  4. Mukamal KJ, Maclure M, Muller JE, Mittleman MA (2005) Binge drinking and mortality after acute myocardial infarction. Circulation 112: 3839-3845.
  5. Tunstall-Pedoe H, Kuulasmaa K, Mähönen M, Tolonen H, Ruokokoski E, et al. (1999) Contribution of trends in survival and coronary-event rates to changes in coronary heart disease mortality: 10-year results from 37 WHO MONICA project populations. Monitoring trends and determinants in cardiovascular disease. Lancet 353: 1547-1557.
  6. Bolego C, Poli A, Paoletti R (2002) Smoking and gender. Cardiovasc Res 53: 568-576.
  7. Erhardt L (2009) Cigarette smoking: an undertreated risk factor for cardiovascular disease. Atherosclerosis 205: 23-32.
  8. Håheim LL, Holme I, Hjermann I, Leren P (1996) Smoking habits and risk of fatal stroke: 18 years follow up of the Oslo Study. J Epidemiol Community Health 50: 621-624.
  9. Hankey GJ (1999) Smoking and risk of stroke. J Cardiovasc Risk 6: 207-211.
  10. Danaei G, Ding EL, Mozaffarian D, Taylor B, Rehm J, et al. (2009) The preventable causes of death in the United States: comparative risk assessment of dietary, lifestyle, and metabolic risk factors. PLoS Med 6: e1000058.
  11. Miller NS, Gold MS (1998) Comorbid cigarette and alcohol addiction: epidemiology and treatment. J Addict Dis 17: 55-66.
  12. Bien TH, Burge R (1990) Smoking and drinking: a review of the literature. Int J Addict 25: 1429-1454.
  13. Grant BF, Hasin DS, Chou SP, Stinson FS, Dawson DA (2004) Nicotine dependence and psychiatric disorders in the United States: results from the national epidemiologic survey on alcohol and related conditions. Arch Gen Psychiatry 61: 1107-1115.
  14. Walitzer KS, Dearing RL (2013) Characteristics of alcoholic smokers, nonsmokers, and former smokers: personality, negative effect, alcohol involvement, and treatment participation. Nicotine & Tobacco Research: official journal of the Society for Research on Nicotine and Tobacco 15: 282-286.
  15. Batel P, Pessione F, Maître C, Rueff B (1995) Relationship between alcohol and tobacco dependencies among alcoholics who smoke. Addiction 90: 977-980.
  16. DiFranza JR, Guerrera MP (1990) Alcoholism and smoking. J Stud Alcohol 51: 130-135.
  17. Fouriezos G, Wise RA (1976) Pimozide-induced extinction of intracranial self-stimulation: response patterns rule out motor or performance deficits. Brain Res 103: 377-380.
  18. Liebman JM, Butcher LL (1973) Effects on self-stimulation behavior of drugs influencing dopaminergic neurotransmission mechanisms. Naunyn Schmiedebergs Arch Pharmacol 277: 305-318.
  19. Wise RA, Rompre PP (1989) Brain dopamine and reward. Annu Rev Psychol 40: 191-225.
  20. Wise RA (1998) Drug-activation of brain reward pathways. Drug Alcohol Depend 51: 13-22.
  21. Mansvelder HD, McGehee DS (2000) Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron 27: 349-357.
  22. Wise RA (1996) Addictive drugs and brain stimulation reward. Annu Rev Neurosci 19: 319-340.
  23. Mansvelder HD, Keath JR, McGehee DS (2002) Synaptic mechanisms underlie nicotine-induced excitability of brain reward areas. Neuron 33: 905-919.
  24. Blackburn JR, Phillips AG, Jakubovic A, Fibiger HC (1986) Increased dopamine metabolism in the nucleus accumbens and striatum following consumption of a nutritive meal but not a palatable non-nutritive saccharin solution. Pharmacol Biochem Behav 25: 1095-1100.
  25. Wise RA, Bozarth MA (1987) A psychomotor stimulant theory of addiction. Psychol Rev 94: 469-492.
  26. McKinzie DL, Rodd-Henricks ZA, Dagon CT, Murphy JM, McBride WJ (1999) Cocaine is self-administered into the shell region of the nucleus accumbens in Wistar rats. Ann N Y Acad Sci 877: 788-791.
  27. Pierce RC, Kumaresan V (2006) The mesolimbic dopamine system: the final common pathway for the reinforcing effect of drugs of abuse? Neurosci Biobehav Rev 30: 215-238.
  28. Yamaguchi T, Sheen W, Morales M (2007) Glutamatergic neurons are present in the rat ventral tegmental area. Eur J Neurosci 25: 106-118.
  29. Lammel S, Lim BK, Ran C, Huang KW, Betley MJ, et al. (2012) Input-specific control of reward and aversion in the ventral tegmental area. Nature 491: 212-217.
  30. Ikemoto S (2007) Dopamine reward circuitry: two projection systems from the ventral midbrain to the nucleus accumbens-olfactory tubercle complex. Brain Res Rev 56: 27-78.
  31. Sturgess JE, Ting AKRA, Podbielski D, Sellings LH, Chen JF, et al. (2010) Adenosine A1 and A2A receptors are not upstream of caffeine's dopamine D2 receptor-dependent aversive effects and dopamine-independent rewarding effects. The European Journal of Neuroscience 32: 143-154.
  32. Vashchinkina E, Panhelainen A, Vekovischeva OY, Aitta-aho T, Ebert B, et al. (2012) GABA site agonist gaboxadol induces addiction-predicting persistent changes in ventral tegmental area dopamine neurons but is not rewarding in mice or baboons. J Neurosci 32: 5310-5320.
  33. Goeders NE, Smith JE (1983) Cortical dopaminergic involvement in cocaine reinforcement. Science 221: 773-775.
  34. Mackey WB, van der Kooy D (1985) Neuroleptics block the positive reinforcing effects of amphetamine but not of morphine as measured by place conditioning. Pharmacol Biochem Behav 22: 101-105.
  35. Spyraki C, Fibiger HC, Phillips AG (1982) Cocaine-induced place preference conditioning: lack of effects of neuroleptics and 6-hydroxydopamine lesions. Brain Res 253: 195-203.
  36. Xi ZX, Stein EA (2002) GABAergic mechanisms of opiate reinforcement. Alcohol Alcohol 37: 485-494.
  37. Margolis EB, Toy B, Himmels P, Morales M, Fields HL (2012) Identification of rat ventral tegmental area GABAergic neurons. PLoS One 7: e42365.
  38. Ting-A-Kee R, Vargas-Perez H, Mabey JK, Shin SI, Steffensen SC, et al. (2013) Ventral tegmental area GABA neurons and opiate motivation. Psychopharmacology (Berl).
  39. Laviolette SR, Gallegos RA, Henriksen SJ, van der Kooy D (2004) Opiate state controls bi-directional reward signaling via GABAA receptors in the ventral tegmental area. Nat Neurosci 7: 160-169.
  40. Di Chiara G, North RA (1992) Neurobiology of opiate abuse. Trends Pharmacol Sci 13: 185-193.
  41. Robinson TE, Berridge KC (1993) The neural basis of drug craving: an incentive-sensitization theory of addiction. Brain Res Brain Res Rev 18: 247-291.
  42. Koob GF, Le Moal M (1997) Drug abuse: hedonic homeostatic dysregulation. Science 278: 52-58.
  43. Bechara A, van der Kooy D (1992) A single brain stem substrate mediates the motivational effects of both opiates and food in nondeprived rats but not in deprived rats. Behav Neurosci 106: 351-363.
  44. Nader KvdK, D. (1994) The motivation produced by morphine and food is isomorphic: approaches to specific motivational stimuli are learned. Psychobiology 22: 68-76.
  45. Koob GF, Volkow ND (2010) Neurocircuitry of addiction. Neuropsychopharmacology 35: 217-238.
  46. Söderpalm B, Ericson M, Olausson P, Blomqvist O, Engel JA (2000) Nicotinic mechanisms involved in the dopamine activating and reinforcing properties of ethanol. Behav Brain Res 113: 85-96.
  47. Brodie MS, Shefner SA, Dunwiddie TV (1990) Ethanol increases the firing rate of dopamine neurons of the rat ventral tegmental area in vitro. Brain Res 508: 65-69.
  48. Gallegos RA, Lee RS, Criado JR, Henriksen SJ, Steffensen SC (1999) Adaptive responses of gamma-aminobutyric acid neurons in the ventral tegmental area to chronic ethanol. J Pharmacol Exp Ther 291: 1045-1053.
  49. Brodie MS, Appel SB (1998) The effects of ethanol on dopaminergic neurons of the ventral tegmental area studied with intracellular recording in brain slices. Alcohol Clin Exp Res 22: 236-244.
  50. Gessa GL, Muntoni F, Collu M, Vargiu L, Mereu G (1985) Low doses of ethanol activate dopaminergic neurons in the ventral tegmental area. Brain Res 348: 201-203.
  51. Ikemoto S, Qin M, Liu ZH (2006) Primary reinforcing effects of nicotine are triggered from multiple regions both inside and outside the ventral tegmental area. The Journal of Neuroscience 26: 723-730.
  52. Rodd-Henricks ZA, McKinzie DL, Crile RS, Murphy JM, McBride WJ (2000) Regional heterogeneity for the intracranial self-administration of ethanol within the ventral tegmental area of female Wistar rats. Psychopharmacology (Berl) 149: 217-224.
  53. Laviolette SR, van der Kooy D (2003) The motivational valence of nicotine in the rat ventral tegmental area is switched from rewarding to aversive following blockade of the alpha7-subunit-containing nicotinic acetylcholine receptor. Psychopharmacology (Berl) 166: 306-313.
  54. Gatto GJ, McBride WJ, Murphy JM, Lumeng L, Li TK (1994) Ethanol self-infusion into the ventral tegmental area by alcohol-preferring rats. Alcohol 11: 557-564.
  55. Rodd ZA, Bell RL, Melendez RI, Kuc KA, Lumeng L, et al. (2004) Comparison of intracranial self-administration of ethanol within the posterior ventral tegmental area between alcohol-preferring and Wistar rats. Alcohol Clin Exp Res 28: 1212-1219.
  56. Saal D, Dong Y, Bonci A, Malenka RC (2003) Drugs of abuse and stress trigger a common synaptic adaptation in dopamine neurons. Neuron 37: 577-582.
  57. Laviolette SR, van der Kooy D (2003) Blockade of mesolimbic dopamine transmission dramatically increases sensitivity to the rewarding effects of nicotine in the ventral tegmental area. Mol Psychiatry 8: 50-59, 9.
  58. Lammel S, Ion DI, Roeper J, Malenka RC (2011) Projection-specific modulation of dopamine neuron synapses by aversive and rewarding stimuli. Neuron 70: 855-862.
  59. Yeomans JS, Mathur A, Tampakeras M (1993) Rewarding brain stimulation: role of tegmental cholinergic neurons that activate dopamine neurons. Behav Neurosci 107: 1077-1087.
  60. Lança AJ, Adamson KL, Coen KM, Chow BL, Corrigall WA (2000) The pedunculopontine tegmental nucleus and the role of cholinergic neurons in nicotine self-administration in the rat: a correlative neuroanatomical and behavioral study. Neuroscience 96: 735-742.
  61. Hunt T, Amit Z (1987) Conditioned taste aversion induced by self-administered drugs: paradox revisited. Neurosci Biobehav Rev 11: 107-130.
  62. Dagher A, Bleicher C, Aston JA, Gunn RN, Clarke PB, et al. (2001) Reduced dopamine D1 receptor binding in the ventral striatum of cigarette smokers. Synapse 42: 48-53.
  63. Nugent FS, Penick EC, Kauer JA (2007) Opioids block long-term potentiation of inhibitory synapses. Nature 446: 1086-1090.
  64. Stobbs SH, Ohran AJ, Lassen MB, Allison DW, Brown JE, et al. (2004) Ethanol suppression of ventral tegmental area GABA neuron electrical transmission involves N-methyl-D-aspartate receptors. J Pharmacol Exp Ther 311: 282-289.
  65. Alkondon M, Pereira EF, Cortes WS, Maelicke A, Albuquerque EX (1997) Choline is a selective agonist of alpha7 nicotinic acetylcholine receptors in the rat brain neurons. Eur J Neurosci 9: 2734-2742.
  66. Alkondon M, Pereira EF, Eisenberg HM, Albuquerque EX (1999) Choline and selective antagonists identify two subtypes of nicotinic acetylcholine receptors that modulate GABA release from CA1 interneurons in rat hippocampal slices. The Journal of Neuroscience 19: 2693-2705.
  67. Alkondon M, Pereira EF, Eisenberg HM, Albuquerque EX (2000) Nicotinic receptor activation in human cerebral cortical interneurons: a mechanism for inhibition and disinhibition of neuronal networks. The Journal of Neuroscience 20: 66-75.
  68. Dohrman DP, Reiter CK (2003) Ethanol modulates nicotine-induced upregulation of nAChRs. Brain Res 975: 90-98.
  69. de Fiebre CM, Marks MJ, Collins AC (1990) Ethanol-nicotine interactions in long-sleep and short-sleep mice. Alcohol 7: 249-257.
  70. Collins AC, Wilkins LH, Slobe BS, Cao JZ, Bullock AE (1996) Long-term ethanol and nicotine treatment elicit tolerance to ethanol. Alcohol Clin Exp Res 20: 990-999.
  71. Biala G, Weglinska B (2010) Rimonabant attenuates sensitization, cross-sensitization and cross-reinstatement of place preference induced by nicotine and ethanol. Pharmacol Rep 62: 797-807.
  72. Biała G, Weglińska B (2004) Calcium channel antagonists attenuate cross-sensitization to the locomotor effects of nicotine and ethanol in mice. Pol J Pharmacol 56: 391-397.
  73. Blomqvist O, Ericson M, Johnson DH, Engel JA, Söderpalm B (1996) Voluntary ethanol intake in the rat: effects of nicotinic acetylcholine receptor blockade or subchronic nicotine treatment. Eur J Pharmacol 314: 257-267.
  74. Ericson M, Blomqvist O, Engel JA, Söderpalm B (1998) Voluntary ethanol intake in the rat and the associated accumbal dopamine overflow are blocked by ventral tegmental mecamylamine. Eur J Pharmacol 358: 189-196.
  75. de Bruin NM, McCreary AC, van Loevezijn A, de Vries TJ, Venhorst J, et al. (2013) A novel highly selective 5-HT6 receptor antagonist attenuates ethanol and nicotine seeking but does not affect inhibitory response control in Wistar rats. Behavioural brain research 236: 157-165.
  76. Drews E, Zimmer A (2010) Modulation of alcohol and nicotine responses through the endogenous opioid system. Prog Neurobiol 90: 1-15.
  77. Proctor WR, Dobelis P, Moritz AT, Wu PH (2011) Chronic nicotine treatment differentially modifies acute nicotine and alcohol actions on GABA(A) and glutamate receptors in hippocampal brain slices. Br J Pharmacol 162: 1351-1363.
  78. Ribeiro-Carvalho A, Lima CS, Medeiros AH, Siqueira NR, Filgueiras CC, et al. (2009) Combined exposure to nicotine and ethanol in adolescent mice: effects on the central cholinergic systems during short and long term withdrawal. Neuroscience 162: 1174-1186.
  79. de Bruin NM, Lange JH, Kruse CG, Herremans AH, Schoffelmeer AN, et al. (2011) SLV330, a cannabinoid CB(1) receptor antagonist, attenuates ethanol and nicotine seeking and improves inhibitory response control in rats. Behav Brain Res 217: 408-415.
  80. Alijanpour S, Rezayof A, Zarrindast MR (2013) Dorsal hippocampal cannabinoid CB1 receptors mediate the interactive effects of nicotine and ethanol on passive avoidance learning in mice. Addict Biol 18: 241-251.
  81. Larsson A, Engel JA (2004) Neurochemical and behavioral studies on ethanol and nicotine interactions. Neurosci Biobehav Rev 27: 713-720.
  82. Gotti C, Clementi F (2004) Neuronal nicotinic receptors: from structure to pathology. Prog Neurobiol 74: 363-396.
  83. Narahashi T, Kuriyama K, Illes P, Wirkner K, Fischer W, et al. (2001) Neuroreceptors and ion channels as targets of alcohol. Alcohol Clin Exp Res 25: 182S-188S.
  84. Ribeiro-Carvalho A, Lima CS, Filgueiras CC, Manhães AC, Abreu-Villaça Y (2008) Nicotine and ethanol interact during adolescence: effects on the central cholinergic systems. Brain Res 1232: 48-60.
  85. Ericson M, Molander A, Löf E, Engel JA, Söderpalm B (2003) Ethanol elevates accumbal dopamine levels via indirect activation of ventral tegmental nicotinic acetylcholine receptors. Eur J Pharmacol 467: 85-93.
  86. Davis TJ, de Fiebre CM (2006) Alcohol's actions on neuronal nicotinic acetylcholine receptors. Alcohol Res Health 29: 179-185.
  87. Criswell HE, Simson PE, Duncan GE, McCown TJ, Herbert JS, et al. (1993) Molecular basis for regionally specific action of ethanol on gamma-aminobutyric acidA receptors: generalization to other ligand-gated ion channels. J Pharmacol Exp Ther 267: 522-537.
  88. Fröhlich R, Patzelt C, Illes P (1994) Inhibition by ethanol of excitatory amino acid receptors and nicotinic acetylcholine receptors at rat locus coeruleus neurons. Naunyn Schmiedebergs Arch Pharmacol 350: 626-631.
  89. Yang X, Criswell HE, Breese GR (1999) Action of ethanol on responses to nicotine from cerebellar interneurons and medial septal neurons: relationship to methyllycaconitine inhibition of nicotine responses. Alcohol Clin Exp Res 23: 983-990.
  90. Cardoso RA, Brozowski SJ, Chavez-Noriega LE, Harpold M, Valenzuela CF, et al. (1999) Effects of ethanol on recombinant human neuronal nicotinic acetylcholine receptors expressed in Xenopus oocytes. J Pharmacol Exp Ther 289: 774-780.
  91. Söderpalm B, Ericson M (2013) Neurocircuitry involved in the development of alcohol addiction: the dopamine system and its access points. Curr Top Behav Neurosci 13: 127-161.
  92. de Fiebre NC, de Fiebre CM (2005) alpha7 Nicotinic acetylcholine receptor knockout selectively enhances ethanol-, but not beta-amyloid-induced neurotoxicity. Neurosci Lett 373: 42-47.
  93. Yu D, Zhang L, Eiselé JL, Bertrand D, Changeux JP, et al. (1996) Ethanol inhibition of nicotinic acetylcholine type alpha 7 receptors involves the amino-terminal domain of the receptor. Mol Pharmacol 50: 1010-1016.
  94. Narahashi T, Aistrup GL, Marszalec W, Nagata K (1999) Neuronal nicotinic acetylcholine receptors: a new target site of ethanol. Neurochem Int 35: 131-141.
  95. Blomqvist O, Ericson M, Engel JA, Söderpalm B (1997) Accumbal dopamine overflow after ethanol: localization of the antagonizing effect of mecamylamine. Eur J Pharmacol 334: 149-156.
  96. Covernton PJ, Connolly JG (1997) Differential modulation of rat neuronal nicotinic receptor subtypes by acute application of ethanol. Br J Pharmacol 122: 1661-1668.
  97. Aistrup GL, Marszalec W, Narahashi T (1999) Ethanol modulation of nicotinic acetylcholine receptor currents in cultured cortical neurons. Mol Pharmacol 55: 39-49.
  98. Larsson A, Svensson L, Söderpalm B, Engel JA (2002) Role of different nicotinic acetylcholine receptors in mediating behavioral and neurochemical effects of ethanol in mice. Alcohol 28: 157-167.
  99. Carr DB, Sesack SR (2000) Projections from the rat prefrontal cortex to the ventral tegmental area: target specificity in the synaptic associations with mesoaccumbens and mesocortical neurons. The Journal of neuroscience: the official journal of the Society for Neuroscience 20: 3864-3873.
  100. Lê AD, Corrigall WA, Harding JW, Juzytsch W, Li TK (2000) Involvement of nicotinic receptors in alcohol self-administration. Alcohol Clin Exp Res 24: 155-163.
  101. Cartier GE, Yoshikami D, Gray WR, Luo S, Olivera BM, et al. (1996) A new alpha-conotoxin which targets alpha3beta2 nicotinic acetylcholine receptors. J Biol Chem 271: 7522-7528.
  102. Picciotto MR, Zoli M, Rimondini R, Léna C, Marubio LM, et al. (1998) Acetylcholine receptors containing the beta2 subunit are involved in the reinforcing properties of nicotine. Nature 391: 173-177.
  103. Maskos U, Molles BE, Pons S, Besson M, Guiard BP, et al. (2005) Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature 436: 103-107.
  104. Salminen O, Whiteaker P, Grady SR, Collins AC, McIntosh JM, et al. (2005) The subunit composition and pharmacology of alpha-Conotoxin MII-binding nicotinic acetylcholine receptors studied by a novel membrane-binding assay. Neuropharmacology 48: 696-705.
  105. Champtiaux N, Han ZY, Bessis A, Rossi FM, Zoli M, et al. (2002) Distribution and pharmacology of alpha 6-containing nicotinic acetylcholine receptors analyzed with mutant mice. The Journal of Neuroscience: Official Journal of the Society for Neuroscience 22: 1208-1217.
  106. Drenan RM, Grady SR, Steele AD, McKinney S, Patzlaff NE, et al. (2010) Cholinergic modulation of locomotion and striatal dopamine release is mediated by alpha6alpha4* nicotinic acetylcholine receptors. The Journal of neuroscience: Official Journal of the Society for Neuroscience 30: 9877-9889.
  107. Zhao-Shea R, Liu L, Soll LG, Improgo MR, Meyers EE, et al. (2011) Nicotine-mediated activation of dopaminergic neurons in distinct regions of the ventral tegmental area. Neuropsychopharmacology 36: 1021-1032.
  108. Hendrickson LM, Gardner P, Tapper AR (2011) Nicotinic acetylcholine receptors containing the α4 subunit are critical for the nicotine-induced reduction of acute voluntary ethanol consumption. Channels (Austin) 5: 124-127.
  109. Cui C, Booker TK, Allen RS, Grady SR, Whiteaker P, et al. (2003) The beta3 nicotinic receptor subunit: a component of alpha-conotoxin MII-binding nicotinic acetylcholine receptors that modulate dopamine release and related behaviors. The Journal of neuroscience : the official journal of the Society for Neuroscience 23: 11045-11053.
  110. Quik M, Polonskaya Y, Gillespie A, Jakowec M, Lloyd GK, et al. (2000) Localization of nicotinic receptor subunit mRNAs in monkey brain by in situ hybridization. J Comp Neurol 425: 58-69.
  111. Gotti C, Clementi F, Fornari A, Gaimarri A, Guiducci S, et al. (2009) Structural and functional diversity of native brain neuronal nicotinic receptors. Biochem Pharmacol 78: 703-711.
  112. Gotti C, Guiducci S, Tedesco V, Corbioli S, Zanetti L, et al. (2010) Nicotinic acetylcholine receptors in the mesolimbic pathway: primary role of ventral tegmental area alpha6beta2* receptors in mediating systemic nicotine effects on dopamine release, locomotion, and reinforcement. The Journal of neuroscience : the official journal of the Society for Neuroscience 30: 5311-5325.
  113. Zoli M, Moretti M, Zanardi A, McIntosh JM, Clementi F, et al. (2002) Identification of the nicotinic receptor subtypes expressed on dopaminergic terminals in the rat striatum. The Journal of neuroscience : the official journal of the Society for Neuroscience 22: 8785-8789.
  114. Quik M, Vailati S, Bordia T, Kulak JM, Fan H, et al. (2005) Subunit composition of nicotinic receptors in monkey striatum: effect of treatments with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine or L-DOPA. Mol Pharmacol 67: 32-41.
  115. Larsson A, Jerlhag E, Svensson L, Söderpalm B, Engel JA (2004) Is an alpha-conotoxin MII-sensitive mechanism involved in the neurochemical, stimulatory, and rewarding effects of ethanol? Alcohol 34: 239-250.
  116. Jerlhag E, Grøtli M, Luthman K, Svensson L, Engel JA (2006) Role of the subunit composition of central nicotinic acetylcholine receptors for the stimulatory and dopamine-enhancing effects of ethanol. Alcohol Alcohol 41: 486-493.
  117. Pons S, Fattore L, Cossu G, Tolu S, Porcu E, et al. (2008) Crucial role of alpha4 and alpha6 nicotinic acetylcholine receptor subunits from ventral tegmental area in systemic nicotine self-administration. The Journal of Neuroscience: the official journal of the Society for Neuroscience 28: 12318-12327.
  118. Exley R, Clements MA, Hartung H, McIntosh JM, Cragg SJ (2008) Alpha6-containing nicotinic acetylcholine receptors dominate the nicotine control of dopamine neurotransmission in nucleus accumbens. Neuropsychopharmacology 33: 2158-2166.
  119. Löf E, Olausson P, deBejczy A, Stomberg R, McIntosh JM, et al. (2007) Nicotinic acetylcholine receptors in the ventral tegmental area mediate the dopamine activating and reinforcing properties of ethanol cues. Psychopharmacology (Berl) 195: 333-343.
  120. Yang K, Buhlman L, Khan GM, Nichols RA, Jin G, et al. (2011) Functional nicotinic acetylcholine receptors containing a6 subunits are on GABAergic neuronal boutons adherent to ventral tegmental area dopamine neurons. J Neurosci 31: 2537-2548.
  121. Champtiaux N, Gotti C, Cordero-Erausquin M, David DJ, Przybylski C, et al. (2003) Subunit composition of functional nicotinic receptors in dopaminergic neurons investigated with knock-out mice. The Journal of Neuroscience 23: 7820-7829.
  122. Charpantier E, Barnéoud P, Moser P, Besnard F, Sgard F (1998) Nicotinic acetylcholine subunit mRNA expression in dopaminergic neurons of the rat substantia nigra and ventral tegmental area. Neuroreport 9: 3097-3101.
  123. Quik M, McIntosh JM (2006) Striatal alpha6* nicotinic acetylcholine receptors: potential targets for Parkinson's disease therapy. J Pharmacol Exp Ther 316: 481-489.
Citation: Taylor DH, Steffensen SC, Wu J (2013) Nicotinic Acetylcholine Receptors in the Ventral Tegmental Area are Important Targets for Nicotine and Ethanol Codependence. Biochem Pharmacol (Los Angel) S1:002.

Copyright: © 2013 Taylor DH, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top