Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jun 18.
Published in final edited form as: Neurotox Res. 2011 Sep 20;21(1):57–69. doi: 10.1007/s12640-011-9275-6

Positive and Negative Effects of Alcohol and Nicotine and Their Interactions: A Mechanistic Review

Laura L Hurley 1, Robert E Taylor 1, Yousef Tizabi 1,
PMCID: PMC3377362  NIHMSID: NIHMS371434  PMID: 21932109

Abstract

Nicotine and alcohol are two of the most commonly abused legal substances. Heavy use of one drug can often lead to, or is predictive of, heavy use of the other drug in adolescents and adults. Heavy drinking and smoking alone are of significant health hazard. The combination of the two, however, can result in synergistic adverse effects particularly in incidences of various cancers (e.g., esophagus). Although detrimental consequences of smoking are well established, nicotine by itself might possess positive and even therapeutic potential. Similarly, alcohol at low or moderated doses may confer beneficial health effects. These opposing findings have generated considerable interest in how these drugs act. Here we will briefly review the negative impact of drinking–smoking co-morbidity followed by factors that appear to contribute to the high rate of co-use of alcohol and nicotine. Our main focus will be on what research is telling us about the central actions and interactions of these drugs, and what has been elucidated about the mechanisms of their positive and negative effects. We will conclude by making suggestions for future research in this area.

Keywords: Alcoholism, Smoking, Alcohol, Nicotine

Introduction

Two of the biggest threats to world health come from the negative effects of using tobacco and alcohol. It has become clear over the last few decades that heavy use of tobacco and/or alcohol leads to serious health consequences such as development of cardio- and cerebro-vascular diseases, gastric ulcers, various cancers, particularly those of the head, neck, esophagus, and even liver (Castellsagué et al. 1999; Franceschi et al. 1990; Johnson and Jennison 1992; Ko and Cho 2000; Koob and Le Moal 2006; Mitrouska et al. 2007; Olsen et al. 1985; Pelucchi et al. 2008). The high use of alcohol and tobacco products likely stems in part from their abundant and legal availability, but also from other factors that can lead to users’ abusive and addictive use. In recent years it has been elucidated that abuse of alcohol and nicotine (from tobacco products) can be attributed in part to genetic, rewarding, and possibly the analgesic effects the drugs have. Furthermore, high incidence of co-morbidity can be linked by these same factors, as well as possible pharmacokinetic and pharmacodynamic interactions (e.g., enhancement of the rewarding and analgesic effects) and counteracting mechanisms that co-use affords the consumer. This article first reviews and evaluates current evidence of these contributory factors to co-use, then looks at how dosing should be considered when evaluating the impacts of these drugs. The main focus will be on central effects of alcohol and nicotine, and how dosing level and combination of these drugs can alter the effects of both drugs. Finally, we will comment on how these research findings potentially impact suggestions for cessation treatment for addicts and general heath management, and suggest lines of study that need to be pursued.

Contributory Factors to Alcohol and Tobacco Co-morbidity

Epidemiological studies have demonstrated that in adolescents and adults high rates of smoking correlate highly with alcohol use, with smoking rates in alcoholics estimated to be at least two times higher than the general population (Falk et al. 2006), and the rate of cigarette consumption to be higher in alcoholic smokers than nonalcoholic smokers (Dawson 2000). Research involving this phenomenon has given rise to four prominent explanations for the high incident rate of alcohol and tobacco co-use. The depth of research in each field is varied, but begins to paint a picture of how alcohol and nicotine possibly interact to contribute to their co-morbidity.

Genetic Influence

In recent years studies have focused on the nature side of addiction over the nurture, by following addiction rates in twins (mono- vs. dizygotic) and siblings raised in different environments (Ball 2008; Enoch and Goldman 2001). It has become clear that there is a strong genetic component to addiction, which may account for addictive behavior more than environmental influences. Current estimations for heritability of all major addictive disorders range between 40 and 80% (Goldman et al. 2005). This suggests strongly that there is an underlying genetic vulnerability for alcohol–nicotine co-morbidity. It is estimated that genetic factors account for about 50% of nicotine or alcohol dependence (Maes et al. 1999; Sartor et al. 2010; True et al. 1999), but this number can vary with gender, race, culture, and religion. The influence of genetic factors on addiction to these drugs appears to vary with the age that drinking or smoking was initiated (Rose 1998), hinting on a possible epigenetic influence. Alcohol–nicotine co-morbidity appears to be equally dependant on genetic factors in men and women, but the same does not appear to be true in nonaddicted users (males more genetically predicated: Han et al. 1999; Kendler et al. 1994, 2000).

Genetic studies of addiction have focused on dysfunctions of several neurotransmitter systems that in some way contribute to alteration of the reward (e.g., dopaminergic) or mood (e.g., glutamatergic, opioidergic, serotonergic, or cholinergic) pathways in the brain. The hypothesis being that alteration to these systems could alter an individual’s experience or tolerance to a drug, and therefore change the likelihood that the person becomes addicted to or have difficulty quitting. However, the search for candidate genes in these pathways (via various methods such as Whole Genome Association) has turned up many genes on various chromosomes that act on and alter a variety of functions including enzymatic activity, protein translation, transcriptional regulation, and receptor function. Changes in these functions could alter susceptibility to alcohol, nicotine, or drug addiction in general (Ducci and Goldman 2008; Uhl 2004; Uhl et al. 2008). Still the genetic basis of alcohol and nicotine addiction is unknown. Some studies suggest that even subtle changes in genetic makeup (different strains of rats) can alter sensitivity to drugs and the rewarding release of dopamine (see “Reward Pathway Activation” section: Cadoni et al. 2009). However, the most promising have been studies of mutations of genes involved in alcohol metabolism (e.g., ADH/ALDH: alcohol/aldehyde dehydrogenase), nicotine metabolism (e.g., CYP2A6: cytochrome P450 2A6) or nicotinic receptor.

It is the belief of some scientists that polymorphisms in ADH/ALDH are the only ones that truly alter the risk of alcoholism (Crabbe et al. 2006). ADH and ALDH are enzymes that metabolize alcohol to acetaldehyde and then to acetate, respectively. Isoform variation of either gene appears to reduce the risk of alcoholism, with additive effects when both vary (Ball 2008; Dick and Foroud 2003; Köhnke 2008). It is unclear why variations cause decreased alcohol abuse and alcoholism, but it might be in part due to increased negative side effects of the alcohol consumption. These side effects are likely tied to accumulation of acetaldehyde, leading to acetaldehyde syndrome-like symptoms, including palpitations, skin flushing, nausea, and severe hangovers (Ball 2008; Luczak et al. 2002; Müller et al. 2010; Tu and Israel 1995). These polymorphisms appear most predominantly in East Asian populations, but studies are finding similar mutations in ADH and ALDH that likely alter rates of alcoholism in non-Asian populations (Liu et al. 2011).

CYP2A6 is part of an enzyme subfamily involved in the oxidative metabolism and detoxification of various compounds. Therefore, polymorphisms within this subfamily can dramatically alter drug effects (Johansson and Ingelman-Sundberg 2011). CYP2A6 specifically, is the primary enzyme responsible for the oxidation of nicotine to cotinine (Benowitz and Jacob 1994; Messina et al. 1997; Nakajima et al. 1996a, b). Polymorphisms of CYP2A6 alters nicotine metabolism resulting in increased or decreased smoking rate depending on the mutation. For example, individuals with inactivated CYP2A6 are less likely to smoke or smoke very little (Messina et al. 1997; Pianezza et al. 1998; Rao et al. 2000), whereas duplication of the gene increases smoking (Rao et al. 2000). This suggests that length of time nicotine is bioavailable in the system due to CYP2A6 mutations can dramatically influence smoking rates and possibly age of smoking initiation and cancer risk (Ariyoshi et al. 2002; Thorgeirsson et al. 2010).

Recent exciting discoveries emanating from genome wide association studies have identified polymorphisms localized to chromosome 15q24 (gene cluster encoding α3, α5, β4 nicotinic acetylcholine receptor (nAChRs) subunits) variants of which are associated with nicotine dependence as well as risk for lung cancer and peripheral arterial disease (Berrettini et al. 2008; Thorgeirsson et al. 2008). Moreover, emerging evidence suggest that nAChRs containing α5, α6, β3, and β4 subunits regulate nicotine intake (Fowler et al. 2011; Frahm et al. 2011; Hoft et al. 2009b). Interestingly, polymorphisms within the cluster encoding for nAChRs are also associated with alcohol dependence and early initiation of alcohol use (Hoft et al. 2009a; Schlaepfer et al. 2008; Wang et al. 2009). See below for more detailed discussion of nicotinic receptors.

Nicotinic Receptors

Nicotinic receptors belong to ionotropic class of receptors. These receptors act by regulating directly the opening of a cation channel in the neuronal membrane (see Changeux et al. 1998; Gotti and Clementi 2004; Rathouz et al. 1996; Wu and Lukas 2011 for reviews). Considerable information on interaction between these receptors and other neurotransmitter systems is now available. Indeed, therapeutic potentials for selective nicotinic receptor agonists in various neuropsychiatric and neurodegenerative disorders have been suggested. Various subtypes of these receptors with distinct anatomical, physiological, and pharmacological characteristics have been identified (see reviews by: Albuquerque et al. 1995; Changeux et al. 1998; Clarke 1995; Conti-Fine et al. 1995; Gotti and Clementi 2004; Lukas and Bencherif 1992; Miwa et al. 2011; Olale et al. 1997; Picciotto 2003; Wu and Lukas 2011). The most predominant and most extensively studied subtype in the brain has a high affinity for cytisine, nicotine or acetylcholine and is formed from α4 to β2 subunits (see Clarke et al. 1985; Flores et al. 1992; Pabreza et al. 1991). This subtype is commonly referred to as high-affinity binding site. The other major class with a high affinity for α-bungarotoxin but low affinity for nicotine is formed from α7 subunits and can be labeled by [125I]α-bungarotoxin. This subtype is commonly referred to as low-affinity binding site. It should be noted that [125I]α-bungarotoxin also binds with high affinity to neuro-muscular nicotinic receptors and in some cases to ganglionic nicotinic receptors (Zigmond and Loring 1988). However, the subunit structures of the nicotinic receptors in the muscle are different from those in the ganglia which are different from those in the CNS.

Further distinction between nicotinic receptor subtypes is evident in their central distribution as well as their physiological roles. For example, [125I]α-bungarotoxin binding sites in the brain are most abundant in hippocampus (see Clarke et al. 1985 for detailed distribution) and are believed to have a prominent role in neuronal growth and survival (De Fiebre et al. 1995). Furthermore, these receptors appear to be involved in cognitive functions, particularly attentional processes (Freedman et al. 1997; Matsuyama et al. 2001; Miwa et al. 2011). A role for α7 receptor subtype in central reward pathway has also been suggested (Schilström et al. 1998; Jones and Wonnacott 2004). High-affinity nicotinic receptors (e.g., α4–β2 or α3 containing receptors), on the other hand, are more prominent in mesolimbic or nigro-striatal pathways and appear to be more involved in rewarding or addictive behavior, locomotor activity and antinociception (Changeux et al. 1998; Damaj et al. 1998; Lindstrom 1997; Olale et al. 1997; Stolerman et al. 1997). Both receptors appear to be involved in neuroprotection as well (Belluardo et al. 2000; Picciotto et al. 2000; Tizabi et al. 2003, 2004). Studies with Flinders and WKY rat models of depression and Fawn-Hooded rat model of alcoholism and depression also suggest involvement of high-affinity nicotinic receptor subtypes in these behavioral characteristics (Tizabi et al. 1999, 2000, 2009).

Epigenetic Influence

Another consideration is that changes are not occurring within the genes themselves, but through activation or deactivation of more than one gene. In other words, it is possible that environmental factors are causing modulation of gene activation through epigenetic changes (i.e., change in gene expression controlled by alteration of DNA methylation and/or chromatin structure). Strong evidence supports the idea that epigenetic mechanisms exist that can alter gene expression in neurons that can lead to changes in behavior including psychiatric disorders such as depression, schizophrenia, and drug addiction (reviewed Tsankova et al. 2007). In terms of drug addiction, these epigenetic changes may provide reasons for why addictive behaviors arise, persist long after drug cessation, and have high rates of relapse. Epigenetic effects of drug abuse has been extensively studied in cocaine (Renthal and Nestor 2008; reviewed Tsankova et al. 2007), but poorly studied in alcohol and nicotine. Nicotine has been shown to have longterm effects after intrauterine exposure: both immediate (Dickson et al. 2011; Lotfipour et al. 2009; Toledo-Rodriguez et al. 2010) and transgenerational (Holloway et al. 2007). Furthermore, it has been seen in humans that nicotine has lasting effects on demethylation of monoamine oxidase-B promoters (altering serotonin metabolism: Launay et al. 2009), and is associated with high CYP2A6 activity due to full demethylation of the gene site (Al Koudsi et al. 2010). Recent studies, using a mouse model to study nicotine abuse in schizophrenics, have demonstrated the likelihood of nicotine playing a role in regulating epigenetic alterations of GABAergic neurons seen in this disease (Maloku et al. 2011). The link may be through activation or alteration of the α4β2 nicotinic receptor subtype (Maloku et al. 2011). Ethanol has been more directly studied, with both chronic and acute exposure causing chromatin changes (Bönsch et al. 2005; Kim and Shukla 2006; Mahadev and Vemuri 1998). Chronic alcohol exposure leads to changes in protein expression in astrocytes and neurons (Mahadev and Vemuri 1998), and withdrawal from chronic exposure can cause chromatin remodeling that leads to withdrawal related anxiety (Pandey et al. 2008; Moonat et al. 2010). In one study no changes were seen in the brain after acute alcohol exposure, but other tissues did show epigenetic changes via histone modification (Kim and Shukla 2006). Another study showed anxiolytic effects of alcohol, possibly in relation to chromatin modification (Pandey et al. 2008). Studies in Drosophila have shown that acute exposure to alcohol caused gene regulation changes that altered drug sensitivity (Ghezzi et al. 2004; Wang et al. 2007). Collectively, these findings suggest epigenetic/genetic mechanisms by which drug tolerance and dependence may develop.

Pharmacokinetic Interactions

There are several pharmacological interactions between alcohol and nicotine that may explain how use of one drug might lead to increased co-use of the other. For example, nicotine and alcohol interact to alter the effect either drug has alone, and chronic use of either or both drugs alters this interaction (reviewed Lajtha and Sershen 2010). Aside from the pharmacokinetic interactions between these two drugs, where either drug may alter the peripheral metabolism of the other (Parnell et al. 2006; Yue et al. 2009), a primary contributory factor to co-use of alcohol and nicotine, is likely the additive or synergistic activation of the reward system. In addition, the analgesic properties of the combination and possible counteraction of adverse effects when the drugs are used together may also promote co-use of alcohol and nicotine.

Reward Pathway Activation

A number of animal and some human studies, provide strong evidence that alcohol and nicotine are interacting to potentiate the rewarding effects of one another through activation of the dopamine reward pathway. In short, this pathway is comprised of dopaminergic neurons in the ventral tegmental area (VTA) which possess nAChRs, that when stimulated cause release of dopamine in several brain areas including the nucleus accumbens (NAcc). Activation of the NAcc has been implicated in changes in emotional and cognitive behaviors, especially in relation to regulation of reward-induced addiction (reviewed in Ikemoto and Panksepp 1999; Willuhn et al. 2010). Specifically, nicotine and ethanol have been shown to increase dopaminergic neuron firing (Kleijn et al. 2011; Mereu et al. 1984, 1987), and dopamine release (Kleijn et al. 2011; Löf et al. 2007; Nisell et al. 1994a, b; Tizabi et al. 2002, 2007).

Nicotine appears to activate mesolimbic reward pathway through nAChRs (especially α4/α6 β2 and α7 subunits: Exley et al. 2011; Gotti et al. 2010; Mansvelder and McGehee 2000; Maskos et al. 2005; Pidoplichko et al. 1997; Pons et al. 2008). Moreover, the sensitivity to nicotine reward may be modulated by dopamine signaling through its interactions with D1 and D2 receptors (Laviolette et al. 2008). However, the exact pre- or post-synaptic nicotinic receptors that activate dopamine release are not known.

Alcohol (ethanol) may influence the reward pathway by a number of mechanisms including interactions with nicotinic receptors (Blomqvist et al. 1993, 1996; Jerlhag et al. 2006; Larsson and Engel 2004). Co-administration of alcohol and nicotine produces an additive release of dopamine in the NAcc (Tizabi et al. 2002, 2007) which may reflect additive rewarding effect and hence a possible mechanism contributing to the co-use of alcohol and nicotine. It is also conceivable and likely that interactions of alcohol and nicotine with other transmitter systems may influence their individual as well as their combined rewarding and addictive nature.

In humans it has been reported that smoking of normal nicotine cigarettes versus denicotinized cigarettes increased alcohol consumption (Barrett et al. 2006), and that alcohol consumption can increase self-reported pleasure derived from cigarette smoking (Rose et al. 2004). The strength of this effect is dynamic and can be altered by a number of factors including age and gender (Acheson et al. 2006; Grant et al. 2006). These influences may stem from alteration of nAChR activation as the brain ages and after long-term smoking (Hellström-Lindahl and Court 2000; Teaktong et al. 2003, 2004), especially if changes occur in the dopamine reward pathway. Chronic smoking globally increases nAChRs, probably due to desensitization to repeated nicotine exposure instead of enhanced receptor function (Grady et al. 1994; Ke et al. 1998; Marks et al. 1993; Sabbagh et al. 2002; Wonnacott 1990), although in some unique cases the up-regulation of receptors may also be associated with an increase in function (Nguyen et al. 2004). One theory postulates that desensitization occurs to protect neurons from uncontrolled excitation, as mice with mutated α7 receptor show decreased desensitization and an increased rate of mortality (Mudo et al. 2007; Wang and Sun 2005). Further research is needed to understand this phenomenon, as desensitization and alteration in numbers vary with receptor subtype and location (Buisson and Bertrand 2001; Fenster et al. 1997; Gentry and Lukas 2002; Olale et al. 1997; Perry et al. 2007; Picciotto et al. 2008; Teaktong et al. 2003, 2004; Yu and Wecker 1994).

It has been shown in rats with moderate ethanol preference that consumption of ethanol increases after nicotine treatment (Blomqvist et al. 1996), and application of a nicotinic antagonist reduces ethanol intake (Bell et al. 2009). This response may be from the fact that ethanol and nicotine together appear to increase and sustain dopamine release significantly over that seen when they are administered alone (Tizabi et al. 2002, 2007). However, pre-treatment with mecamylamine, a nicotinic receptor antagonist, blocked the additive effect of combined alcohol and nicotine on dopamine release (Tizabi et al. 2007), and decreased ethanol intake in rats (Blomqvist et al. 1996; Ericson et al. 1998; Le et al. 2000). Similarly, in some human studies it was shown that use of mecamylamine caused a decrease in euphoric effects of alcohol and the desire to consume it (Chi and de Wit 2003; Young et al. 2005). More recently it has been demonstrated that the FDA approved varenicline, a nicotinic partial agonist at α4–β2 subtype, for smoking cessation may also be effective in reducing alcohol intake as suggested by preclinical as well as limited clinical trials (Ericson et al. 2009; Fucito et al. 2011; Hendrickson et al. 2010; Kamens et al. 2010; McKee et al. 2009; Steensland et al. 2007). Thus, further elucidation of the roles of specific nicotinic receptors in pathways involved in reward enhancing effects of alcohol–nicotine co-use could lead to novel interventions in drinking–smoking co-morbidity (see also Chatterjee and Bartlett 2010).

Analgesia

Pain is a broad term that now refers to both a physical and emotional experience (Price 2000). Beyond the activation of nAChRs and possible dopamine re-enforced pleasure in alcohol–nicotine co-use another contributory factor could be the analgesic (or antinociceptive) effect of such a combination. Nicotine in itself appears to be effective as a reducer of pain and inflammation in many situations (Decker et al. 2001; Yagoubian et al. 2011; Yoshikawa et al. 2006). Nicotine’s analgesic ability was first demonstrated before the identification of nAChRs (Davis et al. 1932), but the activation of the nAChRs has long been suggested as a key component in mediation of analgesic effects of nicotine and other nicotinic agonists (e.g., epibatidine, see below: Damaj et al. 1999; Khan et al. 1998; Simons et al. 2005; Vincler 2005). Many brain nuclei express a diverse number of nAChR subtypes, and areas regulating pain are no exception (Millar and Gotti 2009; Tracey and Mantyh 2007). Nicotine activation of nAchRs containing α4, β2, and possibly α7 subunits appear to be a key part to regulating pain (Damaj et al. 2000; Flores 2000; Gao et al. 2010). In support of this idea, mice with α4 and β2 mutations demonstrate no analgesic effect to nicotine administration (Marubio et al. 1999). Treatment with α4 and β2 agonists show similar, although not complete, analgesic effects to nicotine administration (Flores 2000; Gao et al. 2010), but administration of epibatidine and ABT-594, both α4β2 agonists, produce effects 200 times more potent than morphine (Bannon et al. 1998; Spande et al. 1992). Although it has been suggested that the analgesic effect that comes from activation of the nAChR receptors by nicotine may be due to opioid release (Galeote et al. 2006), the majority or finding suggest that it is independent of the opioid system when given alone (Campbell et al. 2006; Cooley et al. 1990; Damaj et al. 1999; Khan et al. 1998; Rogers and Iwamoto 1993; Tripathi et al. 1982).

Alcohol has historically been used to dull central nervous system responses to pain before development of more sophisticated pain medications. Basic pain tolerance studies have shown that even low doses of alcohol work to alleviate physical pain (James et al. 1978; Perrino et al. 2008; Woodrow and Eltherington 1988). This antinociceptive effect of alcohol appears to be primarily regulated by the opioid system (Boada et al. 1981; Campbell et al. 2006), and may be influenced by factors such as family history of alcoholism and personality (Ralevski et al. 2010). In fact, individual history of alcohol use may be key in how it works as an analgesic, as continued alcohol use may affect pain modulation by altering dopamine and opioid function (Cowen et al. 2004; Cutter and O’Farrell 1987; Koob et al. 1998; Vanderah et al. 2001). This contention is supported by human studies showing a positive relationship between chronic pain and substance abuse, and that alcoholics are more sensitive to pain—especially when sober—compared to non-alcoholics (Askay et al. 2009; Brown and Cutter 1977; Rosenblum et al. 2003). The altered analgesic effect in those that are addicted to nicotine or alcohol, may contribute to co-use of both drugs for their analgesic properties.

Simultaneous administration of alcohol and nicotine causes additive or synergistic analgesic effects (Franklin 1989, 1998), and appears to occur primarily—but not exclusively—by activation of the opioid system (Campbell et al. 2006). Alleviation of pain—emotional and physical—may be what leads to increased use and co-abuse of nicotine and alcohol. This increased use combined with developing tolerance—especially if the user becomes fully or partially cross-tolerant (i.e., the chronic use of one substance resulting in tolerance to effects of the other: Burch et al. 1988; Collins et al. 1988; de Fiebre and Collins 1993)—may be a key component in making the user vulnerable to addiction. Co-use of alcohol and nicotine (smoking) may therefore arise from a user seeking to find additional pain relief.

Counteracting Mechanisms

In addition to the above, one contributing factor to comorbid use of alcohol and nicotine may be the counteracting of adverse effects of one drug by the other drug. Thus, certain deleterious effects of alcohol (e.g., cognitive impairment, subjective intoxication, and sedating properties) appear to be alleviated by tobacco’s nicotine (Ceballos 2006; Perkins et al. 1995). Early studies of this interaction showed that smoking improved alcohol-induced cognitive impairment by shortening temporal distortion in the form of overestimation of time elapsed and improving decrements in divided attention tasks (Leigh and Tong 1976; Leigh et al. 1977). This decrease in negative effects of alcohol by nicotine may actually lead to an increase in alcohol consumption and dependence (Schuckit and Smith 2004). Specific testing of alcohol-induced impairment shows that nicotine in the cigarette smoke is likely what causes the amelioration of impairments or deficits (Al-Rejaie and Dar 2006; Dar et al. 1993, 1994; Gould et al. 2001; Taslim et al. 2011; Tracy et al. 1999). Chronic smokers report feeling less intoxicated than non-smokers after consumption of the same dose of alcohol (Madden et al. 2000). The beneficial effect of nicotine may be limited to acute exposure dosing (Craddock et al. 2003; Ernst et al. 2001; Lawrence et al. 2002; Rezvani and Levin 2001), and may be lost in chronic co-morbid users (Cervilla et al. 2000; Durazzo et al. 2010; Glass et al. 2006, 2009). Thus, chronic consumers of combination of alcohol and cigarettes may show increased cognitive impairment (Durazzo et al. 2010; Glass et al. 2006, 2009). Even acute dosing of nicotine in some animal studies appears to work synergistically with alcohol to cause impairment at doses where alcohol alone caused little to no deficit (Bizarro et al. 2003; Rezvani and Levin 2002). The variations seen in experimentation, and between acute and chronic dosing, are likely due to differential pharmacokinetic interactions causing cross-tolerance or sensitization (Gulick and Gould 2008). A recent study has revealed that α4β2 and α7 nAChR subtypes are key in behavioral cross-tolerance between nicotine and ethanol-induced ataxia, with nitric oxide signaling being a potential mechanism to this effect (Taslim et al. 2011).

Dosing Consideration

Chronic smoking (the primary route of tobacco consumption) and drinking both have well documented long-term impacts on cardiovascular system and smoking in particular, causes a variety of obstructive airway diseases (Koob and Le Moal 2006; Mitrouska et al. 2007). Both habits also have been attributed to cerebrovascular diseases such as transient ischemic attacks and stroke, and can increase the risk of a number of cancers (Castellsagué et al. 1999; Franceschi et al. 1990; Johnson and Jennison 1992; Ko and Cho 2000; Koob and Le Moal 2006; Mitrouska et al. 2007; Olsen et al. 1985; Pelucchi et al. 2008). The multitude and severity of these detriments can be synergistically exacerbated when the two are combined. However, the majority of the populations are not heavy alcohol and/or nicotine consumers, and on average would have only low-level exposure to one or both compounds. Recent evidence shows that alcohol or nicotine alone at low to moderate levels can have health and therapeutic benefits. Specifically, both compounds at low concentrations have demonstrated the ability to act as neuroprotectants (Belmadani et al. 2004; Collins et al. 2009; Ferrea and Winterer 2009; Quik et al. 2008; Ramlochansingh et al. 2011; Tizabi et al. 2003, 2005, 2007; see also below). However, when low doses of alcohol and nicotine are combined this protection appears to be reversed in at least some in vitro studies (Ramlochansingh et al. 2011; Smith et al. 2006). Although, further in vivo interactions between these two compounds are necessary, it is noteworthy that a study in adolescent mice did not detect toxicity by co-administration of alcohol and nicotine (Oliveira-da-Silva et al. 2009, 2010). Thus, it is important to elucidate the extent of acute and chronic interactions between alcohol and nicotine in various in vitro and in vivo paradigms.

Nicotine in low doses has been seen in in vitro cell (primary and immortal) cultures to protect against or attenuate damage induced by β-amyloid, lipopolysaccharide (LPS)-induced inflammation, glutamate, alcohol, N-methyl-d-aspartate (NMDA), hypoxia, and salsolinol (Copeland et al. 2005, 2007; Dajas-Bailador et al. 2002; Das and Tizabi 2009; Guan et al. 2003; Hejmadi et al. 2003; Kihara et al. 1998; Liu and Zhao 2004; Park et al. 2007; Ramlochansingh et al. 2011; Stevens et al. 2003; Tizabi et al. 2003, 2004). The action of this protection is unclear, but it appears to be mediated by activation of nicotinic receptors (discussed above: Dajas-Bailador et al. 2002; Das and Tizabi 2009; Hejmadi et al. 2003; Picciotto and Zoli 2008). The signal transduction mechanisms underlying the neuroprotection may involve direct or indirect nicotinic receptor mediated modulation of calcium and other antiapoptotic mechanisms (Donnelly-Roberts et al. 1996; Kihara et al. 2001; Liu and Zhao 2004; Ren et al. 2005; Stevens et al. 2003; reviewed in Buckingham et al. 2009), but these potential pathways are still poorly understood.

Alcohol given in low to moderate doses also appears to provide neuroprotection. Epidemiological studies show trends that light to moderate drinkers have reduced risk of dementia and cognitive decline in comparison to nondrinkers (reviewed in Collins et al. 2009). Some of these benefits could be attributed to anti-oxidant polyphenols (e.g., resveratrol in red wine), but it is likely that alcohol in moderate levels has its own direct neuroprotective effect. However, empirical data to support this claim is limited, especially in in vivo models. Studies have shown that giving low doses of alcohol (ethanol) protects in vitro and ex vivo neural cultures exposed to toxins that cause neurodegeneration such as HIV-1 glycoprotein gp120 (gp120: Collins et al. 2000), homoquinolinic acid (Cebere and Liljequist 2003), and NMDA (Cebere and Liljequist 2003; Chandler et al. 1993; Wegelius and Korpi 1995). Similarly, pre-treatment of SH-SY5Y cells, a cell line commonly used to model nigral dopaminergic neurons for Parkinson’s disease with low ethanol concentrations caused attenuated salsolinol-induced toxicity (Ramlochansingh et al. 2011). Salsolinol (1-methyl-6,7-dihydroxy-1,2,3,4-tetrahydroisoquinoline) is an endogenous dopamine metabolite that has structural similarity to MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine) which is especially toxic to nigral dopaminergic neurons, a cluster of cells implicated in Parkinson’s disease (Maruyama et al. 2004; Naoi et al. 2004; Storch et al. 2002). Since many Parkinson patients show high levels of salsolinol in their cerebrospinal fluid, it has been suggested that salsolinol might be involved in the etiology or loss of dopamine neurons in at least some of these patients (Maruyama et al. 2004; Storch et al. 2002). SH-SY5Y cells, derived from human neuroblastoma cells express high level of dopaminergic activity and are used extensively as a model to study nigral dopaminergic neurons (Maruyama et al. 2004; Naoi et al. 2004; Storch et al. 2002). Although the exact neuroprotective mechanism of low alcohol concentration is not fully understood, it appears that several mechanisms may be at work.

Alcohol in low to moderate concentration has also been shown to dampen the inflammatory processes within the brain or in culture (Belmadani et al. 2001; Collins et al. 2000), possibly by influencing increased release of heat shock proteins (HSPs: Belmadani et al. 2004; Sivaswamy et al. 2010; reviewed in Collins et al. 2010). Studies in ex vivo models have shown that preconditioning with alcohol for more that 4 days provided significant protection against neurodegeneration caused by gp120 and β-amyloid exposure (Belmadani et al. 2001, 2004). The current theory of how this preconditioning may work is that alcohol exposure translates an upstream signal that causes an increase in select HSPs (reviewed in Collins et al. 2010). Interestingly, HSPs increase differentially in neurons and astrocytes, with HSP70 primarily located in neurons and HSP27 diffusely elevated in both (Sivaswamy et al. 2010). This suggests that alcohol is differentially interacting with each cell type, and/or affecting the communication between them.

Future Considerations

The story of why there is such a high incidence of alcohol and nicotine co-morbidity is still being unraveled, and there are still many gaps in our knowledge about how the interaction of these two drugs may lead to this co-morbidity. Moreover, the full consequences of alcohol–nicotine co-use—especially at low doses—are far from clear. We have discussed that studies have found additive, synergistic, or contradicting effects of nicotine and/or alcohol on some measures. Obviously, differences in models (in vivo vs. in vitro) and in between in vivo (e.g., different species, age, sex, etc.) and in vitro models (e.g., different cell lines) have to be taken into consideration. Thus, future studies are needed to: establish dose equivalency in humans as compared to rodents; evaluate effects of acute versus chronic treatment in both sexes and in different age groups; and address the mechanism of action (e.g., receptor involvement). Moreover, epigenetic influences of the single or combined effects of alcohol and nicotine have yet to be determined. It is expected that future findings will not only enhance our understanding of basic mechanisms leading to co-use of alcohol and nicotine, but also shine more light on possible therapeutic interventions in drug addiction in general and co-morbid condition of alcoholism and smoking in particular.

Acknowledgments

This study was supported by NIH/NIGMS (2 SO6 GM08016-39), NIAAA (P20 AA014643), and NIH-RCMI 2 G12 RR003048.

References

  1. Acheson A, Mahler SV, Chi H, de Wit H. Differential effects of nicotine on alcohol consumption in men and women. Psychopharmacology. 2006;186(1):54–63. doi: 10.1007/s00213-006-0338-y. [DOI] [PubMed] [Google Scholar]
  2. Al Koudsi N, Hoffmann EB, Assadzadeh A, Tyndale RF. Hepatic CYP2A6 levels and nicotine metabolism: impact of genetic, physiological, environmental, and epigenetic factors. Eur J Clin Pharmacol. 2010;66(3):239–251. doi: 10.1007/s00228-009-0762-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Albuquerque EX, Pereira EFR, Castro NG, Alkondon M, Reinhardt S, Schroder H, Maelicke A. Nicotinic receptor function in the mammalian central nervous system. NY Acad Sci. 1995;757:48–69. doi: 10.1111/j.1749-6632.1995.tb17464.x. [DOI] [PubMed] [Google Scholar]
  4. Al-Rejaie S, Dar MS. Antagonism of ethanol ataxia by intracerebellar nicotine: possible modulation by mouse cerebellar nitric oxide and cGMP. Brain Res Bull. 2006;69(2):187–196. doi: 10.1016/j.brainresbull.2005.12.002. [DOI] [PubMed] [Google Scholar]
  5. Ariyoshi N, Miyamoto M, Umetsu Y, Kunitoh H, Dosaka-Akita H, Sawamura Y, et al. Genetic polymorphism of CYP2A6 gene and tobacco-induced lung cancer risk in male smokers. Cancer Epidemiol Biomarkers Prev. 2002;11:890–894. [PubMed] [Google Scholar]
  6. Askay SW, Bombardier CH, Patterson DR. Effect of acute and chronic alcohol abuse on pain management in a trauma center. Expert Rev Neurother. 2009;9:271–277. doi: 10.1586/14737175.9.2.271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Ball D. Addiction science and its genetics. Addiction. 2008;103:360–367. doi: 10.1111/j.1360-0443.2007.02061.x. [DOI] [PubMed] [Google Scholar]
  8. Bannon AW, Decker MW, Curzon P, Buckley MJ, Kim DJ, Radek RJ, et al. ABT-594 [(R)-5-(2-azetidinylmethoxy)-2-chloropyridine]: a novel, orally effective antinociceptive agent acting via neuronal nicotinic acetylcholine receptors: II. In vivo characterization. J Pharmacol Exp Ther. 1998;285:787–794. [PubMed] [Google Scholar]
  9. Barrett SP, Tichauer M, Leyton M, Pihl RO. Nicotine increases alcohol self-administration in non-dependent male smokers. Drug Alcohol Depend. 2006;81(2):197–204. doi: 10.1016/j.drugalcdep.2005.06.009. [DOI] [PubMed] [Google Scholar]
  10. Bell RL, Eiler BJ, Cook JB, Rahman S. Nicotinic receptor ligands reduce ethanol intake by high alcohol-drinking HAD-2 rats. Alcohol. 2009;43:581–592. doi: 10.1016/j.alcohol.2009.09.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Belluardo N, Mudo G, Blum M, Fuxe K. Central nicotinic receptors, neurotrophic factors and neuroprotection. Behav Brain Res. 2000;113:21–34. doi: 10.1016/s0166-4328(00)00197-2. [DOI] [PubMed] [Google Scholar]
  12. Belmadani A, Zou JY, Schipma MJ, Neafsey EJ, Collins MA. Ethanol pre-exposure suppresses HIV-1 glycoprotein 120-induced neuronal degeneration by abrogating endogenous glutamate/Ca2+-mediated neurotoxicity. Neuroscience. 2001;104(3):769–781. doi: 10.1016/s0306-4522(01)00139-7. [DOI] [PubMed] [Google Scholar]
  13. Belmadani A, Kumar S, Schipma M, Collins MA, Neafsey EJ. Inhibition of amyloid-beta-induced neurotoxicity and apoptosis by moderate ethanol preconditioning. Neuroreport. 2004;15:2093–2096. doi: 10.1097/00001756-200409150-00019. [DOI] [PubMed] [Google Scholar]
  14. Benowitz NL, Jacob P., 3rd Metabolism of nicotine to cotinine studied by a dual stable isotope method. Clin Pharmacol Ther. 1994;56:483–593. doi: 10.1038/clpt.1994.169. [DOI] [PubMed] [Google Scholar]
  15. Berrettini W, Yuan X, Tozzi F, Song K, Francks C, Chilcoat H, Waterworth D, Muglia P, Mooser V. Alpha-5/alpha-3 nicotinic receptor subunit alleles increase risk for heavy smoking. Mol Psychiatry. 2008;13(4):368–373. doi: 10.1038/sj.mp.4002154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bizarro L, Patel S, Stolerman IP. Comprehensive deficits in performance of an attentional task produced by co-administering alcohol and nicotine to rats. Drug Alcohol Depend. 2003;72(3):287–295. doi: 10.1016/j.drugalcdep.2003.08.004. [DOI] [PubMed] [Google Scholar]
  17. Blomqvist O, Engel JA, Nissbrandt H, Söderpalm B. The mesolimbic dopamine-activating properties of ethanol are antagonized by mecamylamine. Eur J Pharmacol. 1993;249(2):207–213. doi: 10.1016/0014-2999(93)90434-j. [DOI] [PubMed] [Google Scholar]
  18. Blomqvist O, Ericson M, Johnson DH, Engel JA, Söderpalm B. Voluntary ethanol intake in the rat: effects of nicotinic acetylcholine receptor blockade or subchronic nicotine treatment. Eur J Pharmacol. 1996;314:257–267. doi: 10.1016/s0014-2999(96)00583-3. [DOI] [PubMed] [Google Scholar]
  19. Boada J, Feria M, Sanz E. Inhibitory effect of naloxone on the ethanol-induced antinociception in mice. Pharmacol Res Commun. 1981;13:673–678. doi: 10.1016/s0031-6989(81)80055-0. [DOI] [PubMed] [Google Scholar]
  20. Bönsch D, Greifenberg V, Bayerlein K, Biermann T, Reulbach U, Hillemacher T, et al. Alpha-synuclein protein levels are increased in alcoholic patients and are linked to craving. Alcohol Clin Exp Res. 2005;29(5):763–765. doi: 10.1097/01.alc.0000164360.43907.24. [DOI] [PubMed] [Google Scholar]
  21. Brown RA, Cutter HS. Alcohol, customary drinking behavior, and pain. J Abnorm Psychol. 1977;86:179–188. doi: 10.1037/0021-843X.86.2.179. [DOI] [PubMed] [Google Scholar]
  22. Buckingham SD, Jones AK, Brown LA, Sattelle DB. Nicotinic acetylcholine receptor signalling: roles in Alzheimer’s disease and amyloid neuroprotection. Pharmacol Rev. 2009;61(1):39–61. doi: 10.1124/pr.108.000562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Buisson B, Bertrand D. Chronic exposure to nicotine upregulates human a4b2 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]
  24. Burch JB, de Fiebre CM, Marks MJ, Collins AC. Chronic ethanol or nicotine treatment results in partial cross-tolerance between these agents. Psychopharmacology. 1988;95:452–458. doi: 10.1007/BF00172954. [DOI] [PubMed] [Google Scholar]
  25. Cadoni C, Muto T, Di Chiara G. Nicotine differentially affects dopamine transmission in the nucleus accumbens shell and core of Lewis and Fischer 344 rats. Neuropharmacology. 2009;57(5–6):496–501. doi: 10.1016/j.neuropharm.2009.07.033. [DOI] [PubMed] [Google Scholar]
  26. Campbell VC, Taylor RE, Tizabi Y. Antinociceptive effects of alcohol and nicotine: involvement of the opioid system. Brain Res. 2006;1097(1):71–77. doi: 10.1016/j.brainres.2006.04.054. [DOI] [PubMed] [Google Scholar]
  27. Castellsagué X, Muñoz N, De Stefani E, Victora CG, Castelletto R, Rolón PA, et al. Independent and joint effects of tobacco smoking and alcohol drinking on the risk of esophageal cancer in men and women. Int J Cancer. 1999;82(5):657–664. doi: 10.1002/(sici)1097-0215(19990827)82:5<657::aid-ijc7>3.0.co;2-c. [DOI] [PubMed] [Google Scholar]
  28. Ceballos NA. Tobacco use, alcohol dependence, and cognitive performance. J Gen Psychol. 2006;133:375–388. doi: 10.3200/GENP.133.4.375-388. [DOI] [PubMed] [Google Scholar]
  29. Cebere A, Liljequist S. Ethanol differentially inhibits homoquinolinic acid- and NMDA-induced neurotoxicity in primary cultures of cerebellar granule cells. Neurochem Res. 2003;28(8):1193–1199. doi: 10.1023/a:1024228412198. [DOI] [PubMed] [Google Scholar]
  30. Cervilla JA, Prince M, Mann A. Smoking, drinking, and incident cognitive impairment: a cohort community based study included in the Gospel Oak project. J Neurol Neurosurg Psychiatr. 2000;68:622–626. doi: 10.1136/jnnp.68.5.622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Chandler LJ, Sumners C, Crews FT. Ethanol inhibits NMDA receptor-mediated excitotoxicity in rat primary neuronal cultures. Alcohol Clin Exp Res. 1993;17(1):54–60. doi: 10.1111/j.1530-0277.1993.tb00726.x. [DOI] [PubMed] [Google Scholar]
  32. Changeux JP, Bertrand D, Corringer PG, Dehaene S, Edelstein S, Lena C, et al. Brain nicotinic receptors: structure and regulation, role in learning and reinforcement. Brain Res Rev. 1998;26:198–216. doi: 10.1016/s0165-0173(97)00040-4. [DOI] [PubMed] [Google Scholar]
  33. Chatterjee S, Bartlett SE. Neuronal nicotinic acetylcholine receptors as pharmacotherapeutic targets for the treatment of alcohol use disorders. 2010;9(1):60–76. doi: 10.2174/187152710790966597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chi H, de Wit H. Mecamylamine attenuates the subjective stimulant-like effects of alcohol in social drinkers. Alcohol Clin Exp Res. 2003;27:780–786. doi: 10.1097/01.ALC.0000065435.12068.24. [DOI] [PubMed] [Google Scholar]
  35. Clarke PBS. Nicotinic receptors and cholinergic neurotransmission in the central nervous system. Ann NY Acad Sci. 1995;757:73–83. doi: 10.1111/j.1749-6632.1995.tb17465.x. [DOI] [PubMed] [Google Scholar]
  36. Clarke PBS, Schwartz RD, Paul SM, Pert CB, Pert A. Nicotinic binding in rat brain: autoradiographic comparison of 3H-acetylcholine, 3H-nicotine and 125I-alpha-bungarotoxin. J Neurosci. 1985;5:1307–1315. doi: 10.1523/JNEUROSCI.05-05-01307.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Collins AC, Burch JB, deFieber CM, Marks MJ. Toleranceto and cross-tolerance between ethanol and nicotine. Pharmacol Biochem Behav. 1988;29:365–373. doi: 10.1016/0091-3057(88)90170-0. [DOI] [PubMed] [Google Scholar]
  38. Collins MA, Neafsey EJ, Zou JY. HIV-I gpI20 neurotoxicity in brain cultures is prevented by moderate ethanol pretreatment. Neuroreport. 2000;11(6):1219–1222. doi: 10.1097/00001756-200004270-00015. [DOI] [PubMed] [Google Scholar]
  39. Collins MA, Neafsey EJ, Mukamal KJ, Gray MO, Parks DA, Das DK, et al. Alcohol in moderation, cardioprotection, and neuroprotection: epidemiological considerations and mechanistic studies. Alcohol Clin Exp Res. 2009;33:206–219. doi: 10.1111/j.1530-0277.2008.00828.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Collins MA, Neafsey EJ, Wang K, Achille NJ, Mictchell RM. Moderate ethanol preconditioning of rat brain cultures engenders neuroprotection against dementia-inducing neuroinflammmatory proteins: possible signal mechanisms. Mol Neurobiol. 2010;41:420–425. doi: 10.1007/s12035-010-8138-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Conti-Fine BM, Maelicke A, Reinhardt-Maelicke SR, Chiappinelli V, McLane KE. Binding sites for neurotoxins and cholinergic ligands in peripheral and neuronal nicotinic receptors. Ann NY Acad Sci. 1995;757:133–153. doi: 10.1111/j.1749-6632.1995.tb17470.x. [DOI] [PubMed] [Google Scholar]
  42. Cooley JE, Villarosa GA, Lombardo TW, Moss RA, Fowler SC, Sult S. Effect of pCPA on nicotine-induced analgesia. Pharmacol Biochem Behav. 1990;36:413–415. doi: 10.1016/0091-3057(90)90425-h. [DOI] [PubMed] [Google Scholar]
  43. Copeland RL, Jr, Leggett YA, Kanaan YM, Taylor RE, Tizabi Y. Neuroprotective effects of nicotine against salsolinol-induced cytotoxicity: implications for Parkinson’s disease. Neurotox Res. 2005;8(3–4):289–293. doi: 10.1007/BF03033982. [DOI] [PubMed] [Google Scholar]
  44. Copeland RL, Jr, Das JR, Kanaan YM, Taylor RE, Tizabi Y. Antiapoptotic effects of nicotine in its protection against salsolinol-induced cytotoxicity. Neurotox Res. 2007;12(1):61–69. doi: 10.1007/BF03033901. [DOI] [PubMed] [Google Scholar]
  45. Cowen MS, Chen F, Lawrence AJ. Neuropeptides: implications for alcoholism. J Neurochem. 2004;89:273–285. doi: 10.1111/j.1471-4159.2004.02394.x. [DOI] [PubMed] [Google Scholar]
  46. Crabbe JC, Phillips TJ, Harris RA, Arends MA, Koob GF. Alcohol-related genes: contributions from studies with genetically engineered mice. Addict Biol. 2006;11:195–269. doi: 10.1111/j.1369-1600.2006.00038.x. [DOI] [PubMed] [Google Scholar]
  47. Craddock A, Cheek JA, Tivis R, Nixon SJ. Nicotine’s effects on neurocognitive performance in alcoholics. Alcohol Clin Exp Res. 2003;27:140A. [Google Scholar]
  48. Cutter HS, O’Farrell TJ. Experience with alcohol and the endogenous opioid system in ethanol analgesia. Addict Behav. 1987;12:331–343. doi: 10.1016/0306-4603(87)90047-5. [DOI] [PubMed] [Google Scholar]
  49. Dajas-Bailador FA, Soliakov L, Wonnacott S. Nicotineactivates the extracellular signal-regulated kinase I/2 via the alpha7 nicotinic acetylcholine receptor and protein kinase A, in SH-SY5Y cells and hippocampal neurones. J Neurochem. 2002;80:520–530. doi: 10.1046/j.0022-3042.2001.00725.x. [DOI] [PubMed] [Google Scholar]
  50. Damaj MI, Fei-Yin M, Dukat M, Glassco W, Glennon RA, Martin BR. Antinociceptive responses to nicotinic acetylcholine receptor ligands after systemic and intrathecal administration in mice. J Pharmacol Exp Therap. 1998;284(3):1058–1065. [PubMed] [Google Scholar]
  51. Damaj MI, Glassco W, Aceto MD, Martin BR. Antinociceptive and pharmacological effects of metanicotine, a selective nicotinic agonist. J Pharmacol Exp Ther. 1999;291:390–398. [PubMed] [Google Scholar]
  52. Damaj MI, Meyer EM, Martin BR. The antinociceptive effects of a7 nicotinic agonists in an acute pain model. Neuropharmacology. 2000;39:2785–2791. doi: 10.1016/s0028-3908(00)00139-8. [DOI] [PubMed] [Google Scholar]
  53. Dar MS, Li C, Bowman ER. Central behavioral interactions between ethanol, (–)-nicotine, and (–)-cotinine in mice. Brain Res Bull. 1993;32(1):23–28. doi: 10.1016/0361-9230(93)90314-2. [DOI] [PubMed] [Google Scholar]
  54. Dar MS, Bowman ER, Chunxiao L. Intracerebellar nicotinic-cholinergic participation in the cerebellar adesoinergic modulation of ethanol-induced motor incoordination in mice. Brain Res. 1994;644:117–127. doi: 10.1016/0006-8993(94)90354-9. [DOI] [PubMed] [Google Scholar]
  55. Das RS, Tizabi Y. Additive protective effects of donepezil and nicotine against salsolinol-induced cytotoxicity in SH-SY5Y cells. Neurotox Res. 2009;16:194–204. doi: 10.1007/s12640-009-9040-2. [DOI] [PubMed] [Google Scholar]
  56. Davis L, Pollock LJ, Stone TT. Visceral pain. Surg Gynecol Obstet. 1932;55:418–427. [Google Scholar]
  57. Dawson DA. Drinking as a risk factor for sustained smoking. Drug Alcohol Depend. 2000;9:235–249. doi: 10.1016/s0376-8716(99)00130-1. [DOI] [PubMed] [Google Scholar]
  58. De Fiebre CM, Collins AC. A comparison of the development of tolerance to ethanol and cross-tolerance to nicotine after chronic ethanol treatment in long- and short-sleep mice. Pharma Exp Ther. 1993;266:1398–1406. [PubMed] [Google Scholar]
  59. De Fiebre CM, Meyer EM, Henry JC, Muraskin SI, Kem WR, Papke RL. Characterization of a series of anabaseine-derived compounds reveals that the 3-(4)-dimethylaminocinnamylidine derivative is a selective agonist at neuronal nicotinic α7/125l-α-bungarotoxin receptor subtypes. Mol Pharmacol. 1995;47:164–171. [PubMed] [Google Scholar]
  60. Decker MW, Meyer MD, Sullivan JP. The therapeutic potential of nicotinic acetylcholine receptor agonists for pain control. Expert Opin Investig Drugs. 2001;10(10):1819–1830. doi: 10.1517/13543784.10.10.1819. [DOI] [PubMed] [Google Scholar]
  61. Dick DM, Foroud T. Candidate genes for alcohol dependence: a review of genetic evidence from human studies. Alcohol Clin Exp Res. 2003;27:868–879. doi: 10.1097/01.ALC.0000065436.24221.63. [DOI] [PubMed] [Google Scholar]
  62. Dickson PE, Rogers TD, Lester DB, Miller MM, Matta SG, Chesler EJ, et al. Genotype-dependent effects of adolescent nicotine exposure on dopamine functional dynamics in the nucleus accumbens shell in male and female mice: a potential mechanism underlying the gateway effect of nicotine. Psychopharmacology (Berl) 2011;215(4):631–642. doi: 10.1007/s00213-010-2159-2. [DOI] [PubMed] [Google Scholar]
  63. Donnelly-Roberts DL, Xue IC, Arneric SP, Sullivan JP. In vitro neuroprotective properties of the novel cholinergic channel activator (ChCA), ABT-418. Brain Res. 1996;719:36–44. doi: 10.1016/0006-8993(96)00063-7. [DOI] [PubMed] [Google Scholar]
  64. Ducci F, Goldman D. Genetic approaches to addiction: genes and alcohol. Addiction. 2008;103:1414–1428. doi: 10.1111/j.1360-0443.2008.02203.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Durazzo TC, Fryer SL, Rothlind JC, Vertinski M, Gazdzinski S, Mon A, et al. Measures of learning, memory and processing speed accurately predict smoking status in short-term abstinent treatment-seeking alcohol-dependent individuals. Alcohol Alcohol. 2010;45(6):507–513. doi: 10.1093/alcalc/agq057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Enoch MA, Goldman D. The genetics of alcoholism and alcohol abuse. Curr Psychiatry Rep. 2001;3(2):144–151. doi: 10.1007/s11920-001-0012-3. [DOI] [PubMed] [Google Scholar]
  67. Ericson M, Blomqvist O, Engel JA, Soderpalm B. Voluntary ethanol intake in the rat and the associated accumbal dopamine overflow are blocked by ventral tegmental mecamylamine. Eur J Pharmacol. 1998;358:189–196. doi: 10.1016/s0014-2999(98)00602-5. [DOI] [PubMed] [Google Scholar]
  68. Ericson M, Löf E, Stomberg R, Söderpalm B. The smoking cessation medication varenicline attenuates alcohol and nicotine interactions in the rat mesolimbic dopamine system. J Pharmacol Exp Ther. 2009;329(1):225–230. doi: 10.1124/jpet.108.147058. [DOI] [PubMed] [Google Scholar]
  69. Ernst M, Heishman SJ, Spurgeon L, London ED. Smoking history and nicotine effects on cognitive performance. Neuropsychopharmacology. 2001;25:313–319. doi: 10.1016/S0893-133X(01)00257-3. [DOI] [PubMed] [Google Scholar]
  70. Exley R, Maubourguet N, David V, Eddine R, Evrard A, Pons S, Marti F, Threlfell S, Cazala P, McIntosh JM, Changeux JP, Maskos U, Cragg SJ, Faure P. Distinct contributions of nicotinic acetylcholine receptor subunit alpha4 and subunit alpha6 to the reinforcing effects of nicotine. Proc Natl Acad Sci USA. 2011;108(18):7577–7582. doi: 10.1073/pnas.1103000108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Falk DE, Yi HY, Hiller-Sturmhöfel S. An epidemiologic analysis of co-occurring alcohol and tobacco use and disorders: findings from the National Epidemiologic Survey on Alcohol and Related Conditions. Alcohol Res Health. 2006;29(3):162–171. [PMC free article] [PubMed] [Google Scholar]
  72. 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]
  73. Ferrea S, Winterer G. Neuroprotective and neurotoxic effects of nicotine. Pharmacopsychiatry. 2009;42(6):255–265. doi: 10.1055/s-0029-1224138. [DOI] [PubMed] [Google Scholar]
  74. Flores CM. The promise and pitfalls of a nicotinic cholinergic approach to pain management. Pain. 2000;88:1–6. doi: 10.1016/S0304-3959(00)00389-4. [DOI] [PubMed] [Google Scholar]
  75. 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(1):31–37. [PubMed] [Google Scholar]
  76. Fowler CD, Lu Q, Johnson PM, Marks MJ, Kenny PJ. Habenular α5 nicotinic receptor subunit signalling controls nicotine intake. Nature. 2011;471(7340):597–601. doi: 10.1038/nature09797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Frahm S, Slimak MA, Ferrarese L, Santos-Torres J, Antolin-Fontes B, Auer S, Filkin S, Pons S, Fontaine JF, Tsetlin V, Maskos U, Ibañez-Tallon I. Aversion to nicotine is regulated by the balanced activity of β4 and α5 nicotinic receptor subunits in the medial habenula. Neuron. 2011;70(3):522–535. doi: 10.1016/j.neuron.2011.04.013. [DOI] [PubMed] [Google Scholar]
  78. Franceschi S, Talamini R, Barra S, Barón AE, Negri E, Bidoli E, et al. Smoking and drinking in relation to cancers of the oral cavity, pharynx, larynx, and esophagus in northern Italy. Cancer Res. 1990;50(20):6502–6507. [PubMed] [Google Scholar]
  79. Franklin KB. Analgesia and the neural substrate of reward. Neurosci Biobehav Rev. 1989;13:149–154. doi: 10.1016/s0149-7634(89)80024-7. [DOI] [PubMed] [Google Scholar]
  80. Franklin KB. Analgesia and abuse potential: an accidental association or a common substrate? Pharmacol Biochem Behav. 1998;59:993–1002. doi: 10.1016/s0091-3057(97)00535-2. [DOI] [PubMed] [Google Scholar]
  81. Freedman R, Coon H, Myles-Worsley M, Orr-Urtreger A, Olincy, et al. Linkage of a neurophysiological deficit in schizophrenia to a chromosome 15 locus. Proc Natl Acad Sci USA. 1997;94:587–592. doi: 10.1073/pnas.94.2.587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Fucito LM, Toll BA, Wu R, Romano DM, Tek E, O’Malley SS. A preliminary investigation of varenicline for heavy drinking smokers. Psychopharmacology (Berl) 2011;215(4):655–663. doi: 10.1007/s00213-010-2160-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Galeote L, Kieffer BL, Maldonado R, Berrendero F. Mu-opioid receptors are involved in the tolerance to nicotine antinociception. J Neurochem. 2006;97:416–423. doi: 10.1111/j.1471-4159.2006.03751.x. [DOI] [PubMed] [Google Scholar]
  84. Gao B, Hierl M, Clarkin K, Juan T, Nguyen H, Valk M, et al. Pharmacological effects of nonselective and subtype-selective nicotinic acetylcholine receptor agonists in animal models of persistent pain. Pain. 2010;149(1):33–49. doi: 10.1016/j.pain.2010.01.007. [DOI] [PubMed] [Google Scholar]
  85. Gentry CL, Lukas RJ. Regulation of nicotinic acetylcholine receptor numbers and function by chronic nicotine exposure. Curr Drug Targets CNS Neurol Disord. 2002;1:359–385. doi: 10.2174/1568007023339184. [DOI] [PubMed] [Google Scholar]
  86. Ghezzi A, Al-Hasan YM, Larios LE, Bohm RA, Atkinson NS. slo K(+) channel gene regulation mediates rapid drug tolerance. Proc Natl Acad Sci USA. 2004;101(49):17276–17281. doi: 10.1073/pnas.0405584101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Glass JM, Adams KM, Nigg JT, Wong MM, Puttler LI, Buu A, et al. Smoking is associated with neurocognitive deficits in alcoholism. Drug Alcohol Depend. 2006;82(2):119–126. doi: 10.1016/j.drugalcdep.2005.08.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Glass JM, Buu A, Adams KM, Nigg JT, Puttler LI, Jester JM, et al. Effects of alcoholism severity and smoking on executive neurocognitive function. Addiction. 2009;104(1):38–48. doi: 10.1111/j.1360-0443.2008.02415.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Goldman D, Oroszi G, Ducci F. The genetics of addictions: uncovering the genes. Nat Rev Genetics. 2005;6:521–532. doi: 10.1038/nrg1635. [DOI] [PubMed] [Google Scholar]
  90. Gotti C, Clementi F. Neuronal nicotinic receptors: from structure to pathology. Prog Neurobiol. 2004;74:363–396. doi: 10.1016/j.pneurobio.2004.09.006. [DOI] [PubMed] [Google Scholar]
  91. Gotti C, Guiducci S, Tedesco V, Corbioli S, Zanetti L, Moretti M, et al. Nicotinic acetylcholine receptors in the mesolimbic pathway: primary role of ventral tegmental area alpha6beta2* receptors in mediating systemic nicotine effects on dopamine release, locomotion, and reinforcement. J Neurosci. 2010;30(15):5311–5325. doi: 10.1523/JNEUROSCI.5095-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Gould TJ, Collins AC, Wehner JM. Nicotine enhances latent inhibition and ameliorates ethanol-induced deficits in latent inhibition. Nicotine Tob Res. 2001;3(1):17–24. doi: 10.1080/14622200020032060. [DOI] [PubMed] [Google Scholar]
  93. Grady SR, Marks MJ, Collins AC. Desensitization of nicotine-stimulated [3H]dopamine release from mouse striatal synaptosomes. J Neurochem. 1994;62(4):1390–1398. doi: 10.1046/j.1471-4159.1994.62041390.x. [DOI] [PubMed] [Google Scholar]
  94. Grant JD, Scherrer JF, Lynskey MT, Lyons MJ, Eisen SA, Tsuang MT, et al. Adolescent alcohol use is a risk factor for adult alcohol and drug dependence: evidence from a twin design. Psychol Med. 2006;36(1):109–118. doi: 10.1017/S0033291705006045. [DOI] [PubMed] [Google Scholar]
  95. Guan ZZ, Yu WF, Nordberg A. Dual effects of nicotine on oxidative stress and neuroprotection in PC12 cells. Neurochem Int. 2003;43:243–249. doi: 10.1016/s0197-0186(03)00009-3. [DOI] [PubMed] [Google Scholar]
  96. Gulick D, Gould TJ. Varenicline ameliorates ethanol-induced deficits in learning in C57BL/6 mice. Neurobiol Learn Mem. 2008;90:230–236. doi: 10.1016/j.nlm.2008.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Han C, McGue MK, Iacono WG. Lifetime tobacco, alcohol and other substance use in adolescent Minnesota twins: univariate and multivariate behavioral genetic analyses. Addiction. 1999;94:981–993. doi: 10.1046/j.1360-0443.1999.9479814.x. [DOI] [PubMed] [Google Scholar]
  98. Hejmadi MV, Dajas-Bailador F, Barns SM, Jones B, Wonnacott S. Neuroprotection by nicotine against hypoxia-induced apoptosis in cortical cultures involves activation of multiple nicotinic acetylcholine receptor subtypes. Mol Cell Neurosci. 2003;24:779–786. doi: 10.1016/s1044-7431(03)00244-6. [DOI] [PubMed] [Google Scholar]
  99. Hellström-Lindahl E, Court JA. Nicotinic acetylcholine receptors during prenatal development and brain pathology in human aging. Behav Brain Res. 2000;113:159–168. doi: 10.1016/s0166-4328(00)00210-2. [DOI] [PubMed] [Google Scholar]
  100. Hendrickson LM, Zhao-Shea R, Pang X, Gardner PD, Tapper AR. Activation of alpha4* nAChRs is necessary and sufficient for varenicline-induced reduction of alcohol consumption. J Neurosci. 2010;30(30):10169–10176. doi: 10.1523/JNEUROSCI.2601-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Hoft NR, Corley RP, McQueen MB, Schlaepfer IR, Huizinga D, Ehringer MA. Genetic association of the CHRNA6 and CHRNB3 genes with tobacco dependence in a nationally representative sample. Neuropsychopharmacology. 2009a;34(3):698–706. doi: 10.1038/npp.2008.122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Hoft NR, Corley RP, McQueen MB, Huizinga D, Menard S, Ehringer MA. SNPs in CHRNA6 and CHRNB3 are associated with alcohol consumption in a nationally representative sample. Genes Brain Behav. 2009b;8(6):631–637. doi: 10.1111/j.1601-183X.2009.00495.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Holloway AC, Cuu DQ, Morrison KM, Gerstein HC, Tarnopolsky MA. Transgenerational effects of fetal and neonatal exposure to nicotine. Endocrine. 2007;31(3):254–259. doi: 10.1007/s12020-007-0043-6. [DOI] [PubMed] [Google Scholar]
  104. Ikemoto S, Panksepp J. The role of nucleus accumbens dopamine in motivated behavior: a unifying interpretation with special reference to reward-seeking. Brain Res Brain Res Rev. 1999;31:6–41. doi: 10.1016/s0165-0173(99)00023-5. [DOI] [PubMed] [Google Scholar]
  105. James MF, Duthie AM, Duffy BL, McKeag AM, Rice CP. Analgesic effect of ethyl alcohol. Br J Anaesth. 1978;50(2):139–141. doi: 10.1093/bja/50.2.139. [DOI] [PubMed] [Google Scholar]
  106. Jerlhag E, Grøtli M, Luthman K, Svensson L, Engel JA. Role of the subunit composition of central nicotinic acetylcholine receptors for the stimulatory and dopamine-enhancing effects of ethanol. Alcohol Alcohol. 2006;41(5):486–493. doi: 10.1093/alcalc/agl049. [DOI] [PubMed] [Google Scholar]
  107. Johansson I, Ingelman-Sundberg M. Genetic polymorphism and toxicology—with emphasis on cytochrome p450. Toxicol Sci. 2011;120(1):1–13. doi: 10.1093/toxsci/kfq374. [DOI] [PubMed] [Google Scholar]
  108. Johnson KA, Jennison KM. The drinking-smoking syndrome and social context. Int J Addict. 1992;27(7):749–792. doi: 10.3109/10826089209068767. [DOI] [PubMed] [Google Scholar]
  109. Jones IW, Wonnacott S. Precise localization of alpha7 nicotinic acetylcholine receptors on glutamatergic axon terminals in the rat ventral tegmental area. J Neurosci. 2004;24(50):11244–11252. doi: 10.1523/JNEUROSCI.3009-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Kamens HM, Andersen J, Picciotto MR. Modulation of ethanol consumption by genetic and pharmacological manipulation of nicotinic acetylcholine receptors in mice. Psychopharmacology. 2010;208(4):613–626. doi: 10.1007/s00213-009-1759-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. 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(2):825–840. [PubMed] [Google Scholar]
  112. Kendler KS, Neale MC, Heath AC, Kessler RC, Eaves LJ. A twin-family study of alcoholism in women. Am J Psychiatry. 1994;151:707–715. doi: 10.1176/ajp.151.5.707. [DOI] [PubMed] [Google Scholar]
  113. Kendler KS, Karkowski LM, Neale MC, Prescott CA. Illicit psychoactive substance use, heavy use, abuse, and dependence in a US population-based sample of male twins. Arch Gen Psychiatry. 2000;57:261–269. doi: 10.1001/archpsyc.57.3.261. [DOI] [PubMed] [Google Scholar]
  114. Khan IM, Buerkle H, Taylor P, Yaksh TL. Nociceptive and antinociceptive responses to intrathecally administered nicotinic agonists. Neuropharmacology. 1998;37:1515–1525. doi: 10.1016/s0028-3908(98)00143-9. [DOI] [PubMed] [Google Scholar]
  115. Kihara T, Shimohama S, Urushitani M, Sawada H, Kimura J, Kume T, et al. Stimulation of α4β2 nicotinic acetylcholine receptors inhibits β-amyloid toxicity. Brain Res. 1998;792:331–334. doi: 10.1016/s0006-8993(98)00138-3. [DOI] [PubMed] [Google Scholar]
  116. Kihara T, Shimohama S, Sawada H, Honda K, Nakamizo T, Shibasaki H, Kume T, Akaike A. Alpha 7 nicotinic receptor transduces signals to phosphatidylinositol 3-kinase to block A beta-amyloid-induced neurotoxicity. J Biol Chem. 2001;276(17):13541–13546. doi: 10.1074/jbc.M008035200. [DOI] [PubMed] [Google Scholar]
  117. Kim JS, Shukla SD. Acute in vivo effect of ethanol (binge drinking) on histone H3 modifications in rat tissues. Alcohol Alcohol. 2006;41:126–132. doi: 10.1093/alcalc/agh248. [DOI] [PubMed] [Google Scholar]
  118. Kleijn J, Folgering JH, van der Hart MC, Rollema H, Cremers TI, Westerink BH. Direct effect of nicotine on mesolimbic dopamine release in rat nucleus accumbens shell. Neurosci Lett. 2011;493(1–2):55–58. doi: 10.1016/j.neulet.2011.02.035. [DOI] [PubMed] [Google Scholar]
  119. Ko JK, Cho CH. Alcohol drinking and cigarette smoking: a “partner” for gastric ulceration. Zhonghua Yi Xue Za Zhi (Taipei) 2000;63(12):845–854. [PubMed] [Google Scholar]
  120. Köhnke MD. Approach to the genetics of alcoholism: a review based on pathophysiology. Biochem Pharmacol. 2008;75:160–177. doi: 10.1016/j.bcp.2007.06.021. [DOI] [PubMed] [Google Scholar]
  121. Koob GF, Le Moal M. Nicotine. In: Koob GF, Le Moal M, editors. Neurobiology of addiction. Academic Press; London: 2006. pp. 243–287. [Google Scholar]
  122. Koob GF, Roberts AJ, Schulteis G, Parsons LH, Heyser CJ, Hyytia P, Merlo-Pich E, Weiss F. Neurocircuitry targets in ethanol reward and dependence. Alcohol Clin Exp Res. 1998;22:3–9. [PubMed] [Google Scholar]
  123. Lajtha A, Sershen H. Nicotine: alcohol reward interactions. Neurochem Res. 2010;35(8):1248–1258. doi: 10.1007/s11064-010-0181-8. [DOI] [PubMed] [Google Scholar]
  124. Larsson A, Engel JA. Neurochemical and behavioral studies on ethanol and nicotine interactions. Neurosci Biobehav Rev. 2004;27:713–720. doi: 10.1016/j.neubiorev.2003.11.010. [DOI] [PubMed] [Google Scholar]
  125. Launay JM, Del Pino M, Chironi G, Callebert J, Peoc’h K, Mégnien JL, et al. Smoking induces long-lasting effects through a monoamine-oxidase epigenetic regulation. PLoS One. 2009;4(11):e7959. doi: 10.1371/journal.pone.0007959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Laviolette SR, Lauzon NM, Bishop SF, Sun N, Tan H. Dopamine signaling through D1-like versus D2-like receptors in the nucleus accumbens core versus shell differentially modulates nicotine reward sensitivity. J Neurosci. 2008;28(32):8025–8033. doi: 10.1523/JNEUROSCI.1371-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Lawrence NS, Ross TJ, Stein EA. Cognitive mechanisms of nicotine on visual attention. Neuron. 2002;36:539–548. doi: 10.1016/s0896-6273(02)01004-8. [DOI] [PubMed] [Google Scholar]
  128. Le AD, Corrigall WA, Harding JW, Juzytsch W, Li TK. Involvement of nicotinic receptors in alcohol self-administration. Alcohol Clin Exp Res. 2000;24:155–163. doi: 10.1111/j.1530-0277.2000.tb04585.x. [DOI] [PubMed] [Google Scholar]
  129. Leigh G, Tong JE. Effects of ethanol and tobacco on time judgment. Percept Mot Skills. 1976;43:899–903. doi: 10.2466/pms.1976.43.3.899. [DOI] [PubMed] [Google Scholar]
  130. Leigh G, Tong JE, Campbell JA. Effects of ethanol and tobacco on divided attention. J Stud Alcohol. 1977;38:1233–1239. doi: 10.15288/jsa.1977.38.1233. [DOI] [PubMed] [Google Scholar]
  131. Lindstrom J. Nicotinic acetylcholine receptors in health and disease. Mol Neurobiol. 1997;15(2):193–222. doi: 10.1007/BF02740634. [DOI] [PubMed] [Google Scholar]
  132. Liu Q, Zhao B. Nicotine attenuates beta-amyloid peptide-induced neurotoxicity, free radical and calcium accumulation in hippocampal neuronal cultures. Br J Pharmacol. 2004;141(4):746–754. doi: 10.1038/sj.bjp.0705653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Liu J, Zhou Z, Hodgkinson CA, Yuan Q, Shen PH, Mulligan CJ, et al. Haplotype-based study of the association of alcohol-metabolizing genes with alcohol dependence in four independent populations. Alcohol Clin Exp Res. 2011;35(2):304–316. doi: 10.1111/j.1530-0277.2010.01346.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Löf E, Ericson M, Stomberg R, Söderpalm B. Characterization of ethanol-induced dopamine elevation in the rat nucleus accumbens. Eur J Pharmacol. 2007;555(2–3):148–155. doi: 10.1016/j.ejphar.2006.10.055. [DOI] [PubMed] [Google Scholar]
  135. Lotfipour S, Ferguson E, Leonard G, Perron M, Pike B, Richer L, et al. Orbitofrontal cortex and drug use during adolescence: role of prenatal exposure to maternal smoking and BDNF genotype. Arch Gen Psychiatry. 2009;66(11):1244–1252. doi: 10.1001/archgenpsychiatry.2009.124. [DOI] [PubMed] [Google Scholar]
  136. Luczak SE, Elvine-Kreis B, Shea SH, Carr LG, Wall TL. Genetic risk for alcoholism relates to level of response to alcohol in Asian-American men and women. J Stud Alcohol. 2002;63:74–82. [PubMed] [Google Scholar]
  137. 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]
  138. Madden PA, Bucholz KK, Martin NG, Heath AC. Smoking and the genetic contribution to alcohol-dependence risk. Alcohol Res Health. 2000;24(4):209–214. [PMC free article] [PubMed] [Google Scholar]
  139. Maes HH, Woodard CE, Murrelle L, Meyer JM, Silberg JL, Hewitt JK, et al. Tobacco, alcohol and drug use in eight- to sixteen-year-old twins: the Virginia Twin Study of Adolescent Behavioral Development. J Stud Alcohol. 1999;60:293–305. doi: 10.15288/jsa.1999.60.293. [DOI] [PubMed] [Google Scholar]
  140. Mahadev K, Vemuri MC. Effect of ethanol on chromatin and nonhistone nuclear proteins in rat brain. Neurochem Res. 1998;23(9):1179–1184. doi: 10.1023/a:1020778018149. [DOI] [PubMed] [Google Scholar]
  141. Maloku E, Kadriu B, Zhubi A, Dong E, Pibiri F, Satta R, et al. Selective α(4)β(2) nicotinic acetylcholine receptor agonists target epigenetic mechanisms in cortical GABAergic neurons. Neuropsychopharmacology. 2011;36(7):1366–1374. doi: 10.1038/npp.2011.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Mansvelder HD, McGehee DS. Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron. 2000;27(2):349–357. doi: 10.1016/s0896-6273(00)00042-8. [DOI] [PubMed] [Google Scholar]
  143. Marks MJ, Grady SR, Collins AC. Downregulation of nicotinic receptor function after chronic nicotine infusion. J Pharmacol Exp Ther. 1993;266(3):1268–1276. [PubMed] [Google Scholar]
  144. Marubio LM, del Mar Arroyo-Jimenez M, Cordero-Erausquin M, Lena C, Le Novere N, de Kerchove d’Exaerde A, et al. Reduced antinociception in mice lacking neuronal nicotinic receptor subunits. Nature. 1999;398:805–810. doi: 10.1038/19756. [DOI] [PubMed] [Google Scholar]
  145. Maruyama W, Yi H, Takahashi T, Shimazu S, Ohde H, Yoneda F, et al. Neuroprotective function of R-(–)-1-(benzofuran-2-yl)-2-propylaminopentane, [R-(–)-BPAP], against apoptosis induced by N-methyl(R)salsolinol, an endogenous dopaminergic neurotoxin, in human dopaminergic neuroblastoma SH-SY5Y cells. Life Sci. 2004;75(1):107–117. doi: 10.1016/j.lfs.2003.12.001. [DOI] [PubMed] [Google Scholar]
  146. Maskos U, Molles BE, Pons S, Besson M, Guiard BP, Guilloux JP, et al. Nicotine reinforcement and cognition restored by targeted expression of nicotinic receptors. Nature. 2005;436(7047):103–107. doi: 10.1038/nature03694. [DOI] [PubMed] [Google Scholar]
  147. Matsuyama S, Matsumoto A, Enomoto T, Nishizaki T. Activation of nicotinic acetylcholine receptors induces long-term potentiation in vivo in the intact mouse dentate gyrus. Eur J Neurosci. 2001;12:3741–3747. doi: 10.1046/j.1460-9568.2000.00259.x. [DOI] [PubMed] [Google Scholar]
  148. McKee SA, Harrison EL, O’Malley SS, Krishnan-Sarin S, Shi J, Tetrault JM, Picciotto MR, Petrakis IL, Estevez N, Balchunas E. Varenicline reduces alcohol self-administration in heavy-drinking smokers. Biol Psychiatry. 2009;66(2):185–190. doi: 10.1016/j.biopsych.2009.01.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Mereu G, Fadda F, Gessa GL. Ethanol stimulates the firing rate of nigral dopaminergic neurons in unanesthetized rats. Brain Res. 1984;292:63–69. doi: 10.1016/0006-8993(84)90890-4. [DOI] [PubMed] [Google Scholar]
  150. Mereu G, Yoon KW, Boi V, Gessa GL, Naes L, Westfall TC. Preferential stimulation of ventral tegmental area dopaminergic neurons by nicotine. Eur J Pharmacol. 1987;141:395–399. doi: 10.1016/0014-2999(87)90556-5. [DOI] [PubMed] [Google Scholar]
  151. Messina ES, Tyndale RF, Sellers EM. A major role for CYP2A6 in nicotine C-oxidation by human liver microsomes. J Pharmacol Exp Ther. 1997;282:1608–1614. [PubMed] [Google Scholar]
  152. Millar NS, Gotti C. Diversity of vertebrate nicotinic acetylcholine receptors. Neuropharmacology. 2009;56:237–246. doi: 10.1016/j.neuropharm.2008.07.041. [DOI] [PubMed] [Google Scholar]
  153. Mitrouska I, Bouloukaki I, Siafakas NM. Pharmacological approaches to smoking cessation. Pulm Pharmacol Ther. 2007;20:220–232. doi: 10.1016/j.pupt.2005.10.012. [DOI] [PubMed] [Google Scholar]
  154. Miwa JM, Freedman R, Lester HA. Neural systems governed by nicotinic acetylcholine receptors: emerging hypotheses. Neuron. 2011;70(1):20–33. doi: 10.1016/j.neuron.2011.03.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Moonat S, Starkman BG, Sakharkar A, Pandey SC. Neuroscience of alcoholism: molecular and cellular mechanisms. Cell Mol Life Sci. 2010;67:73–88. doi: 10.1007/s00018-009-0135-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Mudo G, Belluardo N, Fuxe K. Nicotinic receptor agonists as uroprotective/neurotrophic drugs. Progress in molecular mechanisms. J Neural Transm. 2007;114(1):135–147. doi: 10.1007/s00702-006-0561-z. [DOI] [PubMed] [Google Scholar]
  157. Müller DJ, Likhodi O, Heinz A. Neural markers of genetic vulnerability to drug addiction. Curr Top Behav Neurosci. 2010;3:277–299. doi: 10.1007/7854_2009_25. [DOI] [PubMed] [Google Scholar]
  158. Nakajima M, Yamamoto T, Nunoya K, Yokoi T, Nagashima K, Inoue K, et al. Role of human cytochrome P450 2A6 in C-oxidation of nicotine. Drug Metab Dispos. 1996a;11:1212–1217. [PubMed] [Google Scholar]
  159. Nakajima M, Yamamoto T, Nunoya K, Yokoi T, Nagashima K, Inoue K, et al. Characterization of CYP2A6 involved in 3-hydroxylation of cotinine in human liver microsomes. J Pharmacol Exp Ther. 1996b;277:1010–1015. [PubMed] [Google Scholar]
  160. Naoi M, Maruyama W, Nagy GM. Dopamine-derived salsolinol derivatives as endogenous monoamine oxidase inhibitors: occurrence, metabolism and function in human brains. Neurotoxicology. 2004;25(1–2):193–204. doi: 10.1016/S0161-813X(03)00099-8. [DOI] [PubMed] [Google Scholar]
  161. Nguyen HN, Rasmussen BA, Perry D. Binding and functional activity of nicotinic cholinergic receptors in selected rat brain regions are increased following long-term but not short-term nicotine treatment. J Neurochem. 2004;90:40–49. doi: 10.1111/j.1471-4159.2004.02482.x. [DOI] [PubMed] [Google Scholar]
  162. Nisell M, Nomikos GG, Svensson TH. Systemic nicotine-induced dopamine release in the rat nucleus accumbens is regulated by nicotinic receptors in the ventral tegmental area. Synapse. 1994a;16(1):36–44. doi: 10.1002/syn.890160105. [DOI] [PubMed] [Google Scholar]
  163. Nisell M, Nomikos GG, Svensson TH. Infusion of nicotine in the ventral tegmental area or the nucleus accumbens of the rat differentially affects accumbal dopamine release. Pharmacol Toxicol. 1994b;75(6):348–352. doi: 10.1111/j.1600-0773.1994.tb00373.x. [DOI] [PubMed] [Google Scholar]
  164. Olale F, Gerzanich V, Kuryatov A, Wang F, Lindstrom J. Chronic nicotine exposure differentially affects the function of human a3, a4, and a7 neuronal nicotinic receptor subtypes. J Pharmacol Exp Ther. 1997;283:675–683. [PubMed] [Google Scholar]
  165. Oliveira-da-Silva A, Vieira FB, Cristina-Rodrigues F, Filgueiras CC, Manhães AC, AbreuVillaça Y. Increased apoptosis and reduced neuronal and glial densities in the hippocampus due to nicotine and ethanol exposure in adolescent mice. Int J Dev Neurosci. 2009;27(6):539–548. doi: 10.1016/j.ijdevneu.2009.06.009. [DOI] [PubMed] [Google Scholar]
  166. Oliveira-da-Silva A, Manhães AC, Cristina-Rodrigues F, Filgueiras CC, AbreuVillaça Y. Hippocampal increased cell death and decreased cell density elicited by nicotine and/or ethanol during adolescence are reversed during drug withdrawal. Neuroscience. 2010;167(1):163–173. doi: 10.1016/j.neuroscience.2010.01.060. [DOI] [PubMed] [Google Scholar]
  167. Olsen J, Sabreo S, Fasting U. Interaction of alcohol and tobacco as risk factors in cancer of the laryngeal region. J Epidemiol Community Health. 1985;39(2):165–168. doi: 10.1136/jech.39.2.165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Pabreza LA, Dhawan S, Kellar KJ. [3H]cytisine binding to nicotinic cholinergic receptors in brain. Mol Pharmacol. 1991;39(1):9–12. [PubMed] [Google Scholar]
  169. Pandey SC, Ugale R, Zhang H, Tang L, Prakash A. Brain chromatin remodeling: a novel mechanism of alcoholism. J Neurosci. 2008;28(14):3729–3737. doi: 10.1523/JNEUROSCI.5731-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Park HJ, Lee PH, Ahn YW, Choi YJ, Lee G, Lee DY, et al. Neuroprotective effect of nicotine on dopaminergic neurons by anti-inflammatory action. Eur J Neurosci. 2007;26(1):79–89. doi: 10.1111/j.1460-9568.2007.05636.x. [DOI] [PubMed] [Google Scholar]
  171. Parnell SE, West JR, Chen WJ. Nicotine decreases blood alcohol concentrations in adult rats: a phenomenon potentially related to gastric function. Alcohol Clin Exp Res. 2006;30(8):1408–1413. doi: 10.1111/j.1530-0277.2006.00168.x. [DOI] [PubMed] [Google Scholar]
  172. Pelucchi C, Gallus S, Garavello W, Bosetti C, La Vecchia C. Alcohol and tobacco use, and cancer risk for upper aerodigestive tract and liver. Eur J Cancer Prev. 2008;17(4):340–344. doi: 10.1097/CEJ.0b013e3282f75e91. [DOI] [PubMed] [Google Scholar]
  173. Perkins KA, Sexton JE, DiMarco A, Grobe JE, Scierka A, Stiller RL. Subjective and cardiovascular responses to nicotine combined with alcohol in male and female smokers. Psychopharmacology (Berl) 1995;119:205–212. doi: 10.1007/BF02246162. [DOI] [PubMed] [Google Scholar]
  174. Perrino AC, Jr, Ralevski E, Acampora G, Edgecombe J, Limoncelli D, Petrakis IL. Ethanol and pain sensitivity: effects in healthy subjects using an acute pain paradigm. Alcohol Clin Exp Res. 2008;32:952–958. doi: 10.1111/j.1530-0277.2008.00653.x. [DOI] [PubMed] [Google Scholar]
  175. Perry DC, Mao D, Gold AB, McIntosh JM, Pezzullo JC, Kellar KJ. Chronic nicotine differentially regulates alpha6- and beta3-containing nicotinic cholinergic receptors in rat brain. J Pharmacol Exp Ther. 2007;322(1):306–315. doi: 10.1124/jpet.107.121228. [DOI] [PubMed] [Google Scholar]
  176. Pianezza ML, Sellers EM, Tyndale RF. Nicotine metabolism defect reduces smoking. Nature. 1998;393(6687):6750. doi: 10.1038/31623. [DOI] [PubMed] [Google Scholar]
  177. Picciotto MR. Nicotine as a modulator of behavior: beyond the inverted U. Trends Pharm Sci. 2003;24:493–499. doi: 10.1016/S0165-6147(03)00230-X. [DOI] [PubMed] [Google Scholar]
  178. Picciotto MR, Zoli M. Neuroprotection via nAChRs: the role of nAChRs in neurodegenerative disorders such as Alzheimer’s and Parkinson’s disease. Front Biosci. 2008;13:492–504. doi: 10.2741/2695. [DOI] [PubMed] [Google Scholar]
  179. Picciotto MR, Caldarone BJ, King SL, Zachariou V. Nicotinic receptors in the brain. Links between molecular biology and behavior. Neuropsychopharmacology. 2000;22:451–465. doi: 10.1016/S0893-133X(99)00146-3. [DOI] [PubMed] [Google Scholar]
  180. Picciotto MR, Addy NA, Mineur YS, Brunzell DH. It is not “either/or”: activation and desensitization of nicotinic acetylcholine receptors both contribute to behaviors related to nicotine addiction and mood. Prog Neurobiol. 2008;84:329–342. doi: 10.1016/j.pneurobio.2007.12.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Pidoplichko VI, DeBiasi M, Williams JT, Dani JA. Nicotine activates and desensitizes midbrain dopamine neurons. Nature. 1997;390(6658):401–404. doi: 10.1038/37120. [DOI] [PubMed] [Google Scholar]
  182. Pons S, Fattore L, Cossu G, Tolu S, Porcu E, McIntosh JM, Changeux JP, Maskos U, Fratta W. Crucial role of alpha4 and alpha6 nicotinic acetylcholine receptor subunits from ventral tegmental area in systemic nicotine self-administration. J Neurosci. 2008;28(47):12318–12327. doi: 10.1523/JNEUROSCI.3918-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Price DD. Psychological and neural mechanisms of the affective dimension of pain. Science. 2000;288(5472):1769–1772. doi: 10.1126/science.288.5472.1769. [DOI] [PubMed] [Google Scholar]
  184. Quik M, O’Leary K, Tanner CM. Nicotine and Parkinson’s disease: implications for therapy. Mov Disord. 2008;23(12):1641–1652. doi: 10.1002/mds.21900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Ralevski E, Perrino A, Acampora G, Koretski J, Limoncelli D, Petrakis I. Analgesic effects of ethanol are influenced by family history of alcoholism and neuroticism. Alcohol Clin Exp Res. 2010;34(8):1433–1441. doi: 10.1111/j.1530-0277.2010.01228.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Ramlochansingh C, Taylor RE, Tizabi Y. Toxic effects of low alcohol and nicotine combinations in SH-SY5Y cells are apoptotically mediated. Neurotox Res. 2011;20:263–269. doi: 10.1007/s12640-011-9239-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Rao Y, Hoffmann E, Zia M, Bodin L, Zeman M, Sellers EM, et al. Duplications and defects in the CYP2A6 gene: identification, genotyping, and in vivo effects on smoking. Mol Pharmacol. 2000;58:747–755. doi: 10.1124/mol.58.4.747. [DOI] [PubMed] [Google Scholar]
  188. Rathouz MM, Vijayaraghaven S, Berg DK. Elevation of intracellular calcium levels in neurons by nicotinic acetylcholine receptors. Mol Neurobiol. 1996;12(2):117–131. doi: 10.1007/BF02740649. [DOI] [PubMed] [Google Scholar]
  189. Ren K, Puig V, Papke RL, Itoh Y, Hughes JA, Meyer EM. Multiple calcium channels and kinases mediate alpha 7 nicotinic receptor neuroprotection in PC12 cells. J Neurochem. 2005;94:926–933. doi: 10.1111/j.1471-4159.2005.03223.x. [DOI] [PubMed] [Google Scholar]
  190. Renthal W, Nestor EJ. Epigenetic mechanisms in drug addiction. Trends Mol Med. 2008;14(8):341–350. doi: 10.1016/j.molmed.2008.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Rezvani AH, Levin ED. Cognitive effects of nicotine. Biol Psychiatr. 2001;49:258–267. doi: 10.1016/s0006-3223(00)01094-5. [DOI] [PubMed] [Google Scholar]
  192. Rezvani AH, Levin ED. Nicotine-alcohol interactions and cognitive functions in rats. Pharmacol Biochem Behav. 2002;72:865–872. doi: 10.1016/s0091-3057(02)00762-1. [DOI] [PubMed] [Google Scholar]
  193. Rogers DT, Iwamoto ET. Multiple spinal mediators in parenteral nicotine-induced antinociception. J Pharmacol Exp Ther. 1993;267:341–349. [PubMed] [Google Scholar]
  194. Rose RJ. A developmental behavior-genetic perspective on alcoholism risk. Alcohol Health Res World. 1998;22:131–143. [PMC free article] [PubMed] [Google Scholar]
  195. Rose JE, Brauer LH, Behm FM, Cramblett M, Calkins K, Lawhon D. Psychopharmacological interactions between nicotine and ethanol. Nicotine Tob Res. 2004;6(1):133–144. doi: 10.1080/14622200310001656957. [DOI] [PubMed] [Google Scholar]
  196. Rosenblum A, Joseph H, Fong C, Kipnis S, Cleland C, Portenoy RK. Prevalence and characteristics of chronic pain among chemically dependent patients in methadone maintenance and residential treatment facilities. JAMA. 2003;289:2370–2378. doi: 10.1001/jama.289.18.2370. [DOI] [PubMed] [Google Scholar]
  197. Sabbagh MN, Lukas RJ, Sparks DL, Reid RT. The nicotinic acetylcholine receptor, smoking, and Alzheimer’s disease. J Alzheimer’s Dis. 2002;4:317–325. doi: 10.3233/jad-2002-4407. [DOI] [PubMed] [Google Scholar]
  198. Sartor CE, Grant JD, Bucholz KK, Madden PA, Heath AC, Agrawal A, et al. Common genetic contributions to alcohol and cannabis use and dependence symptomatology. Alcohol Clin Exp Res. 2010;34(3):545–554. doi: 10.1111/j.1530-0277.2009.01120.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Schilström B, Svensson HM, Svensson TH, Nomikos GG. Nicotine and food induced dopamine release in the nucleus accumbens of the rat: putative role of alpha7 nicotinic receptors in the ventral tegmental area. Neuroscience. 1998;85:1005–1009. doi: 10.1016/s0306-4522(98)00114-6. [DOI] [PubMed] [Google Scholar]
  200. Schlaepfer IR, Hoft NR, Collins AC, Corley RP, Hewitt JK, Hopfer CJ, Lessem JM, McQueen MB, Rhee SH, Ehringer MA. The CHRNA5/A3/B4 gene cluster variability as an important determinant of early alcohol and tobacco initiation in young adults. Biol Psychiatry. 2008;63(11):1039–1046. doi: 10.1016/j.biopsych.2007.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Schuckit MA, Smith TL. Changes over time in the self-reported level of response to alcohol. Alcohol Alcohol. 2004;39(5):433–438. doi: 10.1093/alcalc/agh081. [DOI] [PubMed] [Google Scholar]
  202. Simons CT, Cuellar JM, Moore JA, Pinkerton KE, Uyeminami D, Carstens MI, et al. Nicotinic receptor involvement in antinociception induced by exposure to cigarette smoke. Neurosci Lett. 2005;389:71–76. doi: 10.1016/j.neulet.2005.07.025. [DOI] [PubMed] [Google Scholar]
  203. Sivaswamy S, Neafsey EJ, Collins MA. Neuroprotective preconditioning of rat brain cultures with ethanol: potential transduction by PKC isoforms and focal adhesion kinase upstream of increases in effector heat shock proteins. Eur J Neurosci. 2010;32(11):1800–1812. doi: 10.1111/j.1460-9568.2010.07451.x. [DOI] [PubMed] [Google Scholar]
  204. Smith AM, Zeve DR, Dohrman DP, Chen WJ. The interactive effect of alcohol and nicotine on NGF-treated pheochromocytoma cells. Alcohol. 2006;39:65–72. doi: 10.1016/j.alcohol.2006.06.010. [DOI] [PubMed] [Google Scholar]
  205. Spande TF, Garraffo HM, Edwards MW, Yeh HJC, Pannell L, Daly JW. A new class of alkaloids from a dendrobatid poison frog: a structure for alkaloid 251F. J Am Chem Soc. 1992;114:3475–3478. doi: 10.1021/np50084a002. [DOI] [PubMed] [Google Scholar]
  206. Steensland P, Simms JA, Holgate J, Richards JK, Bartlett SE. Varenicline, an a4b2 nicotinic acetylcholine receptor partial agonist, selectively decreases ethanol consumption and seeking. Proc Natl Acad Sci USA. 2007;104:12518–12523. doi: 10.1073/pnas.0705368104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Stevens TR, Krueger SR, Fitzsimonds RM, Picciotto MR. Neuroprotection by nicotine in mouse primary cortical cultures involves activation of calcineurin and l-type calcium channel inactivation. J Neurosci. 2003;23:10093–10099. doi: 10.1523/JNEUROSCI.23-31-10093.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Stolerman IP, Chandler CJ, Garcha HS, Newton JM. Selective antagonism of behavioural effects of nicotine by dihydro-β-erythroidine in rats. Psychopharmacology. 1997;129:390–397. doi: 10.1007/s002130050205. [DOI] [PubMed] [Google Scholar]
  209. Storch A, Ott S, Hwang YI, Ortmann R, Hein A, Frenzel S, Matsubara K, Ohta S, Wolf HU, Schwarz J. Selective dopaminergic neurotoxicity of isoquinoline derivatives related to Parkinson’s disease: studies using heterologous expression systems of the dopamine transporter. Biochem Pharmacol. 2002;63(5):909–920. doi: 10.1016/s0006-2952(01)00922-4. [DOI] [PubMed] [Google Scholar]
  210. Taslim N, Soderstrom K, Dar MS. Role of mouse cerebellar nicotinic acetylcholine receptor (nAChR) α(4)β(2)- and α(7) subtypes in the behavioral cross-tolerance between nicotine and ethanol-induced ataxia. Behav Brain Res. 2011;217(2):282–292. doi: 10.1016/j.bbr.2010.10.026. [DOI] [PubMed] [Google Scholar]
  211. Teaktong T, Graham A, Court J, Perry R, Jaros E, Johnson M, et al. Alzheimer’s disease is associated with a selective increase in alpha7 nicotinic acetylcholine receptor immunoreactivity in astrocytes. Glia. 2003;41(2):207–211. doi: 10.1002/glia.10132. [DOI] [PubMed] [Google Scholar]
  212. Teaktong T, Graham AJ, Johnson M, Court JA, Perry EK. Selective changes in nicotinic acetylcholine receptor subtypes related to tobacco smoking: an immunohistochemical study. Neuropathol Appl Neurobiol. 2004;30:243–254. doi: 10.1046/j.0305-1846.2003.00528.x. [DOI] [PubMed] [Google Scholar]
  213. Thorgeirsson TE, Geller F, Sulem P, Rafnar T, Wiste A, Magnusson KP, et al. A variant associated with nicotine dependence, lung cancer and peripheral arterial disease. Nature. 2008;452(7187):638–642. doi: 10.1038/nature06846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Thorgeirsson TE, Gudbjartsson DF, Surakka I, Vink JM, Amin N, Geller F, et al. Sequence variants at CHRNB3–CHRNA6 and CYP2A6 affect smoking behavior. Nat Genet. 2010;42:448–453. doi: 10.1038/ng.573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Tizabi Y, Overstreet DH, Rezvani AH, Louis VA, Clark E, Jr, Janowsky DS, Kling MA. Antidepressant effects of nicotine in an animal model of depression. Psychopharmacology. 1999;142:193–199. doi: 10.1007/s002130050879. [DOI] [PubMed] [Google Scholar]
  216. Tizabi Y, Rezvani AH, Russell LT, Tyler KY, Overstreet DH. Depressive characteristics of FSL rats: Involvement of central nicotinic receptors. Pharmacol Biochem Behav. 2000;66:73–77. doi: 10.1016/s0091-3057(00)00236-7. [DOI] [PubMed] [Google Scholar]
  217. Tizabi Y, Copeland RL, Jr, Louis VA, Taylor RE. Effects of combined systemic alcohol and central nicotine administration into ventral tegmental area on dopamine release in the nucleus accumbens. Alcohol Clin Exp Res. 2002;26(3):394–399. [PubMed] [Google Scholar]
  218. Tizabi Y, Al-Namaeh M, Manaye KF, Taylor RE. Protective effects of nicotine on ethanol-induced toxicity in cultured cerebellar granule cells. Neurotox Res. 2003;5(5):315–321. doi: 10.1007/BF03033151. [DOI] [PubMed] [Google Scholar]
  219. Tizabi Y, Manaye KF, Smoot DT, Taylor RE. Nicotine inhibits ethanol-induced toxicity in cultured cortical cells. Neurotox Res. 2004;6:311–316. doi: 10.1007/BF03033441. [DOI] [PubMed] [Google Scholar]
  220. Tizabi Y, Manaye KF, Taylor RE. Nicotine blocks ethanol-induced apoptosis in primary cultures of rat cerebral cortical and cerebellar granule cells. Neurotox Res. 2005;7(4):319–322. doi: 10.1007/BF03033888. [DOI] [PubMed] [Google Scholar]
  221. Tizabi Y, Bai L, Copeland RL, Jr, Taylor RE. Combined effects of systemic alcohol and nicotine on dopamine release in the nucleus accumbens shell. Alcohol Alcohol. 2007;42(5):413–416. doi: 10.1093/alcalc/agm057. [DOI] [PubMed] [Google Scholar]
  222. Tizabi Y, Getachew B, Rezvani AH, Hauser SR, Overstreet DO. Antidepressant-like effects of nicotine and reduced nicotine binding in Fawn-Hooded rats. Prog Neuropsychopharm Biol Psychiatr. 2009;33(3):398–402. doi: 10.1016/j.pnpbp.2008.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Toledo-Rodriguez M, Lotfipour S, Leonard G, Perron M, Richer L, Veillette S, et al. Maternal smoking during pregnancy is associated with epigenetic modifications of the brain-derived neurotrophic factor-6 exon in adolescent offspring. Am J Med Genet B Neuropsychiatr Genet. 2010;153B(7):1350–1354. doi: 10.1002/ajmg.b.31109. [DOI] [PubMed] [Google Scholar]
  224. Tracey I, Mantyh PW. The cerebral signature for pain perception and its modulation. Neuron. 2007;55:377–391. doi: 10.1016/j.neuron.2007.07.012. [DOI] [PubMed] [Google Scholar]
  225. Tracy HA, Jr, Wayner MJ, Armstrong DL. Nicotine blocks ethanol and diazepam impairment of air righting and ethanol impairment of maze performance. Alcohol. 1999;18:123–130. doi: 10.1016/s0741-8329(98)00074-3. [DOI] [PubMed] [Google Scholar]
  226. Tripathi HL, Martin BR, Aceto MD. Nicotine-induced antinociception in rats and mice: correlation with nicotine brain levels. J Pharmacol Exp Ther. 1982;221:91–96. [PubMed] [Google Scholar]
  227. True WR, Heath AC, Scherrer JF, Xian H, Lin N, Eisen SA, et al. Interrelationship of genetic and environmental influences on conduct disorder and alcohol and marijuana dependence symptoms. Am J Med Genet. 1999;88:391–397. doi: 10.1002/(sici)1096-8628(19990820)88:4<391::aid-ajmg17>3.0.co;2-l. [DOI] [PubMed] [Google Scholar]
  228. Tsankova N, Renthal W, Kumar A, Nestler EJ. Epigenetic regulation in psychiatric disorders. Nat Rev Neurosci. 2007;8:355–367. doi: 10.1038/nrn2132. [DOI] [PubMed] [Google Scholar]
  229. Tu GC, Israel Y. Alcohol consumption by orientals in North America is predicted largely by a single gene. Behav Genet. 1995;25:59–65. doi: 10.1007/BF02197242. [DOI] [PubMed] [Google Scholar]
  230. Uhl GR. Molecular genetics of substance abuse vulnerability: remarkable recent convergence of genome scan results. Ann NY Acad Sci. 2004;1025:1–13. doi: 10.1196/annals.1316.001. [DOI] [PubMed] [Google Scholar]
  231. Uhl GR, Drgon T, Johnson C, Fatusin OO, Liu QR, Contoreggi C, Li CY, Buck K, Crabbe J. “Higher order” addiction molecular genetics: convergent data from genome-wide association in humans and mice. Biochem Pharmacol. 2008;75:98–111. doi: 10.1016/j.bcp.2007.06.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Vanderah TW, Suenaga NM, Ossipov MH, Malan TP, Jr, Lai J, Porreca F. Tonic descending facilitation from the rostral ventromedial medulla mediates opioid-induced abnormal pain and antinociceptive tolerance. J Neurosci. 2001;21(1):279–286. doi: 10.1523/JNEUROSCI.21-01-00279.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Vincler M. Neuronal nicotinic receptors as targets for novel analgesics. Expert Opin Investig Drugs. 2005;14:1191–1198. doi: 10.1517/13543784.14.10.1191. [DOI] [PubMed] [Google Scholar]
  234. Wang H, Sun X. Desensitized nicotinic receptors in brain. Brain Res Brain Res Rev. 2005;48:420–437. doi: 10.1016/j.brainresrev.2004.09.003. [DOI] [PubMed] [Google Scholar]
  235. Wang Y, Krishnan HR, Ghezzi A, Yin JC, Atkinson NS. Drug-induced epigenetic changes produce drug tolerance. PLoS Biol. 2007;5(10):e265. doi: 10.1371/journal.pbio.0050265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Wang JC, Grucza R, Cruchaga C, Hinrichs AL, Bertelsen S, Budde JP, et al. Genetic variation in the CHRNA5 gene affects mRNA levels and is associated with risk for alcohol dependence. Mol Psychiatry. 2009;14(5):501–510. doi: 10.1038/mp.2008.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Wegelius K, Korpi ER. Ethanol inhibits NMDA-induced toxicity and trophism in cultured cerebellar granule cells. Acta Physiol Scand. 1995;154(1):25–34. doi: 10.1111/j.1748-1716.1995.tb09882.x. [DOI] [PubMed] [Google Scholar]
  238. Willuhn I, Wanat MJ, Clark JJ, Phillips PE. Dopamine signaling in the nucleus accumbens of animals self-administering drugs of abuse. Curr Top Behav Neurosci. 2010;3:29–71. doi: 10.1007/7854_2009_27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Wonnacott S. The paradox of nicotine acetylcholine receptor upregulation by nicotine. Trends Pharmacol Sci. 1990;11:216–219. doi: 10.1016/0165-6147(90)90242-z. [DOI] [PubMed] [Google Scholar]
  240. Woodrow KM, Eltherington LG. Feeling no pain: alcohol as an analgesic. Pain. 1988;32(2):159–163. doi: 10.1016/0304-3959(88)90064-4. [DOI] [PubMed] [Google Scholar]
  241. Wu J, Lukas RJ. Naturally-expressed nicotinic acetylcholine receptor subtypes. Biochem Pharmacol. 2011;82(8):800–807. doi: 10.1016/j.bcp.2011.07.067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Yagoubian B, Akkara J, Afzali P, Alfi DM, Olson L, Conell-Price J, et al. Nicotine nasal spray as an adjuvant analgesic for third molar surgery. J Oral Maxillofac Surg. 2011;69(5):1316–1319. doi: 10.1016/j.joms.2010.07.025. [DOI] [PubMed] [Google Scholar]
  243. Yoshikawa H, Kurokawa M, Ozaki N, Nara K, Atou K, Takada E, et al. Nicotine inhibits the production of proinflammatory mediators in human monocytes by suppression of I-kappaB phosphorylation and nuclear factor-kappaB transcriptional activity through nicotinic acetylcholine receptor alpha7. Clin Exp Immunol. 2006;146(1):116–123. doi: 10.1111/j.1365-2249.2006.03169.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Young EM, Mahler S, Chi H, de Wit H. Mecamylamine and ethanol preference in healthy volunteers. Alcohol Clin Exp Res. 2005;29:58–65. doi: 10.1097/01.alc.0000150007.34702.16. [DOI] [PubMed] [Google Scholar]
  245. Yu ZJ, Wecker L. Chronic nicotine administration differentially affects neurotransmitter release from rat striatal slices. J Neurochem. 1994;63:186–194. doi: 10.1046/j.1471-4159.1994.63010186.x. [DOI] [PubMed] [Google Scholar]
  246. Yue J, Khokhar J, Miksys S, Tyndale RF. Differential induction of ethanol-metabolizing CYP2E1 and nicotine-metabolizing CYP2B1/2 in rat liver by chronic nicotine treatment and voluntary ethanol intake. Eur J Pharmacol. 2009;609(1–3):88–95. doi: 10.1016/j.ejphar.2009.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  247. Zigmond RE, Loring RH. Characterization and localization of ganglionic nicotinic receptors using neuronal bungarotoxin. Nato ASI Ser H. 1988;25:31–40. [Google Scholar]

RESOURCES