Skip to main content

Genetics of Tobacco Use

Abstract

The worldwide prevalence of tobacco use is widespread, resulting in nearly 4.5 million deaths every year. Nicotine is addictive and has psychopharmacological effects that maintain the use of tobacco products. Several studies have documented a strong hereditary component to tobacco use. The present article reviews results from twin and adoption studies and proceeds to present association studies of specific genes that may be involved in tobacco use. Cholinergic receptor nicotinic beta polypeptide 2, serotonin receptor and transporter genes, dopamine receptor and transporter genes, and the cytochrome P450A6 gene are reviewed. Linkage studies help to identify regions of the genome that may be worth further investigation. The paper concludes with a discussion of the limitations of genetic research and the future of genetic epidemiology in this domain.

Epidemiology of Tobacco Use and Dependence

Tobacco use and dependence are byproducts of a large, complex web of social, environmental and genetic influences. Tobacco dependence essentially results from prolonged exposure to the pharmacologic effects of nicotine. Although smoking has been popular for more than a century, the addictive qualities of nicotine were not acknowledged in scientific literature until the 1980s. Consequently, despite the knowledge that tobacco use carries considerable health risks [1, 2], more than 1.1 billion people (one-third of the global adult population) use tobacco regularly [3]. Current global estimates indicate that 48% of men and 12% of women smoke cigarettes [4]. Transnational consumption differences indicate that smoking is decreasing in most developed countries yet increasing in most developing nations. In the United States, 25% of the total population aged 12 or older (56.3 million people) smoke cigarettes, 5.4% (12.1 million people) smoke cigars, 3.2% (7.3 million people) use smokeless tobacco, and 1.0% (2.3 million people) smoke tobacco in pipes [5]. Among adult smokers, many wish to stop smoking, but the addictive qualities of nicotine create a complex barrier. In the United States, 70% of current smokers express a desire to quit, yet only 4.7% of those who attempted to quit in the past year were able to maintain abstinence from smoking for 3–12 months [2].

Withdrawal difficulties are evident even in adolescent smokers who try to quit [6]. Indeed, the age of smoking initiation is declining and tobacco use is rising among youth in many countries [7]. There are a number of complex and inter-relating factors that predispose young people to smoke, and these vary among individuals and among different populations. Low prices, easy access, parental smoking, family conflicts, peer pressure, and the positive image of smoking through advertisement are all contributing factors for tobacco use initiation in adolescence.

Characteristics of Tobacco Users

Although historically, adult tobacco use was linked to affluence [8], today, continuing tobacco consumption is associated with societal disadvantage indicators, such as low income and low education [9], and individual factors, such as race, use of alcohol or other illicit drugs, depression, and personality type [10]. There are differences in specific smoking behaviors across races and genders. In the United States, Hispanics and Asians have lower smoking prevalences (19% and 14% respectively) than Caucasians or African-Americans (about 25%) [2]. Reports indicate that African-Americans start smoking regularly at a later age and smoke fewer cigarettes, but smoke those with higher nicotine content [11–13]. In terms of gender, differences may exist in nicotine sensitivity and nicotine requirements. For example, after smoking similar cigarettes, women have significantly lower nicotine levels than men [14]. Also, many studies show that although men are more likely to be smokers, women tend to have greater difficulty quitting [10].

Certain personality or behavioral traits, such as depression, neuroticism, attention deficit and hyperactive disorder (ADHD), and novelty seeking are associated with tobacco use [10, 15]. Novelty-seeking is characterized by stimulation-seeking and sensitivity to reward or reinforcement, and is associated with nicotine dependence not only in adults [16], but also in adolescents [17] and early age of smoking initiation [18]. Children and teenagers with a diagnosis of ADHD engage in more health-threatening behaviors such as smoking, and alcohol and substance abuse. Though the etiology of ADHD is not known, recent studies suggest that maternal smoking during pregnancy is a factor and that there is a strong genetic link to the disorder [19].

In general, individuals who are tobacco-dependent exhibit greater psychiatric comorbidities that may stem from a common etiologic basis involving neuro-transmitter pathways and genetics. A prime example comes from the considerable and consistent evidence that smoking, alcohol, and depression are linked. Compared to never smokers, smokers have more than twice the lifetime prevalence of major depressive disorder [20]. Prevalence of smoking increases with levels of current symptoms of depression [21] and smoking cessation relapse is associated with increased symptoms of depression [20]. Furthermore, individuals with a combined history of alcoholism and major depression are at a high risk to use smoking as a means of mood enhancement [22]. A common neurological thread to these disorders is the involvement of serotonin and dopaminergic depletion and their associated rewards of pleasurable feelings [10, 23, 24], which are discussed in the following section. Thus, genes involved in the dopaminergic system have been studied in relation to alcoholism, depression and many other psychiatric disorders, with positive results [25]. Considering that tobacco dependence is linked to such individual characteristics that are shaped not only by environment, but also by genetics [10], it is not surprising that studies have found a strong hereditary component to tobacco use [26, 27].

Psychoactive Components of Tobacco

Nicotine's addictive characteristics are known to primarily result from its ability to bind to receptors in the brain, which then increase concentrations of neurotransmitters such as dopamine and serotonin in reward regions of the brain. Dopamine, affected by many stimulants and addictive drugs, is involved in the mesolimbic reward pathway, which produces pleasurable feelings. Recent human drug-reversal trials confirm that nicotine mediates reinforcement from smoking via dopamine, and that smoking behavior can be manipulated within the same subjects in opposite directions by alternately stimulating and blocking dopamine [28].

Dopamine is secreted partly by the release of serotonin (5-HT), which is a key neurotransmitter in the reward pathway and a logical candidate to study in relation to smoking behavior. When nicotine binds to its receptors, serotonin secretion from the brain increases [29, 30]. The reward pathway proceeds as serotonin stimulates enkephalin in the hypothalamus, which then inhibits GABA at the substania nigra, finally releasing a specified amount of dopamine at the nucleus accumbens, termed the "reward site" [31]. Another link between nicotine and serotonin has been seen in the rat hippocampus, where smoking clearly reduces serotonin binding to serotonin receptors 1 (5-HT1) and 5-HT1A [32]. Thus, nicotine not only increases serotonin release from the brain, but also decreases its reuptake in the hippocampus. Together, these effects make available a greater amount of serotonin, which ultimately stimulates dopamine release. During withdrawal from nicotine, serotonin release is indeed reduced [30], and changes in mood seen in nicotine withdrawal may partly result from diminished serotonin transmission [33]. Accordingly, it has been hypothesized that people with a functional deficit in this dopamine pathway may be more prone to nicotine dependence [27].

To confirm the notion that smoking and nicotine affect the dopaminergic system in the human brain, neuroimaging techniques, such as positron emission tomography (PET), have been used to directly measure physiologic, pharmacokinetic and pharmacodymanic events in the same person following nicotine administration. These studies have shown that nicotine has rapid pharmacokinetics, changing cerebral blood flow and brain metabolism [34, 35]. Compared to non-smokers, smokers exhibit significantly increased pre-synaptic dopamine transmission in dopaminergic brain regions [36]. Furthermore, smoking may create a vicious cycle of reward and deficiency; a recent study found that smoking seems to inhibit dopamine receptor D3 mRNA expression, thereby diminishing the rewarding effect of dopamine, leading to a reward deficiency state and creating motivation for continued smoking [37]. Indeed, phenotypes of interest have been delineated with the help of PET studies. For example, in response to monetary rewards, striatal and dopaminergic activity is different for smokers and non-smokers. A possible explanation is that a smoker's brain interprets and reacts to rewards differently, and perhaps in a more impulsive and sensation-seeking manner, than a non-smoker's brain [38]. Another phenotypical finding is that changes in cerebral blood flow following nicotine administration are strongly correlated to self-reported craving for cigarettes and addiction as a motivation for smoking [34]. Thus, it seems apparent that nicotine's effects are linked to neuroanatomical systems involved in the dopaminergic pathway, as well as arousal and reward.

Complicating the epidemiology of tobacco dependence is the notion that nicotine may not be the only psychoactive component found in tobacco. Primary evidence for this notion comes from the observation that nicotine can evoke a behavioral response that lasts much longer than the presence of nicotine in plasma [39]. Fowler and colleagues (1996) observed that smoking, but not nicotine, decreases the brain levels of an important enzyme that breaks down dopamine, monoamine oxidase (MAO), by 30–40%, which may result in increases in dopamine levels [40]. Further PET studies indicate that the reduction in MAO B in smokers is most likely to occur gradually and require chronic tobacco smoke exposure [41].

Another reason to speculate that additional components of tobacco use are involved is the recent observation that cotinine, the primary metabolite of nicotine, is bioactive as well [42]. Compared to nicotine, cotinine has a long pharmacological half-life (15–19 vs. 2–3 hours) and may be 5–30 times less potent [39]. Recent studies observed that cotinine evokes nicotine-like pharmacological responses, such as antagonizing the effects of dopamine receptor agonists [39]. Considering that cotinine, like nicotine, can stimulate dopamine release, it is theoretically possible that cotinine could mediate the long-term behavioral actions of nicotine. However, many studies have ascribed no pharmacological action to the compound; even in levels as high as ten times that attained from cigarette smoking, it has been found to have no observable acute or withdrawal effects in humans [43]. Other possible compounds of tobacco may reinforce dependence by irritating the mouth, throat or bronchial tree, and thereby enhancing the addictive potential of nicotine through altered sensory cues [44]. Overall, while many components of tobacco continue to be investigated in their role in addiction, it is clear that nicotine is at the core of neuropharmacological actions for many users.

Essentially, smokers alter their tobacco use to maintain enough levels of nicotine to act on neurotransmitters; therefore, genetic variation in the psychological need for nicotine, the ability to metabolize it, and the pathway to its pleasurable effects, ultimately influences if and how people use tobacco. Indeed, as smokers pass through stages from initiation to dependence, there are many opportunities for genetic factors to influence behavior. Research suggests that certain genes may be unique to one stage of the smoking process, while others influence several smoking characteristics [45–47]. The following paragraphs present major positive findings in twin and adoption studies that predict the strong role of genetics. The review then proceeds to discuss the results of association studies for specific alleles involved in tobacco use, as well as genome-wide scans and linkage studies.

Twin and Adoption Studies on the Genetics of Smoking Behaviors

The twin pair study, which has long been a popular research design to investigate the role of genetics in disease causation [48], provided early evidence that there is a hereditary component to tobacco use. In 1958, there was a report of increased frequency of smoking among pairs of monozygotic male twins, as compared to dizygotic twins [49]; this observation implicated genetics in the etiology of the habit. Since then, innovative research designs in twins have attempted to quantify the role of inheritance in various outcomes, including initiation, intensity (current amount smoked), persistence (years of smoking), dependence and inability to quit. Kendler and colleagues (1999) studied 1,898 female twins in an effort to identify risk factors for initiation and dependence. They found that initiation was associated with low education, religiosity, increased neuroticism, extroversion and mental health comorbidities. Nicotine dependence was associated with low education, extroversion, mastery, self-esteem, increased neuroticism, history of mood disorders and alcohol abuse [50]. A review of data from 14 different twin studies and more than 17,000 twins estimated that genetic factors account for 56% of the variance in smoking initiation and regular tobacco use, environmental (shared, familial) risk factors account for 24% and individual risk factors account for the remaining 20%. The authors note that genetic factors appear to be more prominent in the transition to nicotine dependence, where they account for approximately 70% of the variance, and shared environmental influences seem negligible [47]. Elsewhere, smoking persistence (the number of years of regular smoking) has been found to be substantially due to genetic factors, with similar effects for women and men of all ages [46], although the role of environment, and especially low education, is significant. To estimate the heritability of failed smoking cessation, Xian and colleagues (2003) studied 1,818 twin pairs of smokers that had experienced at least one unsuccessful attempt to quit; they found that genetics accounted for 54% of the variance in risk of failed cessation and 29.7% of the variance in the risk of nicotine withdrawal [51].

In general, twin studies of both adolescents and adults indicate that the role of genetics is greater in dependence and quantity smoked, than in initiation, where environmental influences prevail [45, 52, 53]. Initiation, like persistence, is largely influenced by important interactions between environmental and genetic elements [54]. The notion that both environmental and inheritance factors may co-exist, and the fact that this coexistence is difficult to disentangle, create the potential for confounding or bias in twin studies. To avoid this problem, Osler and colleagues (2001) conducted an adoption study using 840 families [55]. Their finding that smoking status in adult adoptees was significantly associated with the smoking status of the biological full siblings, who were reared separately, supports the results of twin studies and strengthens the notion that there is a genetic influence on smoking [56].

Genetic Association Studies

The fact that smoking is a complex process involving many different molecular targets has led to proposals to study a variety of genes as modulators of smoking behavior. To identify exactly which gene variants are associated with smoking phenotypes, molecular epidemiology has primarily used the candidate gene approach. For complex psychiatric disorders such as substance abuse, no single genetic locus is itself the causal factor; rather, multiple alleles found at various loci interact to produce vulnerability to an outcome. Unfortunately, the ability to identify numerous alleles that each contribute small effects, or that have an effect only when other critical alleles are also present, is poor, and studies are often under-powered [57]. Still, variations in the genome that have functional consequences (e.g., an important protein changes in the amount, location, or structure of expression in carriers of that allele) can be studied with success given adequate sample sizes and plausible mechanisms of action [58].

Advances in technology, together with the unique ability of these genetic variations to facilitate gene identification, have resulted in a recent flood of detection of single nucleotide polymorphisms (SNPs), which are often implicated in genetic association studies. Another recent focus of genetic research is the study of haplotypes, which are specified variants in an entire region of a chromosome, rather than just at one locus or gene. A 'haplotype map' is currently being constructed, to better pinpoint areas of chromosomes that are of interest during future research. Further information and updates on genetic research can be found at the National Center for Biotechnology Information web site: http://www.ncbi.nlm.nih.gov.

To investigate variant alleles and their role in smoking behavior, studies have focused on polymorphisms involved with the pathway of pleasurable sensations conferred by tobacco use, such as the dopamine system, as well as those involved in the metabolization of nicotine, such as genes that code for proteins in the cytochrome P450 family. The prominent outcomes that have been studied are smoking initiation, progression to nicotine dependence, quantity smoked, and persistence of smoking. Increasingly, endophenotypes, such as initial sensitivity to nicotine, are also being considered to further the power and interpretation of studies.

Neurotransmitter Pathways

Since the effects of nicotine on the dopamine pathway are relatively well understood, many studies have investigated genetic variations in the dopaminergic system. Genes for this neurotransmitter's synthesis, degradation, receptors, and transporters are logical candidates. At the same time, serotonin, norepinephrine, GABA (gamma amino butyric acid), opioid, and cannabinoid neurons all modify dopamine metabolism and dopamine neurons. For example, mu-opioid receptors are known to play an important role in mediating the effects of various substances, such as opioids [59, 60], cannabinoids [61] and alcohol [62].

Recently, it was shown that mice lacking mu-opioid receptors do not obtain rewarding effects of nicotine and have reduced withdrawal symptoms, suggesting a functional interaction between nicotine, the opioid system, and dopaminergic mesolimbic activity [63]. It has been proposed that defects in various combinations of the genes for these neurotransmitters result in a Reward Deficiency Syndrome (RDS) and that such individuals are at risk for abuse of substances that provide unnatural rewards [64].

In addition to this dopamine pathway, other neuronal systems, including brainstem cholinergic, GABAergic, noradrenergic and serotonergic nuclei, are also affected by nicotine and may be involved in nicotine dependence [65]. Much of the current knowledge of these systems comes from observations of the behavior of mice with mutations. Thus, the mechanisms of neurotransmitters, such as norephinephrine, or other pathways that are independent of the dopamine pathway, such as those following nicotinic reception in the brainstem, are yet to be determined [65]. While animal models are critical to the understanding of genetic factors in complex behavior, the following discussion concentrates on studies of smoking behavior in human populations and their associations with neurotransmitters with known polymorphisms.

Nicotinic Receptors

Reward pathways are initiated when nicotine binds to nicotinic acetylcholine receptors (nAchRs), which then mediate the positive and negative effects that follow. Two classes of these receptors have been identified and many subunits are expressed in key regions of the brain and the mesolimbic dopamine system. One receptor gene, CHRNB2 (cholinergic receptor, nicotinic, beta polypeptide 2, located on chromosome 1q21) is a functional candidate gene for nicotine dependence, since it seems to be essential for a number of reinforcing effects of nicotine in mice [65, 66]. However, a study of four SNPs in CHRNB2 found no association with the polymorphisms or their estimated haplotypes with smoking initiation or nicotine dependence, using carefully selected samples of non-smokers and regular smokers [66]. Polymorphisms in several other receptor subunits have since then been identified [67, 68]. If new polymorphisms prove to have functional significance, then these genes may be candidates to explore in future research in smoking behavior [56].

Serotonin

Serotonin Synthesis Neurotransmitter, Tryptophan Hydroxylase Gene (TPH)

Tryptophan hydroxylase (TPH) catalyzes the rate-limiting step in serotonin biosynthesis. Although the known polymorphisms in the TPH gene are in non-coding regions and are not thought to have functional consequence [69], two polymorphisms, C218A and C779A, have been linked to personality traits [70] and have been studied in association with smoking (Table 1) [71, 72]. These polymorphisms were not found to be associated with smoking status or progression to nicotine dependence. Significant associations, however, were seen in smoking initiation variables, such as age at onset of smoking [72] and ever having smoked more than an entire cigarette [71]. Considering the lack of functional significance of the polymorphisms and the possibility of false positive results, it is likely that either the TPH markers are in linkage disequilibrium for an unknown functional polymorphism, that the action of TPH on smoking is indirect (i.e., through a personality trait), or that TPH plays a minor role in the large web of causal factors for smoking initiation [71].

Table 1 Epidemiologic studies of tryptophan hydroxylase (TPH) gene, serotonin transporter gene and tobacco use in human populations.

Serotonin Receptor Gene

Although there have been no reports of association studies between serotonin receptor gene polymorphisms and smoking, studies have provided grounds to explore the relationship [33]. For example, the effect of a polymorphism in the 5-HT2C receptor gene (Cys23Ser) on the reward dependence trait was significantly modified by the presence of a dopamine polymorphism (DRD4 VNTR) in individuals [73].

Serotonin Transporter Gene

Unlike the receptor gene, serotonin transporter gene (SLC6A4) polymorphisms have been studied in association with smoking behavior (Table 1). The serotonin transporter gene, on chromosome 17q12, regulates the reuptake of serotonin from the synaptic junction. Gene transcription is modulated by a polymorphism (5-HTTLPR) that results in a long allele (l) or a short allele (s), which occur in 57% and 43% of Caucasians, respectively [74]. The s allele is associated with reduced transcription, resulting in a lower 5-HTT uptake and more available serotonin [75]. In addition, the s allele is associated with neuroticism, an anxiety-related personality trait [74], in samples from the United States, Japan and Israel [15].

The hypothesis that the increased available serotonin in individuals with the s allele would protect against habitual smoking or help in the smoking cessation process has been tested. No association between the 5-HTT polymorphism and smoking was seen in a study of Caucasians and African-Americans [76]. In contrast, a study of Japanese men found that the l allele was observed significantly more often in smokers than in nonsmokers or ex-smokers [77]. Although the subjects in the U.S. study were recruited through ads and may not have been representative of all smokers, the Japanese study differs from other studies by combining the l/l and l/s genotypes (since the l allele frequency tends to be lower in Japanese, at roughly 19%). Therefore, more studies are needed to confirm that there is a main effect of the serotonin transporter gene polymorphism on smoking behavior.

Meanwhile, an interaction between the transporter's 5-HTT polymorphism and neurotic personality has been found in two studies of smoking behavior [15, 78]. For individuals with the s allele, neuroticism is linked to increased nicotine intake, dependence, smoking for stimulation, and smoking to reduce negative affect. In contrast, neuroticism is not an important impediment to smoking cessation for individuals with the l/l genotype, although they may exhibit the personality as well. Thus, rather than influence smoking initiation, these genotypes affect dependence, ability to quit, and motivations for smoking, such as smoking to self-medicate mood disturbances [15]. More work is needed to clarify the role of polymorphisms in serotonin genes, so that drugs used in tobacco cessation, such as the serotonin uptake inhibitor fluoxetine, can be administered to those people who would be most likely to benefit.

Dopamine Transporter Gene (SLC6A3)

Released dopamine is taken up by the presynaptic neuron via the dopamine transporter (DAT), which is the primary mechanism for dopamine clearance from the synapse in the midbrain. The DAT protein is encoded by the locus SLC6A3 on chromosome 5p15.3. At SLC6A3, there is a polymorphism involving a variable number of tandem repeats (VNTR) in the 3'-untranslated region, with repeat numbers ranging from 3 to 11. The most common allele in population studies is the ten-repeat allele (A10, about 70% in European, Caucasian or African-Americans), followed by the nine-repeat allele (A9, about 25% in European, Caucasian or African-Americans). However, there is some regional variation; A10 is seen in more than 90% of Asian or South American populations, and certain African populations have high frequencies of the A7 allele. Although the VNTR polymorphism does not affect the actual protein product, it may affect mRNA localization, transcript stability, regulation of protein synthesis, and expression [79]. Indeed, the polymorphism seems to alter translation of the DAT protein and decrease its availability [80–82]. Under certain conditions, a decrease in transporter proteins decreases the clearance of synaptic dopamine. As a result, the baseline "set point" of synaptic dopamine is higher, thereby increasing sensitivity to drug-induced dopamine surges, or decreasing desire for substances, such as nicotine, that would increase synaptic dopamine [82].

Dopamine Receptor Gene (DRD)

The synaptic levels of dopamine may also be determined by the density of dopamine receptors, which involves the dopamine receptor gene (DRD). Although there are many dopamine receptors (D1–D5), each exhibiting polymorphisms, the D2 dopamine receptor gene has been the focus of research regarding substance abuse. This is primarily due to the lack of evidence that the other polymorphisms have functional significance or are related to other complex psychiatric disorders [83]. The D2 receptor gene is located on chromosome 11q.23, and several polymorphisms have been identified, most notably in the TaqI A allele (A1 and A2) and the TaqI B allele (B1 and B2). Studies suggest that there is linkage disequilibrium between the less common A1 and B1 allele variants [84–86]. There is considerable variation in the frequency of these alleles in different racial and ethnic populations, even within European-American samples [87]. The A1 allele has been correlated with a reduced number of dopamine binding sites in the brain [84, 88–90], and with increased dopamine transporter levels [91]. Heterosis, which refers to the situation where heterozygotes for an allele (e.g., A1/A2) are more likely to exhibit a phenotype than either homozygote (e.g. A1/A1 or A2/A2), was reported in the effect of the DRD2 gene on D2 receptor density [92].

Another DRD polymorphism that seems to have functional significance and has been studied in relation to smoking is the DRD4 VNTR polymorphism, on chromosome 11p15.5. There are three common variants of two (D4.2), four (D4.4), and seven repeats (D4.7) [93], which occur within coding regions. The longer, seven-repeat variant appears to blunt the intracellular response to dopamine in vitro, as compared with the 2 and 4 repeat variants [94]. D4 receptors are structurally and pharmacologically very similar to D2 receptors and are localized to the same limbic brain structures that underlie reward feelings [95]. It has been found that D4 receptors in the nucleus accumbens (the reward site) may modulate excitatory transmission [96] and the sensitization of these pathways [97].

The Role of the SLC6A3 and DRD in Tobacco Use

The hypothesis that people with altered transport or a reduced D2 receptor density have a deficit in their reward system and feel an enhanced reward when exposed to dopaminergic agents, thereby making them more prone to nicotine addiction, has been supported by epidemiologic studies (Table 2) [27, 85, 86, 98–105]. However, the findings regarding behavioral phenotypes associated with certain alleles are contradictory and indicate that the exact role of dopamine genes is complex. Although explanations for inconsistent findings are discussed in a subsequent section, it is worth noting here that heterosis in the effects of dopamine genes may play a role [89, 106, 107].

Table 2 Epidemiologic studies of dopamine receptor gene (DRD2), dopamine transporter gene (SLC6A3) and tobacco use in human populations.

With this in mind, the SLC6A3-9 allele has been found to be associated with an increased susceptibility to tobacco use in some studies [103, 104], but a decreased susceptibility in others [100, 108]. Earlier studies comparing those who have smoked <100 cigarettes in their lifetime (designated as non-smokers) to either former or current smokers found that carrying at least one copy of the 9 allele decreases risk of smoking by almost 50% and is associated with a four-fold higher likelihood of successful smoking cessation [103, 104]. The notion that SLC6A3 is associated with smoking cessation was also supported by a recent case-control analysis using 250 current smokers [109].

Opposing findings, in which the 9 allele significantly increases the likelihood of smoking, may be partly due to the distinction made between never-smokers (never inhaled the smoke from even one cigarette) and non-smokers (smoked 1–100 cigarettes in lifetime) [100]. This distinction is likely to be important, particularly if risk-taking personality is a factor. The notion that the SLC6A3-9 allele may increase risk is supported by a recent investigation into the endophe-notype of self-reported craving; smokers with at least one copy of the 9 allele had significantly greater craving reactions following a laboratory-induced stress [108]. Interestingly, despite disagreement regarding the direction of the association of the SLC6A3-9 allele, studies agree that there seems to be a significant interaction between this gene and DRD2-A1 in increasing susceptibility to tobacco use [103, 108]. An acceptable postulated mechanism is that individuals with certain SLC6A3 genotypes have lower endogenous synaptic dopamine, and therefore a greater need to use substances such as nicotine to stimulate dopamine transmission. This unfavorable effect of decreased dopamine may be especially pronounced in people with DRD2-A1 genotypes, i.e., those with deficient receptor density.

In general, the evidence for the DRD2 TaqI A1 allele is more consistent and points to an increased risk of tobacco use. Studies have found significant main effects with younger age at smoking initiation [86, 105], fewer attempts to quit [86, 104], shorter duration of quit attempts, current or past smoking [27, 85, 105] and craving reactions [108], but not nicotine addiction scales [86]. Recent investigations have found that DRD2 is associated with withdrawal symptoms, such as difficulty in concentrating and sleeping, and that DRD3 is associated with age of initiation, heaviness of smoking, and depression during withdrawal [99]. Meta-analyses based on three studies suggest that the odds ratio for likelihood of drug abuse is 2.4 for individuals possessing an DRD2-A1 allele and 3.3 for those having a DRD2-B1 allele (p < 0.001 in both cases) [110]. The notion that the A1 allele increases susceptibility to tobacco use is supported pharmacologically by randomized trials of bupropion hydrochloride, a weak dopamine reuptake inhibitor effective in decreasing withdrawal symptoms and craving. Studies have found that bupropion attenuates withdrawal-related craving, irritability and anxiety only among subjects with the DRD2-A2/A2 genotype [111, 112]. A plausible biological hypothesis is that dopamine reuptake inhibition during nicotine withdrawal is less pronounced in DRD2-A1 carriers, since they have a limited ability to stimulate postsynaptic dopamine receptors in the brain's reward regions [90].

The association between DRD4 variants and smoking has been studied independently of the DRD2 and SLC6A3 polymorphisms, but results are similarly complicated (Table 3). Regarding the main effects of DRD4 in smoking initiation, dependence and persistence, analyses of Caucasians have found no association [23, 113, 114], but an African-American sample showed that presence of at least one long (L, 6–8 repeats) allele conferred an OR of 7.7 (95% CI 1.5, 39.9) for risk of smoking and a significantly shorter time to first cigarette in the morning [113]. In a combined sample, a significant interaction was seen between depression and DRD4, in which depressed smokers homozygous for the short alleles of DRD4 were more likely to smoke to reduce negative feelings and to smoke in response to stimulation [23]. The notion that smoking cues (e.g., seeing a lit cigarette) are important in the mechanism of D4 variants was further supported by a recent study, in which only individuals with long-repeats alleles showed significant reactions to smoking cues, in particular increased craving and arousal, with decreased positive affect [114]. The exact mechanism of affecting the reaction to smoking cues has been speculated, but requires verification by molecular studies [114].

Table 3 Epidemiologic studies of dopamine receptor gene (DRD4) and tobacco use in human populations.

Of note, a recent study simultaneously analyzed linkage and association of the DRD4 genotype to smoking, alcohol and novelty-seeking behaviors in a sibling-pair design [115]. In the second wave of the study, DRD4 had a significant effect on smoking status, but the authors admit that the number of statistical tests performed and the inconsistency of the gene effect over time invalidate this finding. Thus, with improved methods, considerable power, repeated measures for some phenotypes, and strengths of having gender-, age-, and ethnicity-specific analyses, results showed no evidence that DRD4 variation has any affects on the outcomes. Although family studies and linkage analysis are discussed in other sections of this review, it can be noted here that a combination of linkage and association analysis, as was done by Luciano and colleagues, is likely to offer greater power and validity [115]. Consequently, the argument for a role for DRD4 in tobacco use is considerably weakened.

Dopamine Metabolizers

Polymorphisms in genes that code for dopamine-metabolizing enzymes have also been linked to smoking behavior [116]. For example, the gene for monoamine oxidase A (MAO-A) has a point mutation (G→C) polymorphism causing an MAO-A deficiency. This polymorphism has been linked both to abnormal aggressiveness [117] and elevated brain levels of serotonin and dopamine [118]. A study in smokers found that individuals with the MAO-A 1460 TT/TO polymorphism smoke a greater number of cigarettes compared to those with CC/CT/CO (RR: 2.9, 95% CI: 0.6–5.1, p = 0.013) [116]. The C allele was less frequent in heavy smokers (>20 cig/day) compared to light smokers (<10 cig/day) (RR: 0.3, 95% CI: 0.1–0.7, p = 0.012). Another enzyme, dopamine B-hydroxylase (DBH), was also found to be associated with amount smoked per day. However, this finding reached statistical significance only in women, and the effect was small (those with the DBH 1368 GG genotype smoked 3.8 fewer cig/day). Heavy smokers were more likely to have the DBH 1368A allele, compared to light smokers (RR 2.3, 95% CI 1.1–5.0, P = 0.024).

Since cigarette smoking reduces the activity of MAO-A [40, 119] and other enzymes, the relationship between these factors is complex and needs clarification. Future studies may also like to consider functional polymorphisms in MAO-B and catechol-O-methyltransferase (COMT), which also metabolize dopamine [33].

Nicotine Metabolization Pathways: CYP2A6

Genetically-based individual variation in the metabolization of nicotine presents an additional pathway by which polymorphisms may influence smoking behavior. In humans, approximately 70–80% of nicotine is metabolized into cotinine [42], primarily by the enzyme cytochrome P450 2A6 (CYP2A6) [120, 121]. This enzyme, responsible for the clearance of numerous chemicals, is predominantly expressed in the liver but also in respiratory organs [121]. Considerable interindividual variability in the activity of CYP2A6 has been seen in studies of human liver microsomes [122–125], as well as phenotyping studies using various ethnic populations [126–129]. Specifically, greater than 30-fold variation in nicotine-to-cotinine Vmax values [120] was revealed in a study of human liver microsomes. Although environmental compounds can affect its activity, either inducing the enzyme (e.g., phenobarbital) or decreasing its expression (e.g., nicotine) [130, 131], the CYP2A6 gene also expresses genetic polymorphisms that may contribute to the variation in nicotine metabolism and therefore smoking behavior.

Numerous alleles have been found for the CYP2A6 gene locus on chromosome 19 (19q12 – 19q13.2). The wild-type, which is fully functional to metabolize nicotine, is the CYP2A6*1 allele. The primary null alleles are the *2, *4, *5 alleles. Since these null alleles code for inactivity, individuals homozygous for these alleles cannot metabolize nicotine through the CYP2A6 pathway. Other defective alleles (*6, *7, *9, *10, *11, *12) have reduced activity. Furthermore, a duplicate wild-type allele has been found, which confers increased activity. The frequencies of the defective alleles are low (1–3%) in Caucasian and European populations, but more common in Asian populations, where approximately 15% carry a null allele [132, 133]. Although allele frequencies in African-Americans have not been reported, this group tends to have a generally decreased nicotine clearance, which may be due to decreased CYP2A6 activity [134]. Also, males tend to have faster rates of nicotine metabolism than do women [120, 135, 136].

The finding that smokers adjust the amount and intensity of their smoking to maintain peripheral and central nicotine levels [137, 138] provides the basis for the proposed mechanism by which the defective alleles influence smoking behavior. The hypothesis is that those who carry a variant allele maintain greater levels of nicotine in their system for longer time periods after a given dose of tobacco; therefore, they need less external sources of nicotine. An additional consequence of a slower metabolism of nicotine is that bothersome effects of nicotine may be more pronounced. Thus, people with defective alleles may be more sensitive to tobacco products and less likely to become regular or dependent users [120].

Studies testing this hypothesis confirm that after oral administration of nicotine [139] or cigarette smoking [140], those with the defective alleles, compared to homozygotes for the functional allele, have significantly higher plasma levels of nicotine and lower levels of cotinine (Table 4). Furthermore, studies have found that individuals with defective alleles smoke fewer cigarettes per day and are less likely to become nicotine dependent [141, 142]. Specifically, one study of 296 Caucasian men and women found that smokers who carry the null *2 or *4 alleles smoked fewer cigarettes per day both currently (13.5 + 2.3 vs. 19.5 + 0.7, p < 0.03) and at the time of heaviest smoking (19 vs. 29, p < 0.001), had lower breath carbon monoxide levels, and had lower cotinine levels than did individuals homozygous with the wild-type *1 fully functional allele. Interestingly, people carrying CYP2A61/1 plus duplication had higher carbon monoxide and cotinine levels, but reported smoking fewer cigarettes per day, than did people with either a single copy of the active or defective alleles. The authors hypothesize that those with the duplicate gene smoke more intensely, rather than more frequently, in order to maintain their nicotine levels [141]. A separate study confirmed findings that nicotine and cotinine levels significantly vary by genotype according to level of enzyme activity associated with the allele. For example, people with the 1/1 genotype had the lowest plasma levels of nicotine and highest levels of cotinine, while people carrying the defective *7 allele had intermediate levels, and those carrying the null *4 allele had the highest levels of nicotine and lowest levels of cotinine [139].

Table 4 Epidemiologic studies of cytochrome P450 2A6 (CYP2A6), nicotine metabolization and tobacco use in human populations.

Other cytochrome P450 enzymes that have been found to be involved in the metabolism of nicotine include CYP2B6 and CYP2D6; however, their contributions are minor [120, 121]. In human liver microsomes, CYP2B6 was found to be involved in nicotine C-oxidation at high, but not low, substrate concentrations. CYP2D6, on the other hand, is unlikely to have a role in nicotine metabolism, according to recent findings [121, 143].

In summary, studies have found polymorphisms in CYP2A6 to be associated with risk of tobacco-dependence, amount smoked, various smoking indices, and tobacco-related cancer [133, 144]. The complexity of the role of cytochrome P450 is increased by the fact that environmental chemicals and drugs may induce its activity, therefore affecting the overall clearance of nicotine from the body [120, 131]. Another concern is that these polymorphisms are quite rare in Caucasian populations, thereby limiting the ability to study the role of CYP2A6 in nicotine metabolism and smoking behavior. However, if variant alleles do indeed affect metabolism and smoking behavior, then the efficacy of nicotine-replacement therapy can be greatly increased by identifying and using inhibitors of CYP2A6 [145–147]. Therapeutic use of CYP2A6 inhibitors to mimic the decreased activity of variants may produce the same benefits of the null alleles and provide new approaches to prevention and treatment of tobacco use.

Linkage Studies

While the goal of association studies is to identify specific alleles, linkage analysis aims to identify regions of the genome likely to be involved in susceptibility to tobacco use. In 1999, a complete genome scan revealed that there are many regions with numerous consecutive markers that yield small but positive results of linkage to a dichotomous outcome of nicotine dependence [148]. However, further investigation into regions of chromosomes 2, 4, 10, 16, 17 and 18 failed to find any significant linkage. Despite inconclusive findings, the authors of this study re-analyzed this affected sibling pair data in 2004 for a hypothesis-generating study of nicotine dependence candidate genes, via linkage, epistasis and bioinformatics. The recent analysis suggests that the results of the genome scan were more informative than initially perceived, and indicates a possible etiological relevance for 10 genomic regions. In particular, a broad region on chromosome 2 (2q22.1–2q24.1) and a narrower region on chromosome 10 (10q23.33–10q25.1) nearly reached the rigorous Lander and Krugylak [149] criteria for "significant" linkage. The analysis also searched for genome-wide epistatic interactions and found four correlations of interest, one of which (correlation of D8S1145 with D22S444) was particularly large and unlikely to be due to chance. Interpretation of the epistatic correlations is hampered, however, by the fact that markers in each correlated pair did not have sizeable positive NPL Z scores [150].

To clarify the significance of these results and to form hypotheses on candidate genes in nicotine dependence, the authors looked for agreement between genes located in the 10 regions statistically implicated in their analyses with a list of genes suggested by microarray studies and human association studies. Correspondence was found for genes involved with the mitogen-activated protein kinase (MAPK) signaling system, nuclear factor kappa B (NFKB) complex, neuropeptide Y (NPY) neurotransmission, a nicotine receptor subunit (CHRNA2), the vesicular monoamine transporter (SLC18A2), genes in pathways linked to human anxiety (HTR7, TDO2, and the endozepine-related protein precursor DKFZP434A2417), and the mu-opioid receptor (OPRM1) [150].

Unfortunately, the results of this genome scan are not entirely in agreement with studies using data from the Collaborative Study on the Genetics of Alcoholism (COGA), which also found some evidence for linkage with tobacco use. In COGA, using a dichotomous outcome of ever/never smoker, a sib-pair analysis found support for linkage, particularly to regions on chromosomes 6, 9 and 14 [151]. The evidence for linkage was weaker when the analysis used a cumulative pack-year history as the outcome, although some candidate regions were still noted. Agreement between the findings of COGA and the genome scan of Sullivan and colleagues occurred in areas found to have only weak or equivocal linkage signal in the latter study, on chromosomes 4, 6, 9, 14 and 17. However, the COGA analysis does correlate to other study findings; two of the three markers shared between COGA's ever/never outcome and the pack-years outcome have been previously identified as associated with smoking; one is the locus for DBH, a smoking candidate gene, and the other is for the chromosome 5q marker D5S1354. The chromosome 5 marker was found to be the primary genetic determinant of smoking behavior in a separate COGA study published that year [152]. Interestingly, this locus is very close to the D1 dopamine receptor gene locus, which has been previously been associated with smoking [153]. Although these findings are intriguing, additional linkage and experimental studies need to clarify the relevance of these overlaps.

Linkage studies have potential to support the candidacy of the genes discussed in this review. Unfortunately, the analyses published thus far have not appreciably strengthened the evidence for these particular genes, such as DRD2 or CYP2A6. Disappointing results can be partly explained by the fact linkage studies have low power when testing genes with modest or small effects. A gene with a relative risk of 2 or less requires sample sizes that are rarely attainable (>2,500 families) to obtain 80% power [154]. Inconsistencies may also be due to the fact that the currently available genetic maps are still imprecise, despite the efforts of the Human Genome Initiative toward high-resolution physical maps and large-scale sequencing. As the number of known genes increases with time, directed genomic screening may at times become more appropriate than random genomic screening, which has been the primary method of linkage studies of tobacco use [155]. In the meantime, direct tests of association, which remain powerful despite modest relative risks, are useful tools [154]. To direct genomic screening towards the best candidates, association studies will need to find solutions to their own current limitations, as discussed in the following section.

Limitations and Trends for Future Research in the Genetics of Tobacco Use

Overall, the role of genetics in determining tobacco use seems to be an important yet complex one. Much of the difficulty in identifying key genes is the fact that they are likely to be numerous, each with small effects that together shape various phenotypes associated with smoking. Despite small effects, genetic polymorphisms are of public health significance; their attributable risk is likely to be substantial due to their frequency in the population [154]. Unfortunately, studies of single genes will often suffer from insufficient power to find such small effects. Furthermore, there are limitations of past studies that make attempts to unify their results difficult, and these issues imply certain considerations that are necessary in future studies.

A primary concern here is that the exposure variable used in these studies, namely, one particular gene, may be insufficient by itself to detect true associations. That is, it is necessary to examine gene-gene interactions and gene-environment interactions. For example, while serotonin transporter or DRD4 receptors genes do not seem to have main effects on smoking, investigations considering their interaction with environmental factors found that these gene variants are important with certain personality types or when subjects are presented with smoking cues. In future studies, the insufficient power offered by studying one single locus may be partly overcome by studying haplotypes or the association of multiple polymorphisms within and around an unknown polymorphism of significance. While family studies are needed to construct haplotypes, association studies can use statistical methods to estimate haplotype frequencies. Either a single haplotype, or a subgroup of more than one haplotype, may be of greater frequency in cases than in controls, indicating haplo-type or allelic heterogeneity [58]. Implicated genes may then be investigated in animal models to establishing possible mechanisms of action.

A separate reason for reduced power in past studies may involve the recent finding that heterosis occurs with certain genes and smoking behavior [106]. This finding suggests that in future studies, comparing the frequency of heterozygotes and genotype distribution may have better discriminative power than comparing gene frequency [156]. Furthermore, false negatives may be reduced if gender differences and gender-specific heterosis are considered in the analysis of association studies [106]. In the D2 gene, for example, studies have found that the effects of variants are in opposite directions for males and females [89, 106, 107], and heterosis has been suggested by one study, where male heterozygotes had significantly higher smoking rates than male homozygotes, while female heterozygotes were less likely to smoke than female homozygotes [106].

Inconsistencies across studies are also likely to be due to false positives, as the risk of a type I error is alarmingly high when the candidate gene approach is used. The standard p-value of 0.05 may not be appropriate, and its use has led to controversy over how to view varying results of studies of complex behaviors [157]. The mistake of relying on statistical significance to determine the truth of a hypothesis is exacerbated when one considers other sources of bias that may also lead to false positives. Confounding issues need to be anticipated and properly controlled. Although population stratification was once thought to be an important concern in confounding, it is doubtful that it is the cause of much bias in properly designed genetic association studies [158], and if needed, it can be dealt with in the design phase by matching on race and ethnicity, or in the analysis by testing for its presence, followed by adjustment with genomic control (i.e., inflating the standard chi-square) or cluster analysis methods [159]. Indeed, failure to replicate findings of positive results is most likely to be due to issues in poor study design and execution, and investigator-reporting bias. Most notably, ascertainment and selection of cases and control subjects, sample size estimation, response rates, and laboratory and other measurement error need to be of prime concern for researchers [159].

A particular concern in the studies reviewed here is the choice of outcome variable. Many analyses use a crude dichotomy of smokers vs. non-smokers, paying no regard to light or heavy smoking and using various definitions of non-smokers. Also, it may be important to separate never users of tobacco (i.e., never smoked even one cigarette) from non-users (i.e., smoked more than one but less than 100 cigarettes) [80, 100]. Indeed, some genes (e.g. CYP) rely on exposure to tobacco to determine the level of their expression or if they will even be expressed at all [160], making it is impossible to determine the effects of tobacco use when one has never been exposed. Another limitation of the exposure assessment in these studies is that only cigarette smoking has been taken into account; no information on smokeless tobacco, cigar, or pipe use has been incorporated into tobacco use variables. Misclassification of the outcome can obscure meaningful differences between groups.

Meanwhile, studies that have used more specific smoking outcomes are difficult to compare due to the wide variety of definitions used. Examples of phenotypes used as outcomes include age of initiation, time to nicotine dependence, quantity of use, frequency of use, or ability to quit. To facilitate interpretation of findings and increase statistical power to detect associations, there has recently been a movement towards the use of endophenotypes as outcomes in genetic studies. An ideal smoking endophenotype would be one that is narrowly defined, readily identifiable, empirically related to the clinical manifestation of nicotine dependence, and associated with an underlying biological mechanism [161]. Examples include initial sensitivity to nicotine, abstinence-induced withdrawal and response to different drug challenges [56].

In conclusion, to refine the accepted role of genetics in tobacco behavior, there is a need for studies that examine multiple genetic polymorphisms and that obtain various, detailed smoking-related phenotype information. Problems of misclassification, low power, and false positives are important not only in and of themselves, but also since they may contribute to the separate problem of publication bias, in which there is a tendency to publish only positive findings. Still, positive findings from family studies were able to stimulate association studies of genes and smoking behavior. Thus far, results in the serotonin transporter gene, dopamine transporter gene, dopamine receptor genes D2 and D4, and cytochrome P450 2A6 have provided the most compelling evidence of a connection with smoking behavior. Future association studies, coupled with linkage analysis in family studies, will provide deeper insight into the relationship between genes, the environment and tobacco use. Undoubtedly, knowledge of the underlying biological processes involved in the addiction of various smokers will lead to the identification of different types of smokers. Patient-specific therapy and therefore more effective and efficient treatment for tobacco dependence will soon follow.

References

  1. Annual Smoking-Attributable Mortality, Years of Potential Life Lost, and Economic Costs – United States, 1995–1999. Centers for Disease Control and Prevention. Morbidity and Mortality Weekly Report. 2002, 51: 300-303.

  2. Cigarette smoking among adults – United States, 2000. Centers for Disease Control and Prevention. Morbidity and Mortality Weekly Report. 2002, 51: 642-645.

  3. Tobacco Dependence. World Health Organization Fact Sheet. 1999, 222: [http://www.who.int/inf-fs/en/fact222.html]

  4. WHO Urges Countries to Prevent a Tobacco Epidemic Among Women and Girls. 2001, World Health Organization Press Release WHO/27, [http://www.who.int/inf-pr-2001/en/pr2001-27.html]

  5. Results from the 2001 National Household Survey on Drug Abuse: Summary of National Findings. Office of Applied Studies, National Household Survey on Drug Abuse Series H-17. 2002, Washington, DC.: U.S. Department of Health and Human Services. Substance Abuse and Mental Health Services Administration, I.

  6. McNeill AD, West RJ, Jarvis M, Jackson P, Bryant A: Cigarette withdrawal symptoms in adolescent smokers. Psychopharmacology. 1986, 90: 533-536.

    CAS  PubMed  Google Scholar 

  7. Tobacco Use Among Young People. 1998, World Health Organization

  8. Jarvis MJ: Epidemiology of cigarette smoking and cessation. Journal of Clinical Psychiatry Monograph. 2003, 18: 6-11.

    Google Scholar 

  9. Wilcox NS, Prochaska JO, Velicer WF, Di-Clemente CC: Subject characteristics as predictors of self-change in smoking. Addictive Behaviors. 1985, 10: 407-412.

    CAS  PubMed  Google Scholar 

  10. Lerman C, Niaura R: Applying genetic approaches to the treatment of nicotine dependence. Oncogene. 2002, 21: 7412-7420.

    CAS  PubMed  Google Scholar 

  11. Nelson DE, Giovino GA, Shopland DR, Mowery PD, Mills SL, Eriksen MP: Trends in cigarette smoking among US adolescents, 1974 through 1991. American Journal of Public Health. 1995, 85: 34-40.

    PubMed Central  CAS  PubMed  Google Scholar 

  12. King G: The "race" concept in smoking: a review of the research on African Americans. Social Sciences and Medicine. 1997, 45: 1075-1087.

    CAS  Google Scholar 

  13. King TK, Borrelli B, Black C, Pinto BM, Marcus BH: Minority women and tobacco: implications for smoking cessation interventions. Annals of Behavioral Medicine. 1997, 19: 301-313.

    CAS  PubMed  Google Scholar 

  14. Zeman MV, Hiraki L, Sellers EM: Gender differences in tobacco smoking: higher relative exposure to smoke than nicotine in women. Journal of Women's Health & Gender-Based Medicine. 2002, 11: 147-153.

    CAS  Google Scholar 

  15. Hu S, Brody CL, Fisher C, Gunzerath L, Nelson ML, Sabol SZ, Sirota LA, Marcus SE, Greenberg BD, Murphy DL, Hamer DH: Interaction between the serotonin transporter gene and neuroticism in cigarette smoking behavior. Molecular Psychiatry. 2000, 5: 181-188.

    CAS  PubMed  Google Scholar 

  16. Pomerleau CS, Pomerleau OF, Flessland KA, Basson SM: Relationship of Tridimensional Personality Questionnaire scores and smoking variables in female and male smokers. Journal of Substance Abuse. 1992, 4: 143-154.

    CAS  PubMed  Google Scholar 

  17. Wills TA, Vaccaro D, McNamara G: Novelty seeking, risk taking, and related constructs as predictors of adolescent substance use: an application of Cloninger's theory. Journal of Substance Abuse. 1994, 6: 1-20.

    CAS  PubMed  Google Scholar 

  18. Masse LC, Tremblay RE: Behavior of boys in kindergarten and the onset of substance use during adolescence. Archives of General Psychiatry. 1997, 54: 62-68.

    CAS  PubMed  Google Scholar 

  19. Rowland AS, Lesesne CA, Abramowitz AJ: The epidemiology of attention-deficit/hyperactivity disorder (ADHD): a public health view. Mental Retardation and Developmental Disabilities Research Reviews. 2002, 8: 162-170.

    PubMed  Google Scholar 

  20. Breslau N, Kilbey M, Andreski P: Nicotine dependence, major depression, and anxiety in young adults. Archives of General Psychiatry. 1991, 48: 1069-1074.

    CAS  PubMed  Google Scholar 

  21. Anda RF, Williamson DF, Escobedo LG, Mast EE, Giovino GA, Remington PL: Depression and the dynamics of smoking. A national perspective. Journal of the American Medical Association. 1990, 264: 1541-1545.

    CAS  PubMed  Google Scholar 

  22. Niaura R, Britt DM, Shadel WG, Goldstein M, Abrams D, Brown R: Symptoms of depression and survival experience among three samples of smokers trying to quit. Psychology of Addictive Behaviors. 2001, 15: 13-17.

    CAS  PubMed  Google Scholar 

  23. Lerman C, Caporaso N, Main D, Audrain J, Boyd NR, Bowman ED, Shields PG: Depression and self-medication with nicotine: the modifying influence of the dopamine D4 receptor gene. Health Psychology. 1998, 17: 56-62.

    CAS  PubMed  Google Scholar 

  24. Gilbert DG, Gilbert BO: Personality, psychopathology, and nicotine response as mediators of the genetics of smoking. Behavior Genetics. 1995, 25: 133-147.

    CAS  PubMed  Google Scholar 

  25. Noble EP: D2 dopamine receptor gene in psychiatric and neurologic disorders and its phenotypes. American Journal of Medical Genetics. 2003, 116: 103-125.

    Google Scholar 

  26. Johnson EO, Bree van den MB, Uhl GR, Pickens RW: Indicators of genetic and environmental influences in drug abusing individuals. Drug and Alcohol Dependence. 1996, 41: 17-23.

    CAS  PubMed  Google Scholar 

  27. Noble EP, St Jeor ST, Ritchie T, Syndulko K, St Jeor SC, Fitch RJ, Brunner RL, Sparkes RS: D2 dopamine receptor gene and cigarette smoking: a reward gene?. Medical Hypotheses. 1994, 42: 257-260.

    CAS  PubMed  Google Scholar 

  28. Caskey NH, Jarvik ME, Wirshing WC, Madsen DC, Iwamoto-Schaap PN, Eisenberger NI, Huerta L, Terrace SM, Olmstead RE: Modulating tobacco smoking rates by dopaminergic stimulation and blockade. Nicotine & Tobacco Research. 2002, 4: 259-266.

    CAS  Google Scholar 

  29. Mihailescu S, Palomero-Rivero M, Meade-Huerta P, Maza-Flores A, Drucker-Colin R: Effects of nicotine and mecamylamine on rat dorsal raphe neurons. European Journal of Pharmacology. 1998, 360: 31-36.

    CAS  PubMed  Google Scholar 

  30. Ribeiro EB, Bettiker RL, Bogdanov M, Wurtman RJ: Effects of systemic nicotine on serotonin release in rat brain. Brain Research. 1993, 621: 311-8.

    CAS  PubMed  Google Scholar 

  31. Blum K, Braverman ER, Holder JM, Lubar JF, Monastra VJ, Miller D, Lubar JO, Chen TJ, Comings DE: Reward deficiency syndrome: a biogenetic model for the diagnosis and treatment of impulsive, addictive, and compulsive behaviors. Journal of Psychoactive Drugs. 2000, 32 (Suppl): i-iv. 1–112

    PubMed  Google Scholar 

  32. Dursun SM, Kutcher S: Smoking, nicotine and psychiatric disorders: evidence for therapeutic role, controversies and implications for future research. Medical Hypotheses. 1999, 52: 101-109.

    CAS  PubMed  Google Scholar 

  33. Arinami T, Ishiguro H, Onaivi ES: Polymorphisms in genes involved in neurotransmission in relation to smoking. European Journal of Pharmacology. 2000, 410: 215-226.

    CAS  PubMed  Google Scholar 

  34. Rose JE, Behm FM, Westman EC, Mathew RJ, London ED, Hawk TC, Turkington TG, Coleman RE: PET studies of the influences of nicotine on neural systems in cigarette smokers. American Journal of Psychiatry. 2003, 160: 323-333.

    PubMed  Google Scholar 

  35. Volkow ND, Fowler JS, Ding YS, Wang GJ, Gatley SJ: Imaging the neurochemistry of nicotine actions: studies with positron emission tomography. Nicotine & Tobacco Research. 1999, 1: S127-132. discussion S139–140

    Google Scholar 

  36. Salokangas RK, Vilkman H, Ilonen T, Taiminen T, Bergman J, Haaparanta M, Solin O, Alanen A, Syvalahti E, Hietala J: High levels of dopamine activity in the basal ganglia of cigarette smokers. American Journal of Psychiatry. 2000, 157: 632-634.

    CAS  PubMed  Google Scholar 

  37. Czermak C, Lehofer M, Wagner EM, Prietl B, Gorkiewicz G, Lemonis L, Rohrhofer A, Legl T, Schauenstein K, Liebmann PM: Reduced dopamine D3 receptor expression in blood lymphocytes of smokers is negatively correlated with daily number of smoked cigarettes: a peripheral correlate of dopaminergic alterations in smokers. Nicotine & Tobacco Research. 2004, 6: 49-54.

    CAS  Google Scholar 

  38. Martin-Solch C, Magyar S, Kunig G, Missimer J, Schultz W, Leenders KL: Changes in brain activation associated with reward processing in smokers and nonsmokers. A positron emission tomography study. Experimental Brain Research. 2001, 139: 278-286.

    CAS  PubMed  Google Scholar 

  39. Buccafusco JJ, Terry AV: The potential role of cotinine in the cognitive and neuroprotective actions of nicotine. Life Sciences. 2003, 72: 2931-2942.

    CAS  PubMed  Google Scholar 

  40. Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, MacGregor R, Alexoff D, Shea C, Schlyer D, Wolf AP, Warner D, Zezulkova I, Cilento R: Inhibition of monoamine oxidase B in the brains of smokers. Nature. 1996, 379: 733-736.

    CAS  PubMed  Google Scholar 

  41. Fowler JS, Wang GJ, Volkow ND, Franceschi D, Logan J, Pappas N, Shea C, MacGregor RR, Garza V: Smoking a single cigarette does not produce a measurable reduction in brain MAO B in non-smokers. Nicotine & Tobacco Research. 1999, 1: 325-329.

    CAS  Google Scholar 

  42. Benowitz NL, Jacob P: Metabolism of nicotine to cotinine studied by a dual stable isotope method. Clinical Pharmacology and Therapeutics. 1994, 56: 483-493.

    CAS  PubMed  Google Scholar 

  43. Hatsukami DK, Grillo M, Pentel PR, Oncken C, Bliss R: Safety of cotinine in humans: physiologic, subjective, and cognitive effects. Pharmacology, Biochemistry and Behavior. 1997, 57: 643-650.

    CAS  PubMed  Google Scholar 

  44. Balfour DJK: The psychopharmacology of tobacco dependence. Journal of Clinical Psychiatry Monograph. 2003, 18: 12-21.

    Google Scholar 

  45. Heath AC, Kirk KM, Meyer JM, Martin NG: Genetic and social determinants of initiation and age at onset of smoking in Australian twins. Behavior Genetics. 1999, 29: 395-407.

    CAS  PubMed  Google Scholar 

  46. Madden PA, Heath AC, Pedersen NL, Kaprio J, Koskenvuo MJ, Martin NG: The genetics of smoking persistence in men and women: a multicultural study. Behavior Genetics. 1999, 29: 423-431.

    CAS  PubMed  Google Scholar 

  47. Sullivan PF, Kendler KS: The genetic epidemiology of smoking. Nicotine & Tobacco Research. 1999, 1: S51-57. discussion S69–70

    Google Scholar 

  48. Batra V, Patkar AA, Berrettini WH, Weinstein SP, Leone FT: The genetic determinants of smoking. Chest. 2003, 123: 1730-1739.

    CAS  PubMed  Google Scholar 

  49. Fisher RA: Lung cancer and cigarettes. Nature. 1958, 182: 108.

    CAS  PubMed  Google Scholar 

  50. Kendler KS, Neale MC, Sullivan P, Corey LA, Gardner CO, Prescott CA: A population-based twin study in women of smoking initiation and nicotine dependence. Psychological Medicine. 1999, 29: 299-308.

    CAS  PubMed  Google Scholar 

  51. Xian H, Scherrer J, Madden P, Lyons M, Tsuang M, True W, Eisen S: The heritability of failed smoking cessation and nicotine withdrawal in twins who smoked and attempted to quit. Nicotine & Tobacco Research. 2003, 5: 245-254.

    CAS  Google Scholar 

  52. Koopmans JR, Slutske WS, Heath AC, Neale MC, Boomsma DI: The genetics of smoking initiation and quantity smoked in Dutch adolescent and young adult twins. Behavior Genetics. 1999, 29: 383-393.

    CAS  PubMed  Google Scholar 

  53. Stallings MC, Hewitt JK, Beresford T, Heath AC, Eaves LJ: A twin study of drinking and smoking onset and latencies from first use to regular use. Behavior Genetics. 1999, 29: 409-421.

    CAS  PubMed  Google Scholar 

  54. Legrand LN, McGue M, Iacono WG: Searching for interactive effects in the etiology of early-onset substance use. Behavior Genetics. 1999, 29: 433-444.

    CAS  PubMed  Google Scholar 

  55. Osler M, Holst C, Prescott E, Sorensen TI: Influence of genes and family environment on adult smoking behavior assessed in an adoption study. Genetic Epidemiology. 2001, 21: 193-200.

    CAS  PubMed  Google Scholar 

  56. Lerman C, Berrettini W: Elucidating the role of genetic factors in smoking behavior and nicotine dependence. American Journal of Medical Genetics. 2003, 118B: 48-54.

    PubMed  Google Scholar 

  57. Malhotra AK, Goldman D: Benefits and pitfalls encountered in psychiatric genetic association studies. Biological Psychiatry. 1999, 45: 544-550.

    CAS  PubMed  Google Scholar 

  58. Schork NJ, Fallin D, Thiel B, Xu X, Broeckel U, Jacob HJ, Cohen D: The future of genetic case-control studies. Advances in Genetics. 2001, 42: 191-212.

    CAS  PubMed  Google Scholar 

  59. Shi J, Hui L, Xu Y, Wang F, Huang W, Hu G: Sequence variations in the mu-opioid receptor gene (OPRM1) associated with human addiction to heroin. Human Mutation. 2002, 19: 459-460.

    PubMed  Google Scholar 

  60. Hutcheson DM, Matthes HW, Valjent E, Sanchez-Blazquez P, Rodriguez-Diaz M, Garzon J, Kieffer BL, Maldonado R: Lack of dependence and rewarding effects of deltorphin II in mu-opioid receptor-deficient mice. European Journal of Neuroscience. 2001, 13: 153-161.

    CAS  PubMed  Google Scholar 

  61. Ghozland S, Matthes HW, Simonin F, Filliol D, Kieffer BL, Maldonado R: Motivational effects of cannabinoids are mediated by mu-opioid and kappa-opioid receptors. Journal of Neuroscience. 2002, 22: 1146-1154.

    CAS  PubMed  Google Scholar 

  62. Roberts AJ, McDonald JS, Heyser CJ, Kieffer BL, Matthes HW, Koob GF, Gold LH: mu-Opioid receptor knockout mice do not self-administer alcohol. Journal of Pharmacology and Experimental Therapeutics. 2000, 293: 1002-1008.

    CAS  PubMed  Google Scholar 

  63. Berrendero F, Kieffer BL, Maldonado R: Attenuation of nicotine-induced antinociception, rewarding effects, and dependence in mu-opioid receptor knock-out mice. Journal of Neuroscience. 2002, 22: 10935-10940.

    CAS  PubMed  Google Scholar 

  64. Comings DE, Blum K: Reward deficiency syndrome: genetic aspects of behavioral disorders. Progress in Brain Research. 2000, 126: 325-341.

    CAS  PubMed  Google Scholar 

  65. Picciotto MR, Corrigall WA: Neuronal systems underlying behaviors related to nicotine addiction: neural circuits and molecular genetics. Journal of Neuroscience. 2002, 22: 3338-3341.

    CAS  PubMed  Google Scholar 

  66. Silverman MA, Neale MC, Sullivan PF, Harris-Kerr C, Wormley B, Sadek H, Ma Y, Kendler KS, Straub RE: Haplotypes of four novel single nucleotide polymorphisms in the nicotinic acetylcholine receptor beta2-subunit (CHRNB2) gene show no association with smoking initiation or nicotine dependence. American Journal of Medical Genetics. 2000, 96: 646-653.

    CAS  PubMed  Google Scholar 

  67. Duga S, Solda G, Asselta R, Bonati MT, Dalpra L, Malcovati M, Tenchini ML: Characterization of the genomic structure of the human neuronalnicotinic acetylcholine receptor CHRNA5/A3/B4 gene cluster and identification of novel intragenic polymorphisms. Journal of Human Genetics. 2001, 46: 640-648.

    CAS  PubMed  Google Scholar 

  68. Gault J, Robinson M, Berger R, Drebing C, Logel J, Hopkins J, Moore T, Jacobs S, Meriwether J, Choi MJ, Kim EJ, Walton K, Buiting K, Davis A, Breese C, Freedman R, Leonard S: Genomic organization and partial duplication of the human alpha7 neuronal nicotinic acetylcholine receptor gene (CHRNA7). Genomics. 1998, 52: 173-185.

    CAS  PubMed  Google Scholar 

  69. Nielsen DA, Jenkins GL, Stefanisko KM, Jefferson KK, Goldman D: Sequence, splice site and population frequency distribution analyses of the polymorphic human tryptophan hydroxylase intron 7. Brain Research. Molecular Brain Research. 1997, 45: 145-148.

    CAS  PubMed  Google Scholar 

  70. Manuck SB, Flory JD, Ferrell RE, Dent KM, Mann JJ, Muldoon MF: Aggression and anger-related traits associated with a polymorphism of the tryptophan hydroxylase gene. Biological Psychiatry. 1999, 45: 603-614.

    CAS  PubMed  Google Scholar 

  71. Sullivan PF, Jiang Y, Neale MC, Kendler KS, Straub RE: Association of the tryptophan hydroxylase gene with smoking initiation but not progression to nicotine dependence. American Journal of Medical Genetics. 2001, 105: 479-484.

    CAS  PubMed  Google Scholar 

  72. Lerman C, Caporaso NE, Bush A, Zheng YL, Audrain J, Main D, Shields PG: Tryptophan hydroxylase gene variant and smoking behavior. American Journal of Medical Genetics. 2001, 105: 518-520.

    CAS  PubMed  Google Scholar 

  73. Ebstein RP, Segman R, Benjamin J, Osher Y, Nemanov L, Belmaker RH: 5-HT2C (HTR2C) serotonin receptor gene polymorphism associated with the human personality trait of reward dependence: interaction with dopamine D4 receptor (D4DR) and dopamine D3 receptor (D3DR) polymorphisms. American Journal of Medical Genetics. 1997, 74: 65-72.

    CAS  PubMed  Google Scholar 

  74. Lesch KP, Bengel D, Heils A, Sabol SZ, Greenberg BD, Petri S, Benjamin J, Muller CR, Hamer DH, Murphy DL: Association of anxiety-related traits with a polymorphism in the serotonin transporter gene regulatory region. Science. 1996, 274: 1527-1531.

    CAS  PubMed  Google Scholar 

  75. Heils A, Mossner R, Lesch KP: The human serotonin transporter gene polymorphism – basic research and clinical implications. Journal of Neural Transmission. 1997, 104: 1005-1014.

    CAS  PubMed  Google Scholar 

  76. Lerman C, Shields PG, Audrain J, Main D, Cobb B, Boyd NR, Caporaso N: The role of the serotonin transporter gene in cigarette smoking. Cancer Epidemiolgy, Biomarkers & Prevention. 1998, 7: 253-255.

    CAS  Google Scholar 

  77. Ishikawa H, Ohtsuki T, Ishiguro H, Yamakawa-Kobayashi K, Endo K, Lin YL, Yanagi H, Tsuchiya S, Kawata K, Hamaguchi H, Arinami T: Association between serotonin transporter gene polymorphism and smoking among Japanese males. Cancer Cancer Epidemiolgy, Biomarkers & Prevention. 1999, 8: 831-833.

    CAS  Google Scholar 

  78. Lerman C, Caporaso NE, Audrain J, Main D, Boyd NR, Shields PG: Interacting effects of the serotonin transporter gene and neuroticism in smoking practices and nicotine dependence. Molecular Psychiatry. 2000, 5: 189-192.

    CAS  PubMed  Google Scholar 

  79. Kang AM, Palmatier MA, Kidd KK: Global variation of a 40-bp VNTR in the 3'-untranslated region of the dopamine transporter gene (SLC6A3). Biological Psychiatry. 1999, 46: 151-160.

    CAS  PubMed  Google Scholar 

  80. Vandenbergh DJ, Kozlowski LT, Bennett CJ, Grant MD, Strasser AA, O'Connor R, Stauffer RL, Vogler GP: DAT's not all, but it may be more than we realize. Nicotine & Tobacco Research. 2002, 4: 251-252.

    Google Scholar 

  81. Jacobsen LK, Staley JK, Zoghbi SS, Seibyl JP, Kosten TR, Innis RB, Gelernter J: Prediction of dopamine transporter binding availability by genotype: a preliminary report. American Journal of Psychiatry. 2000, 157: 1700-1703.

    CAS  PubMed  Google Scholar 

  82. Heinz A, Goldman D, Jones DW, Palmour R, Hommer D, Gorey JG, Lee KS, Linnoila M, Weinberger DR: Genotype influences in vivo dopamine transporter availability in human striatum. Neuropsychopharmacology. 2000, 22: 133-139.

    CAS  PubMed  Google Scholar 

  83. Wong AH, Buckle CE, Van Tol HH: Polymorphisms in dopamine receptors: what do they tell us?. European Journal of Pharmacology. 2000, 410: 183-203.

    CAS  PubMed  Google Scholar 

  84. Noble EP, Blum K, Ritchie T, Montgomery A, Sheridan PJ: Allelic association of the D2 dopamine receptor gene with receptor-binding characteristics in alcoholism. Archives of General Psychiatry. 1991, 48: 648-654.

    CAS  PubMed  Google Scholar 

  85. Smith SS, O'Hara BF, Persico AM, Gorelick DA, Newlin DB, Vlahov D, Solomon L, Pickens R, Uhl GR: Genetic vulnerability to drug abuse. The D2 dopamine receptor Taq I B1 restriction fragment length polymorphism appears more frequently in polysubstance abusers. Archives of General Psychiatry. 1992, 49: 723-727.

    CAS  PubMed  Google Scholar 

  86. Spitz MR, Shi H, Yang F, Hudmon KS, Jiang H, Chamberlain RM, Amos CI, Wan Y, Cinciripini P, Hong WK, Wu X: Case-control study of the D2 dopamine receptor gene and smoking status in lung cancer patients. Journal of the National Cancer Institute. 1998, 90: 358-363.

    CAS  PubMed  Google Scholar 

  87. Gelernter J, Southwick S, Goodson S, Morgan A, Nagy L, Charney DS: No association between D2 dopamine receptor (DRD2) "A" system alleles, or DRD2 haplotypes, and posttraumatic stress disorder. Biological Psychiatry. 1999, 45: 620-625.

    CAS  PubMed  Google Scholar 

  88. Pohjalainen T, Rinne JO, Nagren K, Lehikoinen P, Anttila K, Syvalahti EK, Hietala J: The A1 allele of the human D2 dopamine receptor gene predicts low D2 receptor availability in healthy volunteers. Molecular Psychiatry. 1998, 3: 256-260.

    CAS  PubMed  Google Scholar 

  89. Thompson J, Thomas N, Singleton A, Piggott M, Lloyd S, Perry EK, Morris CM, Perry RH, Ferrier IN, Court JA: D2 dopamine receptor gene (DRD2) Taq1 A polymorphism: reduced dopamine D2 receptor binding in the human striatum associated with the A1 allele. Pharmacogenetics. 1997, 7: 479-484.

    CAS  PubMed  Google Scholar 

  90. Jonsson EG, Nothen MM, Grunhage F, Farde L, Nakashima Y, Propping P, Sedvall GC: Polymorphisms in the dopamine D2 receptor gene and their relationships to striatal dopamine receptor density of healthy volunteers. Molecular Psychiatry. 1999, 4: 290-296.

    CAS  PubMed  Google Scholar 

  91. Laine TP, Ahonen A, Rasanen P, Pohjalainen T, Tiihonen J, Hietala J: The A1 allele of the D2 dopamine receptor gene is associated with high dopamine transporter density in detoxified alcoholics. Alcohol and Alcoholism. 2001, 36: 262-265.

    CAS  PubMed  Google Scholar 

  92. Comings DE: Molecular heterosis as the explanation for the controversy about the effect of the DRD2 gene on dopamine D2 receptor density. Molecular Psychiatry. 1999, 4: 213-215.

    CAS  PubMed  Google Scholar 

  93. Van Tol HH, Wu CM, Guan HC, Ohara K, Bunzow JR, Civelli O, Kennedy J, Seeman P, Niznik HB, Jovanovic V: Multiple dopamine D4 receptor variants in the human population. Nature. 1992, 358: 149-152.

    CAS  PubMed  Google Scholar 

  94. Asghari V, Sanyal S, Buchwaldt S, Paterson A, Jovanovic V, Van Tol HH: Modulation of intracellular cyclic AMP levels by different human dopamine D4 receptor variants. Journal of Neurochemistry. 1995, 65: 1157-1165.

    CAS  PubMed  Google Scholar 

  95. Van Tol HH, Bunzow JR, Guan HC, Sunahara RK, Seeman P, Niznik HB, Civelli O: Cloning of the gene for a human dopamine D4 receptor with high affinity for the antipsychotic clozapine. Nature. 1991, 350: 610-614.

    CAS  PubMed  Google Scholar 

  96. Svingos AL, Periasamy S, Pickel VM: Presynaptic dopamine D(4) receptor localization in the rat nucleus accumbens shell. Synapse. 2000, 36: 222-232.

    CAS  PubMed  Google Scholar 

  97. Feldpausch DL, Needham LM, Stone MP, Althaus JS, Yamamoto BK, Svensson KA, Merchant KM: The role of dopamine D4 receptor in the induction of behavioral sensitization to amphetamine and accompanying biochemical and molecular adaptations. Journal of Pharmacology and Experimental Therapeutics. 1998, 286: 497-508.

    CAS  PubMed  Google Scholar 

  98. Yoshida K, Hamajima N, Kozaki K, Saito H, Maeno K, Sugiura T, Ookuma K, Takahashi T: Association between the dopamine D2 receptor A2/A2 genotype and smoking behavior in the Japanese. Cancer Epidemiolgy, Biomarkers & Prevention. 2001, 10: 403-405.

    CAS  Google Scholar 

  99. Vandenbergh DJ, Grant MD, O'Connor R, Strasser AA, Steverns V, Vogler GP, Kozlowski LT: The influence of dopamine receptor genes (DRD2, DRD3, & DRD4) on smoking status: smokers and non-smokers study (SANS). Abstract presented at Society for Research on Nicotine and Tobacco 2003 annual meeting: Nicotine and Tobacco Research. 2003

    Google Scholar 

  100. Vandenbergh DJ, Bennett CJ, Grant MD, Strasser AA, O'Connor R, Stauffer RL, Vogler GP, Kozlowski LT: Smoking status and the human dopamine transporter variable number of tandem repeats (VNTR) polymorphism: failure to replicate and finding that never-smokers may be different. Nicotine & Tobacco Research. 2002, 4: 333-340.

    Google Scholar 

  101. Wu X, Hudmon KS, Detry MA, Chamberlain RM, Spitz MR: D2 dopamine receptor gene polymorphisms among African-Americans and Mexican-Americans: a lung cancer case-control study. Cancer Epidemiolgy, Biomarkers & Prevention. 2000, 9: 1021-1026.

    CAS  Google Scholar 

  102. Bierut LJ, Rice JP, Edenberg HJ, Goate A, Foroud T, Cloninger CR, Begleiter H, Conneally PM, Crowe RR, Hesselbrock V, Li TK, Nurnberger JI, Porjesz B, Schuckit MA, Reich T: Family-based study of the association of the dopamine D2 receptor gene (DRD2) with habitual smoking. American Journal of Medical Genetics. 2000, 90: 299-302.

    CAS  PubMed  Google Scholar 

  103. Lerman C, Caporaso NE, Audrain J, Main D, Bowman ED, Lockshin B, Boyd NR, Shields PG: Evidence suggesting the role of specific genetic factors in cigarette smoking. Health Psychology. 1999, 18: 14-20.

    CAS  PubMed  Google Scholar 

  104. Sabol SZ, Nelson ML, Fisher C, Gunzerath L, Brody CL, Hu S, Sirota LA, Marcus SE, Greenberg BD, Lucas FRt, Benjamin J, Murphy DL, Hamer DH: A genetic association for cigarette smoking behavior. Health Psychology. 1999, 18: 7-13.

    CAS  PubMed  Google Scholar 

  105. Comings DE, Ferry L, Bradshaw-Robinson S, Burchette R, Chiu C, Muhleman D: The dopamine D2 receptor (DRD2) gene: a genetic risk factor in smoking. Pharmacogenetics. 1996, 6: 73-79.

    CAS  PubMed  Google Scholar 

  106. Lee HS: Gender-specific molecular heterosis and association studies: Dopamine D2 receptor gene and smoking. American Journal of Medical Genetics. 2003, 118B: 55-59.

    PubMed  Google Scholar 

  107. Dluzen DE, Anderson LI: Estrogen differentially modulates nicotine-evoked dopamine release from the striatum of male and female rats. Neuroscience Letters. 1997, 230: 140-142.

    CAS  PubMed  Google Scholar 

  108. Erblich J, Lerman C, Self DW, Diaz GA, Bovbjerg DH: Stress-induced cigarette craving: effects of the DRD2 TaqI RFLP and SLC6A3 VNTR polymorphisms. Pharmacogenomics Journal. 2004, 4 (2): 102-109.

    CAS  PubMed  Google Scholar 

  109. Johnstone E, Draper L, York E, Jacob R, Griffiths S, Murphy M, Walton RT: Characterisation of genetic variation in the human dopamine transporter gene and its relevance to smoking behavior. Abstract presented at Society for Research on Nicotine and Tobacco 2003 annual meeting: Nicotine and Tobacco Research. 2003

    Google Scholar 

  110. Uhl G, Elme GI, LaBuda MC, Pickens RW: Genetic Influences in Drug Abuse. Psychopharmacology – The Fourth Generation of Progress: An Official Publication of the American College of Neuropsychopharmacology. Edited by: F B. 2000, Raven Press

    Google Scholar 

  111. Ring HZ, Swan GE, Valdes AM, Jack LM, Ton CC, Curry SJ, McAfee T: Dopamine receptor DRD2 genotype and smoking cessation outcome following treatment with bupropion SR. Abstract presented at Society for Research on Nicotine and Tobacco 2003 annual meeting: Nicotine and Tobacco Research. 2003

    Google Scholar 

  112. David SP, Niaura R, Papandonatos GD, Shadel WG, Burkholder GJ, Britt DM, Day A, Stumpff J, Hutchison K, Murphy M, Johnstone E, Griffiths SE, Walton RT: Does the DRD2-Taq1 A polymorphism influence treatment response to bupropion hydrochloride for reduction of the nicotine withdrawal syndrome?. Nicotine & Tobacco Research. 2003, 5: 935-942.

    CAS  Google Scholar 

  113. Shields PG, Lerman C, Audrain J, Bowman ED, Main D, Boyd NR, Caporaso NE: Dopamine D4 receptors and the risk of cigarette smoking in African-Americans and Caucasians. Cancer Epidemiology, Biomarkers and Prevention. 1998, 7: 453-458.

    CAS  PubMed  Google Scholar 

  114. Hutchison KE, LaChance H, Niaura R, Bryan A, Smolen A: The DRD4 VNTR polymorphism influences reactivity to smoking cues. Journal of Abnormal Psychology. 2002, 111: 134-143.

    PubMed  Google Scholar 

  115. Luciano M, Zhu G, Kirk KM, Whitfield JB, Butler R, Heath AC, Madden PA, Martin NG: Effects of dopamine receptor D4 variation on alcohol and tobacco use and on novelty seeking: multivariate linkage and association analysis. American Journal of Medical Genetics. 2004, 124B: 113-123.

    PubMed  Google Scholar 

  116. McKinney EF, Walton RT, Yudkin P, Fuller A, Haldar NA, Mant D, Murphy M, Welsh KI, Marshall SE: Association between polymorphisms in dopamine metabolic enzymes and tobacco consumption in smokers. Pharmacogenetics. 2000, 10: 483-491.

    CAS  PubMed  Google Scholar 

  117. Brunner HG, Nelen M, Breakefield XO, Ropers HH, van Oost BA: Abnormal behavior associated with a point mutation in the structural gene for monoamine oxidase A. Science. 1993, 262: 578-580.

    CAS  PubMed  Google Scholar 

  118. Shih JC, Chen K, Ridd MJ: Monoamine oxidase: from genes to behavior. Annual Review of Neuroscience. 1999, 22: 197-217.

    PubMed Central  CAS  PubMed  Google Scholar 

  119. Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, Shea C, Alexoff D, MacGregor RR, Schlyer DJ, Zezulkova I, Wolf AP: Brain monoamine oxidase A inhibition in cigarette smokers. Proceedings of the National Academy of Sciences of the United States of America. 1996, 93: 14065-14069.

    PubMed Central  CAS  PubMed  Google Scholar 

  120. Messina ES, Tyndale RF, Sellers EM: A major role for CYP2A6 in nicotine C-oxidation by human liver microsomes. Journal of Pharmacology and Experimental Therapeutics. 1997, 282: 1608-1614.

    CAS  PubMed  Google Scholar 

  121. Yamazaki H, Inoue K, Hashimoto M, Shimada T: Roles of CYP2A6 and CYP2B6 in nicotine C-oxidation by human liver microsomes. Archives of Toxicology. 1999, 73: 65-70.

    CAS  PubMed  Google Scholar 

  122. Shimada T, Yamazaki H, Guengerich FP: Ethnic-related differences in coumarin 7-hydroxylation activities catalyzed by cytochrome P4502A6 in liver microsomes of Japanese and Caucasian populations. Xenobiotica. 1996, 26: 395-403.

    CAS  PubMed  Google Scholar 

  123. Yun CH, Shimada T, Guengerich FP: Purification and characterization of human liver microsomal cytochrome P-450 2A6. Molecular Pharmacology. 1991, 40: 679-685.

    CAS  PubMed  Google Scholar 

  124. Miles JS, McLaren AW, Forrester LM, Glancey MJ, Lang MA, Wolf CR: Identification of the human liver cytochrome P-450 responsible for coumarin 7-hydroxylase activity. Biochemical Journal. 1990, 267: 365-371.

    PubMed Central  CAS  PubMed  Google Scholar 

  125. Camus AM, Geneste O, Honkakoski P, Bereziat JC, Henderson CJ, Wolf CR, Bartsch H, Lang MA: High variability of nitrosamine metabolism among individuals: role of cytochromes P450 2A6 and 2E1 in the dealkylation of N-nitrosodimethylamine and N-nitrosodiethylamine in mice and humans. Molecular Carcinogenesis. 1993, 7: 268-275.

    CAS  PubMed  Google Scholar 

  126. Cholerton S, Idle ME, Vas A, Gonzalez FJ, Idle JR: Comparison of a novel thin-layer chromatographic-fluorescence detection method with a spectrofluorometric method for the determination of 7-hydroxycoumarin in human urine. Journal of Chromatography. 1992, 575: 325-330.

    CAS  PubMed  Google Scholar 

  127. Rautio A, Kraul H, Kojo A, Salmela E, Pelkonen O: Interindividual variability of coumarin 7-hydroxylation in healthy volunteers. Pharmacogenetics. 1992, 2: 227-233.

    CAS  PubMed  Google Scholar 

  128. Iscan M, Rostami H, Guray T, Pelkonen O, Rautio A: Interindividual variability of coumarin 7-hydroxylation in a Turkish population. European Journal of Clinical Pharmacology. 1994, 47: 315-318.

    CAS  PubMed  Google Scholar 

  129. Hadidi H, Irshaid Y, Vagbo CB, Brunsvik A, Cholerton S, Zahlsen K, Idle JR: Variability of coumarin 7- and 3-hydroxylation in a Jordanian population is suggestive of a functional polymorphism in cytochrome P450 CYP2A6. European Journal of Clinical Pharmacology. 1998, 54: 437-441.

    CAS  PubMed  Google Scholar 

  130. Pearce R, Greenway D, Parkinson A: Species differences and interindividual variation in liver microsomal cytochrome P450 2A enzymes: effects on coumarin, dicumarol, and testosterone oxidation. Archives of Biochemistry and Biophysics. 1992, 298: 211-225.

    CAS  PubMed  Google Scholar 

  131. Schoedel KA, Sellers EM, Palmour R, Tyndale RF: Down-regulation of hepatic nicotine metabolism and a CYP2A6-like enzyme in African green monkeys after long-term nicotine administration. Molecular Pharmacology. 2003, 63: 96-104.

    CAS  PubMed  Google Scholar 

  132. Oscarson M, McLellan RA, Gullsten H, Agundez JA, Benitez J, Rautio A, Raunio H, Pelkonen O, Ingelman-Sundberg M: Identification and characterisation of novel polymorphisms in the CYP2A locus: implications for nicotine metabolism. FEBS Letters. 1999, 460: 321-327.

    CAS  PubMed  Google Scholar 

  133. Oscarson M: Genetic polymorphisms in the cytochrome P450 2A6 (CYP2A6) gene: implications for interindividual differences in nicotine metabolism. Drug Metabolism and Disposition. 2001, 29: 91-95.

    CAS  PubMed  Google Scholar 

  134. Benowitz NL, Perez-Stable EJ, Fong I, Modin G, Herrera B, Jacob P: Ethnic differences in N-glucuronidation of nicotine and cotinine. Journal of Pharmacology and Experimental Therapeutics. 1999, 291: 1196-1203.

    CAS  PubMed  Google Scholar 

  135. Benowitz NL, Jacob P: Daily intake of nicotine during cigarette smoking. Clinical Pharmacology and Therapeutics. 1984, 35: 499-504.

    CAS  PubMed  Google Scholar 

  136. Beckett AH, Gorrod JW, Jenner P: The effect of smoking on nicotine metabolism in vivo in man. Journal of Pharmacy and Pharmacology. 1971, 23: 62S-67S.

    CAS  PubMed  Google Scholar 

  137. Russel MSH: Nicotine intake and its regulation by smokers: Tobacco smoking and nicotine. Advances in Behavioral Biology. Edited by: Martin WR, Vauhon GR, Iwamoto ET, David L. 1997, New York: Plenum Press, 25-50.

    Google Scholar 

  138. McMorrow MJ, Foxx RM: Nicotine's role in smoking: an analysis of nicotine regulation. Psychological Bulletin. 1983, 93: 302-327.

    CAS  PubMed  Google Scholar 

  139. Xu C, Goodz S, Sellers EM, Tyndale RF: CYP2A6 genetic variation and potential consequences. Advanced Drug Delivery Reviews. 2002, 54: 1245-1256.

    CAS  PubMed  Google Scholar 

  140. Kitagawa K, Kunugita N, Katoh T, Yang M, Kawamoto T: The significance of the homozygous CYP2A6 deletion on nicotine metabolism: a new genotyping method of CYP2A6 using a single PCR-RFLP. Biochemical and Biophysical Research Communications. 1999, 262: 146-151.

    CAS  PubMed  Google Scholar 

  141. Rao Y, Hoffmann E, Zia M, Bodin L, Zeman M, Sellers EM, Tyndale RF: Duplications and defects in the CYP2A6 gene: identification, genotyping, and in vivo effects on smoking. Molecular Pharmacology. 2000, 58: 747-755.

    CAS  PubMed  Google Scholar 

  142. Pianezza ML, Sellers EM, Tyndale RF: Nicotine metabolism defect reduces smoking. Nature. 1998, 393: 750.

    CAS  PubMed  Google Scholar 

  143. Caporaso NE, Lerman C, Audrain J, Boyd NR, Main D, Issaq HJ, Utermahlan B, Falk RT, Shields P: Nicotine metabolism and CYP2D6 phenotype in smokers. Cancer Epidemiolgy, Biomarkers & Prevention. 2001, 10: 261-263.

    CAS  Google Scholar 

  144. Miyamoto M, Umetsu Y, Dosaka-Akita H, Sawamura Y, Yokota J, Kunitoh H, Nemoto N, Sato K, Ariyoshi N, Kamataki T: CYP2A6 gene deletion reduces susceptibility to lung cancer. Biochemical and Biophysical Research Communications. 1999, 261: 658-660.

    CAS  PubMed  Google Scholar 

  145. Zhang W, Kilicarslan T, Tyndale RF, Sellers EM: Evaluation of methoxsalen, tranylcypromine, and tryptamine as specific and selective CYP2A6 inhibitors in vitro. Drug Metabolism and Disposition. 2001, 29: 897-902.

    CAS  PubMed  Google Scholar 

  146. Tyndale RF, Sellers EM: Genetic variation in CYP2A6-mediated nicotine metabolism alters smoking behavior. Therapeutic Drug Monitoring. 2002, 24: 163-171.

    CAS  PubMed  Google Scholar 

  147. Sellers EM, Kaplan HL, Tyndale RF: Inhibition of cytochrome P450 2A6 increases nicotine's oral bioavailability and decreases smoking. Clinical Pharmacology and Therapeutics. 2000, 68: 35-43.

    CAS  PubMed  Google Scholar 

  148. Straub RE, Sullivan PF, Ma Y, Myakishev MV, Harris-Kerr C, Wormley B, Kadambi B, Sadek H, Silverman MA, Webb BT, Neale MC, Bulik CM, Joyce PR, Kendler KS: Susceptibility genes for nicotine dependence: a genome scan and followup in an independent sample suggest that regions on chromosomes 2, 4, 10, 16, 17 and 18 merit further study. Molecular Psychiatry. 1999, 4: 129-144.

    CAS  PubMed  Google Scholar 

  149. Lander E, Kruglyak L: Genetic dissection of complex traits: guidelines for interpreting and reporting linkage results. Nature Genetics. 1995, 11: 241-247.

    CAS  PubMed  Google Scholar 

  150. Sullivan PF, Neale BM, Oord Van Den E, Miles MF, Neale MC, Bulik CM, Joyce PR, Straub RE, Kendler KS: Candidate genes for nicotine dependence via linkage, epistasis, and bioinformatics. American Journal of Medical Genetics. 2004, 126B: 23-36.

    PubMed  Google Scholar 

  151. Bergen AW, Korczak JF, Weissbecker KA, Goldstein AM: A genome-wide search for loci contributing to smoking and alcoholism. Genetic Epidemiology. 1999, 17 (Suppl 1): S55-60.

    PubMed  Google Scholar 

  152. Duggirala R, Almasy L, Blangero J: Smoking behavior is under the influence of a major quantitative trait locus on human chromosome 5q. Genetic Epidemiology. 1999, 17 (Suppl 1): S139-144.

    PubMed  Google Scholar 

  153. Comings DE, Gade R, Wu S, Chiu C, Dietz G, Muhleman D, Saucier G, Ferry L, Rosenthal RJ, Lesieur HR, Rugle LJ, MacMurray P: Studies of the potential role of the dopamine D1 receptor gene in addictive behaviors. Molecular Psychiatry. 1997, 2: 44-56.

    CAS  PubMed  Google Scholar 

  154. Risch N, Merikangas K: The future of genetic studies of complex human diseases. Science. 1996, 273: 1516-1517.

    CAS  PubMed  Google Scholar 

  155. Haines JL: Genomic screening. Approaches to Gene Mapping in Complex Human Diseases. Edited by: Haines JL, Pericak-Vance MA. 1998, New York: Wiley-Liss, Inc, 243-251.

    Google Scholar 

  156. Comings DE: Why different rules are required for polygenic inheritance: lessons from studies of the DRD2 gene. Alcohol. 1998, 16: 61-70.

    CAS  PubMed  Google Scholar 

  157. Buckland PR: Genetic association studies of alcoholism-problems with the candidate gene approach. Alcohol and Alcoholism. 2001, 36: 99-103.

    CAS  PubMed  Google Scholar 

  158. Wacholder S, Rothman N, Caporaso N: Population stratification in epidemiologic studies of common genetic variants and cancer: quantification of bias. Journal of the National Cancer Institute. 2000, 92: 1151-1158.

    CAS  PubMed  Google Scholar 

  159. Cardon LR, Palmer LJ: Population stratification and spurious allelic association. Lancet. 2003, 361: 598-604.

    PubMed  Google Scholar 

  160. Kozlowski LT: Rehabilitating a genetic perspective in the study of tobacco and alcohol use. British Journal of Addiction. 1991, 86: 517-520.

    CAS  PubMed  Google Scholar 

  161. Hutchison KE, McGeary J, Smolen A, Bryan A, Swift RM: The DRD4 VNTR polymorphism moderates craving after alcohol consumption. Health Psychology. 2002, 21: 139-146.

    PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by National Iinstitutes of Health grants K23-DE00420-04 and T32 DE07151.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Athanasios I. Zavras.

Additional information

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Maserejian, N.N., Zavras, A.I. Genetics of Tobacco Use. Tob. Induced Dis. 2, 81 (2004). https://doi.org/10.1186/1617-9625-2-2-81

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/1617-9625-2-2-81

Keywords