Did you arrive here by via search engine?
Click here to view the original version of this article

Click to Print This Page
(This section will not print)

New Frontiers in the Treatment and Management of Smoking Cessation

Course Authors

Eric Siu, M.Sc., Nael Al Koudsi, H.B.Sc., Man Ki Ho, H.B.Sc., Rachel F. Tyndale, M.Sc., Ph.D.

Dr. Tyndale is a Scientist at the Centre for Addiction and Mental Health and a Professor of Pharmacology, University of Toronto. Mr. Siu, Mr. Al Koudsi, and Miss Ho are doctoral students in the Department of Pharmacology at the University of Toronto.

Within the past 12 months, Dr. Tyndale has been the Chief Scientific Officer and a shareholder of Nicogen Inc. Mr. Siu, Mr. Al Koudsi and Miss Ho report no commercial conflicts of interest.

This activity is made possible by an unrestricted educational grant from Pfizer. Pfizer

Estimated course time: 3 hour(s).

Albert Einstein College of Medicine – Montefiore Medical Center designates this enduring material activity for a maximum of 1.0 AMA PRA Category 1 Credit(s)™. Physicians should claim only the credit commensurate with the extent of their participation in the activity.

In support of improving patient care, this activity has been planned and implemented by Albert Einstein College of Medicine-Montefiore Medical Center and InterMDnet. Albert Einstein College of Medicine – Montefiore Medical Center is jointly accredited by the Accreditation Council for Continuing Medical Education (ACCME), the Accreditation Council for Pharmacy Education (ACPE), and the American Nurses Credentialing Center (ANCC), to provide continuing education for the healthcare team.

 
Learning Objectives

Upon completion of this Cyberounds®, you should be able to:

  • Discuss the epidemiology of smoking among U.S. population

  • Discuss the neurophysiological basis of nicotine addiction, including the role of the reward pathway

  • Discuss the genetic contribution to smoking behaviours and how genetic variation in neurotransmitter pathways might contribute

  • Discuss pharmacotherapeutic treatment strategies that promote smoking cessation.

 

Dr. Tyndale and colleagues will discuss the unlabeled use of rimonabant and NicVAX for smoking cessation.

The gender gap in smoking prevalence has dramatically narrowed in recent years...

According to the World Health Organization (WHO), tobacco smoking is responsible for the death of approximately 5 million people each year.(1) If current trends continue, by the year 2030 it is estimated that 10 million people will die per year as a result of smoking.(2) In the United States alone, tobacco use accounts for approximately $157 billion in health care costs and annual loss of productivity.(3)

In spite of increasingly strict tobacco control policies and widespread knowledge of the numerous adverse health effects associated with cigarette smoking, approximately 21% of adults in the United States are currently smokers (Figure 1).(4),(5) One of the national health objectives is to lower smoking prevalence to under 12% in adults by the year 2010, and while there has been an overall decline in recent years it is not occurring at a rate that will allow this objective to be met.(6)

Large differences in smoking prevalence exist among American racial/ethnic groups, with American Indians/Alaska Natives having the highest rates at 33.4%, followed by non-Hispanic whites (22.2%) and non-Hispanic blacks (20.2%), with the lowest rates found among Hispanics (15.0%) and Asians (11.3%).(4) The gender gap in smoking prevalence has dramatically narrowed in recent years, with overall current rates of 23.4% in men and 18.5% in women (Figure 1).(2),(4),(5)

Figure 1. The Change in Smoking Prevalence in Men and Women From 1955 to 2005.*

Figure 1

*Cokkinides, V. et al. Progress and opportunities in tobacco control. CA Cancer J Clin 2006;56(3):135-142.

Neurobiology of Nicotine Dependence

While the majority (~80%) of smokers express a desire to quit smoking, less than 5% are able to maintain abstinence unaided as a consequence of the highly addictive properties of inhaled cigarette nicotine, the primary psychoactive substance in cigarettes.(7),(8) Cigarettes, unlike other nicotine delivery systems (e.g., patch or gum), are particularly addictive because smoked nicotine is rapidly absorbed and reaches the brain (site of action) within 10-20 seconds;(9) the rapid delivery intensifies addictiveness at least in part by enhancing the learning of the behavior.

In the brain, nicotine mediates its psychoactive properties, which include mood elevation, decreased anxiety, decreased appetite, increased arousal and cognitive enhancement, by stimulating nicotinic acetylcholine receptors (nAChRs).(10) Nicotinic AChRs are pentameric ligand, ion gated-channels containing combinations of α (α2 - α7) and β (β2 - β4) subunits that form either homomeric (e.g., α7 nAChR) or heteromeric (e.g. α4β2 nAChR) receptors that differ in their affinity for nicotine.(11),(12)

Of the many nAChR subunits, α4 and β2 subunits are the most widely expressed, high affinity subunits in the brain.(13) Animal studies indicate that the activation of the α4 subunit is required for nicotine-induced reward, tolerance, and sensitization,(14) whereas the β2-containing nAChRs are involved in mediating the reinforcing properties of nicotine but not the expression of nicotine withdrawal symptoms.(15),(16) These studies showed that distinct molecular and neural mechanisms mediate differing components of nicotine dependence.

The main mechanism by which nicotine is thought to establish and maintain dependence is by activating the mesocorticolimbic dopamine system, also referred to as the brain reward center (Figure 2).(11),(17) Nicotine activates nAChRs on dopaminergic neurons in the ventral tegmental area (VTA) that innervate the nucleus accumbens (NAc), resulting in sustained dopamine release.(17),(18) Other addictive drugs (e.g., cocaine) and natural rewards (e.g., sex) also promote dopamine release in the NAc. In addition, because of the predominantly post-synaptic localization of the nicotinic receptors, nicotine modulates the release of other neurotransmitters such as serotonin and γ-aminobutyric acid (GABA) which can have direct psychoactive properties (e.g,. serotonin) and indirect effects on dopamine release.(19)

...Distinct molecular and neural mechanisms mediate differing components of nicotine dependence.

Figure 2. The Mesocorticolimbic Dopamine System.

Figure 2

The Genetics of Nicotine Addiction

Cigarette smoking is a complex behavior that includes a number of stages such as initiation, experimentation, regular use, dependence, cessation and relapse.(20),(21) While environmental factors such as the influence of family members, peers and culture undoubtedly affect one's smoking behaviors at these stages, genetics also plays a substantial role. Twin studies have been used as a tool to determine the heritability of complex traits such as smoking. The degree of concordance of a given phenotype between genetically identical monozygotic twins compared to dizygotic twins, who share only approximately half of their genes, is an indication of a genetic contribution to that phenotype.(22)

From numerous twin studies, the heritability of "ever" smoking has been estimated in the range of 11-78%.(23),(24),(25),(26),(27),(28),(29),
(30),(31) This large variation may be due, at least in part, to the broadness of this phenotype, which encompasses those who have only tried cigarettes once to regular heavy smokers. Once smoking is initiated, the heritability for persistence to regular smoking has been estimated at 28-84%,(23),(26),(27),(28),(30),(31),(32),(34) for number of cigarettes smoked per day at 45-86%(33),(34),(35),(36),(37),(38),(39) and for dependence on nicotine at 31-75%.(30),(34),(40) A genetic influence on nicotine withdrawal symptoms and smoking cessation has also been identified with heritability estimated at 26-48% and 50-58% respectively.(33),(41),(42),(43) Taken together, these studies suggest a substantial genetic contribution to most aspects of smoking including the ability to quit smoking.

While heritability studies can demonstrate a genetic influence in smoking behaviors, advances in our understanding of the human genome have allowed for the localization of specific genes involved. Similar to other complex disorders, multiple interacting genetic factors are likely involved in the etiology of smoking behaviors, with each gene contributing a small effect.(34),(44),(45) Genome-wide linkage analyses are used to locate areas in the genome with high linkage disequilibrium to a specific phenotype such as cigarette consumption.

Some of the loci associated with smoking phenotypes are near genes of known biological relevance, including the α2-nicotinic acetylcholine receptor (CHRNA2) and α1A-adrenergic receptor (ADRA1A) on chromosome 8, the dopamine receptor (D1) on chromosome 5, the μ1-opioid receptor (OPRM1) on chromosome 6 and the serotonin receptor 5A on chromosome 7.(46) Genes present in loci identified from these genome-wide linkage studies, as well as those with biological relevance, can be further examined for their impact on smoking using, for example, case-control association studies.

Nicotine's addictive properties are a result of the activation of nAChRs in the brain and the consequent impact on a number of downstream neurotransmitter systems (e.g., dopamine, serotonin and GABA). Therefore, genes encoding receptors, transporters and enzymes involved in the synthesis or degradation of these neurotransmitters are plausible candidates for gene association studies with smoking phenotypes. Here we provide a few examples of gene association studies and briefly discuss some of the discrepancies in the literature. For an extensive review see (44),(47),(48).

...Studies suggest a substantial genetic contribution to most aspects of smoking including the ability to quit smoking.

Dopaminergic System

The dopaminergic system has been studied extensively because of its pivotal role in mediating nicotine dependence. Of the five different dopamine receptor subtypes (D1-D5), a single nucleotide polymorphism 10 kb downstream of the DRD2 gene (TaqI A1 allele),which encodes the dopamine D2 receptor, has received considerable attention. The TaqI A1 variant lies within a neighboring gene (ANKK1) and is linked to another variant in the D2 receptor (TaqI B1).(49),(50),(51) TaqI A1 allele has been associated with reduced dopamine D1 receptor expression in the striatum,(52),(53),(54) an increased risk for current smoking, an earlier age of smoking onset and shorter abstinent periods from smoking. (55),(56) However, other studies have shown conflicting results.(57),(58),(59)

Interestingly, the TaqI A1 allele has also been shown to influence abstinence rates on nicotine replacement therapy (NRT) and bupropion(60),(61),(62) Johnstone et al. (2004) found that transdermal nicotine patch is significantly more effective than placebo for smokers with at least one TaqI A1 allele.(61) A follow-up of this study supported the association between TaqI A1 and cessation at 6 and 12 months of follow up but only among women.(62) In a recent bupropion cessation study, Swan et al. (2005) reported a trend for lower cessation rates (at 12-month follow up) among women with the TaqI A1 allele.(60) Thus, it seems that women with the TaqI A1 allele might benefit more from NRT compared to bupropion.

Another interesting genetic variant (-141C Ins/Del) in DRD2 has also been associated with different cessation rates for NRT and bupropion. Individuals homozygous for the -141C allele had higher abstinence rates on bupropion (35%) compared to placebo (19%), while those with the (-141C Ins/Del allele had higher abstinence rates on nicotine nasal spray (42%) or patch (50%) compared to placebo (33%) or bupropion (20%).(63) This is an example of:

  1. how response to specific treatments may be differentially altered by genetic variation in the same gene and
  2. how genetics could potentially be used to improve cessation by individualizing treatment based on genotype (e.g., improving efficacy by treating individuals with the -141C Ins allele with bupropion and those with the -141C Del allele with NRT).

Other important players in the dopaminergic system include the dopamine transporter and enzymes involved in either the synthesis (e.g., tyrosine hydroxylase) or metabolism/degradation (e.g., dopamine β-hydroxylase, monoamine oxidase and catechol-O-methyltransferase) of dopamine. A number of studies have found significant associations between several smoking phenotypes (e.g., craving or dependence) and the genes encoding the dopamine transporter(64),(65),(66) and neurotransmitter enzymes.(67),(68),(69),(70),(71)

Nicotinic Receptors

Nicotine's addictive properties are a result of the activation of nAChRs in the brain...

Li et al. (2005) recently reported an association between genetic variations in CHRNA4, the gene encoding the α4 subunit of the nAChR, with the following phenotypes: smoking quantity, heaviness of smoking index or Fagerstrom test for nicotine dependence.(72) However, there are no reported significant associations between genetic variations in CHRNB2, the gene encoding the β2 subunit of nicotinic receptor, and nicotine dependence.(72),(73),(74),(75)

Other Candidate Genes

The serotonin and the opioid systems have both been shown to play an important role in nicotine dependence.(76),(77) Thus, it is not surprising that associations have been reported between genetic variants in genes encoding serotonin or opioid receptors with different smoking phenotypes and cessation rates.(78),(79) For example, a polymorphism (102T>C) in the serotonin receptor 2A (5-HT2A) gene has been associated with lower 5-HT2A expression in the brain(80) and current smoking in a Brazilian population.(79)

Limitations and Significance of Genetic Association Studies

Lack of replication in genetic association studies is still common for a number of reasons such as:

  1. Racial admixture: the inclusion of individuals of different ethnicities can influence the results since different ethnic groups have unique genetic variants and differing frequencies of variants.
  2. Inconsistency among studies in genotype assessment and phenotypes (e.g., smokers versus nicotine dependence) being measured.

Nonetheless, gene-association studies provide insight into the neurobiological mechanisms of nicotine dependence and provide the rationale for future use of genetics to identify individuals who might be at a greater risk for smoking or smoking related illnesses. In addition, pharmacogenetics holds the promise of improving the efficacy of smoking cessation treatments by individualizing the type, dosage and duration of treatment based on an individual's genotype/phenotype.

Established Drug Treatments

Nicotine Replacement Therapies

Nicotine replacement products (NRTs) are one of the first-line treatments for nicotine dependence approved by the US Federal Drug Administration (FDA). Currently, there are several nicotine delivery devices available on the market - these include gum, transdermal patch, vapor inhaler, nasal spray, lozenge and sublingual tablet. NRT treatments facilitate cessation by delivering nicotine without the exposure to carcinogenic compounds found in cigarette smoke. The primary mechanisms of action of NRTs is to maintain the desirable emotional states (e.g., mood, attention) achieved from smoking and reduce withdrawal symptoms associated with abstinence by providing steady state amounts of nicotine.(81),(82) The use of nicotine replacement products has led to varying degrees of success in long-term smoking cessation.(83),(84) Overall, NRTs increase the odds for abstinence by 1.5- to 2-fold compared to placebo at 6 months after target quit date.(84) In addition, various types of NRTs appear to be equally effective and behavioral counseling does not substantially improve cessation rates beyond the NRT effect.(84) More extensive discussions on each of the NRTs can be found elsewhere.(85),(86),(87)

Bupropion (Zyban®)

Both smoking and depression are highly associated in that they are influenced by dopamine levels.(88),(89) The first non-nicotine product approved as a first-line therapy for smoking cessation is the atypical anti-depressant bupropion. The main mechanism of action of bupropion in smoking cessation appears to be related to craving reduction and alleviation of certain symptoms associated with withdrawal in abstinent smokers; however, other less understood mechanisms may also be involved.(90),(91),(92),(93)

The serotonin and the opioid systems have both been shown to play an important role in nicotine dependence.

Bupropion reduces negative withdrawal symptoms which may be mediated by binding to striatal dopamine transporters and preventing the reuptake of dopamine.(94),(95) Bupropion can also potentially attenuate the rewarding effects of nicotine by acting as an indirect functional antagonist of the nAChRs.(96),(97),(98) Though the precise mechanism of action is unclear, bupropion can also act on the noradrenergic system where it reduces the firing rates of noradrenergic neurons in the locus coeruleus,(99) which contains neuronal projections to the hippocampus, a region implicated in drug-dependence.(100) Finally, bupropion metabolites, in particular hydroxybupropion, have been demonstrated to be be pharmacologically active, although the individual effects of these metabolites on nicotine dependence have not been clarified.(101),(102),(103)

Overall, most studies have demonstrated bupropion to be effective at improving smoking cessation(91),(104),(105),(106),(107),(108),(109),
(110),(111)(112) while some studies showed no significant improvement compared to placebo;(113),(114) (113, 114)[see review (108)]. With respect to safety, a large amount of data have been gathered over the past few years concerning the use of bupropion for smoking cessation in a community setting. The drug caused few serious side effects with the primary reasons for discontinuation being insomnia, nausea/vomiting and dizziness.(115)

Drugs in Development

Varenicline (ChantixTM)

Varenicline has recently been approved by the FDA as a new smoking cessation drug. It is a partial agonist of the α4β2 nAChR; its structure is based on the alkaloid extract cytisine from the plant Cytisus laburnum L.(116) Varenicline is hypothesized to reduce nicotine reward during smoking and to stimulate dopamine release to reduce craving and withdrawal during abstinence.(116),(117) Two separate Phase III 12-week treatment studies found that varenicline more than doubled the continuous abstinence rates up to one year compared to placebo.(118),(119) Varenicline was also found to be significantly more efficacious in maintaining continuous abstinence up to one year compared to bupropion.(118),(119) The most common adverse event related to varenicline was nausea.(118),(119)

Nicotine Vaccines

Nicotine vaccines are a new approach being investigated for smoking cessation. The principle of this strategy is to stimulate antibody production which will prevent nicotine from entering the brain.(120),(121) An advantage of the nicotine vaccines is that daily drug administration is not required. A peer-reviewed safety study found that the 30-day continuous abstinence rates were 4-fold higher in patients receiving the highest vaccine (NicVAX) dose.(122) In a phase II study, Cytos Biotechnology reported that high responders (high antibody production) were twice as likely to be continuously abstinent from smoking compared to placebo at 6 and 12 months. Careful interpretation of these results is needed as sample sizes were small in both studies. Most of the reported side effects were mild to moderate and were resolved without medical treatment.(122),(123)

Rimonabant (SR141716)

Rimonabant is a cannabinoid receptor antagonist.(124) In animal studies it was found that rimonabant prevented intravenous nicotine self-administration and blocked the release of dopamine in response to nicotine.(124) In a phase III study (STRATUS-US), it was found that patients receiving the highest dose of rimonabant had the greatest abstinence rate compared to those who were on a lower dose or placebo.(125) In contrast, another study (STRATUS-Europe) found no significant increase relative to placebo in cessation for either dosing group.(126) In a third study (STRATUS-Worldwide), significantly more patients who received the highest dose remained smoke-free for the entire year compared to placebo.(126) The most common side effects were nausea and upper respiratory tract infection.(125) Rimonabant has not been approved for use as a smoking cessation product in the USA and the moderate benefits obtained from this drug may limit its widespread use.

Metabolism Modulators

Another novel approach to smoking cessation is through inhibition of nicotine metabolism. In humans the majority of nicotine is metabolically inactivated to cotinine by the CYP2A6 enzyme.(126),(127) Alterations in the amount or function (e.g., due to genetic variations or functional inhibition) of this enzyme significantly affect nicotine metabolism.(128),
(129)(130),(131) As smokers titrate their smoking to maintain nicotine levels(8),(132),(133) variation in the activity of the enzyme can alter smoking behaviors;(134),(135),(136),(137),(138),(139),(140) inhibiting CYP2A6 might represent a novel approach to smoking cessation and smoking reduction products.

Varenicline has recently been approved by the FDA as a new smoking cessation drug.

Summary

Nicotine is the primary psychoactive substance responsible for establishing and maintaining tobacco dependence. Like other drugs of abuse (e.g. cocaine), nicotine acts via the mesolimbic dopaminergic pathway stimulating feelings of reward. Genetic linkage studies have identified a host of genes (e.g. nicotinic receptor genes) that may contribute to the variable risk for tobacco dependence. The first-line pharmacotherapies for tobacco dependence include nicotine replacement therapies, bupropion (Zyban®) and the newly approved varenicline. In addition, genetic studies may provide additional therapeutic targets for smoking cessation and opportunities in the future for optimization of smoking cessation therapy based on genetics.


Footnotes

1Why is tobacco a public health priority? 2006. In World Health Organization: Tobacco Free Initiative.
2The World Bank. 1999. Curbing the Epidemic:Governments and the Economics of Tobacco Control. Washington DC: The World Bank.
3Centers for Disease Control and Prevention. 2002. Annual Smoking-Attributable Mortality, Years of Potential Life Lost, and Economic Costs --- United States, 1995--1999. Morbidity and Mortality Weekly Report 51: 300-3.
4Centers for Disease Control and Prevention. 2005. Cigarette Smoking Among Adults: United States, 2004. In Morbidity and Mortality Weekly Report, pp. 1121-48.
5Cokkinides V, Bandi P, Ward E, Jemal A, Thun M. 2006. Progress and Opportunities in Tobacco Control. CA Cancer J Clin 56: 135-42.
6U.S. Department of Health and Human Services. 2000. Healthy People 2010, 2nd ed. With Understanding and Improving Health and Objectives for Improving Health (2 vols.). ed. DC Washington.
7Diagnostic and Statistical Manual - Text Revision IV. 2000. American Psychiatric Associations.
8Baillie AJ, Mattick RP, Hall W. 1995. Quitting smoking: estimation by meta-analysis of the rate of unaided smoking cessation. Aust J Public Health. 19: 129-31.
9Benowitz NL. 1988. Drug therapy. Pharmacologic aspects of cigarette smoking and nicotine addition. N Engl J Med 319: 1318-30.
10Picciotto MR. 2003. Nicotine as a modulator of behavior: beyond the inverted U. Trends Pharmacol Sci 24: 493-9.
11Laviolette SR, van der Kooy D. 2004. The neurobiology of nicotine addiction: bridging the gap from molecules to behaviour. Nat Rev Neurosci 5: 55-65.
12Wonnacott S, Sidhpura N, Balfour DJ. 2005. Nicotine: from molecular mechanisms to behaviour. Curr Opin Pharmacol 5: 53-9.
13Mathieu-Kia AM, Kellogg SH, Butelman ER, Kreek MJ. 2002. Nicotine addiction: insights from recent animal studies. Psychopharmacology (Berl) 162: 102-18.
14Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, et al. 2004. Nicotine activation of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science 306: 1029-32.
15Picciotto MR, Zoli M, Rimondini R, Lena C, Marubio LM, et al. 1998. Acetylcholine receptors containing the beta2 subunit are involved in the reinforcing properties of nicotine. Nature 391: 173-7.
16Besson M, David V, Suarez S, Cormier A, Cazala P, et al. 2006. Genetic dissociation of two behaviors associated with nicotine addiction: Beta-2 containing nicotinic receptors are involved in nicotine reinforcement but not in withdrawal syndrome. Psychopharmacology (Berl) 187: 189-99.
17Balfour DJ. 2004. The neurobiology of tobacco dependence: a preclinical perspective on the role of the dopamine projections to the nucleus accumbens [corrected]. Nicotine Tob Res 6: 899-912.
18Pidoplichko VI, Noguchi J, Areola OO, Liang Y, Peterson J, et al. 2004. Nicotinic cholinergic synaptic mechanisms in the ventral tegmental area contribute to nicotine addiction. Learn Mem 11: 60-9.
19Seth P, Cheeta S, Tucci S, File SE. 2002. Nicotinic--serotonergic interactions in brain and behaviour. Pharmacol Biochem Behav 71: 795-805.
20Malaiyandi V, Sellers EM, Tyndale RF. 2005. Implications of CYP2A6 genetic variation for smoking behaviors and nicotine dependence. Clinical Pharmacology & Therapeutics 77: 145.
21Mayhew KP, Flay BR, Mott JA. 2000. Stages in the development of adolescent smoking. Drug and Alcohol Dependence 59: 61.
22Winerman L. 2004. A second look at twin studies. Monitor on Psychology 35: 46.
23True WR, Heath AC, Scherrer JF, Waterman B, Goldberg J, et al. 1997. Genetic and environmental contributions to smoking. Addiction 92: 1277-88.
24Heath AC, Martin NG. 2003. Genetic models for the natural history of smoking: evidence for a genetic influence on smoking persistence. Addictive Behaviors 18: 19-34.
25Kendler KS, Neale MC, Sullivan P, Corey LA, Gardner CO, Prescott CA. 1999. A population-based twin study in women of smoking initiation and nicotine dependence. Psychol Med 29: 299-308.
26Heath AC, Cates, R., Martin, N. G., Meyer, J., Hewitt, J. K., Neale, M. C. & Eaves, L. J. 1993. Genetic contribution to risk of smoking initiation: comparisons across birth cohorts and across cultures. Journal of Substance Abuse 5: 221-46.
27Madden PAF, Heath AC, Pedersen NL, Kaprio J, Koskenvuo MJ, Martin NG. 1999. The Genetics of Smoking Persistence in Men and Women: A Multicultural Study. Behavior Genetics 29: 423
28Maes HH, Sullivan PF, Bulik CM, Neale MC, Prescott CA, et al. 2004. A twin study of genetic and environmental influences on tobacco initiation, regular tobacco use and nicotine dependence. Psychol Med 34: 1251-61
29Sullivan PF, Kendler KS. 1999. The genetic epidemiology of smoking. Nicotine & Tobacco Research 1.
30Vink JM, Willemsen G, Boomsma DI. 2005. Heritability of Smoking Initiation and Nicotine Dependence. Behavior Genetics 35: 397.
31Li MD, Cheng R, Ma JZ, Swan GE. 2003. A meta-analysis of estimated genetic and environmental effects on smoking behavior in male and female adult twins. Addiction 98: 23-31.
32Hamilton AS, Lessov-Schlaggar CN, Cockburn MG, Unger JB, Cozen W, Mack TM. 2006. Gender Differences in Determinants of Smoking Initiation and Persistence in California Twins. Cancer Epidemiol Biomarkers Prev 15: 1189-97.
33Pergadia ML, Heath AC, Martin NG, PA. M. 2006. Genetic analyses of DSM-IV nicotine withdrawal in adult twins. Psychol Med 36: 963-72.
34Lessov CN, Swan GE, Ring HZ, Khroyan TV, Lerman C. 2004. Genetics and Drug Use as a Complex Phenotype. Substance Use & Misuse 39: 1515.
35Koopmans JR, Slutske WS, Heath AC, Neale MC, Boomsma DI. 1999. The genetics of smoking initiation and quantity smoked in Dutch adolescent and young adult twins. Behav Genet 29: 383-93.
36Swan GE, Carmelli D, Rosenman RH, Fabsitz RR, Christian JC. 1990. Smoking and alcohol consumption in adult male twins: genetic heritability and shared environmental influences. J Subst Abuse 2: 39-50.
37Swan GE, Carmelli, D., Cardon, L. R. 1996. The consumption of tobacco, alcohol, and coffee in Caucasian male twins: a multivariate genetic analysis. J Subst Abuse 8: 19-31.
38Swan GE, Carmelli D, Cardon LR. 1997. Heavy consumption of cigarettes, alcohol and coffee in male twins. J Stud Alcohol 58: 182-90.
39Vink JM, Beem AL, Posthuma D, Neale MC, Willemsen G, Kendler KS SP, Boomsma DI. 2004. Linkage analysis of smoking initiation and quantity in Dutch sibling pairs. Pharmacogenomics J 4: 274-82.
40True WR, Xian H, Scherrer JF, Madden PAF, Bucholz KK, et al. 1999. Common Genetic Vulnerability for Nicotine and Alcohol Dependence in Men. Arch Gen Psychiatry 56: 655-61.
41Broms U, Silventoinen K, Madden PA, Heath AC, Kaprio J. 2006. Genetic architecture of smoking behavior: a study of Finnish adult twins. Twins Res Hum Genet 9: 64-72.
42Xian H, Scherrer J, Madden P, Lyons M, Tsuang M, et al. 2003. The heritability of failed smoking cessation and nicotine withdrawal in twins who smoked and attempted to quit. Nicotine & Tobacco Research 5: 245.
43Lessov CN, Martin NG, Statham DJ, Todorov AA, Slutske WS, et al. 2004. Defining nicotine dependence for genetic research: evidence from Australian twins. Psychol Med 34: 865-79.
44Tyndale RF. 2003. Genetics of alcohol and tobacco use in humans. Ann Med 35: 94-121.
45Hall W, Madden P, Lynskey M. 2002. The genetics of tobacco use: methods, findings and policy implications. Tob Control 11: 119-24.
46Swan GE, Hops H, Wilhelmsen KC, Lessov-Schlaggar CN, Cheng LS, et al. 2006. A genome-wide screen for nicotine dependence susceptibility loci. American Journal of Medical Genetics Part B: Neuropsychiatric Genetics 141B: 354-60.
47Ho MK, Tyndale RF. In press. Overview of the Pharmacogenomics of Cigarette Smoking. Pharmacogenomics J.
48Al Koudsi N, Tyndale RF. 2005. Genetic influences on smoking: a brief review. Ther Drug Monit 27: 704-9.
49Neville MJ, Johnstone EC, Walton RT. 2004. Identification and characterization of ANKK1: a novel kinase gene closely linked to DRD2 on chromosome band 11q23.1. Hum Mutat 23: 540-5.
50Hauge XY, Grandy DK, Eubanks JH, Evans GA, Civelli O, Litt M. 1991. Detection and characterization of additional DNA polymorphisms in the dopamine D2 receptor gene. Genomics 10: 527-30.
51Kidd KK, Morar B, Castiglione CM, Zhao H, Pakstis AJ, et al. 1998. A global survey of haplotype frequencies and linkage disequilibrium at the DRD2 locus. Hum Genet 103: 211-27.
52Jonsson EG, Nothen MM, Grunhage F, Farde L, Nakashima Y, et al. 1999. Polymorphisms in the dopamine D2 receptor gene and their relationships to striatal dopamine receptor density of healthy volunteers. Mol Psychiatry 4: 290-6.
53Pohjalainen T, Rinne JO, Nagren K, Lehikoinen P, Anttila K, et al. 1998. The A1 allele of the human D2 dopamine receptor gene predicts low D2 receptor availability in healthy volunteers. Mol Psychiatry 3: 256-60.
54Thompson J, Thomas N, Singleton A, Piggott M, Lloyd S, et al. 1997. D2 dopamine receptor gene (DRD2) Taq1 A polymorphism: reduced dopamine D2 receptor binding in the human striatum associated with the A1 allele. Pharmacogenetics 7: 479-84.
55Noble EP, St Jeor ST, Ritchie T, Syndulko K, St Jeor SC, et al. 1994. D2 dopamine receptor gene and cigarette smoking: a reward gene? Med Hypotheses 42: 257-60.
56Comings DE, Ferry L, Bradshaw-Robinson S, Burchette R, Chiu C, Muhleman D. 1996. The dopamine D2 receptor (DRD2) gene: a genetic risk factor in smoking. Pharmacogenetics 6: 73-9.
57Bierut LJ, Rice JP, Edenberg HJ, Goate A, Foroud T, et al. 2000. Family-based study of the association of the dopamine D2 receptor gene (DRD2) with habitual smoking. Am J Med Genet 90: 299-302.
58Spitz MR, Shi H, Yang F, Hudmon KS, Jiang H, et al. 1998. Case-control study of the D2 dopamine receptor gene and smoking status in lung cancer patients. J Natl Cancer Inst 90: 358-63.
59Johnstone EC, Yudkin P, Griffiths SE, Fuller A, Murphy M, Walton R. 2004. The dopamine D2 receptor C32806T polymorphism (DRD2 Taq1A RFLP) exhibits no association with smoking behaviour in a healthy UK population. Addict Biol 9: 221-6.
60Swan GE, Valdes AM, Ring HZ, Khroyan TV, Jack LM, et al. 2005. Dopamine receptor DRD2 genotype and smoking cessation outcome following treatment with bupropion SR. Pharmacogenomics J 5: 21-9.
61Johnstone EC, Yudkin PL, Hey K, Roberts SJ, Welch SJ, et al. 2004. Genetic variation in dopaminergic pathways and short-term effectiveness of the nicotine patch. Pharmacogenetics 14: 83-90.
62Yudkin P, Munafo M, Hey K, Roberts S, Welch S, et al. 2004. Effectiveness of nicotine patches in relation to genotype in women versus men: randomised controlled trial. Bmj 328: 989-90.
63Lerman C, Jepson C, Wileyto EP, Epstein LH, Rukstalis M, et al. 2006. Role of functional genetic variation in the dopamine D2 receptor (DRD2) in response to bupropion and nicotine replacement therapy for tobacco dependence: results of two randomized clinical trials. Neuropsychopharmacology 31: 231-42.
64Erblich J, Lerman C, Self DW, Diaz GA, Bovbjerg DH. 2005. Effects of dopamine D2 receptor (DRD2) and transporter (SLC6A3) polymorphisms on smoking cue-induced cigarette craving among African-American smokers. Mol Psychiatry 10: 407-14.
65Lerman C, Caporaso NE, Audrain J, Main D, Bowman ED, et al. 1999. Evidence suggesting the role of specific genetic factors in cigarette smoking. Health Psychol 18: 14-20.
66Sabol SZ, Nelson ML, Fisher C, Gunzerath L, Brody CL, et al. 1999. A genetic association for cigarette smoking behavior. Health Psychol 18: 7-13.
67Olsson C, Anney R, Forrest S, Patton G, Coffey C, et al. 2004. Association between dependent smoking and a polymorphism in the tyrosine hydroxylase gene in a prospective population-based study of adolescent health. Behav Genet 34: 85-91.
68Anney RJ, Olsson CA, Lotfi-Miri M, Patton GC, Williamson R. 2004. Nicotine dependence in a prospective population-based study of adolescents: the protective role of a functional tyrosine hydroxylase polymorphism. Pharmacogenetics 14: 73-81.
69McKinney EF, Walton RT, Yudkin P, Fuller A, Haldar NA, et al. 2000. Association between polymorphisms in dopamine metabolic enzymes and tobacco consumption in smokers. Pharmacogenetics 10: 483-91.
70Ito H, Hamajima N, Matsuo K, Okuma K, Sato S, et al. 2003. Monoamine oxidase polymorphisms and smoking behaviour in Japanese. Pharmacogenetics 13: 73-9.
71Colilla S, Lerman C, Shields PG, Jepson C, Rukstalis M, et al. 2005. Association of catechol-O-methyltransferase with smoking cessation in two independent studies of women. Pharmacogenet Genomics 15: 393-8.
72Li MD, Beuten J, Ma JZ, Payne TJ, Lou XY, et al. 2005. Ethnic- and gender-specific association of the nicotinic acetylcholine receptor alpha4 subunit gene (CHRNA4) with nicotine dependence. Hum Mol Genet 14: 1211-9.
73Lueders KK, Hu S, McHugh L, Myakishev MV, Sirota LA, Hamer DH. 2002. Genetic and functional analysis of single nucleotide polymorphisms in the beta2-neuronal nicotinic acetylcholine receptor gene (CHRNB2). Nicotine Tob Res 4: 115-25.
74Silverman MA, Neale MC, Sullivan PF, Harris-Kerr C, Wormley B, et al. 2000. 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. Am J Med Genet 96: 646-53.
75Feng Y, Niu T, Xing H, Xu X, Chen C, et al. 2004. A common haplotype of the nicotine acetylcholine receptor alpha 4 subunit gene is associated with vulnerability to nicotine addiction in men. Am J Hum Genet 75: 112-21.
76Walters CL, Cleck JN, Kuo YC, Blendy JA. 2005. Mu-opioid receptor and CREB activation are required for nicotine reward. Neuron 46: 933-43.
77Ji SP, Zhang Y, Van Cleemput J, Jiang W, Liao M, et al. 2006. Disruption of PTEN coupling with 5-HT2C receptors suppresses behavioral responses induced by drugs of abuse. Nat Med 12: 324-9.
78Lerman C, Wileyto EP, Patterson F, Rukstalis M, Audrain-McGovern J, et al. 2004. The functional mu opioid receptor (OPRM1) Asn40Asp variant predicts short-term response to nicotine replacement therapy in a clinical trial. Pharmacogenomics J 4: 184-92.
79do Prado-Lima PA, Chatkin JM, Taufer M, Oliveira G, Silveira E, et al. 2004. Polymorphism of 5HT2A serotonin receptor gene is implicated in smoking addiction. Am J Med Genet B Neuropsychiatr Genet 128: 90-3.
80Polesskaya OO, Sokolov BP. 2002. Differential expression of the \"C\" and \"T\" alleles of the 5-HT2A receptor gene in the temporal cortex of normal individuals and schizophrenics. J Neurosci Res 67: 812-22.
81Benowitz NL. 1993. Nicotine replacement therapy. What has been accomplished--can we do better? Drugs 45: 157-70.
82Schuh LM, Schuh KJ, Henningfield JE. 1996. Pharmacologic Determinants of Tobacco Dependence. Am J Ther 3: 335-41.
83Silagy C, Lancaster T, Stead L, Mant D, Fowler G. 2004. Nicotine replacement therapy for smoking cessation. Cochrane Database Syst Rev: CD000146.
84Stead L, Lancaster T. 2005. Nicotine replacement therapy for smoking cessation: Cochrane systematic review. Int J Epidemiol 34: 1001-2; discussion 2, 3.
85Frishman WH, Mitta W, Kupersmith A, Ky T. 2006. Nicotine and non-nicotine smoking cessation pharmacotherapies. Cardiol Rev 14: 57-73.
86Henningfield JE, Fant RV, Buchhalter AR, Stitzer ML. 2005. Pharmacotherapy for nicotine dependence. CA Cancer J Clin 55: 281-99; quiz 322-3, 5.
87Mitrouska I, Bouloukaki I, Siafakas NM. 2006. Pharmacological approaches to smoking cessation. Pulm Pharmacol Ther.
88Paperwalla KN, Levin TT, Weiner J, Saravay SM. 2004. Smoking and depression. Med Clin North Am 88: 1483-94, x-xi.
89Quattrocki E, Baird A, Yurgelun-Todd D. 2000. Biological aspects of the link between smoking and depression. Harv Rev Psychiatry 8: 99-110.
90Durcan MJ, Deener G, White J, Johnston JA, Gonzales D, et al. 2002. The effect of bupropion sustained-release on cigarette craving after smoking cessation. Clin Ther 24: 540-51.
91Jorenby DE, Leischow SJ, Nides MA, Rennard SI, Johnston JA, et al. 1999. A controlled trial of sustained-release bupropion, a nicotine patch, or both for smoking cessation. N Engl J Med 340: 685-91.
92Shiffman S, Johnston JA, Khayrallah M, Elash CA, Gwaltney CJ, et al. 2000. The effect of bupropion on nicotine craving and withdrawal. Psychopharmacology (Berl) 148: 33-40.
93Tashkin D, Kanner R, Bailey W, Buist S, Anderson P, et al. 2001. Smoking cessation in patients with chronic obstructive pulmonary disease: a double-blind, placebo-controlled, randomised trial. Lancet 357: 1571-5.
94Argyelan M, Szabo Z, Kanyo B, Tanacs A, Kovacs Z, et al. 2005. Dopamine transporter availability in medication free and in bupropion treated depression: a 99mTc-TRODAT-1 SPECT study. J Affect Disord 89: 115-23.
95Learned-Coughlin SM, Bergstrom M, Savitcheva I, Ascher J, Schmith VD, Langstrom B. 2003. In vivo activity of bupropion at the human dopamine transporter as measured by positron emission tomography. Biol Psychiatry 54: 800-5.
96Alkondon M, Albuquerque EX. 2005. Nicotinic receptor subtypes in rat hippocampal slices are differentially sensitive to desensitization and early in vivo functional up-regulation by nicotine and to block by bupropion. J Pharmacol Exp Ther 313: 740-50.
97Fryer JD, Lukas RJ. 1999. Noncompetitive functional inhibition at diverse, human nicotinic acetylcholine receptor subtypes by bupropion, phencyclidine, and ibogaine. J Pharmacol Exp Ther 288: 88-92.
98Slemmer JE, Martin BR, Damaj MI. 2000. Bupropion is a nicotinic antagonist. J Pharmacol Exp Ther 295: 321-7.
99Cooper BR, Wang CM, Cox RF, Norton R, Shea V, Ferris RM. 1994. Evidence that the acute behavioral and electrophysiological effects of bupropion (Wellbutrin) are mediated by a noradrenergic mechanism. Neuropsychopharmacology 11: 133-41.
100Vorel SR, Liu X, Hayes RJ, Spector JA, Gardner EL. 2001. Relapse to cocaine-seeking after hippocampal theta burst stimulation. Science 292: 1175-8.
101Bondarev ML, Bondareva TS, Young R, Glennon RA. 2003. Behavioral and biochemical investigations of bupropion metabolites. Eur J Pharmacol 474: 85-93.
102Damaj MI, Carroll FI, Eaton JB, Navarro HA, Blough BE, et al. 2004. Enantioselective effects of hydroxy metabolites of bupropion on behavior and on function of monoamine transporters and nicotinic receptors. Mol Pharmacol 66: 675-82.
103Tyndale RF, Benowitz NL, Sellers E, Swan GE, Wong SSY. 2006. CYP2A6 gene amplification enhances nicotine metabolism increasing smoking. Submitted.
104Ahluwalia JS, Harris KJ, Catley D, Okuyemi KS, Mayo MS. 2002. Sustained-release bupropion for smoking cessation in African Americans: a randomized controlled trial. Jama 288: 468-74.
105Aubin HJ, Lebargy F, Berlin I, Bidaut-Mazel C, Chemali-Hudry J, Lagrue G. 2004. Efficacy of bupropion and predictors of successful outcome in a sample of French smokers: a randomized placebo-controlled trial. Addiction 99: 1206-18.
106Gold PB, Rubey RN, Harvey RT. 2002. Naturalistic, self-assignment comparative trial of bupropion SR, a nicotine patch, or both for smoking cessation treatment in primary care. Am J Addict 11: 315-31.
107Hays JT, Hurt RD, Rigotti NA, Niaura R, Gonzales D, et al. 2001. Sustained-release bupropion for pharmacologic relapse prevention after smoking cessation. a randomized, controlled trial. Ann Intern Med 135: 423-33.
108Hughes J, Stead L, Lancaster T. 2004. Antidepressants for smoking cessation. Cochrane Database Syst Rev: CD000031.
109Hurt RD, Sachs DP, Glover ED, Offord KP, Johnston JA, et al. 1997. A comparison of sustained-release bupropion and placebo for smoking cessation. N Engl J Med 337: 1195-202.
110Jamerson BD, Nides M, Jorenby DE, Donahue R, Garrett P, et al. 2001. Late-term smoking cessation despite initial failure: an evaluation of bupropion sustained release, nicotine patch, combination therapy, and placebo. Clin Ther 23: 744-52.
111Swan GE, McAfee T, Curry SJ, Jack LM, Javitz H, et al. 2003. Effectiveness of bupropion sustained release for smoking cessation in a health care setting: a randomized trial. Arch Intern Med 163: 2337-44.
112Tonnesen P, Tonstad S, Hjalmarson A, Lebargy F, Van Spiegel PI, et al. 2003. A multicentre, randomized, double-blind, placebo-controlled, 1-year study of bupropion SR for smoking cessation. J Intern Med 254: 184-92.
113Zellweger JP, Boelcskei PL, Carrozzi L, Sepper R, Sweet R, Hider AZ. 2005. Bupropion SR vs placebo for smoking cessation in health care professionals. Am J Health Behav 29: 240-9.
114Tashkin D, Kanner RE, Bailey WC, Buist AS, Anderson PJ, et al. 1997. A multicenter evaluation of the effects of bupropion hydroxhloride sustained release tablets (BUP SR) versus placebo in a population of smokers with chronic obstructive pulmonary disease (COPD). Presented at Proceedings of the 7 Annual Meeting of the Society for Research on Nicotine and Tobacco, Seattle, Washington.
115Boshier A, Wilton LV, Shakir SA. 2003. Evaluation of the safety of bupropion (Zyban) for smoking cessation from experience gained in general practice use in England in 2000. Eur J Clin Pharmacol 59: 767-73.
116Coe JW, Brooks PR, Vetelino MG, Wirtz MC, Arnold EP, et al. 2005. Varenicline: an alpha4beta2 nicotinic receptor partial agonist for smoking cessation. J Med Chem 48: 3474-7.
117Papke RL, Heinemann SF. 1994. Partial agonist properties of cytisine on neuronal nicotinic receptors containing the beta 2 subunit. Mol Pharmacol 45: 142-9.
118Gonzales D, Rennard SI, Nides M, Oncken C, Azoulay S, et al. 2006. Varenicline, an alpha4beta2 nicotinic acetylcholine receptor partial agonist, vs sustained-release bupropion and placebo for smoking cessation: a randomized controlled trial. Jama 296: 47-55.
119Jorenby DE, Hays JT, Rigotti NA, Azoulay S, Watsky EJ, et al. 2006. Efficacy of varenicline, an alpha4beta2 nicotinic acetylcholine receptor partial agonist, vs placebo or sustained-release bupropion for smoking cessation: a randomized controlled trial. Jama 296: 56-63.
120Hieda Y, Keyler DE, Ennifar S, Fattom A, Pentel PR. 2000. Vaccination against nicotine during continued nicotine administration in rats: immunogenicity of the vaccine and effects on nicotine distribution to brain. Int J Immunopharmacol 22: 809-19.
121Pentel PR, Malin DH, Ennifar S, Hieda Y, Keyler DE, et al. 2000. A nicotine conjugate vaccine reduces nicotine distribution to brain and attenuates its behavioral and cardiovascular effects in rats. Pharmacol Biochem Behav 65: 191-8.
122Hatsukami DK, Rennard S, Jorenby D, Fiore M, Koopmeiners J, et al. 2005. Safety and immunogenicity of a nicotine conjugate vaccine in current smokers. Clin Pharmacol Ther 78: 456-67.
123Maurer P, Jennings GT, Willers J, Rohner F, Lindman Y, et al. 2005. A therapeutic vaccine for nicotine dependence: preclinical efficacy, and Phase I safety and immunogenicity. Eur J Immunol 35: 2031-40.
124Cohen C, Perrault G, Voltz C, Steinberg R, Soubrie P. 2002. SR141716, a central cannabinoid (CB(1)) receptor antagonist, blocks the motivational and dopamine-releasing effects of nicotine in rats. Behav Pharmacol 13: 451-63.
125Anthenelli RM, Despres JP. 2004. Effects of rimonabant in the reduction of major cardiovascular risk factors. Results from the STRATUS-US trial (smoking cessation in smokers motivated to quit). Presented at American College of Cardiology 53rd Annual Scientific Session, Rapid News Summaries, New Orleans, LA.
126Anthenelli RM. 2005. Rimonabant in smoking abstinence: the STRATUS-program. Presented at Proceedings of the 11 Annual Meeting of the Society for Research on Nicotine and Tobacco, Prague, Czech Republic.
127Benowitz NL, Jacob P, 3rd. 1994. Metabolism of nicotine to cotinine studied by a dual stable isotope method. Clin Pharmacol Ther 56: 483-93.
128Messina ES, Tyndale RF, Sellers EM. 1997. A major role for CYP2A6 in nicotine C-oxidation by human liver microsomes. J Pharmacol Exp Ther 282: 1608-14.
129Fukami T, Nakajima M, Yoshida R, Tsuchiya Y, Fujiki Y, et al. 2004. A novel polymorphism of human CYP2A6 gene CYP2A6*17 has an amino acid substitution (V365M) that decreases enzymatic activity in vitro and in vivo. Clin Pharmacol Ther 76: 519-27.
130Xu C, Rao YS, Xu B, Hoffmann E, Jones J, et al. 2002. An in vivo pilot study characterizing the new CYP2A6*7, *8, and *10 alleles. Biochem Biophys Res Commun 290: 318-24.
131Yamanaka H, Nakajima M, Nishimura K, Yoshida R, Fukami T, et al. 2004. Metabolic profile of nicotine in subjects whose CYP2A6 gene is deleted. Eur J Pharm Sci 22: 419-25.
132Yoshida R, Nakajima M, Watanabe Y, Kwon JT, Yokoi T. 2002. Genetic polymorphisms in human CYP2A6 gene causing impaired nicotine metabolism. Br J Clin Pharmacol 54: 511-7.
133Henningfield JE, Goldberg SR. 1988. Pharmacologic determinants of tobacco self-administration by humans. Pharmacol Biochem Behav 30: 221-6.
134Russell MA, Feyerabend C. 1978. Cigarette smoking: a dependence on high-nicotine boli. Drug Metab Rev 8: 29-57.
135Malaiyandi V, Lerman C, Benowitz NL, Jepson C, Patterson F, Tyndale RF. 2006. Impact of CYP2A6 genotype on pretreatment smoking behaviour and nicotine levels from and usage of nicotine replacement therapy. Mol Psychiatry.
136Rao Y, Hoffmann E, Zia M, Bodin L, Zeman M, et al. 2000. Duplications and defects in the CYP2A6 gene: identification, genotyping, and in vivo effects on smoking. Mol Pharmacol 58: 747-55.
137Schoedel KA, Hoffmann EB, Rao Y, Sellers EM, Tyndale RF. 2004. Ethnic variation in CYP2A6 and association of genetically slow nicotine metabolism and smoking in adult Caucasians. Pharmacogenetics 14: 615-26.
138Gu DF, Hinks LJ, Morton NE, Day IN. 2000. The use of long PCR to confirm three common alleles at the CYP2A6 locus and the relationship between genotype and smoking habit. Ann Hum Genet 64: 383-90.
139Minematsu N, Nakamura H, Furuuchi M, Nakajima T, Takahashi S, et al. 2006. Limitation of cigarette consumption by CYP2A6*4, *7 and *9 polymorphisms. Eur Respir J 27: 289-92.
140Draper AJ, Madan A, Parkinson A. 1997. Inhibition of coumarin 7-hydroxylase activity in human liver microsomes. Arch Biochem Biophys 341: 47-61.