Topic Highlight Open Access
Copyright ©2006 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Gastroenterol. Dec 14, 2006; 12(46): 7433-7439
Published online Dec 14, 2006. doi: 10.3748/wjg.v12.i46.7433
The genetics of nicotine dependence: Relationship to pancreatic cancer
Stewart L MacLeod, Department of Pediatrics, University of Arkansas for Medical Sciences, and Arkansas Children’s Hospital Research Institute, Arkansas Center for Birth Defects Research and Prevention, Little Rock, AR, United States
Parimal Chowdhury, Department of Physiology, University of Arkansas for Medical Sciences, Little Rock, AR, United States
Author contributions: All authors contributed equally to the work.
Correspondence to: Stewart MacLeod, PhD, Department of Pediatrics, UAMS College of Medicine, Arkansas Center for Birth Defects Research and Prevention, 1120 Marshall St. Mail Slot 512-40, Little Rock, AR 72202, United States. macleodstewartl@uams.edu
Telephone: +1-501-3644586 Fax: +1-501-3645107
Received: July 29, 2006
Revised: September 15, 2006
Accepted: September 29, 2006
Published online: December 14, 2006

Abstract

Smoking of tobacco products continues to be a major cause of worldwide health problems. Epidemiological studies have shown that tobacco smoking is the greatest risk factor for the development of pancreatic cancer. Smokers who are able to quit smoking can reduce their risk of pancreatic cancer by nearly 50% within two years, however, their risk of developing pancreatic cancer remains higher than that of non-smokers for 10 years. Nicotine is the major psychoactive substance in tobacco, and is responsible for tobacco dependence and addiction. Recent evidence suggests that individuals have genetically based differences in their ability to metabolize nicotine, as well as genetic differences in the psychological reward pathways that may influence individual response to smoking initiation, dependence, addiction and cessation. Numerous associations have been reported between smoking behavior and genetic polymorphisms in genes that are responsible for nicotine metabolism. In addition, polymorphisms in genes that encode neurotransmitters and transporters that function in psychological reward pathways have been implicated in differences in smoking behavior. However, there is a large degree of between-study variability that demonstrates the need for larger, well-controlled case-control studies to identify target genes and deduce mechanisms that account for the genetic basis of inter-individual differences in smoking behavior. Understanding the genetic factors that increase susceptibility to tobacco addiction may result in more effective tobacco cessation programs which will, in turn, reduce the incidence of tobacco related disease, including pancreatic cancer.

Key Words: Smoking, Pancreatic cancer, Nicotine dep-endence, Genetics, Polymorphism, Nicotine metabolism



INTRODUCTION

The use of tobacco products constitutes the most preventable cause of premature death worldwide. It is estimated that half of all Americans who continue to smoke will die of smoking-related diseases[1], including cancer, emphysema and heart disease. The U.S. Surgeon General reported in 2004 that cigarette smoking has caused more than 12 million premature deaths in the United States since the publication of the Surgeon Generals report on smoking and health in 1964[2]. Although the dangers inherent in smoking are well known, the use of tobacco continues because these products are delivery systems for nicotine, an extremely addictive drug. Long term smoking is the most established risk factor for lung cancer, however smoking also increases the risk of developing cancers of the esophagus, uterine cervix, kidney, bladder, stomach and pancreas[2]. This review will briefly discuss the effects of cigarette smoke components on the pancreas, and then focus on the genetic basis for nicotine addiction that is responsible for tobacco product use.

Pancreatic cancer is a relatively rare tumor; the lifetime risk of dying from pancreatic cancer is between 1% to 2% of the general population. However, pancreatic tumors are extremely aggressive, with a 5-year survival rate of less than 5% and a mortality rate of nearly 100%[3,4]. Since there are presently no tests for early screening, and once detected, therapeutic choices are limited, the best hope for reducing mortality from pancreatic cancer is prevention. Smoking of tobacco products is the most important risk factor for developing pancreatic cancer[3-6], with risk increasing with higher levels of tobacco use and more years of exposure. A 40 year epidemiological study of British physicians determined that the pancreatic cancer rate for smokers was 35 per 100 000 person years, while the risk for ex-smokers and non-smokers was reduced to 23 and 16 per 100 000 person years respectively[7]. A large prospective study by Fuchs et al[8] determined that current smokers had a 2.5 times greater relative risk of developing pancreatic cancer compared to study subjects who had never smoked. They also found that although smoking cessation reduces the risk of pancreatic cancer by 48% within 2 years, smokers who quit have an increased risk for approximately 10 years compared to non-smokers. Smoking, as a single risk factor or in combination with alcohol, increases the risk of idiopathic pancreatitis, a precancerous condition[3,5,9,10]. Clearly, one way to reduce the risk of pancreatic cancer is to develop more effective smoking cessation strategies; however, smoking is a complex behavior, which is likely influenced by both environmental and genetic factors. There is wide inter-individual variation in the risk of becoming dependent on nicotine. Genetic factors that influence nicotine metabolism and the psychological reward system of nicotine use are both thought to be important in the development of nicotine dependence.

MECHANISMS OF TOBACCO INDUCED PANCREATIC CANCER

Cigarette smoke is a complex mixture of carcinogenic compounds and nicotine, many of which have deleterious effects on the exocrine pancreas. Nicotine forms carcinogenic N-nitroso compounds during processing[11-14], which induce pancreatic cancer in the Syrian golden hamster model of pancreatic carcinogenesis[15,16]. A particular component of tobacco smoke, 4-(methylnitrosamino)-1-butanone (NNK), is implicated in pancreatic carcinogenesis by its ability to form DNA adducts[17] which have been associated with activating RAS mutations that are found in most human pancreatic adenocarcinomas[18]. NNK also induces proliferation in pancreatic ductal epithelial cells by stimulating EGF mediated signal transduction pathways through binding to β-adrenergic receptors[19,20].

Nicotine, although not carcinogenic, exerts toxic effects on the pancreas. Exposure of rat pancreatic acini to nicotine results in increased protein secretion[21]. We have demonstrated that nicotine induces cytoplasmic vacuolization, cellular edema and increases the cellular amylase content in the exocrine pancreas[22]. We also demonstrated that the increase in pancreatic enzymes in nicotine treated rats was accompanied by reduced CCK mediated enzyme secretion, which may be the causative factor in nicotine induced pancreatic cell pathology. Exposure of rats to tobacco smoke produced fibrosis and scarring of pancreatic acinar structures, characteristic of chronic pancreatitis[23]. In humans, nicotine inhibits the secretion of bicarbonate and affects the composition of pancreatic secretions[24,25], and in patients with pancreatitis, nicotine exposure resulted in increased pancreatic enzyme secretion, including amylase[26] and lipase[27].

THE GENETICS OF NICOTINE DEPENDENCE

Nicotine is the major psychoactive component of tobacco that is responsible for dependence through a nicotine stimulated reward system that is thought to be mediated by the dopaminergic system of the brain. Recent research has identified behavioral, environmental and genetic factors that influence the various stages of smoking behavior, including smoking initiation, development of addiction and smoking cessation. Early evidence for the influence of heredity on tobacco dependence came from studies of tobacco use among families[28], adopted siblings and among mono and dizygotic twins[29,30]. In a study of mono and dizygotic twins, True et al[31] found that 50% of the risk of smoking initiation and 70% of the risk for continuing to smoke were due to genetic factors.

POLYMORPHIC GENES INVOLVED IN NICOTINE METABOLISM

Individuals with increased tolerance to nicotine because of a greater capacity to metabolize the drug may experience fewer adverse reactions to their first encounter with nicotine, and therefore may have a greater propensity to continue using tobacco products. Conversely, slow metabolizers of nicotine would be expected to smoke less and would be less likely to become nicotine dependent. Smokers tend to adjust their smoking behavior in order to maintain a certain level of nicotine in the brain[32], so that an individual’s capacity to metabolize nicotine will influence their intake and exposure. Polymorphic expression of genes that are responsible for nicotine metabolism may be responsible for the wide variability in nicotine tolerance between individuals.

The hepatic enzymes cytochrome P450 2A6 (CYP2A6) and cytochrome P450 2D6 are the major isoforms responsible for the metabolism of nicotine to cotinine, however hepatic CYP2A6 is responsible for 90% of the first pass metabolism of nicotine[33,34]. Both of these enzymes are polymorphic in the human population, with genetic differences that are responsible for high and low activity alleles.

CYTOCHROME P450 2A6 (CYP2A6)

The CYP2A6 gene is polymorphic in the human population, with large inter-individual differences in the levels of hepatic CYP2A6 protein and enzyme activity[34-36]. A deletional allele was found to be responsible for low or non-existent CYP2A6 activity[36]. A number of studies have found that individuals with genetically determined slow or absent CYP 2A6 activity have a reduced risk of becoming smokers, and those who do smoke tend to smoke fewer cigarettes per day and have a higher smoking cessation success rate[37-40]. Pianneza et al[41] reported that a tobacco dependent population had an under representation of low activity CYP2A6 alleles, and that those smokers who had low activity alleles tended to smoke fewer cigarettes per week, suggesting a role for CYP2A6 in nicotine tolerance and dependence. However, other studies have failed to detect an association between genetically low CYP2A6 activity and nicotine use or dependence[42,43]. One reason for this discrepancy was the use in early studies of a genotyping protocol that overestimated the number of low activity alleles in the study population[44,45]. A meta-analysis by Carter et al[46] also failed to find a significant association between CYP2A6 genotype and smoking behavior. In another meta-analysis, Munafo et al[47] concluded that the influence of individual genes on smoking behavior may be subtle, and that larger studies that have the power to detect the effects of multiple genes on smoking behavior will be necessary. However, in a number of studies, they found strong evidence that the reduced activity allele of CYP2A6 was associated with smokers who were able to quit smoking.

Recently, the use of CYP2A6 inhibitors has been explored as a chemoprevention strategy for smoking cessation. A study by von Weymarn et al[48] reported that the benzyl and phenylethyl isothiocyanates that are found in cruciferous vegetables such as broccoli and cabbage, were effective competitive inhibitors of both CYP2A6 and CYP 2A13. CYP2A13 is the enzyme responsible for the activation of the tobacco procarcinogen NNK to its ultimate carcinogenic form. This report suggests that inhibition of CYP2A6 may convert the phenotype of smokers to one which confers less metabolic tolerance to nicotine, leading to fewer cigarettes smoked per day. The resultant reduction in nicotine metabolism may possibly increase success with smoking cessation. At the same time, inhibition of CYP2A13, which is found primarily in the lung, may result in less activation of NNK, potentially protecting individuals who continue to smoke from developing lung cancer. Sellers et al[49] reported that another CYP2A6 inhibitor, methoxsalen, was effective in increasing the bioavailability of nicotine in smokers, resulting in a decrease in the number of cigarettes they smoked per day. These studies demonstrate that the use of CYP2A6 inhibitors may be a useful strategy to reduce tobacco exposure and may have the potential to increase the success rate of smoking cessation programs.

CYTOCHROME P450 2D6 (CYP2D6)

Caporaso et al[50] determined inter-individual differences in CYP2D6 phenotype by measuring the metabolism of dextromethorphan, a CYP2D6 substrate. They concluded that polymorphisms in CYP2D6 were not major determinants of nicotine metabolism in smokers except in ultrametabolizers. These are individuals who have a duplication of the CYP2D6 gene that is present in 3%-8% of Caucasians and up to 30% of other ethnic groups[51]. This duplication results in the production of high levels of functional CYP2D6 protein and results in the increased metabolism of CYP2D6 substrates, including nicotine. In a case-control study of lung and larynx cancer, the CYP2D6 gene duplication was found in 13% of cancer patients compared to 6% of healthy control subjects. The frequency of a genetic polymorphism that codes for a high activity CYP2D6 allele called CYP2D6*9 was also higher in cases compared to controls[52]. Approximately 3%-10% of Caucasians are CYP2D6 poor metabolizers, due to inheritance of two defective alleles. Saarikowski et al[53] found the same proportion of poor metabolizers in groups of smokers and never-smokers, however among men, a trend toward more poor metabolizers in the non-smoking group was observed. They also found twofold more ultrametabolizers among heavy smokers compared to non-smokers. Overall, the results of these studies suggest that CYP2D6 affects nicotine metabolism among individuals who have high activity due to gene duplication, however the influence of the low activity allele remains controversial.

GENES INVOLVED IN NICOTINE DEPENDENCE

Nicotine is thought to induce a euphoric state in users and by that is thought to be the result of activation of the mesolimbic dopaminergic reward system in the nucleus accumbens of the brain[54-56]. Nicotine binds to nicotinic receptors that, when activated, enhance dopamine release in areas of the brain that are thought to be involved in reward[55,57]. The involvement of the dopaminergic system in the reinforcement activity of nicotine may be related to the highly addictive properties of the drug[56]. Genetic polymorphisms in genes that affect this reward system, including dopamine receptors and transporters, nicotinic receptors and serotonin receptors may modulate an individual’s risk of becoming nicotine dependent.

DOPAMINE RECEPTOR GENE POLYMORPHISMS

The human dopamine D2 receptor (DRD2) has a TaqI polymorphism with two minor alleles termed the TaqIA allele (A1 and A2) and the TaqIB allele (B1 and B2). The TaqI*A1 allele has been shown to be associated with reduced expression of dopamine D2 receptor in the striatum[58-61]. It has been hypothesized that subjects with reduced numbers of dopamine receptors may compensate for this deficiency by using nicotine to increase brain dopamine levels. The presence of the DRD2 TaqI allele has been associated with an earlier age of smoking initiation[62], increased risk of being a current smoker, and reduced duration of smoking abstinence[63]. Spitz et al[64], conducted a case control study of lung cancer patients and found that a greater percentage of chronic smokers had the B1B2 genotype compared with non-smokers, whether they were cancer cases or controls. The least common A1 or B1 alleles were associated with individuals who were younger when they started smoking and had attempted to quit smoking fewer times compared with smokers with the more common DRD2 alleles. Other studies have failed to find an association between the DRD2 TaqI allele and smoking behavior. Bierut et al[65] analyzed a family study by the transmission disequilibrium test and found no difference in the frequency of DRD2 alleles transmitted to habitual smokers. In a small British study, Singleton et al[66] found no increase in the DRD2 TaqI allele in smokers compared to non-smokers, and Johnstone et al duplicated these findings in a larger study[67]. Munafo et al[47] conducted a meta-analysis of the genetic basis for smoking behavior and concluded that there is some evidence for an association between the DRD2 TaqI*A1 allele and smoking behavior, but larger, better designed studies in a variety of populations are needed to confirm this relationship.

Other dopamine receptors are genetically polymorphic, but have received less attention than the DRD2 gene with respect to smoking behavior. A 5’ polymorphism of no known function, located in the dopamine D1 receptor gene (DRD1), has been associated with smoking[68]. The DRD3 receptor is genetically polymorphic and is highly expressed in the nucleus accumbens, however, no association with smoking behavior has been reported[69].

A polymorphism characterized as a variable number of tandem repeats has been reported in the dopamine D4 receptor (DRD4)[70,71]. The receptor containing 7 repeats has been characterized as having a reduced response to dopamine binding[72]. Shields et al[73] compared DRD4 genotype for a population of smokers compared to non- smokers and found that African American smokers had a higher incidence of the DRD4 allele containing 7 repeats than African American non-smokers. In addition, African American smokers with this allele had an earlier age of smoking initiation and less success at smoking cessation than African Americans with shorter repeat sequences. The same analysis of a Caucasian population showed no association with smoking status. This data suggests that individual genotypes may be a factor in the success of smoking cessation strategies, and more effective strategies may need to be tailored to an individual’s genetic background. The human DRD5 gene has at least 4 mis-sense or nonsense polymorphisms, however, no association with smoking behavior has been reported[74].

DOPAMINE TRANSPORTER POLYMORPHISMS

The dopamine transporter gene (SLC6AC) has a variable number of tandem repeats polymorphism in the 3’ noncoding region of the gene[75] that is associated with reduced transporter levels in the brain[76]. The role of this polymorphism in determining smoking behavior is not clear, however, in some studies, the polymorphism containing 9 repeats was found to be associated with greater levels of cigarette craving among African American smokers[77]. Other studies found that individuals with the SLC6AC 9 repeat allele were less likely to be smokers, especially if they also had the DRD2 Taq1 A1 allele[78]. These individuals were less likely to start smoking at an early age, and those who did smoke had greater periods of smoking cessation. Sabol et al[79] confirmed these findings in a diverse population of smokers, non-smokers, and ex-smokers, however Vandenbergh et al[80] failed to replicate these results in spite of another report on the function of the 9 repeat sequence as a transcriptional enhancer[81,82].

CATABOLISM OF DOPAMINE

The action of dopamine is terminated by the action of catabolic enzymes, primarily catecholamine-O-methyltransferase (COMT) with lesser roles for monoamine oxidase A and B (MAO) and dopamine β hydroxylase (DBH). Genetic polymorphisms have been discovered in these enzymes, and the effects on smoking behavior have been tested in a number of studies. In a study of smokers, McKinney et al[83] reported that polymorphisms in DBH and MAO were related to maintenance of nicotine levels and predicted the quantity of cigarettes smoked. No association between the amount of tobacco consumed and the functional COMT A1947G single nucleotide polymorphism (SNP), which results in the substitution of a methionine for valine at codon 108 (Met108Val) in the COMT protein, was detected in this study. These results were unexpected in light of a report that the G allele at this locus results in a three to fourfold increase in COMT activity[84], and a reported association of this allele with addiction to other drugs[85]. The lack of association between this SNP and smoking initiation, persistence and cessation was replicated in a larger study of current smokers, ex-smokers and lifetime non-smokers[86]. More recent studies include the finding of a positive association between the high activity COMT allele and nicotine dependence in a Caucasian population, however, these results were not replicated in a second independent study by the same researchers[87]. In a case-control study of women, those who were homozygous for the lower activity Met allele were more likely to be ex-smokers rather than current smokers, and in a nicotine replacement clinical trial reported by the same group, Met homozygotes at the COMT locus had more success at smoking cessation[88]. Bueten et al[89] analyzed five allelic variants in the COMT gene, including the Met108Val SNP, and found a significant association with nicotine dependence. The lack of reproducibility of this data, even when the same researchers analyze independent study populations, may be due to the lack of power to detect relatively small effects of COMT on smoking behavior. These results demonstrate the need for large, adequately powered replication studies to determine the genetic basis of smoking behavior and nicotine addiction.

SEROTONIN TRANSPORTER PROMOTER POLYMORPHISMS

Nicotine increases the secretion of serotonin in the brain[90] therefore, the serotonergic system may have a function in determining smoking behavior. Lower serotonin reuptake has been associated with an increased risk of depression as well as increased impulsive or aggressive behavior[91]. This combination of behavioral traits is termed neuroticism, and has been associated with increased incidence of smoking, nicotine dependence, and difficulty in quitting smoking[92]. Evidence for a genetic link to neuroticism and smoking behavior centers around a 44-bp deletion/insertion polymorphism that corresponds to short (S) and long (L) versions of the serotonin transporter gene (5-hydroxytryptamine transporter or 5-HTT) promoter[93]. Functional characterization of this polymorphism has demonstrated that the short promoter variant reduces the transcriptional activity of the gene and results in decreased 5-HTT expression and decreased serotonin uptake[94]. Hu et al[92] found a relationship between the genotype for the 5-HTT promoter polymorphism and degree of neuroticism and smoking behavior. This finding was confirmed by another study which reported that smokers who were heterozygous or homozygous for the 5-HTT S-allele were more likely to be dependent on nicotine than subjects who were homozygous for the L allele[95]. However in a Japanese population, the presence of the S allele was associated with non-smokers or ex-smokers, indicating that individuals who were homozygous for the S allele were less likely to begin smoking, or were more successful at smoking cessation[96]. Other studies were unable to detect any association between the 5-HTT promoter polymorphism and smoking behavior in Caucasian or African American populations, in spite of significant differences in the distribution of 5-HTT promoter alleles between racial groups[97].

CONCLUSIONS

The variability of results reported for most candidate genes that are hypothesized to affect smoking behavior demonstrates the need for larger, well controlled studies designed to define the genetic basis for inter-individual differences in nicotine and dopamine metabolism as they relate to smoking behavior and nicotine dependence. Results of these future studies will be useful in identifying individuals who are at increased risk of becoming dependent on nicotine and will also facilitate the development of smoking cessation strategies that are targeted to individual differences in nicotine and neurotransmitter action and metabolism. Considering that smoking is the greatest risk factor for pancreatic cancer, reduction of tobacco use through both abstinence programs and successful smoking cessation is the best hope for reducing the risk of developing this devastating disease.

Footnotes

S- Editor Liu Y L- Editor Alpini GD E- Editor Bai SH

References
1.  Peto R, Lopez A, Boreham J, Thun M.  Mortality from smoking in developed countries 1950-2000. London: Oxford University Press 1994; .  [PubMed]  [DOI]  [Cited in This Article: ]
2.  United States Department of Health and Human Services The Health Consequences of Smoking. A Report of the Surgeon General. Rockville: Public Health Services, Centers for Disease Control and Prevention and Health Promotion, Office on Smoking and Health 2004; .  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Lowenfels AB, Maisonneuve P. Risk factors for pancreatic cancer. J Cell Biochem. 2005;95:649-656.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 116]  [Cited by in F6Publishing: 120]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
4.  Lowenfels AB, Maisonneuve P. Epidemiology and risk factors for pancreatic cancer. Best Pract Res Clin Gastroenterol. 2006;20:197-209.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 333]  [Cited by in F6Publishing: 354]  [Article Influence: 19.7]  [Reference Citation Analysis (0)]
5.  Talamini G, Bassi C, Falconi M, Frulloni L, Di Francesco V, Vaona B, Bovo P, Rigo L, Castagnini A, Angelini G. Cigarette smoking: an independent risk factor in alcoholic pancreatitis. Pancreas. 1996;12:131-137.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 96]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
6.  Talamini G, Falconi M, Bassi C, Mastromauro M, Salvia R, Pederzoli P. Chronic pancreatitis: relationship to acute pancreatitis and pancreatic cancer. JOP. 2000;1:69-76.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Doll R, Peto R, Wheatley K, Gray R, Sutherland I. Mortality in relation to smoking: 40 years' observations on male British doctors. BMJ. 1994;309:901-911.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1118]  [Cited by in F6Publishing: 1143]  [Article Influence: 38.1]  [Reference Citation Analysis (0)]
8.  Fuchs CS, Colditz GA, Stampfer MJ, Giovannucci EL, Hunter DJ, Rimm EB, Willett WC, Speizer FE. A prospective study of cigarette smoking and the risk of pancreatic cancer. Arch Intern Med. 1996;156:2255-2260.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 178]  [Cited by in F6Publishing: 179]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
9.  Talamini G, Bassi C, Falconi M, Sartori N, Salvia R, Rigo L, Castagnini A, Di Francesco V, Frulloni L, Bovo P. Alcohol and smoking as risk factors in chronic pancreatitis and pancreatic cancer. Dig Dis Sci. 1999;44:1303-1311.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 175]  [Cited by in F6Publishing: 187]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
10.  Malfertheiner P, Schütte K. Smoking--a trigger for chronic inflammation and cancer development in the pancreas. Am J Gastroenterol. 2006;101:160-162.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 59]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
11.  Anderson LM, Hecht SS, Kovatch RM, Amin S, Hoffmann D, Rice JM. Tumorigenicity of the tobacco-specific carcinogen 4-(methyl-nitrosamino)-1-(3-pyridyl)-1-butanone in infant mice. Cancer Lett. 1991;58:177-181.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 17]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
12.  Hecht SS, Hoffmann D. N-nitroso compounds and man: sources of exposure, endogenous formation and occurrence in body fluids. Eur J Cancer Prev. 1998;7:165-166.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Hoffmann D, Hecht SS. Nicotine-derived N-nitrosamines and tobacco-related cancer: current status and future directions. Cancer Res. 1985;45:935-944.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Hoffmann D, Lavoie EJ, Hecht SS. Nicotine: a precursor for carcinogens. Cancer Lett. 1985;26:67-75.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 31]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
15.  Pour PM, Runge RG, Birt D, Gingell R, Lawson T, Nagel D, Wallcave L, Salmasi SZ. Current knowledge of pancreatic carcinogenesis in the hamster and its relevance to the human disease. Cancer. 1981;47:1573-1589.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 5]  [Reference Citation Analysis (0)]
16.  Pour PM, Salmasi SZ, Runge RG. Selective induction of pancreatic ductular tumors by single doses of N-nitrosobis(2-oxopropyl)amine in Syrian golden hamsters. Cancer Lett. 1978;4:317-323.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 47]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
17.  Wang M, Abbruzzese JL, Friess H, Hittelman WN, Evans DB, Abbruzzese MC, Chiao P, Li D. DNA adducts in human pancreatic tissues and their potential role in carcinogenesis. Cancer Res. 1998;58:38-41.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Li D, Firozi PF, Zhang W, Shen J, DiGiovanni J, Lau S, Evans D, Friess H, Hassan M, Abbruzzese JL. DNA adducts, genetic polymorphisms, and K-ras mutation in human pancreatic cancer. Mutat Res. 2002;513:37-48.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 62]  [Cited by in F6Publishing: 66]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
19.  Askari MD, Tsao MS, Schuller HM. The tobacco-specific carcinogen, 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone stimulates proliferation of immortalized human pancreatic duct epithelia through beta-adrenergic transactivation of EGF receptors. J Cancer Res Clin Oncol. 2005;131:639-648.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 76]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
20.  Schuller HM. Mechanisms of smoking-related lung and pancreatic adenocarcinoma development. Nat Rev Cancer. 2002;2:455-463.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 136]  [Cited by in F6Publishing: 123]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
21.  Majumdar AP, Davis GA, Dubick MA, Geokas MC. Nicotine stimulation of protein secretion from isolated rat pancreatic acini. Am J Physiol. 1985;248:G158-G163.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Chowdhury P, Doi R, Tangoku A, Rayford PL. Structural and functional changes of rat exocrine pancreas exposed to nicotine. Int J Pancreatol. 1995;18:257-264.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Wittel UA, Pandey KK, Andrianifahanana M, Johansson SL, Cullen DM, Akhter MP, Brand RE, Prokopczyk B, Batra SK. Chronic pancreatic inflammation induced by environmental tobacco smoke inhalation in rats. Am J Gastroenterol. 2006;101:148-159.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 112]  [Cited by in F6Publishing: 106]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
24.  Brown P. The influence of smoking on pancreatic function in man. Med J Aust. 1976;2:290-293.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Bynum TE, Solomon TE, Johnson LR, Jacobson ED. Inhibition of pancreatic secretion in man by cigarette smoking. Gut. 1972;13:361-365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 78]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
26.  Milnerowicz H, Sliwińska M, Jabłonowska M, Milnerowicz S. [Effect of tobacco smoking on amylase activity in patients with pancreatitis]. Przegl Lek. 2004;61:1071-1072.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Sliwińska-Mossoń M, Milnerowicz H. [Influence of tobacco smoking on lipase activity in patients with pancreatitis]. Przegl Lek. 2005;62:1058-1061.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Cheng LS, Swan GE, Carmelli D. A genetic analysis of smoking behavior in family members of older adult males. Addiction. 2000;95:427-435.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 19]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
29.  Carmelli D, Swan GE, Robinette D, Fabsitz R. Genetic influence on smoking--a study of male twins. N Engl J Med. 1992;327:829-833.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 232]  [Cited by in F6Publishing: 238]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
30.  Swan GE, Carmelli D, Cardon LR. The consumption of tobacco, alcohol, and coffee in Caucasian male twins: a multivariate genetic analysis. J Subst Abuse. 1996;8:19-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 93]  [Cited by in F6Publishing: 96]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
31.  True WR, Heath AC, Scherrer JF, Waterman B, Goldberg J, Lin N, Eisen SA, Lyons MJ, Tsuang MT. Genetic and environmental contributions to smoking. Addiction. 1997;92:1277-1287.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 163]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
32.  McMorrow MJ, Foxx RM. Nicotine's role in smoking: an analysis of nicotine regulation. Psychol Bull. 1983;93:302-327.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 86]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
33.  Messina ES, Tyndale RF, Sellers EM. A major role for CYP2A6 in nicotine C-oxidation by human liver microsomes. J Pharmacol Exp Ther. 1997;282:1608-1614.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Nakajima M, Yamamoto T, Nunoya K, Yokoi T, Nagashima K, Inoue K, Funae Y, Shimada N, Kamataki T, Kuroiwa Y. Role of human cytochrome P4502A6 in C-oxidation of nicotine. Drug Metab Dispos. 1996;24:1212-1217.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Nunoya K, Yokoi T, Kimura K, Inoue K, Kodama T, Funayama M, Nagashima K, Funae Y, Green C, Kinoshita M. A new deleted allele in the human cytochrome P450 2A6 (CYP2A6) gene found in individuals showing poor metabolic capacity to coumarin and (+)-cis-3,5-dimethyl-2-(3-pyridyl)thiazolidin-4-one hydrochloride (SM-12502). Pharmacogenetics. 1998;8:239-249.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 76]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
36.  Nunoya KI, Yokoi T, Kimura K, Kainuma T, Satoh K, Kinoshita M, Kamataki T. A new CYP2A6 gene deletion responsible for the in vivo polymorphic metabolism of (+)-cis-3,5-dimethyl-2-(3-pyridyl)thiazolidin-4-one hydrochloride in humans. J Pharmacol Exp Ther. 1999;289:437-442.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  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. Mol Pharmacol. 2000;58:747-755.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Schoedel KA, Hoffmann EB, Rao Y, Sellers EM, Tyndale RF. Ethnic variation in CYP2A6 and association of genetically slow nicotine metabolism and smoking in adult Caucasians. Pharmacogenetics. 2004;14:615-626.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 216]  [Cited by in F6Publishing: 209]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]
39.  Ando M, Hamajima N, Ariyoshi N, Kamataki T, Matsuo K, Ohno Y. Association of CYP2A6 gene deletion with cigarette smoking status in Japanese adults. J Epidemiol. 2003;13:176-181.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 26]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
40.  Fujieda M, Yamazaki H, Saito T, Kiyotani K, Gyamfi MA, Sakurai M, Dosaka-Akita H, Sawamura Y, Yokota J, Kunitoh H. Evaluation of CYP2A6 genetic polymorphisms as determinants of smoking behavior and tobacco-related lung cancer risk in male Japanese smokers. Carcinogenesis. 2004;25:2451-2458.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 150]  [Cited by in F6Publishing: 141]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
41.  Pianezza ML, Sellers EM, Tyndale RF. Nicotine metabolism defect reduces smoking. Nature. 1998;393:750.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 288]  [Cited by in F6Publishing: 303]  [Article Influence: 11.7]  [Reference Citation Analysis (0)]
42.  London SJ, Idle JR, Daly AK, Coetzee GA. Genetic variation of CYP2A6, smoking, and risk of cancer. Lancet. 1999;353:898-899.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 84]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
43.  Loriot MA, Rebuissou S, Oscarson M, Cenée S, Miyamoto M, Ariyoshi N, Kamataki T, Hémon D, Beaune P, Stücker I. Genetic polymorphisms of cytochrome P450 2A6 in a case-control study on lung cancer in a French population. Pharmacogenetics. 2001;11:39-44.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 70]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
44.  Oscarson M, Gullstén H, Rautio A, Bernal ML, Sinues B, Dahl ML, Stengård JH, Pelkonen O, Raunio H, Ingelman-Sundberg M. Genotyping of human cytochrome P450 2A6 (CYP2A6), a nicotine C-oxidase. FEBS Lett. 1998;438:201-205.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 108]  [Cited by in F6Publishing: 101]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
45.  Oscarson M, McLellan RA, Gullstén H, Agúndez JA, Benítez 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 Lett. 1999;460:321-327.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 137]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
46.  Carter B, Long T, Cinciripini P. A meta-analytic review of the CYP2A6 genotype and smoking behavior. Nicotine Tob Res. 2004;6:221-227.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 28]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
47.  Munafò M, Clark T, Johnstone E, Murphy M, Walton R. The genetic basis for smoking behavior: a systematic review and meta-analysis. Nicotine Tob Res. 2004;6:583-597.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 207]  [Cited by in F6Publishing: 188]  [Article Influence: 9.4]  [Reference Citation Analysis (0)]
48.  von Weymarn LB, Chun JA, Hollenberg PF. Effects of benzyl and phenethyl isothiocyanate on P450s 2A6 and 2A13: potential for chemoprevention in smokers. Carcinogenesis. 2006;27:782-790.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 67]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
49.  Sellers EM, Kaplan HL, Tyndale RF. Inhibition of cytochrome P450 2A6 increases nicotine's oral bioavailability and decreases smoking. Clin Pharmacol Ther. 2000;68:35-43.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 125]  [Cited by in F6Publishing: 131]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
50.  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 Epidemiol Biomarkers Prev. 2001;10:261-263.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Agúndez JA, Gallardo L, Ledesma MC, Lozano L, Rodriguez-Lescure A, Pontes JC, Iglesias-Moreno MC, Poch J, Ladero JM, Benítez J. Functionally active duplications of the CYP2D6 gene are more prevalent among larynx and lung cancer patients. Oncology. 2001;61:59-63.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 31]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
52.  Boustead C, Taber H, Idle JR, Cholerton S. CYP2D6 genotype and smoking behaviour in cigarette smokers. Pharmacogenetics. 1997;7:411-414.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 22]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
53.  Saarikoski ST, Sata F, Husgafvel-Pursiainen K, Rautalahti M, Haukka J, Impivaara O, Järvisalo J, Vainio H, Hirvonen A. CYP2D6 ultrarapid metabolizer genotype as a potential modifier of smoking behaviour. Pharmacogenetics. 2000;10:5-10.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 45]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
54.  Wise RA, Rompre PP. Brain dopamine and reward. Annu Rev Psychol. 1989;40:191-225.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1286]  [Cited by in F6Publishing: 1246]  [Article Influence: 35.6]  [Reference Citation Analysis (0)]
55.  Pidoplichko VI, Noguchi J, Areola OO, Liang Y, Peterson J, Zhang T, Dani JA. Nicotinic cholinergic synaptic mechanisms in the ventral tegmental area contribute to nicotine addiction. Learn Mem. 2004;11:60-69.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 147]  [Cited by in F6Publishing: 154]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
56.  Balfour DJ. The neurobiology of tobacco dependence: a preclinical perspective on the role of the dopamine projections to the nucleus accumbens [corrected]. Nicotine Tob Res. 2004;6:899-912.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 168]  [Cited by in F6Publishing: 161]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
57.  Brazell MP, Mitchell SN, Gray JA. Effect of acute administration of nicotine on in vivo release of noradrenaline in the hippocampus of freely moving rats: a dose-response and antagonist study. Neuropharmacology. 1991;30:823-833.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 70]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
58.  Noble EP, Blum K, Ritchie T, Montgomery A, Sheridan PJ. Allelic association of the D2 dopamine receptor gene with receptor-binding characteristics in alcoholism. Arch Gen Psychiatry. 1991;48:648-654.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 482]  [Cited by in F6Publishing: 502]  [Article Influence: 15.2]  [Reference Citation Analysis (0)]
59.  Pohjalainen T, Rinne JO, Någren K, Lehikoinen P, Anttila K, Syvälahti EK, Hietala J. The A1 allele of the human D2 dopamine receptor gene predicts low D2 receptor availability in healthy volunteers. Mol Psychiatry. 1998;3:256-260.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 530]  [Cited by in F6Publishing: 508]  [Article Influence: 19.5]  [Reference Citation Analysis (0)]
60.  Jönsson EG, Nöthen MM, Grünhage 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. Mol Psychiatry. 1999;4:290-296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 540]  [Cited by in F6Publishing: 524]  [Article Influence: 21.0]  [Reference Citation Analysis (0)]
61.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 413]  [Cited by in F6Publishing: 420]  [Article Influence: 15.6]  [Reference Citation Analysis (0)]
62.  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. Med Hypotheses. 1994;42:257-260.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 150]  [Cited by in F6Publishing: 146]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
63.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 194]  [Cited by in F6Publishing: 201]  [Article Influence: 7.2]  [Reference Citation Analysis (0)]
64.  Spitz MR, Shi H, Yang F, Hudmon KS, Jiang H, Chamberlain RM, Amos CI, Wan Y, Cinciripini P, Hong WK. Case-control study of the D2 dopamine receptor gene and smoking status in lung cancer patients. J Natl Cancer Inst. 1998;90:358-363.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 159]  [Cited by in F6Publishing: 167]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
65.  Bierut LJ, Rice JP, Edenberg HJ, Goate A, Foroud T, Cloninger CR, Begleiter H, Conneally PM, Crowe RR, Hesselbrock V. Family-based study of the association of the dopamine D2 receptor gene (DRD2) with habitual smoking. Am J Med Genet. 2000;90:299-302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
66.  Singleton AB, Thomson JH, Morris CM, Court JA, Lloyd S, Cholerton S. Lack of association between the dopamine D2 receptor gene allele DRD2*A1 and cigarette smoking in a United Kingdom population. Pharmacogenetics. 1998;8:125-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 45]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
67.  Johnstone EC, Yudkin P, Griffiths SE, Fuller A, Murphy M, Walton R. The dopamine D2 receptor C32806T polymorphism (DRD2 Taq1A RFLP) exhibits no association with smoking behaviour in a healthy UK population. Addict Biol. 2004;9:221-226.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 22]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
68.  Comings DE, Gade R, Wu S, Chiu C, Dietz G, Muhleman D, Saucier G, Ferry L, Rosenthal RJ, Lesieur HR. Studies of the potential role of the dopamine D1 receptor gene in addictive behaviors. Mol Psychiatry. 1997;2:44-56.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 169]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
69.  Arinami T, Ishiguro H, Onaivi ES. Polymorphisms in genes involved in neurotransmission in relation to smoking. Eur J Pharmacol. 2000;410:215-226.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 62]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
70.  Asghari V, Schoots O, van Kats S, Ohara K, Jovanovic V, Guan HC, Bunzow JR, Petronis A, Van Tol HH. Dopamine D4 receptor repeat: analysis of different native and mutant forms of the human and rat genes. Mol Pharmacol. 1994;46:364-373.  [PubMed]  [DOI]  [Cited in This Article: ]
71.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 667]  [Cited by in F6Publishing: 707]  [Article Influence: 22.1]  [Reference Citation Analysis (0)]
72.  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. J Neurochem. 1995;65:1157-1165.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 626]  [Cited by in F6Publishing: 680]  [Article Influence: 23.4]  [Reference Citation Analysis (0)]
73.  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 Epidemiol Biomarkers Prev. 1998;7:453-458.  [PubMed]  [DOI]  [Cited in This Article: ]
74.  Cravchik A, Gejman PV. Functional analysis of the human D5 dopamine receptor missense and nonsense variants: differences in dopamine binding affinities. Pharmacogenetics. 1999;9:199-206.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 55]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
75.  Vandenbergh DJ, Persico AM, Hawkins AL, Griffin CA, Li X, Jabs EW, Uhl GR. Human dopamine transporter gene (DAT1) maps to chromosome 5p15.3 and displays a VNTR. Genomics. 1992;14:1104-1106.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 506]  [Cited by in F6Publishing: 533]  [Article Influence: 16.7]  [Reference Citation Analysis (0)]
76.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 430]  [Cited by in F6Publishing: 444]  [Article Influence: 18.5]  [Reference Citation Analysis (0)]
77.  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 J. 2004;4:102-109.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 49]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
78.  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 Psychol. 1999;18:14-20.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 211]  [Cited by in F6Publishing: 213]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
79.  Sabol SZ, Nelson ML, Fisher C, Gunzerath L, Brody CL, Hu S, Sirota LA, Marcus SE, Greenberg BD, Lucas FR 4th. A genetic association for cigarette smoking behavior. Health Psychol. 1999;18:7-13.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 184]  [Cited by in F6Publishing: 188]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
80.  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 Tob Res. 2002;4:333-340.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 70]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
81.  Michelhaugh SK, Fiskerstrand C, Lovejoy E, Bannon MJ, Quinn JP. The dopamine transporter gene (SLC6A3) variable number of tandem repeats domain enhances transcription in dopamine neurons. J Neurochem. 2001;79:1033-1038.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 127]  [Cited by in F6Publishing: 131]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
82.  Miller GM, Madras BK. Polymorphisms in the 3'-untranslated region of human and monkey dopamine transporter genes affect reporter gene expression. Mol Psychiatry. 2002;7:44-55.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 165]  [Cited by in F6Publishing: 170]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
83.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 101]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
84.  Spielman RS, Weinshilboum RM. Genetics of red cell COMT activity: analysis of thermal stability and family data. Am J Med Genet. 1981;10:279-290.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 91]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
85.  Vandenbergh DJ, Rodriguez LA, Miller IT, Uhl GR, Lachman HM. High-activity catechol-O-methyltransferase allele is more prevalent in polysubstance abusers. Am J Med Genet. 1997;74:439-442.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
86.  David SP, Johnstone E, Griffiths SE, Murphy M, Yudkin P, Mant D, Walton R. No association between functional catechol O-methyl transferase 1947A > G polymorphism and smoking initiation, persistent smoking or smoking cessation. Pharmacogenetics. 2002;12:265-268.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 36]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
87.  Redden DT, Shields PG, Epstein L, Wileyto EP, Zakharkin SO, Allison DB, Lerman C. Catechol-O-methyl-transferase functional polymorphism and nicotine dependence: an evaluation of nonreplicated results. Cancer Epidemiol Biomarkers Prev. 2005;14:1384-1389.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 19]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
88.  Colilla S, Lerman C, Shields PG, Jepson C, Rukstalis M, Berlin J, DeMichele A, Bunin G, Strom BL, Rebbeck TR. Association of catechol-O-methyltransferase with smoking cessation in two independent studies of women. Pharmacogenet Genomics. 2005;15:393-398.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 58]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
89.  Beuten J, Payne TJ, Ma JZ, Li MD. Significant association of catechol-O-methyltransferase (COMT) haplotypes with nicotine dependence in male and female smokers of two ethnic populations. Neuropsychopharmacology. 2006;31:675-684.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 111]  [Cited by in F6Publishing: 106]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
90.  Mihailescu S, Palomero-Rivero M, Meade-Huerta P, Maza-Flores A, Drucker-Colín R. Effects of nicotine and mecamylamine on rat dorsal raphe neurons. Eur J Pharmacol. 1998;360:31-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 92]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
91.  Lesch KP, Bengel D, Heils A, Sabol SZ, Greenberg BD, Petri S, Benjamin J, Müller 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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3630]  [Cited by in F6Publishing: 3400]  [Article Influence: 121.4]  [Reference Citation Analysis (0)]
92.  Hu S, Brody CL, Fisher C, Gunzerath L, Nelson ML, Sabol SZ, Sirota LA, Marcus SE, Greenberg BD, Murphy DL. Interaction between the serotonin transporter gene and neuroticism in cigarette smoking behavior. Mol Psychiatry. 2000;5:181-188.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 85]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
93.  Heils A, Teufel A, Petri S, Stöber G, Riederer P, Bengel D, Lesch KP. Allelic variation of human serotonin transporter gene expression. J Neurochem. 1996;66:2621-2624.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1523]  [Cited by in F6Publishing: 1501]  [Article Influence: 53.6]  [Reference Citation Analysis (0)]
94.  Batra V, Patkar AA, Berrettini WH, Weinstein SP, Leone FT. The genetic determinants of smoking. Chest. 2003;123:1730-1739.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 116]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
95.  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. Mol Psychiatry. 2000;5:189-192.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 99]  [Cited by in F6Publishing: 102]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
96.  Ishikawa H, Ohtsuki T, Ishiguro H, Yamakawa-Kobayashi K, Endo K, Lin YL, Yanagi H, Tsuchiya S, Kawata K, Hamaguchi H. Association between serotonin transporter gene polymorphism and smoking among Japanese males. Cancer Epidemiol Biomarkers Prev. 1999;8:831-833.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  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 Epidemiol Biomarkers Prev. 1998;7:253-255.  [PubMed]  [DOI]  [Cited in This Article: ]