Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-04-30T22:27:54.067Z Has data issue: false hasContentIssue false

Effects of Genetic Variants in the Nicotine Metabolism Pathway on Smoking Cessation

Published online by Cambridge University Press:  01 January 2024

Huijie Li
Affiliation:
Shandong Provincial Hospital Affiliated to Shandong First Medical University, Jinan, China Department of Epidemiology, Shandong University, Jinan, China
Qiang Wang
Affiliation:
Department of Epidemiology, Weifang Medical University, Weifang, China
Suyun Li
Affiliation:
Department of Epidemiology and Health Statistics, Qingdao University, Qingdao, China
Chongqi Jia*
Affiliation:
Department of Epidemiology, Shandong University, Jinan, China
*
Correspondence should be addressed to Chongqi Jia; cjl@sph.com.cn
Rights & Permissions [Opens in a new window]

Abstract

Background. We aimed to investigate the associations of various genetic variants in the nicotine metabolism pathway with smoking cessation (SC) in the Chinese Han population. Method. A case-control study was conducted where 363 successful smoking quitters were referred to as cases, and 345 failed smoking quitters were referred to as controls. A total of 42 genetic variants in 10 genes were selectedand genotyped. The weighted gene score was applied to analyze the whole gene effect. Logistic regression was used to explore associations of each genetic variant and gene score with smoking cessation. Results. Our study found that the variants CYP2A64, rs11726322, rs12233719, and rs3100 were associated with a higher probability of quitting smoking, while rs3760657 was associated with a lower probability of quitting smoking. Moreover, the gene scores of CYP2D6, FMO3, UGT2B10, UGT1A9, UGT2B7, and UGT2B15 were shown to exert a positive effect, while the gene score of CYP2B6 was detected to exert a negative effect on successful smoking cessation. Conclusion. This study revealed that genetic variants in the nicotine metabolic pathway were associated with smoking cessation in the Chinese Han population.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © 2022 Huijie Li et al.

1. Introduction

The usage of tobacco remains the leading cause of preventable death worldwide [1]. It is also associated with an increased risk of various diseases, including respiratory problems, cardiovascular disorders, and cancers [Reference Benowitz2Reference van Osch, Jochems, van Schooten, Bryan and Zeegers4]. In humans, considerable individual variation is observed in the rate of nicotine metabolism, which contributes to differences in various smoking behaviors [Reference von Weymarn, Retzlaff and Murphy5Reference Kyerematen, Morgan, Chattopadhyay, deBethizy and Vesell7]. Compared to the slow metabolizers, a larger amount of cigarette consumption is associated with rapid metabolism of nicotine [Reference Pomerleau, Benowitz, Jacob and Pomerleau8Reference Wassenaar, Dong, Wei, Amos, Spitz and Tyndale11], which leads to a higher degree of nicotine dependence [Reference Schnoll, George, Hawk, Cinciripini, Wileyto and Tyndale10, Reference Audrain-McGovern, Koudsi, Rodriguez, Wileyto, Shields and Tyndale12, Reference Sofuoglu, Herman, Nadim and Jatlow13], increased risk of lung cancer [Reference Wassenaar, Dong, Wei, Amos, Spitz and Tyndale11, Reference Tamaki, Arai and Sugimura14, Reference Yuan, Nelson and Butler15], and a lower probability of success in quitting smoking [Reference Chenoweth, O'Loughlin, Sylvestre and Tyndale16Reference Vaz, Coleman, Cooper, Aveyard and Leonardi-Bee19].

The metabolism of nicotine involves multiple polymorphic catalytic enzymes, whose genetic variabilities were reported to influence nicotine metabolism [Reference Lessov-Schlaggar, Benowitz, Jacob and Swan20Reference Murphy24]. Considering the importance of nicotine metabolism in SC, we wondered whether genetic variants in the nicotine metabolic pathway contributed to the individual variations in SC. Several studies [Reference Chenoweth, O'Loughlin, Sylvestre and Tyndale16, Reference Xu, Goodz, Sellers and Tyndale25] have reported the relationship between gene variants in genes encoding phase I drug metabolic enzymes, especially CYP2A6, and SC. However, to our knowledge, data on the relationship between genetic variants encoding phase II drug metabolic enzymes and smoking behaviors have been lacking so far.

Hence, our study aims to investigate the associations between the genetic variants in the whole nicotine metabolic pathway and SC in the Chinese Han population.

2. Methods

2.1. Related Definitions

According to the WHO data [26], some of the definitions are as follows: smokers: who have/had smoked 100 or more cigarettes (or the equivalent amount of tobacco) during their lifetime; successful smoking quitters: those who comply with the criterion of smokers but have not smoked at all continuously for two years or more during the survey period [Reference Qian, Cai, Gao, Tang, Xu and Critchley27]; failed smoking quitters: those who comply with the criterion of smokers and had quit smoking but relapsed or did not smoke at all continuously for less than two years during the survey period.

2.2. Subjects

This community-based study was conducted among people of 17 villages belonging to three counties (Pingyin, Ju’nan, and Liangshan) of Shandong, China, between April and May 2013. Male participants aged 18 years or more who spontaneously quit smoking anytime were interviewed face to face by well-trained investigators. They were asked to complete a questionnaire designed based on the Global Adult Tobacco Survey (GATS) Core Questionnaire with optional questions. All subjects recruited were of Han Chinese ethnicity. A total of 708 blood samples, including 363 for successful smoking quitters and 345 for failed smoking quitters, were collected successfully and genotyped. Baseline characteristics of all the subjects are provided in Table 1.

TABLE 1: The characteristics of the subjects.

Values are expressed as mean ± SD or frequency (%). Age group: young = age <45, middle = 45≤ age <65, old = age ≥65. Education: low = below the primary school, middle = primary and middle school, high = high school or more. Pack-year = CPD/20 × number of years smoked. CPD: cigarettes per day.

In this study, a community-based case-control analysis was performed, where successful smoking quitters were referred to as cases, and failed smoking quitters were referred to as controls. This study was approved by the Ethics Review Committee of Shandong University, where all subjects provided informed consent.

2.3. Selection of the Genetic Variants

The genetic variants in this study were selected based on the following criteria:

(1) Identification of TagSNPs by Haploview software 4.2 based on the Chinese Han Beijing (CHB) population data of HapMap (HapMap Data Rel 27 Phase II + III, Feb 09, on NCBI B36 assembly, dbSNP b126). TagSNPs were selected if pairwise r 2 > 0.8 and minor allele frequency (MAF) > 0.05. (2) Functional relevance and commonality [MAF> 0.05, based on the CHB population data of the dbSNP database and 1000 Genomes (https://browser.1000.genomes.org/index.html)]. (3) Genetic variants showing a significant relationship with smoking behaviors in previous studies. Overall, a total of 41 SNPs and one gene deletion allele among 10 genes were selected. Table 2 presents detailed information on these 42 genetic variants.

TABLE 2: Genetic variants investigated in this study.

MAF: minor allele frequency; HWE: Hardy-Weinberg Equilibrium. aThe HWE P value shows the Hardy-Weinberg Equilibrium test in control subjects.

2.4. SNPs Genotyping

Blood samples were used to extract genomic DNA by a DNA isolation kit (TIANGEN, China). Next, the SNP genotypes were determined using the Matrix-Assisted Laser Desorption/Ionization Time of Flight Mass Spectrometry (MALDI-TOF–MS) of the MassARRAY system (Sequenom Inc., San Diego, CA, USA). The Assay Design 3.1 software (Sequenom Inc., San Diego, CA, USA) was used to design PCR primers. Genotyping was performed by the Bio Miao Biological Technology (Beijing) Co., Ltd., without any knowledge about the case or control status.

2.5. CYP2A6 Genotyping

The presence of CYP2A6 ∗4 (whole gene deletion) was detected by a two-step allelic-specific PCR assay [Reference Tamaki, Arai and Sugimura14, Reference Oscarson, McLellan and Gullsten28]. The first PCR reaction (PCR I) was performed using the primers 2Aex7F and 2A6R1 (Table 3). All primers were all synthesized by Sangon Biotech Co., Ltd. (Shanghai, China). The total reaction mixture (50 μl) included 4 μl of genomic DNA, 1.5 μl of each primer (10 μM), 25 μl of 2 × Taq PCR Master Mix (BBI), and 18 μl of ddH2O. The PCR program was as follows: initial denaturation step at 94°C for 4 min, followed by 35 cycles of denaturation at 94°C for 30 s, annealing at 56°C for 30 s, extension at 72°C for 2.5 min, and then a final extension at 72°C for 10 min.

TABLE 3: Primers used for amplification of CYP2A6 4 (gene deletion).

F forward primer, R reverse primer.

Next, the allele-specific PCR reaction was performed (PCR II), which involved a PCR mixture containing 2 μL of PCR I product, 2 μL of primer 2A6ex8F (10 μM) or primer 2A7ex8F (10 μM), 2 μL of primer 2A6R2, 25 μL of 2 × Taq PCR Master Mix (BBI), and 19 μL of ddH2O, making up a total volume of 50 μL. The amplification was performed as per the following program: initial denaturation at 94°C for 4 min, followed by 30 cycles of denaturation at 94°C for 30 s, annealing at 52°C for 45 s, extension at 72°C for 1.5 min, and then a final extension at 72°C for 10 min. The amplified PCR products were analyzed on a 1.5% agarose gel (BBI) and stained with ethidium bromide.

As per the definition by Tamaki et al. [Reference Tamaki, Arai and Sugimura14], the CYP2A-specific 1,181-bp product amplified using the 2A6ex8F/2A6R2 primer pair alone indicated the presence of wildtype CYP2A6 (CYP2A6 non 4/non 4). Similarly, the product amplified using the primer pair 2A7ex8F/2A6R2 alone indicated a CYP2A6 deletion (CYP2A64/∗4∗). However, if one individual sample showed product amplification in both reactions, it indicated heterozygosity (CYP2A6 non 4/ 4).

2.6. Construction of the Weighted Gene Score

To increase the power of the test and explore the effect of the whole genes on SC, the weighted gene score of each gene was calculated as the sum of each genotype multiplied by its weight, which was then divided by the sum of the weights [Reference Burgess and Thompson29, Reference Lin, Song and Lim30]. The risk score was calculated as per the following equation:

(1) weighted gene score = w 1 × SNP 1 + w 2 × SNP 2 + + w k × SNP k w 1 + w 2 + + w k ,

where SNP i has a value of 0, 1, or 2 according to the number of minor alleles for SNP; W k is the weight of SNP calculated using logistic regression, where smoking cessation is the dependent variable, and each variant is the independent variable; K is the number of SNPs in each gene [Reference Burgess and Thompson29, Reference Palmer, Lawlor and Harbord31]. Since we analyzed only one genetic variant in the CYP2A6 gene (CYP2A6 4), we did not calculate the gene score of CYP2A6.

2.7. Statistical Analysis

To describe the demographic characteristics of participants, the frequency and percentage of categorical variables were calculated along with the mean and standard deviation of metric variables. Additionally, Pearson’s χ 2 test was used to compare the differences between the categorical variables of the case and control groups, while the Hardy-Weinberg Equilibrium (HWE) was used to compare differences among the control groups. If the variance between the groups was homogeneous, a one-way analysis of variance was used to test the differences in means of metric variables between the groups. Otherwise, the Kruskal-Wallis equality-of-populations rank test was used.

The associations between the genetic variants and SC were evaluated using the odds ratios (ORs) and 95% confidence intervals (CIs), which were first calculated using univariate logistic regression, and then by multiple logistic regression analysis with adjustments being done for age, occupation, education level, marital status, age of smoking onset and pack-year. The association studies were analyzed among four genetic models, including codominant, additive, dominant, and recessive models. Furthermore, the Akaike information criterion (AIC) was utilized to determine the best genetic model for each SNP.

The complete gene effect on SC was analyzed using the logistic regression, where odds ratios (ORs) and multivariate-adjusted ORs (adjusted for age, occupation, education level, marital status, age of smoking onset, and pack-year) were calculated for each gene score and smoking cessation.

To estimate the value of power, we used a range of minor allele frequencies (MAF) in genetic variants, including 0.1, 0.2, 0.3, 0.4, and 0.5. We assumed an odds ratio (OR) of 1.5, the population risk of 11.7% [Reference Yang, Jason and Yang32], and the type I error rate (α) of 0.05. Our results showed that a total sample size of 708 subjects, including 363 cases and 345 controls, provided a power of more than 80% for 0.2, 0.3, 0.4, and 0.5 of MAF and more than 60% power for 0.1 of MAF under the additive inheritance model.

While power calculation was performed using the program QUANTO 1.2.4, the other statistical analyses were carried out using the STATA/SE version 15.1 (Stata Corporation, College Station, TX, USA). All reported probabilities (P value) were two-sided, and a P value less than 0.05 was considered statistically significant.

3. Results

3.1. Participant Characteristics

The demographic characteristics of both the groups, including 363 successful smoking quitters (cases) and 345 failed smoking quitters (controls), are summarized in Table 1. Compared to the failed smoking quitters, the successful smoking quitters showed significantly higher age and shorter smoking duration (P < 0.001). Moreover, the distribution of the marital status between the case and control groups showed significant differences (P < 0.05).

3.2. Associations of Genetic Variants in the Nicotine Metabolic Pathway with Smoking Cessation

Table 2 lists the minor allele frequency (MAF) for each genetic variant in all the subjects, along with the P value of the control subjects from the HWE test. Figure 1 presents the genotyping assay results obtained from the two-step allelic-specific PCR assay.

FIGURE 1: (a) First-step PCR results (1,961-bp) (b) second-step PCR results (1,181-bp). Lanes 1 and 2 are wildtype CYP2A6 (non 4/non 4); lanes 3 and 4 are heterozygous deletions (non 4/4); lanes 5 and 6 are the CYP2A6 deletion genotype (4/4).

The frequencies of the genotype and the ORs (95% CI) for the codominant, additive, dominant, and recessive models are presented in Supplementary Table 1. The univariate analyses showed that CYP2A6 4 was correlated with an increased possibility of SC in the codominant [∗4/4 vs non 4/non 4, 2.016 (1.270–3.201)], additive [non 4/non 4 vs non 4/4 vs 4/4, 1.353 (1.095–1.672)], dominant [non 4/4 + 4/4 vs non 4/non 4, 1.442 (1.041–1.998) ], and recessive models [4/4 vs non 4/non 4 + non 4/4, 1.976 (1.252–3.117)]. Besides, rs11726322 of the UGT2B10 gene, rs12233719 of the UGT2B7 gene, and rs3100 of the UGT2B15 gene were also found to exert a protective effect on SC among different genetic models. However, a negative association was detected between rs3760657 of CYP2B6 gene and SC in the codominant [AG vs AA, 0.669 (0.488–0.918)], additive [AA vs AG vs GG, 0.773 (0.599–0.997)], and dominant [AG + GG vs AA, 0.691 (0.599–0.936)] genetic models.

The multiple logistic regression analysis, adjusted for the potential confounders, showed significant associations of CYP2A6 4, rs11726322, rs12233719, and rs3100 with the increased possibility of SC, as well as rs3760657 with the reduced possibility of SC. Based on the Akaike information criterion (AIC), the corresponding optimal recessive model was CYP2A6 4, rs11726322, and rs12233719, while the optimal dominant model was rs3760657 and the additive model was rs3100. However, no significant relationship was found between other genetic variants and SC among the four genetic models.

3.3. Association between the Gene Scores and Smoking Cessation

The results of the association between the gene scores and SC are presented in Table 4. Positive significant associations were observed between the gene scores of CYP2D6, FMO3, UGT2B10, UGT1A9, UGT2B7, and UGT2B15 and SC. However, a negative association was found between CYP2B6 gene score and SC. The results remained significant even after adjusting for age, occupation, education level, marital status, and age of onset of smoking. Moreover, the total gene score for all the selected genetic variants in the nicotine metabolic pathway showed a positive association with SC both before (OR = 3.311, 95% CI: 2.342–4.679) and after adjusting the above-mentioned potential confounders (OR = 3.411, 95% CI: 2.382–4.884).

TABLE 4: Association between the gene scores and smoking cessation.

Values are expressed as mean ± SD. a indicates adjustment for age, occupation, education level, marital status, age of smoking onset, and pack-year.

4. Discussion

Smoking behavior is a complex trait with a multigenic etiology, which is influenced by both environmental and genetic factors. A genetic influence with heritability has been identified in smoking cessation (SC), which is estimated at 50–58% [Reference Broms, Silventoinen, Madden, Heath and Kaprio33Reference Xian, Scherrer and Madden36]. In this study, we investigated the associations between various genetic variants in the nicotine metabolic pathway and SC in the Chinese Han population. Our results showed that CYP2A6 4, rs11726322, rs12233719, and rs3100 were associated with a higher probability while rs3760657 was associated with a lower probability of quitting smoking. Moreover, the gene scores of CYP2D6, FMO3, UGT2B10, UGT1A9, UGT2B7, and UGT2B15 were shown to exert a positive effect on successful SC, while the gene score of CYP2B6 was detected to exert a negative effect.

Nicotine is primarily metabolized by the following three pathways: cytochrome P450 (CYPs)-catalyzed C-oxidation, UDP-glucuronosyltransferases (UGTs)-catalyzed glucuronidation, and flavin-containing monooxygenase 3 (FMO3)-catalyzed N-oxidation [Reference Benowitz, Hukkanen and Jacob37]. In smokers, 70–80% of nicotine is converted to cotinine before metabolization to other metabolites [Reference Benowitz, Jacob, Fong and Gupta38]. The C-oxidation of nicotine to cotinine occurs via a two-step mechanism, where nicotine is first oxidized to the nicotine-Δ-1′ (5′)-iminium ion mediated by CYPs, which is followed by its conversion to cotinine by cytosolic aldehyde oxidase 1 (AOX1) [Reference Tricker39]. Subsequently, cotinine is further oxidized by CYPs to trans-3′-hydroxycotinine (3HC), which accounts for 27–40% of the nicotine dose recovered in urine. This is the main nicotine metabolite detected in the urine of smokers [Reference Tricker39, Reference Hukkanen, Jacob and Benowitz40]. CYP2A6 is the major CYP enzyme involved in the C-oxidation of nicotine. However, in some individuals, other CYPs, including CYP2B6 and CYP2D6, also contribute minorly [Reference Messina, Tyndale and Sellers41Reference Yamazaki, Inoue, Hashimoto and Shimada43]. The gene encoding the CYP2A6 enzyme is highly polymorphic and results in extensive interindividual variations in the CYP2A6 enzyme activity, affecting the rate of metabolism of nicotine [Reference Xu, Goodz, Sellers and Tyndale25, Reference Kitagawa, Kunugita, Kitagawa and Kawamoto44]. Nakajima et al. [Reference Nakajima, Yamagishi, Yamamoto, Yamamoto, Kuroiwa and Yokoi45] first reported that the poor metabolism of nicotine to cotinine was attributed to the whole deletion of the CYP2A6 gene (CYP2A6 4) in humans. Compared to other ethnic populations, the CYP2A6 4 was reported to have a high allelic frequency in Asians (11–24%) [Reference Schoedel, Hoffmann, Rao, Sellers and Tyndale21, Reference Oscarson, McLellan and Gullsten28, Reference Nakajima and Yokoi46, Reference Nakajima, Kwon and Tanaka47]. In this study, CYP2A6 4 was correlated with a higher probability of quitting smoking. However, the deviation from HWE in the control group for CYP2A6 4 was observed in our study; our result may be biased and needs to be further confirmed.

For in vitro C-oxidation of nicotine, CYP2B6 is responsible for an approximately 10% catalytic efficiency of the CYP2A6 enzyme [Reference Ray, Tyndale and Lerman48]. While CYP2A6 is primarily expressed in the liver, CYP2B6 is expressed at higher levels in the brain. Higher brain activity for CYP2B6 may cause altered sensitivity to centrally acting drugs, which may contribute to influencing the localized metabolism of nicotine in the brains of human smokers [Reference Miksys, Lerman, Shields, Mash and Tyndale49].

Approximately 4–7% of absorbed nicotine is excreted in the urine as nicotine N′-oxide [Reference Byrd, Chang, Greene and deBethizy50, Reference Mwenifumbo and Tyndale51], which is converted by flavin-containing monooxygenase 3 (FMO3) [Reference Nakajima and Yokoi46, Reference Cashman, Park, Yang, Wrighton, Jacob and Benowitz52]. A common decrease in the function of FMO3 alleles exerts modest effects on N-oxidation activity caused by a slight modulation of protein levels and/or function and is more likely to contribute to general population variation in FMO3 [Reference Cashman, Zhang, Leushner and Braun53]. In 2014, Chenoweth et al. [Reference Chenoweth, Zhu, Sanderson Cox, Ahluwalia, Benowitz and Tyndale54] showed that FMO3 E158K (rs2266782) modestly influenced the systemic nicotine metabolism within the CYP2A6 subgroups. However, in this study, we did not observe any significant relationship between FMO3 alleles and SC.

Nicotine, cotinine, and trans-3′-hydroxycotinine undergo further phase II detoxification reactions catalyzed by the UDP-glucuronosyltransferase (UGTs) family of enzymes via the conjugation with glucuronic acid [Reference Yamanaka, Nakajima and Katoh55]. In smokers, 3–5% of absorbed nicotine is excreted in the urine as nicotine N-glucuronide, 16–17% as cotinine N-glucuronide, as well as an additional 8–9% as trans-3′-hydroxycotinine-O-glucuronide [Reference Byrd, Chang, Greene and deBethizy50, Reference Mwenifumbo and Tyndale51]. The UDP-glucuronosyltransferases (UGTs) include a large family of conjugation enzymes, detoxifying a wide variety of both endogenous and exogenous substrates [Reference Papageorgiou and Court56]. Kuehl and Murphy [Reference Kuehl and Murphy57] reported that UGT1A4 and UGT1A9, the UGTs isoforms, were responsible for nicotine and cotinine N-glucuronidation. Although glucuronidation of trans-3′-hydroxycotinine to O-glucuronide occurs mainly via UGT2B7 and UGT2B10 [Reference Chen, Giambrone and Lazarus58], it is partly glucuronidated by UGT1A9 and UGT2B15 [Reference Nakajima, Tanaka, Kwon and Yokoi59]. A change in the gene encoding enzymes (i.e., point mutation, deletion, and gene conversion) responsible for chemical metabolism may lead to overproduction, underproduction, malfunction, or absence of the protein, finally resulting in alterations in the functioning of the enzyme [Reference Tricker39]. Many UGT gene variants were reported to influence the glucuronidation of tobacco-related compounds, including UGT1A4 [Reference Ehmer, Vogel, Schutte, Krone, Manns and Strassburg60, Reference Mori, Maruo, Iwai, Sato and Takeuchi61], UGT2B7 [Reference Wassenaar, Conti and Das62], and UGT2B10 [Reference Wassenaar, Conti and Das62, Reference Chen, Giambrone and Dluzen63].

Several studies have reported a significant impact of nicotine metabolism on the probability of success in quitting smoking [Reference Chenoweth, O'Loughlin, Sylvestre and Tyndale16Reference Vaz, Coleman, Cooper, Aveyard and Leonardi-Bee19]. Evidence suggests that smokers adapt their smoking behavior to maintain desired levels of nicotine in the body. The level of nicotine in the body is determined according to the intake of nicotine and the rate of metabolism by the liver. Based on the above-mentioned associations between genetic polymorphisms and enzymatic activity, we infer that genetic variants in the nicotine metabolic pathway may contribute to the individual variability in SC.

In our study, we not only analyzed single variants but also used gene scores to analyze the whole effect of the genes in the nicotine metabolic pathway on SC. A gene score is important to model multifactorial polygenic traits, particularly when the gene score consists of many common variants with small effects [Reference Burgess and Thompson29]. The gene score may explain a considerable proportion of variations in the risk factor, even when none of the variants can explain it individually, thus, making it a very popular method for genetic association studies [Reference Burgess and Thompson29].

One of the strengths of our study is the exploration of the relationship between the gene variants encoding phase II drug metabolic enzymes, including UGT1A9, UGT2B10, and UGT2B7, and smoking behavior. To our knowledge, data on such a relationship is rarely available so far. Most researchers have focused on genes encoding phase I drug metabolic enzymes, especially CYP2A6. Our study reveals that rs11726322 of UGT2B10, rs12233719 of UGT2B7, and rs3100 of UGT2B15 may influence the susceptibility of SC, which has not been reported in previous publications. Additionally, we employed the method of gene score for the association analysis, which increased the power and simplicity of the test.

However, we acknowledge some limitations in our study. Firstly, since smoking behaviors have a complex etiology generated by many factors, some other uninvestigated variables may alter the results. Secondly, this was a community-based case-control study and may have some biases, including selection and recall bias, which may distort the results.

5. Conclusions

This study revealed that five single variants (CYP2A6 4, rs11726322, rs12233719, rs3100, and rs3760657) and gene scores of CYP2B6, CYP2D6, FMO3, UGT2B10, UGT1A9, UGT2B7, and UGT2B15 might influence the susceptibility of SC in the Chinese Han population. Further examinations using larger sample size and other ethnic groups may be required to confirm our findings.

Data Availability

All data generated or analyzed during this study are included within the article and its supplementary information files.

Ethical Approval

The study procedure conformed to the ethical guidelines of the Declaration of Helsinki, and approval for the study was obtained from the Ethics Review Committee of Shandong University.

Consent

A written informed consent was obtained from the patients enrolled.

Conflicts of Interest

The authors declare that there are no conflicts of interest regarding the publication of this paper.

Acknowledgments

The authors thank the local Health Bureau, local Center for Disease Control and Prevention, other relevant governments and persons, the investigators, and respondents for their support of this research. This study was funded by a grant from the National Natural Science Foundation of China (no. 81273150) and a grant from the Natural Science Foundation of Shandong Province (ZR2020QH303).

Supplementary Materials

Supplementary Table 1: associations of genetic variants in nicotine metabolism pathway with smoking cessation under different genetic models. (Supplementary Materials)

References

World Health Organization, Who Report on the Global Tobacco Epidemic 2019: Offer Help to Quit Tobacco Use, World Health Organization, Geneva, Switzerland, 2019.Google Scholar
Benowitz, N. L., “Nicotine addiction,New England Journal of Medicine, vol. 362, no. 24, pp. 22952303, 2010.CrossRefGoogle ScholarPubMed
Mons, U., Muezzinler, A., Gellert, C. et al., “Impact of smoking and smoking cessation on cardiovascular events and mortality among older adults: meta-analysis of individual participant data from prospective cohort studies of the chances consortium,BMJ, vol. 350, Article ID h1551, 2015.CrossRefGoogle ScholarPubMed
van Osch, F. H., Jochems, S. H., van Schooten, F. J., Bryan, R. T., and Zeegers, M. P., “Quantified relations between exposure to tobacco smoking and bladder cancer risk: a meta-analysis of 89 observational studies,International Journal of Epidemiology, vol. 45, no. 3, pp. 857870, 2016.CrossRefGoogle Scholar
von Weymarn, L. B., Retzlaff, C., and Murphy, S. E., “CYP2A6- and CYP2A13-catalyzed metabolism of the nicotine d5′(1′) Iminium ion,Journal of Pharmacology and Experimental Therapeutics, vol. 343, no. 2, pp. 307315, 2012.CrossRefGoogle Scholar
Haley, N. J., Sepkovic, D. W., and Hoffmann, D., “Elimination of cotinine from body fluids: disposition in smokers and nonsmokers,American Journal of Public Health, vol. 79, no. 8, pp. 10461048, 1989.CrossRefGoogle ScholarPubMed
Kyerematen, G. A., Morgan, M. L., Chattopadhyay, B., deBethizy, J. D., and Vesell, E. S., “Disposition of nicotine and eight metabolites in smokers and nonsmokers: Identification in smokers of two metabolites that are longer lived than cotinine,Clinical Pharmacology & Therapeutics, vol. 48, no. 6, pp. 641651, 1990.CrossRefGoogle ScholarPubMed
Pomerleau, C., Benowitz, N., Jacob, P., and Pomerleau, O., “Nicotine metabolite ratio as a predictor of cigarette consumption,Nicotine & Tobacco Research, vol. 5, no. 5, pp. 621624, 2003.CrossRefGoogle Scholar
Chenoweth, M. J., Novalen, M., Hawk, L. W. et al., “Known and novel sources of variability in the nicotine metabolite ratio in a large sample of treatment-seeking smokers,Cancer Epidemiology, Biomarkers & Prevention, vol. 23, no. 9, pp. 17731782, 2014.CrossRefGoogle Scholar
Schnoll, R. A., George, T. P., Hawk, L., Cinciripini, P., Wileyto, P., and Tyndale, R. F., “The relationship between the nicotine metabolite ratio and three self-report measures of nicotine dependence across sex and race,Psychopharmacology (Berl), vol. 231, no. 12, pp. 25152523, 2014.CrossRefGoogle ScholarPubMed
Wassenaar, C. A., Dong, Q., Wei, Q., Amos, C. I., Spitz, M. R., and Tyndale, R. F., “Relationship between cyp2a6 and chrna5-chrna3-chrnb4 variation and smoking behaviors and lung cancer risk,Journal of the National Cancer Institute: Journal of the National Cancer Institute, vol. 103, no. 17, pp. 13421346, 2011.CrossRefGoogle ScholarPubMed
Audrain-McGovern, J., Koudsi, N. A., Rodriguez, D., Wileyto, E. P., Shields, P. G., and Tyndale, R. F., “The role of cyp2a6 in the emergence of nicotine dependence in adolescents,Pediatrics, vol. 119, no. 1, pp. e26474, 2007.CrossRefGoogle ScholarPubMed
Sofuoglu, M., Herman, A. I., Nadim, H., and Jatlow, P., “Rapid nicotine clearance is associated with greater reward and heart rate increases from intravenous nicotine,Neuropsychopharmacology, vol. 37, no. 6, pp. 15091516, 2012.CrossRefGoogle ScholarPubMed
Tamaki, Y., Arai, T., Sugimura, H. et al., “Association between cancer risk and drug-metabolizing enzyme gene (cyp2a6, cyp2a13, cyp4b1, sult1a1, gstm1, and gstt1) polymorphisms in cases of lung cancer in Japan,Drug Metabolism and Pharmacokinetics, vol. 26, no. 5, pp. 516522, 2011.CrossRefGoogle ScholarPubMed
Yuan, J. M., Nelson, H. H., Butler, L. M. et al., “Genetic determinants of cytochrome p450 2a6 activity and biomarkers of tobacco smoke exposure in relation to risk of lung cancer development in the shanghai cohort study,International Journal of Cancer, vol. 138, no. 9, pp. 21612171, 2016.CrossRefGoogle ScholarPubMed
Chenoweth, M. J., O'Loughlin, J., Sylvestre, M. P., and Tyndale, R. F., “Cyp2a6 slow nicotine metabolism is associated with increased quitting by adolescent smokers,Pharmacogenetics and Genomics, vol. 23, no. 4, pp. 232235, 2013.CrossRefGoogle ScholarPubMed
Chenoweth, M. J., Schnoll, R. A., Novalen, M. et al., “The nicotine metabolite ratio is associated with early smoking abstinence even after controlling for factors that influence the nicotine metabolite ratio,Nicotine & Tobacco Research, vol. 18, no. 4, pp. 491495, 2016.CrossRefGoogle ScholarPubMed
Schnoll, R. A., Patterson, F., Wileyto, E. P., Tyndale, R. F., Benowitz, N., and Lerman, C., “Nicotine metabolic rate predicts successful smoking cessation with transdermal nicotine: a validation study,Pharmacology Biochemistry and Behavior, vol. 92, no. 1, pp. 611, 2009.CrossRefGoogle ScholarPubMed
Vaz, L. R., Coleman, T., Cooper, S., Aveyard, P., and Leonardi-Bee, J., “The nicotine metabolite ratio in pregnancy measured bytrans-3′-hydroxycotinine to cotinine ratio: characteristics and relationship with smoking cessation: table 1,Nicotine & Tobacco Research, vol. 17, no. 11, pp. 13181323, 2015.CrossRefGoogle Scholar
Lessov-Schlaggar, C. N., Benowitz, N. L., Jacob, P., and Swan, G. E., “Genetic influences on individual differences in nicotine glucuronidation,Twin Research and Human Genetics, vol. 12, no. 5, pp. 507513, 2009.CrossRefGoogle ScholarPubMed
Schoedel, K. A., Hoffmann, E. B., Rao, Y., Sellers, E. M., and Tyndale, R. F., “Ethnic variation in cyp2a6 and association of genetically slow nicotine metabolism and smoking in adult caucasians,Pharmacogenetics, vol. 14, no. 9, pp. 615626, 2004.CrossRefGoogle ScholarPubMed
Yoshida, R., Nakajima, M., Watanabe, Y., Kwon, J. T., and Yokoi, T., “Genetic polymorphisms in human cyp2a6 gene causing impaired nicotine metabolism,British Journal of Clinical Pharmacology, vol. 54, no. 5, pp. 511517, 2002.CrossRefGoogle ScholarPubMed
Bloom, A. J., von Weymarn, L. B., Martinez, M., Bierut, L. J., Goate, A., and Murphy, S. E., “The contribution of common ugt2b10 and cyp2a6 alleles to variation in nicotine glucuronidation among european americans,Pharmacogenetics and Genomics, vol. 23, no. 12, pp. 706716, 2013.CrossRefGoogle ScholarPubMed
Murphy, S. E., “Nicotine metabolism and smoking: ethnic differences in the role of p450 2a6,Chemical Research in Toxicology, vol. 30, no. 1, pp. 410419, 2017.CrossRefGoogle ScholarPubMed
Xu, C., Goodz, S., Sellers, E. M., and Tyndale, R. F., “Cyp2a6 genetic variation and potential consequences,Advanced Drug Delivery Reviews, vol. 54, no. 10, pp. 12451256, 2002.CrossRefGoogle ScholarPubMed
World Health Organization, Guidelines for Controlling and Monitoring the Tobacco Epidemic, World Health Organization, Geneva, Switzerland, 1998.Google Scholar
Qian, J., Cai, M., Gao, J., Tang, S., Xu, L., and Critchley, J. A., “Trends in smoking and quitting in China from 1993 to 2003: national health service survey data,Bulletin of the World Health Organization, vol. 88, no. 10, pp. 769776, 2010.CrossRefGoogle Scholar
Oscarson, M., McLellan, R. A., Gullsten, H. et al., “Characterisation and pcr-based detection of a cyp2a6 gene deletion found at a high frequency in a Chinese population,FEBS Letters, vol. 448, no. 1, pp. 105110, 1999.CrossRefGoogle Scholar
Burgess, S. and Thompson, S. G., “Use of allele scores as instrumental variables for mendelian randomization,International Journal of Epidemiology, vol. 42, no. 4, pp. 11341144, 2013.CrossRefGoogle ScholarPubMed
Lin, X., Song, K., Lim, N. et al., “Risk prediction of prevalent diabetes in a swiss population using a weighted genetic score--the colaus study,Diabetologia, vol. 52, no. 4, pp. 600608, 2009.CrossRefGoogle Scholar
Palmer, T. M., Lawlor, D. A., Harbord, R. M. et al., “Using multiple genetic variants as instrumental variables for modifiable risk factors,Statistical Methods in Medical Research, vol. 21, no. 3, pp. 223242, 2012.CrossRefGoogle ScholarPubMed
Yang, G., Jason, H., and Yang, Y., Global Adult Tobacco Survey (Gats) china 2010 Country Report, Chinese Three Gorge Press, Beijing, China, 2011.Google Scholar
Broms, U., Silventoinen, K., Madden, P. A. F., Heath, A. C., and Kaprio, J., “Genetic architecture of smoking behavior: a study of Finnish adult twins,Twin Research and Human Genetics, vol. 9, no. 1, pp. 6472, 2006.CrossRefGoogle ScholarPubMed
Hardie, T. L., Moss, H. B., and Lynch, K. G., “Genetic correlations between smoking initiation and smoking behaviors in a twin sample,Addictive Behaviors, vol. 31, no. 11, pp. 20302037, 2006.CrossRefGoogle Scholar
Xian, H., Scherrer, J. F., Madden, P. A. F. et al., “Latent class typology of nicotine withdrawal: genetic contributions and association with failed smoking cessation and psychiatric disorders,Psychological Medicine, vol. 35, no. 3, pp. 409419, 2005.CrossRefGoogle ScholarPubMed
Xian, H., Scherrer, J. F., Madden, P. A. et al., “The heritability of failed smoking cessation and nicotine withdrawal in twins who smoked and attempted to quit,Nicotine & Tobacco Research, vol. 5, no. 2, pp. 245254, 2003.CrossRefGoogle ScholarPubMed
Benowitz, N. L., Hukkanen, J., and Jacob, P., “Nicotine chemistry, metabolism, kinetics and biomarkers,Handbook of Experimental Pharmacology, vol. 192, pp. 2960, 2009.CrossRefGoogle Scholar
Benowitz, N. L., Jacob, P., Fong, I., and Gupta, S., “Nicotine metabolic profile in man: comparison of cigarette smoking and transdermal nicotine,Journal of Pharmacology and Experimental Therapeutics, vol. 268, no. 1, pp. 296303, 1994.Google ScholarPubMed
Tricker, A. R., “Nicotine metabolism, human drug metabolism polymorphisms, and smoking behaviour,Toxicology, vol. 183, no. 1-3, pp. 151173, 2003.CrossRefGoogle ScholarPubMed
Hukkanen, J., Jacob, P., and Benowitz, N. L., “Metabolism and disposition kinetics of nicotine,Pharmacological Reviews, vol. 57, no. 1, pp. 79115, 2005.CrossRefGoogle ScholarPubMed
Messina, E. S., Tyndale, R. F., and Sellers, E. M., “A major role for cyp2a6 in nicotine c-oxidation by human liver microsomes,Journal of Pharmacology and Experimental Therapeutics, vol. 282, no. 3, pp. 16081614, 1997.Google Scholar
Caporaso, N. E., Lerman, C., Audrain, J. et al., “Nicotine metabolism and cyp2d6 phenotype in smokers,Cancer Epidemiology, Biomarkers & Prevention, vol. 10, no. 3, pp. 261263, 2001.Google ScholarPubMed
Yamazaki, H., Inoue, K., Hashimoto, M., and Shimada, T., “Roles of cyp2a6 and cyp2b6 in nicotine c-oxidation by human liver microsomes,Archives of Toxicology, vol. 73, no. 2, pp. 6570, 1999.CrossRefGoogle ScholarPubMed
Kitagawa, K., Kunugita, N., Kitagawa, M., and Kawamoto, T., “Cyp2a66, a novel polymorphism in cytochrome p450 2a6, has a single amino acid substitution (r128q) that inactivates enzymatic activity,Journal of Biological Chemistry, vol. 276, no. 21, pp. 1783017835, 2001.CrossRefGoogle Scholar
Nakajima, M., Yamagishi, S., Yamamoto, H., Yamamoto, T., Kuroiwa, Y., and Yokoi, T., “Deficient cotinine formation from nicotine is attributed to the whole deletion of the cyp2a6 gene in humans,Clinical Pharmacology & Therapeutics, vol. 67, no. 1, pp. 5769, 2000.CrossRefGoogle Scholar
Nakajima, M. and Yokoi, T., “Interindividual variability in nicotine metabolism: C-oxidation and glucuronidation,Drug Metabolism and Pharmacokinetics, vol. 20, no. 4, pp. 227235, 2005.CrossRefGoogle ScholarPubMed
Nakajima, M., Kwon, J. T., Tanaka, N. et al., “Relationship between interindividual differences in nicotine metabolism and cyp2a6 genetic polymorphism in humans,Clinical Pharmacology & Therapeutics, vol. 69, no. 1, pp. 7278, 2001.CrossRefGoogle ScholarPubMed
Ray, R., Tyndale, R. F., and Lerman, C., “Nicotine dependence pharmacogenetics: role of genetic variation in nicotine-metabolizing enzymes,Journal of Neurogenetics, vol. 23, no. 3, pp. 252261, 2009.CrossRefGoogle ScholarPubMed
Miksys, S., Lerman, C., Shields, P. G., Mash, D. C., and Tyndale, R. T., “Smoking, alcoholism and genetic polymorphisms alter cyp2b6 levels in human brain,Neuropharmacology, vol. 45, no. 1, pp. 122132, 2003.CrossRefGoogle ScholarPubMed
Byrd, G. D., Chang, K. M., Greene, J. M., and deBethizy, J. D., “Evidence for urinary excretion of glucuronide conjugates of nicotine, cotinine, and trans-3′-hydroxycotinine in smokers,Drug Metabolism & Disposition, vol. 20, no. 2, pp. 192197, 1992.Google ScholarPubMed
Mwenifumbo, J. C. and Tyndale, R. F., “Molecular genetics of nicotine metabolism,Handbook of Experimental Pharmacology, vol. 192, pp. 235259, 2009.CrossRefGoogle Scholar
Cashman, J. R., Park, S. B., Yang, Z. C., Wrighton, S. A., Jacob, P., and Benowitz, N. L., “Metabolism of nicotine by human liver microsomes: stereoselective formation of trans-nicotine n′-oxide,Chemical Research in Toxicology, vol. 5, no. 5, pp. 639646, 1992.CrossRefGoogle ScholarPubMed
Cashman, J. R., Zhang, J., Leushner, J., and Braun, A., “Population distribution of human flavin-containing monooxygenase form 3: gene polymorphisms,Drug Metabolism & Disposition, vol. 29, no. 12, pp. 16291637, 2001.Google ScholarPubMed
Chenoweth, M. J., Zhu, A. Z., Sanderson Cox, L., Ahluwalia, J. S., Benowitz, N. L., and Tyndale, R. F., “Variation in p450 oxidoreductase (por) a503v and flavin-containing monooxygenase (fmo)-3 e158k is associated with minor alterations in nicotine metabolism, but does not alter cigarette consumption,Pharmacogenetics and Genomics, vol. 24, no. 3, pp. 172176, 2014.CrossRefGoogle Scholar
Yamanaka, H., Nakajima, M., Katoh, M. et al., “Trans-3′-hydroxycotinine o- and n-glucuronidations in human liver microsomes,Drug Metabolism & Disposition, vol. 33, no. 1, pp. 2330, 2005.CrossRefGoogle ScholarPubMed
Papageorgiou, I. and Court, M. H., “Identification and validation of the microrna response elements in the 3′-untranslated region of the udp glucuronosyltransferase (ugt) 2b7 and 2b15 genes by a functional genomics approach,Biochemical Pharmacology, vol. 146, pp. 199213, 2017.CrossRefGoogle ScholarPubMed
Kuehl, G. E. and Murphy, S. E., “N-glucuronidation of nicotine and cotinine by human liver microsomes and heterologously expressed udp-glucuronosyltransferases,Drug Metabolism & Disposition, vol. 31, no. 11, pp. 13611368, 2003.CrossRefGoogle ScholarPubMed
Chen, G., Giambrone, N. E., and Lazarus, P., “Glucuronidation of trans-3′-hydroxycotinine by ugt2b17 and ugt2b10,Pharmacogenetics and Genomics, vol. 22, no. 3, pp. 183190, 2012.CrossRefGoogle ScholarPubMed
Nakajima, M., Tanaka, E., Kwon, J. T., and Yokoi, T., “Characterization of nicotine and cotinine n-glucuronidations in human liver microsomes,Drug Metabolism & Disposition, vol. 30, no. 12, pp. 14841490, 2002.CrossRefGoogle ScholarPubMed
Ehmer, U., Vogel, A., Schutte, J. K., Krone, B., Manns, M. P., and Strassburg, C. P., “Variation of hepatic glucuronidation: novel functional polymorphisms of the udp-glucuronosyltransferase ugt1a4,Hepatology, vol. 39, no. 4, pp. 970977, 2004.CrossRefGoogle ScholarPubMed
Mori, A., Maruo, Y., Iwai, M., Sato, H., and Takeuchi, Y., “Udp-glucuronosyltransferase 1a4 polymorphisms in a Japanese population and kinetics of clozapine glucuronidation,Drug Metabolism & Disposition, vol. 33, no. 5, pp. 672675, 2005.CrossRefGoogle Scholar
Wassenaar, C. A., Conti, D. V., Das, S. et al., “Ugt1a and ugt2b genetic variation alters nicotine and nitrosamine glucuronidation in european and african american smokers,Cancer Epidemiology, Biomarkers & Prevention, vol. 24, no. 1, pp. 94104, 2015.CrossRefGoogle ScholarPubMed
Chen, G., Giambrone, N. E., Dluzen, D. F. et al., “Glucuronidation genotypes and nicotine metabolic phenotypes: importance of functional ugt2b10 and ugt2b17 polymorphisms,Cancer Research, vol. 70, no. 19, pp. 75437552, 2010.CrossRefGoogle ScholarPubMed
Figure 0

TABLE 1: The characteristics of the subjects.

Figure 1

TABLE 2: Genetic variants investigated in this study.

Figure 2

TABLE 3: Primers used for amplification of CYP2A64 (gene deletion).

Figure 3

FIGURE 1: (a) First-step PCR results (1,961-bp) (b) second-step PCR results (1,181-bp). Lanes 1 and 2 are wildtype CYP2A6 (non 4/non 4); lanes 3 and 4 are heterozygous deletions (non 4/4); lanes 5 and 6 are the CYP2A6 deletion genotype (4/4).

Figure 4

TABLE 4: Association between the gene scores and smoking cessation.

Supplementary material: File

Li et al. supplementary material
Download undefined(File)
File 91.2 KB