Abstract
Alcohol dependence frequently co-occurs with cigarette smoking, another common addictive behavior. Evidence from genetic studies demonstrates that alcohol dependence and smoking cluster in families and have shared genetic vulnerability. Recently a candidate gene study in nicotine dependent cases and nondependent smoking controls reported strong associations between a missense mutation (rs16969968) in exon 5 of the CHRNA5 gene and a variant in the 3′-UTR of the CHRNA3 gene and nicotine dependence. In this study we performed a comprehensive association analysis of the CHRNA5–CHRNA3–CHRNB4 gene cluster in the Collaborative Study on the Genetics of Alcoholism (COGA) families to investigate the role of genetic variants in risk for alcohol dependence. Using the family-based association test, we observed that a different group of polymorphisms, spanning CHRNA5-CHRNA3, demonstrate association with alcohol dependence defined by Diagnostic and Statistical Manual of Mental Disorders, 4th edn (DSM-IV) criteria. Using logistic regression we replicated this finding in an independent case-control series from the family study of cocaine dependence. These variants show low linkage disequilibrium with the SNPs previously reported to be associated with nicotine dependence and therefore represent an independent observation. Functional studies in human brain reveal that the variants associated with alcohol dependence are also associated with altered steady-state levels of CHRNA5 mRNA.
Keywords: nicotinic acetylcholine receptors, mRNA expression, alcohol dependence, polymorphism
Introduction
It is well established that alcohol and tobacco use are highly correlated in humans. Current smokers are more likely to drink heavily and to binge drink than those who have never smoked, and alcoholics smoke more heavily and endorse nicotine withdrawal symptoms at a higher rate than nonalcoholics.1–5 The National Longitudinal Epidemiologic Survey reported that early onset smoking was a significant predictor of lifetime drinking and subsequent progression to lifetime alcohol abuse and dependence.6 Both alcohol dependence and habitual smoking are transmitted in families and genetic factors contribute to the development of both of these disorders.7–14
Evidence from electrophysiological, pharmacological and neurochemical studies suggest that ethanol may interact with nicotinic acetylcholine receptors (nAChR).15–18 The nAChR gene family has 11 known subunits (α2, α3, α4, α5, α6, α7, α8, α9 and β2, β3, β4); these subunits form pentameric receptors with different combinations of subunits.19,20 The effects of ethanol on nAChRs depend on the receptor subunit composition. Studies using different nAChR subtype compositions expressed in Xenopus oocytes demonstrate that ethanol tends to increase nicotine responsiveness in α2β2, α3β2 and α4β2 receptor subtypes, whereas low concentrations of ethanol inhibit homomeric α7-receptor function.16,21 Ethanol also modulates nAChR potentiation. The combinations of α2β4 and α4β2 are the most sensitive receptors to potentiation by ethanol. The α4β4 and α2β2 combinations are slightly less sensitive and the α3β2 and α3β4 combinations are insensitive to ethanol.16
Several genetic association studies involving addiction in humans have focused on the genes encoding the major nAChR subunits expressed in the brain (α4 and β2). A family-based study in human populations reported genetic variants in CHRNA4 and CHRNB2 that are significantly associated with a protective effect against nicotine addiction.22 The involvement of CHRNA4, but not CHRNB2, in nicotine addiction was confirmed in another family-based study.23 A population-based study among ethnically diverse young adults reported that a polymorphism located immediately upstream of CHRNB2 was associated with initial subjective response to both alcohol and tobacco.24
Recently a comprehensive genome-wide association study and a candidate gene study using nicotine dependent smokers as cases and nondependent smokers as controls demonstrated significant association between several genetic variants in nicotine receptors and nicotine dependence.25,26 As CHRNA5, CHRNA3 and CHRNB4 genes cluster together on chromosome 15q, we performed a comprehensive association analysis with this gene cluster in the families of the Collaborative Study on the Genetics of Alcoholism (COGA) to investigate the role of genetic variants in these three nAChRs in risk for alcohol dependence. We also confirmed our findings in an independent dataset.
Materials and methods
Study subjects
Alcohol-dependent probands, defined by meeting lifetime criteria for both Diagnostic and Statistical Manual of Mental Disorders, 3rd ed (DSM-IIIR) alcohol dependence27 and Feighner criteria for definite alcoholism28 were systematically recruited from alcohol-treatment units. Families in which two additional first-degree relatives also met lifetime criteria for alcohol dependence were invited to participate in the genetic protocol. A total of 262 families including 2309 individuals were selected for the genetic study and an average of 4.6 alcohol-dependent individuals per pedigree were genotyped;29,30 (http://www.niaaagenetics.org/coga_instruments/resources.html). Among these pedigrees, 298 individuals from 35 pedigrees are of African-American descent and 8 pedigrees are of mixed ancestry (by self-report).
All subjects were assessed using the Semi-Structured Assessment for the Genetics of Alcoholism (SSAGA).31,32 Affected individuals were those who were alcohol dependent by DSM-IV criteria. Unaffected individuals were those who drank but had no more than two DSM-IV symptoms of alcohol dependence and were not dependent on any illicit substance. When multiple interviews were available, we required consistency in all interviews for affected/unaffected status.
Genotyping assays
dbSNP (http://www.ncbi.nlm.nih.gov/SNP/) was used to identify polymorphisms within and flanking the genes encoding CHRNA5, CHRNA3 and CHRNB4 on the long arm of chromosome 15.We used Sequenom MassArray technology (http://www.sequenom.com), homogenous MassEXTEND (hME) or iPLEX assays for genotyping of single nucleotide polymorphisms (SNPs). PCR primers, termination mixes and multiplexing capabilities were determined with Sequenom MassARRAY Assay Designer software v3.1.2.2. Standard procedures were used to amplify PCR products; unincorporated nucleotides were deactivated with shrimp alkaline phosphatase. A primer extension reaction was then carried out with the mass extension primer and the appropriate termination mix (hME) or terminator (iPLEX). The primer extension products were then cleaned with resin and spotted onto a silicon SpectroChip. The chip was scanned with a mass spectrometry workstation (Bruker), and the resulting genotype spectra were analyzed with the Sequenom SpectroTYPER software v3.4. Call rates greater than 90% and HWE P-value > 0.05 were set as quality control measures. For the 22-bp insertion/deletion (indel) polymorphism (rs3841324), PCR primers (primer sequences available on request) were selected using the MacVector 6.5.3 program (Accelrys) to yield a 166-bp or 188-bp genomic fragment containing the indel. The nonsynonymous coding SNP in exon 5, rs16969968, originally identified by sequencing 40 individuals from COGA families, was genotyped using an RFLP assay with TaqαI restriction enzyme. Genotypes for rs3841324 and rs16969968 were detected by electrophoresis on a 2% agarose gel.
Statistical analyses
Linkage disequilibrium (LD) between markers was computed using the program Haploview.33 The family-based association test (FBAT)34,35 was used to examine association between the variants and alcohol dependence, defined by DSM-IV criteria. FBAT builds on the original time domain transmission (TDT) method36 in which alleles transmitted to affected offspring are compared with the expected distribution of alleles among offspring. In particular, the method puts tests of different genetic models, tests of different sampling designs, tests involving different disease phenotypes, tests with missing parents and tests of different null hypotheses all into the same framework. Similar in spirit to a classical TDT test, the approach compares the genotype distribution observed in the cases to its expected distribution under the null hypothesis, with the null hypothesis being no linkage and no association, or no association in the presence of linkage. Here, the expected distribution is derived using Mendel’s law of segregation and conditioned on the sufficient statistics for any nuisance parameters under the null. Because conditioning eliminates all nuisance parameters, the technique avoids confounding due to model misspecification as well as admixture or population stratification.37,38 The data shown in Table 1 are from the standard FBAT analysis with age and gender as covariates. To correct for multiple testing we used false discovery rate (FDR) with a default threshold of 0.05.39
Table 1.
NP | Chromosome position | Allele | MAF | Nf | P-value | FDR q-value |
---|---|---|---|---|---|---|
rs1979906a | 76629344 | A/G | 0.44 | 171 | 0.008 | 0.036 |
rs880395a | 76631411 | A/G | 0.43 | 162 | 0.088 | 0.114 |
rs7164030a | 76631716 | A/G | 0.44 | 149 | 0.113 | 0.129 |
rs905739b | 76632165 | C/T | 0.22 | 120 | 0.179 | 0.155 |
rs2036527c | 76638670 | C/T | 0.33 | 145 | 0.166 | 0.155 |
rs3841324a | 76644877 | L/S | 0.44 | 172 | 0.005 | 0.030 |
rs503464b | 76644951 | A/T | 0.22 | 133 | 0.192 | 0.157 |
rs684513b | 76645455 | C/G | 0.2 | 116 | 0.277 | 0.194 |
rs667282b | 76650527 | C/T | 0.23 | 136 | 0.352 | 0.229 |
rs17486278c | 76654537 | A/C | 0.32 | 138 | 0.078 | 0.112 |
rs601079a | 76656634 | A/T | 0.43 | 180 | 0.026 | 0.059 |
rs680244a | 76658343 | A/G | 0.44 | 175 | 0.003 | 0.030 |
rs621849a | 76659916 | A/G | 0.44 | 175 | 0.004 | 0.030 |
rs569207b | 76660174 | A/G | 0.23 | 133 | 0.121 | 0.129 |
rs692780a | 76663560 | C/G | 0.38 | 172 | 0.013 | 0.039 |
rs16969968c | 76669980 | A/G | 0.34 | 154 | 0.177 | 0.155 |
rs514743a | 76671282 | A/T | 0.38 | 176 | 0.117 | 0.129 |
rs578776b | 76675455 | C/T | 0.28 | 146 | 0.81 | 0.417 |
rs6495307a | 76677376 | C/T | 0.43 | 172 | 0.01 | 0.036 |
rs12910984b | 76678682 | A/G | 0.23 | 132 | 0.345 | 0.229 |
rs1051730c | 76681394 | C/T | 0.32 | 147 | 0.016 | 0.042 |
rs3743078b | 76681814 | C/G | 0.24 | 137 | 0.671 | 0.381 |
rs3743077a | 76681951 | A/G | 0.42 | 176 | 0.08 | 0.112 |
rs938682b | 76683602 | C/T | 0.23 | 133 | 0.602 | 0.353 |
rs11637630b | 76686774 | A/G | 0.23 | 132 | 0.228 | 0.166 |
rs7177514b | 76694461 | C/G | 0.24 | 128 | 0.589 | 0.353 |
rs6495308b | 76694711 | C/T | 0.24 | 132 | 0.87 | 0.417 |
rs8042059b | 76694914 | A/C | 0.23 | 131 | 0.851 | 0.417 |
rs8042374b | 76695087 | A/G | 0.23 | 121 | 0.432 | 0.271 |
rs3743075a | 76696507 | A/G | 0.39 | 175 | 0.207 | 0.157 |
rs3743073a | 76696594 | A/C | 0.39 | 175 | 0.199 | 0.157 |
rs1878399a | 76699058 | C/G | 0.43 | 175 | 0.061 | 0.107 |
rs17487223c | 76711042 | C/T | 0.35 | 144 | 0.065 | 0.107 |
rs950776 | 76713073 | C/T | 0.35 | 164 | 0.872 | 0.417 |
rs11636605 | 76715933 | A/G | 0.22 | 112 | 0.975 | 0.455 |
rs9920506 | 76718112 | A/G | 0.19 | 102 | 0.716 | 0.383 |
rs3813567 | 76721606 | C/T | 0.22 | 125 | 0.698 | 0.383 |
rs17487514 | 76740840 | C/T | 0.3 | 156 | 0.052 | 0.105 |
rs1996371 | 76743861 | A/G | 0.39 | 154 | 0.155 | 0.155 |
Abbreviations: COGA, Collaborative Study on the Genetics of Alcoholism; FBAT, family-based association test; FDR, false discovery rate; MAF: minor allele frequency; Nf: number of informative families; SNP, single nucleotide polymorphism. The data shown here are from the standard analysis with age and gender as covariates.
Represents SNPs that are highly correlated (r2≥0.7) with a 22-bp indel (rs3841324) in CHRNA5.
Represents SNPs that are highly correlated with rs578776 maps to the 3′-UTR of CHRNA3.
Represents SNPs that are highly correlated with a missense mutation (rs16969968) in exon 5 of CHRNA5.
Replication study with the family study of cocaine dependence data set
Study subjects
Unrelated cases and matched unrelated controls within the candidate-gene study of the family study of cocaine dependence (FSCD) were used for this study.40 Cocaine dependent subjects were recruited from publicly and privately funded inpatient and outpatient chemical dependency treatment centers in the St Louis area. Eligibility requirements included meeting DSM-IV criteria for cocaine dependence, being 18 years of age or older, speaking fluent English and having a full sibling within five years of their age who was willing to participate in the family arm of the study. Control subjects were recruited through driver’s license records maintained by the Missouri Family Registry at Washington University in St Louis for research purposes. Controls were matched to cocaine dependent subjects based on age, ethnicity, gender and zip code. Control subjects were not dependent on alcohol or drugs, including nicotine, but did use at least alcohol because non-substance using individuals are considered phenotypically unknown. The project was approved by the Washington University IRB and all subjects provided informed consent. All participants completed a modified version of the SSAGA.31,32
Genotyping assays
Genotyping for the FSCD study was conducted by the Center for Inherited Disease Research (CIDR) using a custom SNP array on an Illumina platform. Details of genotyping procedures are available at the CIDR website (http://www.cidr.jhmi.edu/index.html). Additional genotyping was performed by Sequenom assays described above.
Statistical analysis
LD between markers was computed using the program COCAPHASE.41 We used logistic regression42 to examine the association between the SNPs and DSM-IV alcohol dependence. For analysis, we selected those cases who were comorbid for DSM-IV alcohol and cocaine dependence and compared them with all of the study controls. This subset included 451 unrelated individuals of European-American descent (207 alcohol-dependent cases and 244 controls) and 424 unrelated individuals of African-American descent (185 alcohol-dependent cases and 239 controls). Separate logistic regression models were run for the European and African-American subjects as well as a combined analysis that incorporated all samples and included race as a covariate and a genotype by race interaction term. Three logistic regression models were examined for each variant to test for additive effects and evidence of dominant or recessive modes of inheritance. The additive effect of a SNP was modeled using an ordinal measure of the number of copies of the risk allele. The dominant and recessive effects of a SNP were modeled using dichotomous indicator variables. For each variant, the model with the strongest association with DSM-IV alcohol dependence, based on the adjusted odds ratio and the magnitude of the corresponding P-value, is reported in Table 2.
Table 2.
SNP | European Americans | African Americans | Combined dataset | |||||
---|---|---|---|---|---|---|---|---|
MAF | OR (CI) | P-value | MAF | OR (CI) | P-value | OR (CI) | P-value | |
rs880395 | 0.46 | 1.32 (1.02, 1.71) | 0.037 | 0.23 | 1.25 (0.89, 1.76) | 0.193 | 1.29 (1.05, 1.58) | 0.015 |
rs7164030 | 0.47 | 1.35 (1.04, 1.75) | 0.023 | 0.23 | 1.24 (0.88, 1.74) | 0.221 | 1.30 (1.06, 1.60) | 0.012 |
rs3841324 | 0.47 | 1.35 (1.04, 1.76) | 0.022 | 0.22 | 1.24 (0.88, 1.75) | 0.219 | 1.31 (1.07, 1.62) | 0.01 |
rs588765 | 0.46 | 1.36 (1.05, 1.77) | 0.020 | 0.31 | 1.61 (1.16, 2.23) | 0.004 | 1.45 (1.19, 1.78) | 0.0003 |
rs6495306 | 0.46 | 1.36 (1.05, 1.77) | 0.020 | 0.31 | 1.55 (1.12, 2.15) | 0.008 | 1.43 91.17, 1.76) | 0.001 |
rs601079 | 0.46 | 1.34 (1.03, 1.74) | 0.028 | 0.43 | 1.37 (1.02, 1.84) | 0.034 | 1.35 (1.11, 1.64) | 0.002 |
rs680244 | 0.46 | 1.35 (1.04, 1.75) | 0.023 | 0.43 | 1.36 (1.02, 1.82) | 0.039 | 1.35 (1.11, 1.64) | 0.002 |
rs692780 | 0.4 | 1.37 (1.05, 1.78) | 0.022 | 0.24 | 1.47 (1.04, 2.08) | 0.029 | 1.41 (1.14, 1.74) | 0.002 |
rs11637635 | 0.40 | 1.33 (1.02, 1.74) | 0.032 | 0.28 | 1.61 (1.16, 2023) | 0.005 | 1.44 (1.17, 1.77) | 0.001 |
rs555018 | 0.46 | 1.36 (1.05,1.77) | 0.020 | 0.31 | 1.58 (1.14, 2.20) | 0.006 | 1.44 (1.18, 1.77) | 0.0005 |
rs647041 | 0.46 | 1.33 (1.02, 1.72) | 0.033 | 0.30 | 1.61 (1.16, 2.25) | 0.005 | 1.43 (1.17, 1.75) | 0.001 |
rs615470 | 0.40 | 1.29 (0.99, 1.69) | 0.055 | 0.42 | 1.48 (1.10, 1.99) | 0.009 | 1.38 (1.13, 1.68) | 0.001 |
rs6495307 | 0.45 | 1.30 (1.00, 1.69) | 0.046 | 0.42 | 1.47 (1.09, 1.97) | 0.012 | 1.37 (1.13, 1.67) | 0.001 |
rs3743077 | 0.46 | 1.34 (1.03, 1.74) | 0.027 | 0.14 | 1.18 (0.75, 1.86) | 0.471 | 1.38 (1.10, 1.72) | 0.005 |
rs2869546 | 0.41 | 1.34 (1.03, 1.74) | 0.030 | 0.39 | 1.18 (0.89, 1.57) | 0.256 | 1.26 (1.04, 1.53) | 0.018 |
rs3743075 | 0.4 | 1.29 (0.99, 1.68) | 0.058 | 0.45 | 1.10 (0.83, 1.47) | 0.511 | 1.20 (0.99, 1.46) | 0.063 |
rs3743074 | 0.41 | 1.32 (1.02, 1.73) | 0.038 | 0.45 | 1.09 (0.82, 1.45) | 0.561 | 1.21 (1.00, 1.47) | 0.053 |
rs1878399 | 0.46 | 1.32 (1.02, 1.71) | 0.038 | 0.26 | 1.09 (0.79, 1.51) | 0.601 | 1.22 (1.00, 1.50) | 0.052 |
Abbreviations: CI, confidence interval; FSCD, family study of cocaine dependence; MAF, minor allele frequency; OR, odd ratio; SNP, single nucleotide polymorphism. P-values < 0.05 are in bold face.
Gene expression analyses
Postmortem brain tissues derived from frontal cortex and cerebellum of 48 unrelated, nondemented elderly European Americans were obtained from the brain bank of the Alzheimer’s Disease Research Center at Washington University in St Louis (http://alzheimer.wustl.edu/). We used Qiagen’s DNeasy Blood & Tissue Kit and RNeasy Lipid Tissue kit (http://www.qiagen.com) to extract DNA and total RNA from brain tissue, respectively. A cDNA library was prepared from total RNA using the High Capacity cDNA Archive Kit (http://www.appliedbiosystems.com).
Genomic DNA from all subjects was genotyped for the SNP, rs588765, which exhibits the lowest P-value in the African-American population. The polymorphism, rs3841324 was also genotyped because of its location within the promoter of the CHRNA5 gene. Gene expression levels were analyzed by real-time PCR using an ABI-7500 real-time PCR system. A TaqMan assay (Hs00181248_m1, ABI) was used for quantifying the expression level of CHRNA5 in the frontal cortex. The expression levels of CHRNA3 were determined in cerebellum using a TaqMan assay (Hs00609519, ABI) due to the low expression levels of CHRNA3 in frontal cortex. Primers and a TaqMan probe (sequences available on request) for the reference gene, GAPDH were designed over exon–exon boundaries using the Primer Express 3 (ABI) program.
Each real-time PCR run included within-plate duplicates and each experiment was performed twice for each sample. Correction for sample-to-sample variation was done by simultaneously amplifying GAPDH as a reference. Real-time data were analyzed using the comparative Ct method.43 The Ct values of each sample were normalized with the Ct value for the housekeeping gene, GADPH and were corrected for the PCR efficiency of each assay,43 although the efficiency of all reactions was close to 100%. Only the samples with an s.e. < 0.15 were analyzed. We used the nonparametric Mann–Whitney U-statistic to test for evidence of differential expression in samples of different genotypes.
Results
Forty-one single nucleotide polymorphisms (SNPs) and an indel within and flanking this cluster of nAChRs were genotyped (Figure 1a). Each of the polymorphisms was in Hardy–Weinberg equilibrium in the founders. Three SNPs that had a minor allele frequency (MAF) less than 5% were removed from the analyses. Using pair-wise linkage disequilibrium analysis, we observed three groups of highly correlated variants tagged by three putative functional polymorphisms, a 22 bp indel (rs3841324) in the promoter region of the CHRNA5 gene, a missense mutation (rs16969968) in exon 5 of the CHRNA5 gene, and a SNP (rs578776) in the 3′-UTR of the CHRNA3 gene, respectively (Figure 1a; Table 1).
Genetic variants in the CHRNA5–CHRNA3–CHRNB4 gene cluster are associated with alcohol dependence in the COGA data set
Eight variants spanning 52 kb in the CHRNA5–CHRNA3–CHRNB4 gene cluster were significantly associated with alcohol dependence using FBAT analysis adjusted for age and gender (Table 1). Seven of these variants are highly correlated (r2≥0.7, Figure 1b). These associated polymorphisms include rs1979906 in the region upstream of the CHRNA5 gene, a 22 bp indel (rs3841324) in the promoter region of CHRNA5, four intronic SNPs in the CHRNA5 gene and one intronic SNP in the CHRNA3 gene (Figure 1a; Table 1). The eighth SNP (rs1051730), a synonymous variant in exon 5 of the CHRNA3 gene has a lesser degree of linkage disequilibrium (r2≤0.4). To determine whether this association was driven by nicotine dependence, we also analyzed the association with habitual smoking as a covariate and found that the association of these polymorphisms with alcohol dependence is independent of smoking status. These polymorphisms remain significant after adjusting for multiple testing using the FDR.39
In contrast, we observed no association between alcohol dependence and either of the SNPs previously reported to be associated with nicotine dependence: the missense mutation (rs16969968) in CHRNA5 and rs578776 located within the 3′-UTR of CHRNA3.26,44 A similar pattern of association was seen for all polymorphisms across the gene cluster in affected only analyses and in analyses without covariates.
Replication of the association with alcohol dependence in an independent data set from the FSCD
To further examine the genetic contribution of variants in this gene cluster to risk for alcohol dependence we genotyped 18 polymorphisms showing high linkage disequilibrium with each other (r2≥0.7 in European Americans), in an independent data set from the FSCD (Table 2). Using logistic regression analysis we confirmed the association between each of these SNPs and alcohol dependence in the subjects of European descent. In the African-American subset, 10 of the 18 polymorphisms, spanning CHRNA5 and CHRNA3, showed significant association with alcohol dependence. Six of these polymorphisms have a lower MAF in African Americans (< 0.31 vs 0.4) but slightly higher odds ratios than in the European American sample (Table 2). To test whether the genetic associations observed show statistically significant differences between the two racial groups we repeated the logistic regression including a genotype by race interaction term. None of the variants showed significant evidence of genetic heterogeneity by race.
Allelic differences in expression of CHRNA5 in human frontal cortex
To examine whether the variants associated with alcohol dependence have a direct effect on gene expression, we performed quantitative real-time PCR analysis with human brain tissue from nondemented elderly European Americans. Genomic DNA from all subjects was genotyped for multiple variants in CHRNA5. Genotypes were in Hardy–Weinberg equilibrium. We then examined CHRNA5 mRNA expression with SNP, rs588765 in 13 samples homozygous for the major allele (CC), 9 samples homozygous for the minor allele (TT), and 21 heterozygous samples (CT). Subjects homozygous for the minor allele of rs588765 showed a 2-fold increase (TT = 1.88 ± 1.2; CC = 0.98 ± 1.04; P = 0.025) in CHRNA5 mRNA expression compared to subjects homozygous for the major allele (Figure 2). Similar results were obtained with rs3841324 (SS = 2.21 ± 1.15; LL = 0.89 ± 0.89; P = 0.02). Heterozygotes for both rs588765 and rs3841324 showed lower expression levels for CHRNA5, compared with homozygotes for the minor allele (P = 0.0003 and P = 0.02 for rs3841324 and rs588765, respectively). However, no significant differences were found between heterozygotes and homozygotes for the major allele (Figure 2). In contrast CHRNA3 expression in the same individuals showed no association with alleles of rs588765 or rs3841324 (data not shown). This expression data suggest that the functional allele explaining the association with alcohol dependence regulates CHRNA5 expression. However, in these European American samples available for our expression studies we are not able to determine which variant is responsible for the expression differences.
To validate this observation using an independent data set and methodology we tested for association between variability in CHRNA5 mRNA expression in lymphoblastoid cell lines derived from CEPH families using the Affymetrix HG Focus panel (Genetic Analysis Workshop 15; www.gaworkshop.org/gaw15.htm).45,46 We retrieved genotypes for rs588765 from the HapMap database and examined the difference in CHRNA5 mRNA expression in subjects with different genotypes at rs588765 in 14 genotyped trios. Using SOLAR VC quantitative analysis of the CHRNA5 mRNA levels with additive genetic effects, we detected significant differences in expression in subjects of different genotypes (P = 0.02). This SNP accounted for approximately 14% of the variance in CHRNA5 gene expression in this system.
Discussion
Alcohol dependence frequently co-occurs with tobacco use. Evidence from electrophysiological and neurochemical studies suggest that ethanol and nicotine share important mechanisms of action in the brain reward pathways and nAChRs.17 A study using the COGA data set demonstrated that the clinical characteristics of alcohol use history are associated with smoking status; current smokers and nicotine-dependent subjects had a greater severity of alcohol dependence.3
Although the main nAChRs in the brain are believed to contain α4 and β2 subunits, a recent comprehensive candidate gene study implicated SNPs in the genes encoding the α3, α5, and β3 subunits, including a missense mutation, rs16969968, in the CHRNA5 gene as risk factors for nicotine dependence.26 Given the shared genetic vulnerability for nicotine and alcohol dependence, detected through twin studies,9,14 we undertook a detailed analysis of the CHRNA5–CHRNA3–CHRNB4 gene cluster in the COGA data set. Although the association between the missense mutation in CHRNA5 and nicotine dependence has been replicated in several data sets, including the COGA data set,44 we saw no evidence of association between this SNP and alcohol dependence, suggesting that the effect of this polymorphism is specific to nicotine. Similarly, the group of SNPs in linkage disequilibrium with rs578776 shows strong association with nicotine dependence26 but shows no evidence of association with alcohol dependence in the COGA data set.
In this study we provide replicated evidence of association for 16 correlated variants with alcohol dependence in two data sets of European descent. The strong association detected in the FSCD data set, where all cases are both cocaine dependent and alcohol dependent, raises the possibility that these variants may also be associated with alcohol related phenotypes, such as illicit drug dependence and antisocial personality disorder. With lower levels of LD among the polymorphisms in the African-American population (Table 3), we observed a narrower region of association extending from rs588765 in intron 1 of CHRNA5 to rs6495307 in intron 5 of CHRNA3. The SNPs showing the lowest P-values in the African-American sample have a MAF of 30% compared to 46% in the European-American sample. As we do not see genetic association with the indel in African Americans, this suggests that further studies should prioritize SNPs in high LD with rs588765. Although there are differences in the LD between the polymorphisms in the European-American and African-American subjects from the FSCD data set, none of these variants demonstrated significant evidence of genetic heterogeneity between the two populations. Thus, though there are some differences in significance and effect sizes between the variants, we cannot definitively determine which polymorphism is causing the functional changes.
Table 3.
D’ r2 |
rs880395 | rs7164030 | rs3841324 | rs588765 | rs6495306 | rs601079 | rs680244 | rs692780 | rs11637635 | rs555018 | rs647041 | rs615470 | rs6495307 | rs3743077 | rs2869546 | rs3743075 | rs3743074 | rs1878399 |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
rs880395 | 1.00 | 0.39 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.87 | 1.00 | 1.00 | 1.00 | 1.00 | |
rs7164030 | 1.00 | 0.39 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.87 | 1.00 | 1.00 | 1.00 | 1.00 | |
rs3841324 | 0.13 | 0.13 | 0.38 | 0.37 | 0.24 | 0.26 | 0.35 | 0.31 | 0.38 | 0.37 | 1.00 | 0.26 | 1.00 | 0.22 | 1.00 | 1.00 | 0.33 | |
rs588765 | 0.68 | 0.68 | 0.10 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.69 | 0.70 | 0.69 | 1.00 | |
rs6495306 | 0.67 | 0.67 | 0.09 | 1.00 | 0.97 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.66 | 0.65 | 0.64 | 1.00 | |
rs601079 | 0.38 | 0.38 | 0.02 | 0.55 | 0.54 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.41 | 1.00 | 0.19 | 0.15 | 0.15 | 1.00 | |
rs680244 | 0.38 | 0.38 | 0.03 | 0.56 | 0.56 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.42 | 1.00 | 0.20 | 0.16 | 0.16 | 1.00 | |
rs692780 | 0.86 | 0.86 | 0.10 | 0.77 | 0.75 | 0.42 | 0.42 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | |
rs11637635 | 0.70 | 0.70 | 0.07 | 1.00 | 0.91 | 0.52 | 0.52 | 0.82 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.73 | 0.75 | 0.74 | 0.79 | |
rs555018 | 0.67 | 0.67 | 0.09 | 1.00 | 1.00 | 0.56 | 0.56 | 0.76 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.68 | 0.67 | 0.67 | 1.00 | |
rs647041 | 0.71 | 0.71 | 0.09 | 1.00 | 0.95 | 0.54 | 0.54 | 0.79 | 1.00 | 1.00 | 1.00 | 1.00 | 1.00 | 0.72 | 0.68 | 0.67 | 1.00 | |
rs615470 | 0.33 | 0.33 | 0.01 | 0.50 | 0.51 | 0.13 | 0.14 | 0.40 | 0.49 | 0.50 | 0.48 | 1.00 | 1.00 | 1.00 | 0.83 | 0.83 | 0.63 | |
rs6495307 | 0.34 | 0.34 | 0.02 | 0.51 | 0.52 | 0.15 | 0.16 | 0.38 | 0.47 | 0.51 | 0.49 | 0.85 | 1.00 | 1.00 | 0.80 | 0.80 | 1.00 | |
rs3743077 | 0.38 | 0.38 | 0.49 | 0.38 | 0.38 | 0.21 | 0.21 | 0.45 | 0.38 | 0.38 | 0.40 | 0.84 | 0.19 | 1.00 | 1.00 | 1.00 | 1.00 | |
rs2869546 | 0.34 | 0.34 | 0.02 | 0.24 | 0.23 | 0.03 | 0.04 | 0.39 | 0.26 | 0.24 | 0.26 | 1.00 | 0.82 | 0.18 | 1.00 | 1.00 | 1.00 | |
rs3743075 | 0.27 | 0.27 | 0.24 | 0.19 | 0.17 | 0.02 | 0.02 | 0.30 | 0.21 | 0.18 | 0.18 | 1.00 | 0.50 | 0.14 | 0.77 | 1.00 | 1.00 | |
rs3743074 | 0.27 | 0.27 | 0.24 | 0.18 | 0.16 | 0.02 | 0.02 | 0.30 | 0.20 | 0.17 | 0.17 | 0.98 | 0.48 | 0.14 | 0.76 | 1.00 | 1.00 | |
rs1878399 | 0.67 | 0.67 | 0.07 | 0.80 | 0.78 | 0.45 | 0.45 | 0.73 | 0.57 | 0.79 | 0.83 | 0.83 | 0.40 | 0.37 | 0.47 | 0.38 | 0.37 |
Using brain mRNA from European Americans we observed an association between the minor alleles of rs588765 and rs3841324 (r2 = 0.8 in European Americans) and higher levels of CHRNA5 mRNA. This association with CHRNA5 expression was replicated in the GAW15 data set with rs588765.45,46 Several in-vitro studies have examined 5′ regulatory sequences of CHRNA5.47–49 Functional characterization of CHRNA5 using luciferase assays in human cell lines has previously demonstrated that the −240/+53 region, which contains the rs3841324 indel, is the core promoter.47,48 A second study has directly tested the effect of the indel using a standardized reporter gene assay system and reported that it alters promoter activity in HEK293t cells by 1.5-fold.49 However, our genetic data in African Americans does not support a primary role of this polymorphism in influencing CHRNA5 expression in the brain. No in-vitro functional experiments have been performed using other putative regulatory regions of CHRNA5. Further examination on CHRNA5 mRNA expression in African-American subjects may be very helpful in narrowing down the likely functional variant because of the lower levels of LD across CHRNA5 in this population.
In the mammalian brain, nAChRs include homopentameric α7 receptors and a variety of heteropentamers, but predominantly α4β2*, where the asterisk denotes the presence of another subunit. In many brain areas, the α5 polypeptide completes the α4β2* complex, with stoichiometry α42β22α5.20,50,51 The α5 subunit contributes one-fifth of the channel lining and therefore can influence channel properties.20 Studies using heterologous expression systems have demonstrated that addition of the α5 subunit has profound effects on the functional properties of α3β4 receptors.52 CHRNA5 increases the rate of desensitization of α3β4 receptors as well as the Ca2+ permeability and decreases the efficacy of nicotine on α3β4 nAchRs.
In conclusion, we provide replicated evidence of association between multiple polymorphisms within the CHRNA5 and CHRNA3 genes and alcohol dependence. Furthermore, we demonstrate that the minor alleles of these polymorphisms are associated with higher CHRNA5 mRNA levels in human frontal cortex. These results suggest that although variation in CHRNA5 influences risk for both alcohol dependence and nicotine dependence, different polymorphisms and different mechanisms of action are responsible for these effects on risk.
Acknowledgments
The Collaborative Study on the Genetics of Alcoholism (COGA)—co-principal investigators B Porjesz, V Hesselbrock, H Edenberg, L Bierut—includes nine different centers where data collection, analysis and storage take place. The nine sites and principal investigators and co-investigators are University of Connecticut (V Hesselbrock), Indiana University (HJ Edenberg, J Nurnberger Jr, PM Conneally, T Foroud), University of Iowa (S Kuperman, R Crowe), SUNY Downstate (B Porjesz), Washington University in St Louis (L Bierut, A Goate, J Rice), University of California at San Diego (M Schuckit), Howard University (R Taylor), Rutgers University (J Tischfield), Southwest Foundation (L Almasy). Zhaoxia Ren serves as the National Institute on Alcohol Abuse and Alcoholism (NIAAA) Staff Collaborator. This national collaborative study is supported by the NIH Grant U10AA008401 from the NIAAA and the National Institute on Drug Abuse (NIDA).
In memory of Henri Begleiter and Theodore Reich, Principal and co-principal investigators of COGA since its inception; we are indebted to their leadership in the establishment and nurturing of COGA, and acknowledge with great admiration their seminal scientific contributions to the field.
We are grateful to the families for their participation in the studies at St Louis Alzheimer’s Disease Research Center (ADRC), Washington University. Funding for the research at ADRC was provided by grants from the National Institute on Aging: P50 AG05681 and P01 AG03991 to John C Morris. The Family Study of Cocaine Dependence (FSCD) data collection was funded by the National Institute of Drug Abuse (NIDA), R01 DA13423 to Laura J Bierut. Genotyping and analysis of the FSCD data was supported through the Case Control Candidate Gene Study of Addiction, NIDA grant R01 DA19963 to Laura J Bierut.
Financial Disclosures:
Drs LJ Bierut, AM Goate, AJ Hinrichs, JP Rice, SF Saccone and JC Wang are listed as inventors on a patent (US 20070258898) held by Perlegen Sciences Inc., covering the use of certain SNPs, including rs16969968 in diagnosing, prognosing and treating addiction. Dr Bierut has acted as a consultant for Pfizer Inc. in 2008.
References
- 1.Istvan J, Matarazzo JD. Tobacco, alcohol, and caffeine use: a review of their interrelationships. Psychol Bull. 1984;95:301–326. [PubMed] [Google Scholar]
- 2.Marks JL, Hill EM, Pomerleau CS, Mudd SA, Blow FC. Nicotine dependence and withdrawal in alcoholic and nonalcoholic ever-smokers. J Subst Abuse Treat. 1997;14:521–527. doi: 10.1016/s0740-5472(97)00049-4. [DOI] [PubMed] [Google Scholar]
- 3.Daeppen JB, Smith TL, Danko GP, Gordon L, Landi NA, Nurnberger JI, Jr, et al. Clinical correlates of cigarette smoking and nicotine dependence in alcohol-dependent men and women. The Collaborative Study Group on the Genetics of Alcoholism. Alcohol Alcohol. 2000;35:171–175. doi: 10.1093/alcalc/35.2.171. [DOI] [PubMed] [Google Scholar]
- 4.John U, Meyer C, Rumpf HJ, Hapke U. Probabilities of alcohol high-risk drinking, abuse or dependence estimated on grounds of tobacco smoking and nicotine dependence. Addiction. 2003;98:805–814. doi: 10.1046/j.1360-0443.2003.00381.x. [DOI] [PubMed] [Google Scholar]
- 5.John U, Meyer C, Rumpf HJ, Schumann A, Thyrian JR, Hapke U. Strength of the relationship between tobacco smoking, nicotine dependence and the severity of alcohol dependence syndrome criteria in a population-based sample. Alcohol Alcohol. 2003;38:606–612. doi: 10.1093/alcalc/agg122. [DOI] [PubMed] [Google Scholar]
- 6.Grant BF. Age at smoking onset and its association with alcohol consumption and DSM-IV alcohol abuse and dependence: results from the National Longitudinal Alcohol Epidemiologic Survey. J Subst Abuse. 1998;10:59–73. doi: 10.1016/s0899-3289(99)80141-2. [DOI] [PubMed] [Google Scholar]
- 7.Heath AC, Bucholz KK, Madden PA, Dinwiddie SH, Slutske WS, Bierut LJ, et al. Genetic and environmental contributions to alcohol dependence risk in a national twin sample: consistency of findings in women and men. Psychol Med. 1997;27:1381–1396. doi: 10.1017/s0033291797005643. [DOI] [PubMed] [Google Scholar]
- 8.True WR, Heath AC, Scherrer JF, Waterman B, Goldberg J, Lin N, et al. Genetic and environmental contributions to smoking. Addiction. 1997;92:1277–1287. [PubMed] [Google Scholar]
- 9.True WR, Xian H, Scherrer JF, Madden PA, Bucholz KK, Heath AC, et al. Common genetic vulnerability for nicotine and alcohol dependence in men. Arch Gen Psychiatry. 1999;56:655–661. doi: 10.1001/archpsyc.56.7.655. [DOI] [PubMed] [Google Scholar]
- 10.Bierut LJ, Schuckit MA, Hesselbrock V, Reich T. Co-occurring risk factors for alcohol dependence and habitual smoking. Alcohol Res Health. 2000;24:233–241. [PMC free article] [PubMed] [Google Scholar]
- 11.Grucza RA, Bierut LJ. Co-occurring risk factors for alcohol dependence and habitual smoking: update on findings from the Collaborative Study on the Genetics of Alcoholism. Alcohol Res Health. 2006;29:172–178. [PMC free article] [PubMed] [Google Scholar]
- 12.Kendler KS, Neale MC, Sullivan P, Corey LA, Gardner CO, Prescott CA. A population-based twin study in women of smoking initiation and nicotine dependence. Psychol Med. 1999;29:299–308. doi: 10.1017/s0033291798008022. [DOI] [PubMed] [Google Scholar]
- 13.Prescott CA, Kendler KS. Genetic and environmental contributions to alcohol abuse and dependence in a population-based sample of male twins. Am J Psychiatry. 1999;156:34–40. doi: 10.1176/ajp.156.1.34. [DOI] [PubMed] [Google Scholar]
- 14.Madden PA, Heath AC. Shared genetic vulnerability in alcohol and cigarette use and dependence. Alcohol Clin Exp Res. 2002;26:1919–1921. doi: 10.1097/01.ALC.0000040960.15151.30. [DOI] [PubMed] [Google Scholar]
- 15.Aistrup GL, Marszalec W, Narahashi T. Ethanol modulation of nicotinic acetylcholine receptor currents in cultured cortical neurons. Mol Pharmacol. 1999;55:39–49. doi: 10.1124/mol.55.1.39. [DOI] [PubMed] [Google Scholar]
- 16.Cardoso RA, Brozowski SJ, Chavez-Noriega LE, Harpold M, Valenzuela CF, Harris RA. Effects of ethanol on recombinant human neuronal nicotinic acetylcholine receptors expressed in Xenopus oocytes. J Pharmacol Exp Ther. 1999;289:774–780. [PubMed] [Google Scholar]
- 17.Larsson A, Engel JA. Neurochemical and behavioral studies on ethanol and nicotine interactions. Neurosci Biobehav Rev. 2004;27:713–720. doi: 10.1016/j.neubiorev.2003.11.010. [DOI] [PubMed] [Google Scholar]
- 18.Butt CM, King NM, Stitzel JA, Collins AC. Interaction of the nicotinic cholinergic system with ethanol withdrawal. J Pharmacol Exp Ther. 2004;308:591–599. doi: 10.1124/jpet.103.059758. [DOI] [PubMed] [Google Scholar]
- 19.Gotti C, Fornasari D, Clementi F. Human neuronal nicotinic receptors. Prog Neurobiol. 1997;53:199–237. doi: 10.1016/s0301-0082(97)00034-8. [DOI] [PubMed] [Google Scholar]
- 20.Lindstrom JM. Nicotinic acetylcholine receptors of muscles and nerves: comparison of their structures, functional roles, and vulnerability to pathology. Ann N Y Acad Sci. 2003;998:41–52. doi: 10.1196/annals.1254.007. [DOI] [PubMed] [Google Scholar]
- 21.Davis TJ, de Fiebre CM. Alcohol’s actions on neuronal nicotinic acetylcholine receptors. Alcohol Res Health. 2006;29:179–185. [PMC free article] [PubMed] [Google Scholar]
- 22.Feng Y, Niu T, Xing H, Xu X, Chen C, Peng S, et al. A common haplotype of the nicotine acetylcholine receptor alpha 4 subunit gene is associated with vulnerability to nicotine addiction in men. Am J Hum Genet. 2004;75:112–121. doi: 10.1086/422194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Li MD, Beuten J, Ma JZ, Payne TJ, Lou XY, Garcia V, et al. Ethnic- and gender-specific association of the nicotinic acetylcholine receptor alpha4 subunit gene (CHRNA4) with nicotine dependence. Hum Mol Genet. 2005;14:1211–1219. doi: 10.1093/hmg/ddi132. [DOI] [PubMed] [Google Scholar]
- 24.Ehringer MA, Clegg HV, Collins AC, Corley RP, Crowley T, Hewitt JK, et al. Association of the neuronal nicotinic receptor beta2 subunit gene (CHRNB2) with subjective responses to alcohol and nicotine. Am J Med Genet B Neuropsychiatr Genet. 2007;144:596–604. doi: 10.1002/ajmg.b.30464. [DOI] [PubMed] [Google Scholar]
- 25.Bierut LJ, Madden PA, Breslau N, Johnson EO, Hatsukami D, Pomerleau OF, et al. Novel genes identified in a high-density genome wide association study for nicotine dependence. Hum Mol Genet. 2007;16:24–35. doi: 10.1093/hmg/ddl441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Saccone SF, Hinrichs AL, Saccone NL, Chase GA, Konvicka K, Madden PA, et al. Cholinergic nicotinic receptor genes implicated in a nicotine dependence association study targeting 348 candidate genes with 3713 SNPs. Hum Mol Genet. 2007;16:36–49. doi: 10.1093/hmg/ddl438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.American Psychiatric Association. Diagnostic and statistical manual of mental disorders. 3rd edn. Washington, DC: American Psychiatric Press; 1987. (revised) [Google Scholar]
- 28.Feighner JP, Robins E, Guze SB, Woodruff RA, Jr, Winokur G, Munoz R. Diagnostic criteria for use in psychiatric research. Arch Gen Psychiatry. 1972;26:57–63. doi: 10.1001/archpsyc.1972.01750190059011. [DOI] [PubMed] [Google Scholar]
- 29.Nurnberger JI, Jr, Wiegand R, Bucholz K, O’Connor S, Meyer ET, Reich T, et al. A family study of alcohol dependence: coaggregation of multiple disorders in relatives of alcohol-dependent probands. Arch Gen Psychiatry. 2004;61:1246–1256. doi: 10.1001/archpsyc.61.12.1246. [DOI] [PubMed] [Google Scholar]
- 30.Reich T, Edenberg HJ, Goate A, Williams JT, Rice JP, Van Eerdewegh P, et al. Genome-wide search for genes affecting the risk for alcohol dependence. Am J Med Genet. 1998;81:207–215. [PubMed] [Google Scholar]
- 31.Bucholz KK, Cadoret R, Cloninger CR, Dinwiddie SH, Hesselbrock VM, Nurnberger JI, Jr, et al. A new, semi-structured psychiatric interview for use in genetic linkage studies: a report on the reliability of the SSAGA. J Stud Alcohol. 1994;55:149–158. doi: 10.15288/jsa.1994.55.149. [DOI] [PubMed] [Google Scholar]
- 32.Hesselbrock M, Easton C, Bucholz KK, Schuckit M, Hesselbrock V. A validity study of the SSAGA—a comparison with the SCAN. Addiction. 1999;94:1361–1370. doi: 10.1046/j.1360-0443.1999.94913618.x. [DOI] [PubMed] [Google Scholar]
- 33.Barrett JCFB, Maller J, Daly MJ. Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics. 2005;21:263–265. doi: 10.1093/bioinformatics/bth457. [DOI] [PubMed] [Google Scholar]
- 34.Laird NM, Horvath S, Xu X. Implementing a unified approach to family-based tests of association. Genet Epidemiol. 2000;19(Suppl 1):S36–S42. doi: 10.1002/1098-2272(2000)19:1+<::AID-GEPI6>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
- 35.Horvath S, Xu X, Laird NM. The family based association test method: strategies for studying general genotype–phenotype associations. Eur J Hum Genet. 2001;9:301–306. doi: 10.1038/sj.ejhg.5200625. [DOI] [PubMed] [Google Scholar]
- 36.Spielman RS, McGinnis RE, Ewens WJ. Transmission test for linkage disequilibrium: the insulin gene region and insulin-dependent diabetes mellitus (IDDM) Am J Hum Genet. 1993;52:506–516. [PMC free article] [PubMed] [Google Scholar]
- 37.Rabinowitz D, Laird N. A unified approach to adjusting association tests for population admixture with arbitrary pedigree structure and arbitrary missing marker information. Hum Hered. 2000;50:211–223. doi: 10.1159/000022918. [DOI] [PubMed] [Google Scholar]
- 38.Lazzeroni LC, Lange K. A conditional inference framework for extending the transmission/disequilibrium test. Hum Hered. 1998;48:67–81. doi: 10.1159/000022784. [DOI] [PubMed] [Google Scholar]
- 39.Storey JD. A direct approach to false discovery rates. J R Stat Soc B. 2002;64:479–498. [Google Scholar]
- 40.Bierut LJ, Strickland JR, Thompson JR, Afful SE, Cottler LB. Drug use and dependence in cocaine dependent subjects, community-based individuals, and their siblings. Drug Alcohol Depend. 2008;95:14–22. doi: 10.1016/j.drugalcdep.2007.11.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Dudbridge F. Pedigree disequilibrium tests for multilocus haplotypes. Genet Epidemiol. 2003;25:115–221. doi: 10.1002/gepi.10252. [DOI] [PubMed] [Google Scholar]
- 42.Hosmer DW, Lemeshow S. Applied Logistic Regression. New York: Wiley; 1989. [Google Scholar]
- 43.Muller PY, Janovjak H, Miserez AR, Dobbie Z. Processing of gene expression data generated by quantitative real-time RT-PCR. Biotechniques. 2002;32:1372–1374. 1376, 1378–1379. [PubMed] [Google Scholar]
- 44.Bierut LJ, Stitzel JA, Wang JC, Hinrichs AL, Grucza RA, Xuei X, et al. Variants in the nicotinic receptors alter the risk for nicotine dependence. Am J Psychiatry. 2008 doi: 10.1176/appi.ajp.2008.07111711. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Morley M, Molony CM, Weber TM, Devlin JL, Ewens KG, Spielman RS, et al. Genetic analysis of genome-wide variation in human gene expression. Nature. 2004;430:743–747. doi: 10.1038/nature02797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Cheung VG, Spielman RS, Ewens KG, Weber TM, Morley M, Burdick JT. Mapping determinants of human gene expression by regional and genome-wide association. Nature. 2005;437:1365–1369. doi: 10.1038/nature04244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Flora A, Schulz R, Benfante R, Battaglioli E, Terzano S, Clementi F, et al. Transcriptional regulation of the human alpha5 nicotinic receptor subunit gene in neuronal and non-neuronal tissues. Eur J Pharmacol. 2000;393:85–95. doi: 10.1016/s0014-2999(00)00040-6. [DOI] [PubMed] [Google Scholar]
- 48.Flora A, Schulz R, Benfante R, Battaglioli E, Terzano S, Clementi F, et al. Neuronal and extraneuronal expression and regulation of the human alpha5 nicotinic receptor subunit gene. J Neurochem. 2000;75:18–27. doi: 10.1046/j.1471-4159.2000.0750018.x. [DOI] [PubMed] [Google Scholar]
- 49.Buckland PR, Hoogendoorn B, Coleman SL, Guy CA, Smith SK, O’Donovan MC. Strong bias in the location of functional promoter polymorphisms. Hum Mutat. 2005;26:214–223. doi: 10.1002/humu.20207. [DOI] [PubMed] [Google Scholar]
- 50.Nashmi R, Lester HA. CNS localization of neuronal nicotinic receptors. J Mol Neurosci. 2006;30:181–184. doi: 10.1385/JMN:30:1:181. [DOI] [PubMed] [Google Scholar]
- 51.Gotti C, Clementi F. Neuronal nicotinic receptors: from structure to pathology. Prog Neurobiol. 2004;74:363–396. doi: 10.1016/j.pneurobio.2004.09.006. [DOI] [PubMed] [Google Scholar]
- 52.Gerzanich V, Wang F, Kuryatov A, Lindstrom J. Alpha 5 Subunit alters desensitization, pharmacology, Ca + + permeability and Ca + + modulation of human neuronal alpha 3 nicotinic receptors. J Pharmacol Exp Ther. 1998;286:311–320. [PubMed] [Google Scholar]