Skip to main content
Human Brain Mapping logoLink to Human Brain Mapping
. 2015 May 4;36(8):3007–3019. doi: 10.1002/hbm.22824

CREB‐BDNF pathway influences alcohol cue‐elicited activation in drinkers

Jiayu Chen 1,, Kent E Hutchison 1,2, Vince D Calhoun 1,3, Eric D Claus 1, Jessica A Turner 1,4, Jing Sui 1,5, Jingyu Liu 1,3
PMCID: PMC4969622  NIHMSID: NIHMS804803  PMID: 25939814

Abstract

Alcohol use disorder (AUD) is suggested to have polygenic risk factors and also exhibits neurological complications, strongly encouraging a translational study to explore the associations between aggregates of genetic variants and brain function alterations related to alcohol use. In this study, we used a semiblind multivariate approach, parallel independent component analysis with multiple references (pICA‐MR) to investigate relationships of genome‐wide single nucleotide polymorphisms with alcohol cue‐elicited brain activations in 315 heavy drinkers, where pICA‐MR assesses multiple reference genes for their architecture and functional influences on neurobiological conditions. The genetic component derived from the cAMP‐response element‐binding protein and ‐brain derived neurotrophic factor (CREB‐BDNF) pathway reference was significantly associated (r = −0.38, P = 3.98 × 10−12) with an imaging component reflecting hyperactivation in precuneus, superior parietal lobule, and posterior cingulate for drinkers with more severe alcohol dependence symptoms. The highlighted brain regions participate in many cognitive processes and have been robustly implicated in craving‐related studies. The genetic factor highlighted the CREB and BDNF references, as well as other genes including GRM5, GRM7, GRID1, GRIN2A, PRKCA, and PRKCB. Ingenuity Pathway Analysis indicated that the genetic component was enriched in synaptic plasticity, GABA, and protein kinase A signaling. Collectively, our findings suggest that genetic variations in various neural plasticity and signaling pathways partially explain the variance of precuneus reactivity to alcohol cues which appears to be associated with AUD severity. Hum Brain Mapp 36:3007–3019, 2015. © 2015 Wiley Periodicals, Inc.

Keywords: alcohol use disorder, single nucleotide polymorphism, functional magnetic resonance imaging, parallel independent component analysis with multiple references, precuneus, cAMP‐response element‐binding, brain derived neurotrophic factor

INTRODUCTION

Alcohol use disorders (AUD) present a substantial health and economic issue. The lifetime prevalence is estimated to be 17.8% for alcohol abuse and 12.5% for alcohol dependence [Hasin et al., 2007]. Genetic factors have been shown to affect liability to AUD, with the heritability estimated to be 40–60% while the remainder of variance is likely attributable to environmental factors [Mayfield et al., 2008; Prescott and Kendler, 1999]. Great efforts have been made toward unraveling the genetic etiology of AUD. Candidate gene and unbiased genome‐wide association studies (GWAS) provided evidence for a number of susceptibility variants, highlighting genes involved in various neural signaling pathways, including dopaminergic [Conner et al., 2005], glutamatergic [Schumann et al., 2008], and GABAergic [Bierut et al., 2010] systems. Genes encoding alcohol dehydrogenase (ADH) enzymes playing a key role in alcohol metabolism are also implicated in AUD vulnerability [Luo et al., 2007].

Despite the growing knowledge on susceptibility loci contributing to individual differences in drinking behavior, the genetic findings generally have modest effect sizes at best [Heath et al., 2011; Kapoor et al., 2013]. For instance, in a large GWAS of alcohol dependence where thousands of participants were included for investigations [Bierut et al., 2010], no single nucleotide polymorphism (SNP) could pass the genome‐wide significance threshold of 5.00 × 10−8 [Stranger et al., 2011]. Instead the highlighted 15 SNPs yielded suggestive associations with P < 10−5, yet none of them replicated findings of a previous GWAS [Treutlein et al., 2010]. This is a common challenge in complex trait mapping. Indeed, like many other complex disorders, AUD is suggested to have polygenic risk factors [Johnson et al., 2006; Salvatore et al., 2014], such that the underlying genetic architecture involves many variants with modest individual effect sizes, which may interact to confer the liability. Understanding the mechanism of genetic effects becomes even more complicated due to phenotypic heterogeneity where genetic variants can exert influences on various phenotypes through different biological mechanisms [Wong and Schumann, 2008].

In this work, we used a novel semiblind multivariate approach, parallel independent component analysis with multiple references (pICA‐MR) [Chen et al., 2014], to investigate the genetic basis underlying brain function related to AUD. Specifically, functional magnetic resonance imaging (fMRI) data were collected from participants exposed to a well‐established alcohol cue paradigm [Claus et al., 2011; Filbey et al., 2008]. The method, pICA‐MR is an extension of pICA‐R [Chen et al., 2013] to assess multiple genetic references in one single analysis. PICA‐MR extracts independent components from the imaging and genetic modalities separately to assess aggregate effects of multiple variables, posing a promising model for polygenicity. It also enhances intermodality associations, providing a translational framework for exploring genetic underpinnings of neuronal functions, which might ultimately lead to clinical manifestations of the disorder. In addition, genetic references (i.e., multiple sets of SNPs) are incorporated to pinpoint components of particular attribute and pICA‐MR delineates the references' architecture and functional influence in a data‐driven manner. This method is expected to help characterize mechanisms of interest in high‐dimensional complex data.

MATERIALS AND METHODS

Participants

A total of 326 participants enrolled in the study to investigate genetic and neurobiological traits related to AUD [Claus et al., 2011]. The University of New Mexico Human Research Review Committee approved the study. All the participants were recruited from the greater Albuquerque metropolitan region and provided written informed consent. The inclusion criterion was based on alcohol consumption, requiring participants to drink at least five times in the past month with at least five (for men) or four (for women) drinks per drinking occasion. The exclusion criteria included a history of severe alcohol withdrawal, brain‐related medical problems, or symptoms of psychosis. In addition, participants were required to be sober during the data collection, with a breath alcohol concentration of 0.00. Sixty‐one and eighty percentage of participants met Diagnostic and Statistical Manual of Mental Disorders Fourth Edition (DSM‐IV) criteria for current alcohol dependence and alcohol abuse, while 70 and 93% met lifetime criteria for alcohol dependence and alcohol abuse. After preprocessing, 315 participants (220 males and 95 females) for which good quality fMRI and SNP data were collected, were included in the analysis. Table 1 provides the demographic information.

Table 1.

Demographic information of participants

Number of participants
Male (220) Female (95)
Caucasian 99 43
African American 4 2
Asian 2 0
Latino 54 28
Native American 13 3
Mixed 47 19
Unreported 1 0
Age of participants
Male Female
Mean ± SD 31.74 ± 9.43 32.52 ± 10.58
Range 21–56 21–55

Data Collection and Preprocessing

Behavioral assessment

Participants completed a variety of questionnaires at a baseline assessment appointment, including the Alcohol Dependence Scale (ADS) [Skinner and Horn, 1984], the Alcohol Use Disorder Identification Test (AUDIT) [Babor et al., 2001], and the Impaired Control Scale (ICS) for alcohol [Heather et al., 1998]. Symptom counts for current/lifetime alcohol dependence/abuse were tallied from the Structured Clinical Interview for DSM Disorders. Years of regular drinking, age at first drink, and age that regular drinking began were derived from a drinking history questionnaire. Depression was assessed using the Beck Depression Inventory. We excluded relatively incomplete measures where data were missing for more than 25 participants. Finally, a total of 23 behavioral measures were investigated for associations with identified imaging and genetic components, as listed in Table 2. The missing ratio was no greater than 4/315. It should be noted that most of these behavioral measures showed significant associations with age, except for AgeFirstDrink, EStress‐tot, and Stress‐tot.

Table 2.

Alcohol dependence assessment

Assessment Sub‐category Description Score range Score mean ± SD
ADS ADS‐con Loss of behavior control 0–18 7.87 ± 4.16
ADS‐obs Obsessive drinking style 0–5 1.07 ± 1.33
ADS‐per Psychoperceptual withdrawal 0–13 2.00 ± 2.34
ADS‐phy Psychophysical withdrawal 0–10 2.63 ± 2.16
ADS‐tot Total ADS 1–43 13.57 ± 8.24
AUDIT AUDIT‐tot Total AUDIT score 4–39 19.02 ± 7.75
AUDIT‐consump Alcohol consumption total 0–12 9.21 ± 2.06
AUDIT‐dep Alcohol dependence total 0–12 3.78 ± 3.04
AUDIT‐probs Alcohol problems total 0–16 5.96 ± 3.99
ICS ICS‐tot Total ICS 0–96 45.65 ± 21.24
ICS‐ac Attempted control 0–20 8.13 ± 4.90
ICS‐fc Failed control 0–40 19.12 ± 9.64
ICS‐pc Perceived control 0–40 18.40 ± 9.67
Alcohol symptom count PA‐count Past alcohol abuse symptom count 0–4 2.01 ± 1.24
CA‐count Current alcohol abuse symptom count 0–4 1.66 ± 1.19
PD‐count Past alcohol dependence symptom count 0–7 3.46 ± 2.21
CD‐count Current alcohol dependence symptom count 0–7 3.50 ± 2.44
Drinking history AgeDrink Probable age that regular drinking first occurred 11–39.75 18.96 ± 4.40
YearsDrink Probable number of years of regular drinking 1–41 13.01 ± 9.17
AgeFirstDrink Probable age of first drink 4–28 14.39 ± 2.88
Stress EStress‐tot Total early stress for ages before 19 0–12 3.04 ± 2.61
Stress‐tot Early stress total (all ages reported) 0–12 3.46 ± 2.77
BDI‐tot Total Beck Depression Inventory 0–47 11.65 ± 9.81

Functional MRI

Brain activation data were collected from 317 participants during an alcohol craving task as described in [Claus et al., 2011; Filbey et al., 2008], see Supporting Information Figure S1 for the schematic of a single taste cue trial. Participants were exposed to 1 ml of alcoholic (individual preferred) or juice (litchi) beverages pseudorandomly presented during the MRI scans. Each taste cue trial sequentially consisted of a 2‐s “Ready” prompt, a 24‐s taste cue presentation, and a 16‐s washout period. During the cue presentation, participants tasted the presented beverage (second 1–10 and 12–22) and then swallowed (second 10–12 and 22–24). No stimuli were presented during the washout and participants viewed the word “Rest”. Two 9‐min runs were conducted for each participant, with a single run spanning 12 trials, six for each tastant. A 3T Siemens Trio was used for the data collection. The echo‐planar gradient‐echo pulse sequence was configured as follows: TR = 2 s, TE = 29 ms, flip angle = 75°, voxel size = 3.75 × 3.75 × 4.55 mm3. The collected fMRI data were preprocessed with Statistical Parametric Mapping 5 (SPM5, http://www.fil.ion.ucl.ac.uk/spm). Standard motion correction was performed and images were normalized to the Montreal Neurological Institute template [Jenkinson et al., 2002] and resliced to 3 × 3 × 3 mm3. An 8‐mm full‐width half‐maximum Gaussian kernel was used for spatial smoothing. Finally, alcohol versus juice contrast images were extracted from a total of 54,937 voxels with no missing values and used for subsequent association analyses. One participant outlier was excluded due to a low correlation (>3‐SD) with the mean activation across all the participants.

SNP data

Saliva samples were collected from 324 participants for DNA extraction. Genotyping for all participants was performed at the Mind Research Network using the Illumina Infinium Human 1M‐Duo assay spanning 1,199,187 SNP loci. BeadStudio was used to make the final genotype calls. A series of standard quality control procedures were then performed with PLINK [Purcell et al., 2007]. Specifically, SNPs and participants were first examined for a genotyping rate threshold of 95%; SNPs were excluded if they deviated from Hardy–Weinberg Equilibrium with a threshold of 10−6 or failed to be missing at random with a threshold of 10−10; Two participants were excluded due to high heterozygosity (3‐SD greater than the mean); Another two participants were excluded due to relatedness with identity‐by‐descent values greater than 0.1875; Minor allele frequency threshold was set to 0.05. After the quality control, discrete numbers were assigned to the categorical genotypes: 0 for no minor allele, 1 for one minor allele, and 2 for two minor alleles. Subsequently, we replaced the missing genotypes using high linkage disequilibrium (LD) loci if available (correlation > 0.80), excluded 18,809 SNPs with a missing ratio greater than 1%, and replaced the rest missing genotypes with the major alleles of individual loci. The resulting 717,129 autosomal SNPs were then used for population stratification correction by principal component analysis (PCA), and three principal components (PC1, 2, and 4) differed significantly among ethnic groups (P = 9.85 × 10−79, 3.23 × 10−86 and 3.21 × 10−55, respectively) while exhibiting no significant associations with drinking behavior measures. Then the data reconstructed from these three ethnicity‐related PCs were subtracted from the original data to eliminate the influence of population structure. In the corrected SNP data, a Q‐Q plot for P‐values of association with the AUDIT‐tot score tested against a uniform distribution showed no clear indication of population structure (Supporting Information Fig. S2).

Association Analysis

The fMRI contrast images were analyzed in conjunction with the SNP data using pICA‐MR [Chen et al., 2014], which extends the pICA [Liu et al., 2009] and pICA‐R [Chen et al., 2013] (Fusion ICA Toolbox, http://mialab.mrn.org/software/fit) approach to accommodate multiple genetic references. Figure 1 shows the flowchart of pICA‐MR. As a multivariate approach, pICA‐MR decomposes the two datasets, X1 and X2, into linear combinations of underlying components separately and in parallel, as depicted in Eq. (1). S, A, and W denote the component, mixing, and unmixing matrices, respectively. The subscript d runs from 1 to 2, denoting two data modalities.

Xd=AdSdSd=WdXd, Ad=Wd1, d = 1, 2 (1)

Figure 1.

Figure 1

A flowchart of pICA‐MR.

The data decomposition primarily builds on infomax ICA [Amari, 1998; Bell and Sejnowski, 1995], where independent sets of covarying variables are organized into different components. The mathematical model is described in Eqs. (2) and (3). F 1 is the objective function of infomax for the imaging modality (modality 1), where independence among components is optimized by maximizing the entropy (H). f y(Y) is the probability density function of the sigmoid function Y, which is chosen based on the form of the expected underlying source distribution [Lee et al., 1999]. E is the expected value and W0 is the bias vector. In contrast to F1, the objective function for the genetic modality (modality 2) F2, is modified based on Infomax so that components are not only independent but closely resemble the reference matrix r. Each row of r represents a reference vector, which is of the same dimension as the SNP data with nonzero elements representing a group of reference loci likely contributing in a coordinated manner. PICA‐MR then calculates the Euclidean distances between each of the reference vectors ri (ith row of r) and all the components (only using nonzero reference loci ri, a subvector of ri) to determine the closest one as the corresponding constrained component, whose distance ( S2,kiri22) is further minimized. S2,ki denotes a subvector of the ri‐constrained component S2,ki (the ki th row of S2) and is computed by W2,ki (the kith row of W2) multiplying X2 (a submatrix of X2). ǁ·ǁ2 represents the L2‐norm Euclidian distance, and λ is the weight parameter. Note that pICA‐MR incorporates the reference constraint in a way similar to pICA‐R, where the distance between the component and reference vector is optimized for the reference loci only. This is to enable a semiblind decomposition which is partially constrained to highlight the presumed causal loci in the resulting components while still allowing the remaining loci to show their own importance driven by the data. As in pICA‐R, the weight parameter λ is adaptively adjusted to avoid false positives due to the reference constraint dominating over the independence metric. In the case when λ is completely de‐emphasized, pICA‐MR converges to regular pICA. Conversely, in contrast to pICA‐R dealing with one single reference (vector), in pICA‐MR, multiple reference vectors can constrain the same or different components, reflecting functional architecture in a data‐driven manner. Finally, the intermodality association function F3 maximizes the correlations computed over the columns of the loading matrices A1 and A2.

Yd=11+eUd,Ud=WdXd+Wd0F1=max{H(Y1)}=max{E[ln fy1(Y1)]}F2=max{λH(Y2)+(1λ)[i=1Idist2(ri,|S2,ki|)]} =max{λ(E[ln fy2(Y2)])+(1λ)(i=1I|||W2,kiX2|ri||22)}  (2)
F3=max{i,jCorr2(A1,i,A2,j)}=max{i,jCov2(A1,i,A2,j)Var(A1,i)Var(A2,j)} (3)

The three objective functions are solved through iteratively updating the unmixing matrices W1 and W2 using gradient optimization. The updating rules are similar to [Chen et al., 2013]. Cautions have been exercised to avoid overfitting and false positive findings. First, a threshold of correlation is imposed such that only component pairs exhibiting above‐threshold correlations are further optimized. In this work, we used a conservative threshold of 0.25 for the sample size of 315, requiring that the component pair showed a significant correlation before the optimization. In addition, an adaptive learning is incorporated through monitoring the components' independence. The optimization of correlation and distance will be depressed if it compromises the metric of independence [Chen et al., 2013; Liu et al., 2009]. Simulations suggest that pICA‐MR effectively identifies similarity among multiple references and the detection power is comparable to pICA‐R and hence significantly improved compared to blind methods [Chen et al., 2014]. It should be noted that currently pICA‐MR is configured to apply references only to the genetic modality, as the fMRI modality is less challenging and in general, the regular ICA is competent. However, the reference constraint can be extended to both modalities if needed.

In this work, the number of fMRI components was estimated by minimum description length (MDL) [Rissanen, 1978] on uncorrelated voxels, yielding the most efficient representation of the original data. MDL proves an effective metric for neuroimaging data as signals of interest in general carry much more variance than noises. However, it is less applicable to SNP data, as most genetic factors account for small amounts of variance, except for those related to population structure. Therefore, we developed an approach to estimate the component number based on components' consistency for SNP data, such that the most stable decomposition could be obtained [Chen et al., 2012].

Six sets of strongly hypothesized genes were selected based on previously reported associations with AUD, as listed in Table 3. Each set consisted of a few genes which might affect brain function in a coordinated manner, and was tested separately. CHRNA3, CHRNA5, and CHRNB4 are located closely in 15q24‐25 and are among the most replicated findings for smoking behaviors [Bierut, 2010; Caporaso et al., 2009] while also implicated in risk of alcohol dependence [Wang et al., 2009]. Cytogenetic band 4p12 hosting GABA receptor genes were shown to present moderate odds ratios for alcohol dependence in a GWAS [Bierut et al., 2010] and independent contributions from individual receptors have been suggested [Enoch et al., 2009]. Another cluster of GABA receptor genes resides in 5q34. Both mouse and human studies provided evidence for their modulatory roles in alcohol dependence [Radel et al., 2005]. ADH is a primary enzyme involved in alcohol metabolism and variants in the gene cluster ADH1AADH1BADH1C have been found associated with risk of alcoholism [Edenberg and Foroud, 2006; Luo et al., 2007]. The opioidergic system is considered as mediating drug‐induced feelings and playing an important role in substance rewarding properties [Gianoulakis, 2001]. Related polymorphisms have been identified as associated with AUD [Filbey et al., 2008b; Zhang et al., 2008]. cAMP‐response element‐binding protein (CREB) is a key transcriptional factor for neuronal growth and regulates the expression of brain derived neurotrophic factor (BDNF). These two genes have been found to interact in a variety of brain regions and play a critical role in addiction [Carlezon et al., 2005; Crews et al., 2007; Pandey, 2003]. A reference matrix was generated for each gene set with each row representing a group of reference SNPs derived from a single gene based on LD structure (r 2 > 0.2, [Ripke et al., 2011]), as described in [Chen et al., 2013]. Specifically, in this work, most of the reference genes spanned tens of SNPs forming a single LD block in our data, which were directly used to generate the reference matrix. One exception was CREB5 hosting 228 SNPs, for which multiple LD blocks were identified and represented in multiple reference matrices. The associations of all the reference SNPs with four major behavioral measures (ADS‐tot, AUDIT‐tot, ICS‐tot, and BDI‐tot) were provided in Supporting Information Table S1.

Table 3.

Tested genetic references

Reference Genes Literature
15q24‐25 CHRNA3, CHRNA5, CHRNB4 Bierut, 2010; Caporaso et al., 2009; Wang et al., 2009
4p12 GABRA4, GABRA2, GABRG1, GABRB1 Bierut et al., 2010; Enoch et al., 2009
5q34 GABRB2, GABRA6, GABRA1, GABRG2 Radel et al., 2005
4q23 ADH1A, ADH1B, ADH1C Edenberg et al., 2006; Luo et al., 2007
Opioid system OPRM1, OPRK1, OPRD1 Filbey et al., 2008; Zhang et al., 2008
CREB‐BDNF CREB1, CREB5, BDNF Carlezon et al., 2005; Crews et al., 2007; Pandey, 2003

To assess the fidelity of the identified association, we applied 10‐fold validation and permutation test. In the 10‐fold validation, the participants were divided into 10 subsets. Then in each of the 10 runs, nine subsets were included and pICA‐MR was applied to the 90% of the original data to extract components and evaluate the correlations. This was to confirm that the observed SNP‐fMRI association was consistently presented by all the participants, instead of being raised by specific subsets. In the 1,000‐run permutation test, all the participants were included and permuted in each run and sent for pICA‐MR analysis. Then based on the top associated component pair of each run, we calculated the tail probability to evaluate the significance level of the identified SNP‐fMRI association. This permutation test investigated the possibility of the identified SNP‐fMRI association occurring in randomly rearranged participants. It also helped evaluate overfitting which may manifest as producing inflated correlations from permuted data.

To understand the functional influences, the identified SNP and fMRI components were further investigated for associations with all the behavioral measures listed in Table 2. Multiple regression analysis was conducted to control for sex and population structure. Particularly, the covariates of population structure included both the self‐reported race/ethnicity (Table 1) and the top three PCs extracted from the corrected SNP data [Ripke et al., 2011]. False discovery rate (FDR) was applied to account for multiple tests given the moderate to high correlations among most of the behavioral measures.

RESULTS

The fMRI (315 × 54,937) and SNP (315 × 717,129) data were analyzed with pICA‐MR. The number of components was estimated to be 15 for fMRI and 11 for SNP, respectively. Among the tested gene sets, CREB‐BDNF elicited a significant SNP‐fMRI correlation of −0.38 (P = 3.98 × 10−12), passing the Bonferroni threshold of 5.05 × 10−5 which corrected for the six tested genetic references and the combinations of all extracted components (15 sMRI × 11 SNP) in each test. The three genes BDNF, CREB1, and CREB5 were identified to constrain the same SNP component. Supporting Information Table S2 summarizes the recruited reference SNPs, which consisted of all the genotyped loci in BDNF (15 SNPs) and CREB1 (20 SNPs), and an LD block spanning 20 SNPs in CREB5. After regressing out controlling variables (age, sex, population structure), the SNP‐fMRI association remained significant, exhibiting a correlation of −0.35 (P = 2.19 × 10−10), as shown in Figure 2a. The identified SNP‐fMRI pair exhibited stable correlations in the 10‐fold validation, ranging from 0.23 to 0.33 with a median of 0.27. In the 1,000‐run permutation, only one permuted sample exhibited a SNP‐fMRI association stronger than that observed in the original data, yielding a significant P‐value of 0.001 for our finding. Other tested gene sets presented insignificant or marginally significant SNP‐fMRI associations, as summarized in Supporting Information Table S3.

Figure 2.

Figure 2

Scatter plots of: (a) the fMRI and SNP loadings; (b) the fMRI loading and CD‐count. [Color figure can be viewed in the online issue, which is available at http://wileyonlinelibrary.com.]

The identified fMRI component, with FDR control, was significantly associated with a number of behavioral measures, including CD‐count, ICS‐fc, ICS‐tot, and PD‐count, but not YearsDrink. The most significant association was observed from CD‐count, exhibiting a correlation of 0.24 (P = 1.73 × 10−5) after regressing out sex and population structure, as shown in Figure 2b. In addition, the correlation remained significant (r = 0.18, P = 1.31 × 10−3) after age was further regressed out. This, together with the fact that no significant association was observed between the fMRI component and YearsDrink, suggested that the variance in activation reflected more current dependence symptoms rather than cumulative alcohol use. The fMRI loadings also showed significant associations with ICS‐fc, ICS‐tot, and PD‐count (correlations of 0.24, 0.23, and 0.21, respectively), all of which highly correlated with CD‐count (correlations of 0.68, 0.72, and 0.82, respectively). Due to this collinearity, the individual effect could not be disentangled and we chose to focus on the most significantly associated symptom CD‐count in the following discussion. Figure 3 shows the spatial map of the identified fMRI component thresholded at |Z| > 2. Table 4 summarizes the Talairach atlas labels [Lancaster et al., 2000] of the mapped brain network, including precuneus, superior, and inferior parietal lobules, as well as posterior cingulate cortex (PCC). These highlighted brain regions remained highly consistent with a |Z| threshold ranging from 2 to 3.

Figure 3.

Figure 3

Spatial map of brain network for the identified fMRI component (|Z| > 2). [Color figure can be viewed in the online issue, which is available at http://wileyonlinelibrary.com.]

Table 4.

Talairach labels of identified brain regions (|Z| > 2)

Brain region Brodmann area L/R volume (cm3) L/R random effects, max Z (x,y,z)
Precuneus 7, 19, 39, 31 16.8/14.1 9.18(0,−58,61)/9.41(3,−58,64)
Superior Parietal Lobule 7, 5 8.9/7.6 8.55(−3,−67,56)/8.72(6,−64,58)
Postcentral Gyrus 7, 5, 3, 2, 40, 1 5.3/4.5 8.06(0,−46,66)/9.03(3,−52,66)
Inferior Parietal Lobule 40, 7, 39 3.5/3.5 4.28(−39,−49,61)/5.11(39,−52,58)
Cuneus 19, 18, 7, 30 2.7/3.7 4.53(0,‐82,40)/4.59(27,−83,37)
Paracentral Lobule 5, 4, 6, 7 2.9/1.8 7.28(0,−46,63)/5.71(3,−37,68)
Posterior Cingulate 29, 30, 23 1.6/1.0 3.25(−6,−41,5)/2.99(6,−41,5)

With FDR control, the identified SNP component did not exhibit any significant association with the behavioral measures. However, a trend of correlation was observed with AUDIT‐dep (r = −0.14, P = 1.14 × 10−2, uncorrected). Figure 4 shows a Manhattan plot of weights of loci for the identified SNP component, where the z‐score threshold of 3.13 is marked to present the selection of top 2,020 SNPs (for details, see Supporting Information Fig. S3). Supporting Information Table S4 provides a summary of all the selected top contributing SNPs, including SNP position, hosting gene, and z‐scored component weight. One thousand and nineteen out of the 2,020 top contributing SNPs were mapped to 457 unique genes, which were used for Ingenuity Pathway Analysis (IPA: Ingenuity® Systems, http://www.ingenuity.com). IPA revealed a number of enriched canonical pathways, including synaptic long term depression (LTD, 1.70 × 10−5) and potentiation (LTP, 5.89 × 10−3), CREB Signaling in Neurons (6.31 × 10−4), protein kinase A (PKA) signaling (1.26 × 10−2), as well as GABA receptor signaling (2.24 × 10−2), as summarized in Table 5. IPA also indicated a significant enrichment of psychiatric disorders in our finding, including bipolar disorder (7.56 × 10−4), schizophrenia (5.50 × 10−3), and major depression (4.37 × 10−2). The identified genes were also significantly overrepresented in neuritogenesis (2.81 × 10−4) and other developmental functions (see Supporting Information Table S5 for a complete summary of pathway analyses). As there is no golden rule to select top contributing SNPs, we also tested top 1,000–5,000 SNPs and similar pathway analyses results were obtained (Supporting Information Table S6).

Figure 4.

Figure 4

Manhattan plot for the identified SNP component. The black line represents the z‐score threshold of 3.13. [Color figure can be viewed in the online issue, which is available at http://wileyonlinelibrary.com.]

Table 5.

Pathway analyses

IPA canonical pathway Genes P‐value
Synaptic long term depression GNA14,ITPR1,GRM5,GRM7,GNAI3,GRID1,PLB1, RYR3,LYN,GNAT2,PPP2R1B,PRKCB,PRKCA 1.70E‐05
CREB signaling in neurons GRM5,GRM7,GNAI3,GRIN2A,GRID1,GNAT2, PIK3CD,GNA14,ITPR1,CREB5,PRKCA,PRKCB 6.31E‐04
Axonal guidance signaling LRRC4C,ITSN1,KALRN,RAC1,GNA14,ROBO1, ADAMTS2,GNAI3,SRGAP3,NTRK3,DCC,ADAM19,RTN4, ADAM23,GNAT2,PIK3CD,WNT5B,PRKCA,PRKCB 5.13E‐03
Synaptic long term potentiation GRM5,GRM7,GRIN2A,GNA14,ITPR1,CREB5,PRKCA, PRKCB 5.89E‐03
Protein kinase A signaling PTPN7,PTPRD,PTPN3,ITPR1,NFKB1,CREB5,PDE1C, GNAI3,HHAT,ADD3,RYR3,DCC,PTPRS,PDE8B, PRKCB,PRKCA 1.26E‐02
GABA receptor signaling GABRG3,GABRR3,AP2M1,AP1B1 2.24E‐02

To further confirm the genetic influence on brain function alterations, we performed a regression analysis between the fMRI and SNP loadings while controlling for age, sex, population structure as well as associated behavioral measures of CD‐count, ICS‐fc, ICS‐tot, and PD‐count. The SNP component still showed a significant regression effect (P = 9.18 × 10−10) on the fMRI component.

DISCUSSION

In this work, pICA‐MR was used to identify SNP‐fMRI associations under the guidance of genetic references. Compared to pICA‐R which deals with reference SNPs from a single gene, pICA‐MR has been extended to accommodate a set of genes which might contribute in a coordinated manner. Simulations showed reliable convergence for all the tested scenarios [Chen et al., 2014] and recommended that pICA‐MR reference selection should be based on a relatively strong hypothesis which deserves further investigation in the following aspects. First, pICA‐MR examines reference SNPs in a multivariate framework for potential aggregate effects. Second, pICA‐MR investigates how the references constrain the same or different genetic components and their further associations with imaging components. This may help further our understanding of genetic influence across various disrupted neural mechanisms involved in a complex disorder. Last but not least, pICA‐MR performs a semi‐blind decomposition where the genetic references serve as guidance and other variables are allowed to show their own importance driven by the data. In this way, we expect to obtain a more complete picture of the underlying genetic architecture. Specifically, in this work, we tested six strongly hypothesized gene sets consistently implicated in alcohol studies. While there exists a large library of alcohol related genes, they will be left for future investigations.

The fMRI loadings exhibited a positive correlation with CD‐count, together with positive activations of the component, indicating that participants experiencing more severe alcohol dependence symptoms had higher regional activations when exposed to the taste of alcohol. The brain network comprised precuneus, superior and inferior parietal lobules, and PCC, as shown in Figure 3 and listed in Table 4. The brain network covered the midline, which might be attributable to spatial smoothing using an 8‐mm Gaussian kernel. Precuneus belongs to associative cortices and is known to be involved in a wide range of highly integrated tasks, including episodic memory retrieval, self‐referential processes, and consciousness [Cavanna and Trimble, 2006]. Its involvement in a variety of processes builds on its anatomical wide‐spread connections with both adjacent areas such as superior parietal lobule and PCC, and frontal lobes including prefrontal cortex and anterior cingulate cortex [Cavanna and Trimble, 2006]. Although not generally targeted for AUD, precuneus and parietal regions have been robustly implicated in craving studies, where activation elicited by drug‐related cues has been found associated with severity of dependence [Claus et al., 2011; Liu et al., 2013; Park et al., 2007; Tapert et al., 2004]. The drug‐cue‐elicited activation in precuneus is possibly a reflection of its recruitment in episodic memory retrieval, as the triggering of craving can be considered as a conditioned response where the recollection of past experience as episodic memories serves as a conditioned cue [Robbins et al., 2008]. PCC is a key player in the salience network [Sutherland et al., 2012] and frequently implicated in the processing of drug‐related stimuli [Tapert et al., 2004; Wrase et al., 2007]. Its functional alteration has proven to underlie the concurrent use of alcohol and tobacco [Liu et al., in press]. Of particular interest, as shown in a meta‐analysis on fMRI studies of alcohol cue reactivity, brain activation in precuneus and PCC, instead of the mesolimbic system, is selectively enhanced and most effectively differentiates cases from controls [Schacht et al., 2013]. Overall, the identified brain network echoes considerable similar findings and deserves more attention to elucidate the neuropathology of addiction.

The associated genetic component elicited by the three genes, BDNF, CREB1, and CREB5, negatively correlated with the fMRI component, indicating that participants carrying lower loadings on the SNP component presented higher brain activation in the identified precuneus and parietal regions. Pathway analyses delineated a complex genetic architecture emphasizing neural plasticity and signaling pathways based on the 457 genes highlighted in the component. A meta‐analysis of mouse models identified 3,800 genes differentially expressed between models of high and low amounts of alcohol consumption [Mulligan et al., 2006], suggesting that thousands of genes are involved in the pathology of AUD. Our findings appeared to be in line with the animal study and further suggested that the CREB‐BDNF pathway likely serves as a hub of the polygenetic effect related to AUD, as illustrated below in several associated canonical pathways.

CREB functions as a transcription factor and is well known for its role in neural plasticity and long‐term memory [Carlezon et al., 2005]. BDNF is a CREB regulated gene and active in synaptic plasticity [Bramham and Messaoudi, 2005]. Together with several glutamatergic genes (GRM5, GRM7, GRID1, and GRIN2A) and protein kinase C genes (PRKCA and PRKCB), they signify the CREB signaling, synaptic LTD, and synaptic LTP pathways. Synaptic LTP and LTD are two forms of synaptic plasticity which enhances or weakens, respectively, the synchronized stimulations between neurons, thus allowing the refinement of neuronal circuits underlying learning and memory [Malenka and Bear, 2004]. It is commonly recognized that synaptic plasticity plays an important role in the development of addiction, through which use of drug progresses from impulsive to compulsive behavior [Kauer and Malenka, 2007]. A meta‐analysis further indicated that a genetic component might affect addiction through regulating synaptic plasticity [Li et al., 2008], where LTD and LTP are among the top enriched pathways for the 396 addiction‐related genes implicated in two or more independent studies. Particularly, one SNP in BDNF (rs6265_A or Val66Met, “A” represents the minor allele) has been identified as predicting relapse in alcohol dependence, where minor allele carriers showed decreased vulnerability to relapse [Wojnar et al., 2009]. This is consistent with our finding where the same SNP exhibited a positive weight, indicating that minor allele carriers presented lower brain activation which was associated with less severe alcohol dependence symptoms. The SNP in GRIN2A (rs4628972_G) and three SNPs in PRKCA (rs17688881_C, rs721429_A and rs7217618_C) presented negative weights, indicating that the participants carrying more minor alleles showed higher brain activations and more severe alcohol dependence symptoms. The opposite was observed for the rest (rs1000061_G in GRM5, rs1353832_C in GRM7, rs1863824_C in GRID1, rs8077110_T in PRKCA, and rs880824_A in PRKCB) which all presented positive weights.

As a major inhibitory neurotransmitter in the central nervous system, GABAergic signaling is implicated in addiction [Bierut et al., 2010]. It has been reported that chronic cocaine uses decrease GABAergic synapse function, such that LTP induction is not effectively suppressed at excitatory synapses [Liu et al., 2005]. Polymorphisms in GABA receptor genes are consistently identified as susceptibility loci to alcohol dependence [Bierut et al., 2010], including a number of SNPs in GABRG3 [Dick et al., 2004]. In our finding, all the three identified SNPs in GABRG3 (rs12439549_G, rs4438262_G, and rs3922613_G) and two SNPs (rs1874864_G and rs7638369_T) in GABRR3 contributed with positive weights, while the rest (rs1844934_T, rs1688378_A, and rs1492054_C) exhibited negative weights. Another pathway, cAMP‐PKA pathway, is noteworthy here, which is a primary signaling cascade modulating numerous cellular events in neurons, including synaptic plasticity [Waltereit and Weller, 2003]. It is documented that activation of the cAMP‐PKA signaling leads to increased activity of CREB [Ron and Jurd, 2005], and genetic mutation reducing cAMP‐PKA signaling results in increased sensitivity to the sedative effects of ethanol [Wand et al., 2001]. The detail of how each SNP in this pathway contributes to the genetic component can be found in Supporting Information Table S4.

It is noted that the identified genetic component was not specific to AUD. The aforementioned canonical pathways and neurodevelopmental functions such as neuritogenesis and axonal guidance (Supporting Information Table S5) are part of much broader neural processes. IPA also implicated that the associated genes were overrepresented in other neuropsychiatric diseases, including bipolar disorder, schizophrenia, and major depression. This suggests a genetic basis for the comorbidity among these disorders, which has been the focus of recent large‐scale genetic studies [Johnson et al., 2009; Lee et al., 2013]. Overall, the genetic component delineates a relatively general substrate whose disruptions may interact with other factors to trigger various diseases including AUD, which is in accord with the fact that the identified brain network also participates in various tasks not specific to those related to AUD. Future studies that combine genetics and neuroimaging across a heterogeneous sample of psychiatric disorders will be the best approach for understanding the degree to which there is overlap in genetic contributions or brain circuits among comorbid conditions.

In summary, using a novel semiblind multivariate approach, we demonstrate gene interactions within the polygenic model of AUD selected by the CREB‐BDNF gene set. The extracted SNP component exhibits enrichment in neural plasticity and signaling pathways. The captured brain network, highlighting the precuneus, superior parietal, and PCC regions, is related to alcohol dependence symptom. The significant SNP‐fMRI association indicates that the genetic factor affecting neural plasticity may influence precuneus, superior parietal, and PCC responses to alcohol cues. We speculate one likely mechanism is through learning and memory function executed in these brain regions that is implicated in addiction [Courtney et al., 2014; Robbins et al., 2008] and modulated by neural plasticity.

Supporting information

Supporting Information

Supporting Information Table 1.

Supporting Information Table 4.

ACKNOWLEDGMENT

The authors would like to thank MRN Genetics Lab for their help in collecting the genetic data.

Disclosure/Conflict of Interest: All authors declare that they have no conflict of interest.

REFERENCES

  1. Amari S (1998): Natural gradient works efficiently in learning. Neural Comput 10:251–276. [Google Scholar]
  2. Babor TF, Higgins‐Biddle JC, Sauders JB, Monteiro MG (2001): The Alcohol Use Disorders Identification Test: Guide for Use in Primary Care. Geneva, Switzerland: World Health Organization. [Google Scholar]
  3. Bell AJ, Sejnowski TJ (1995): An information‐maximization approach to blind separation and blind deconvolution. Neural Comput 7:1129–1159. [DOI] [PubMed] [Google Scholar]
  4. Bierut LJ (2010): Convergence of genetic findings for nicotine dependence and smoking related diseases with chromosome 15q24‐25. Trends Pharmacol Sci 31:46–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bierut LJ, Agrawal A, Bucholz KK, Doheny KF, Laurie C, Pugh E, Fisher S, Fox L, Howells W, Bertelsen S, Hinrichs AL, Almasy L, Breslau N, Culverhouse RC, Dick DM, Edenberg HJ, Foroud T, Grucza RA, Hatsukami D, Hesselbrock V, Johnson EO, Kramer J, Krueger RF, Kuperman S, Lynskey M, Mann K, Neuman RJ, Nothen MM, Nurnberger JI, Jr. , Porjesz B, Ridinger M, Saccone NL, Saccone SF, Schuckit MA, Tischfield JA, Wang JC, Rietschel M, Goate AM, Rice JP (2010): A genome‐wide association study of alcohol dependence. Proc Natl Acad Sci USA 107:5082–5087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bramham CR, Messaoudi E (2005): BDNF function in adult synaptic plasticity: The synaptic consolidation hypothesis. Prog Neurobiol 76:99–125. [DOI] [PubMed] [Google Scholar]
  7. Caporaso N, Gu FY, Chatterjee N, Sheng‐Chih J, Yu K, Yeager M, Chen C, Jacobs K, Wheeler W, Landi MT, Ziegler RG, Hunter DJ, Chanock S, Hankinson S, Kraft P, Bergen AW (2009): Genome‐wide and candidate gene association study of cigarette smoking behaviors. PloS One 4:e4653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Carlezon WA, Duman RS, Nestler EJ (2005): The many faces of CREB. Trends Neurosci 28:436–445. [DOI] [PubMed] [Google Scholar]
  9. Cavanna AE, Trimble MR (2006): The precuneus: A review of its functional anatomy and behavioural correlates. Brain 129:564–583. [DOI] [PubMed] [Google Scholar]
  10. Chen J, Calhoun VD, Pearlson GD, Perrone‐Bizzozero N, Sui J, Turner JA, Bustillo JR, Ehrlich S, Sponheim SR, Canive JM, Ho BC, Liu J (2013): Guided exploration of genomic risk for gray matter abnormalities in schizophrenia using parallel independent component analysis with reference. Neuroimage 83C:384–396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chen J, Calhoun VD, Ulloa AE, Liu J (2014): Parallel ICA with Multiple References: A Semi‐blind Multivariate Approach In: 2014 Annual International Conference of the IEEE Engineering in Medicine and Biology Society (Embc), Chicago, IL: pp 6659–6662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Chen J, Calhoun VD, Liu J (2012): ICA Order Selection Based on Consistency: Application to Genotype Data In: 2012 Annual International Conference of the IEEE Engineering in Medicine and Biology Society (Embc), San Diego, CA: pp 360–363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Claus ED, Ewing SW, Filbey FM, Sabbineni A, Hutchison KE (2011): Identifying neurobiological phenotypes associated with alcohol use disorder severity. Neuropsychopharmacology 36:2086–2096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Conner BT, Noble EP, Berman SM, Ozkaragoz T, Ritchie T, Antolin T, Sheen C (2005): drd2 genotypes and substance use in adolescent children of alcoholics. Drug Alcohol Depend 79:379–387. [DOI] [PubMed] [Google Scholar]
  15. Courtney KE, Ghahremani DG, London ED, Ray LA (2014): The association between cue‐reactivity in the precuneus and level of dependence on nicotine and alcohol. Drug Alcohol Depend 141:21–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Crews F, He J, Hodge C (2007): Adolescent cortical development: A critical period of vulnerability for addiction. Pharmacol Biochem Behav 86:189–199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Dick DM, Edenberg HJ, Xuei XL, Goate A, Kuperman S, Schuckit M, Crowe R, Smith TL, Porjesz B, Begleiter H, Foroud T (2004): Association of gabrg3 with alcohol dependence. Alcohol Clin Exp Res 28:4–9. [DOI] [PubMed] [Google Scholar]
  18. Edenberg HJ, Foroud T (2006): The genetics of alcoholism: Identifying specific genes through family studies. Addict Biol 11:386–396. [DOI] [PubMed] [Google Scholar]
  19. Enoch MA, Hodgkinson CA, Yuan Q, Albaugh B, Virkkunen M, Goldman D (2009): Gabrg1 and gabra2 as independent predictors for alcoholism in two populations. Neuropsychopharmacology 34:1245–1254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Filbey FM, Claus E, Audette AR, Niculescu M, Banich MT, Tanabe J, Du YP, Hutchison KE (2008. a): Exposure to the taste of alcohol elicits activation of the mesocorticolimbic neurocircuitry. Neuropsychopharmacology 33:1391–1401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Filbey FM, Ray L, Smolen A, Claus ED, Audette A, Hutchison KE (2008b): Differential neural response to alcohol priming and alcohol taste cues is associated with drd4 VNTR and oprm1 genotypes. Alcohol Clin Exp Res 32:1113–1123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Gianoulakis C (2001): Influence of the endogenous opioid system on high alcohol consumption and genetic predisposition to alcoholism. J Psychiatry Neurosci 26:304–318. [PMC free article] [PubMed] [Google Scholar]
  23. Hasin DS, Stinson FS, Ogburn E, Grant BF (2007): Prevalence, correlates, disability, and comorbidity of DSM‐IV alcohol abuse and dependence in the united states: Results from the national epidemiologic survey on alcohol and related conditions. Arch Gen Psychiatry 64:830–842. [DOI] [PubMed] [Google Scholar]
  24. Heath AC, Whitfield JB, Martin NG, Pergadia ML, Goate AM, Lind PA, McEvoy BP, Schrage AJ, Grant JD, Chou YL, Zhu R, Henders AK, Medland SE, Gordon SD, Nelson EC, Agrawal A, Nyholt DR, Bucholz KK, Madden PAF, Montgomery GW (2011): A Quantitative‐trait Genome‐wide association study of alcoholism risk in the community: Findings and implications. Biol Psychiatry 70:513–518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Heather N, Booth P, Luce A (1998): Impaired control scale: Cross‐validation and relationships with treatment outcome. Addiction 93:761–771. [DOI] [PubMed] [Google Scholar]
  26. Jenkinson M, Bannister P, Brady M, Smith S (2002): Improved optimization for the robust and accurate linear registration and motion correction of brain images. Neuroimage 17:825–841. [DOI] [PubMed] [Google Scholar]
  27. Johnson C, Drgon T, Liu QR, Walther D, Edenberg H, Rice J, et al. (2006): Pooled association genome scanning for alcohol dependence using 104,268 SNPs: Validation and use to identify alcoholism vulnerability loci in unrelated individuals from the collaborative study on the genetics of alcoholism. Am J Med Genet B Neuropsychiatr Genet 141B:844–853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Johnson C, Drgon T, McMahon FJ, Uhl GR (2009): Convergent genome wide association results for bipolar disorder and substance dependence. Am J Med Genet B Neuropsychiatr Genet 150B:182–190. [DOI] [PubMed] [Google Scholar]
  29. Kapoor M, Wang JC, Wetherill L, Le N, Bertelsen S, Hinrichs AL, Budde J, Agrawal A, Bucholz K, Dick D, Harari O, Hesselbrock V, Kramer J, Nurnberger JI, Rice J, Saccone N, Schuckit M, Tischfield J, Porjesz B, Edenberg HJ, Bierut L, Foroud T, Goate A (2013): A meta‐analysis of two genome‐wide association studies to identify novel loci for maximum number of alcoholic drinks. Hum Genet 132:1141–1151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Kauer JA, Malenka RC (2007): Synaptic plasticity and addiction. Nat Rev Neurosci 8:844–858. [DOI] [PubMed] [Google Scholar]
  31. Lancaster JL, Woldorff MG, Parsons LM, Liotti M, Freitas CS, Rainey L, Kochunov PV, Nickerson D, Mikiten SA, Fox PT (2000): Automated talairach atlas labels for functional brain mapping. Hum Brain Mapp 10:120–131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Lee SH, Ripke S, Neale BM, Faraone SV, Purcell SM, Perlis RH, Mowry BJ, Thapar A, Goddard ME, Witte JS, Absher D, Agartz I, Akil H, Amin F, Andreassen OA, Anjorin A, Anney R, Anttila V, Arking DE, Asherson P, Azevedo MH, Backlund L, Badner JA, Bailey AJ, Banaschewski T, Barchas JD, Barnes MR, Barrett TB, Bass N, Battaglia A, Bauer M, Bayes M, Bellivier F, Bergen SE, Berrettini W, Betancur C, Bettecken T, Biederman J, Binder EB, Black DW, Blackwood DHR, Bloss CS, Boehnke M, Boomsma DI, Breen G, Breuer R, Bruggeman R, Cormican P, Buccola NG, Buitelaar JK, Bunney WE, Buxbaum JD, Byerley WF, Byrne EM, Caesar S, Cahn W, Cantor RM, Casas M, Chakravarti A, Chambert K, Choudhury K, Cichon S, Cloninger CR, Collier DA, Cook EH, Coon H, Cormand B, Corvin A, Coryell WH, Craig DW, Craig IW, Crosbie J, Cuccaro ML, Curtis D, Czamara D, Datta S, Dawson G, Day R, De Geus EJ, Degenhardt F, Djurovic S, Donohoe GJ, Doyle AE, Duan JB, Dudbridge F, Duketis E, Ebstein RP, Edenberg HJ, Elia J, Ennis S, Etain B, Fanous A, Farmer AE, Ferrier IN, Flickinger M, Fombonne E, Foroud T, Frank J, Franke B, Fraser C, Freedman R, Freimer NB, Freitag CM, Friedl M, Frisen L, Gallagher L, Gejman PV, Georgieva L, Gershon ES, Geschwind DH, Giegling I, Gill M, Gordon SD, Gordon‐Smith K, Green EK, Greenwood TA, Grice DE, Gross M, Grozeva D, Guan WH, Gurling H, De Haan L, Haines JL, Hakonarson H, Hallmayer J, Hamilton SP, Hamshere ML, Hansen TF, Hartmann AM, Hautzinger M, Heath AC, Henders AK, Herms S, Hickie IB, Hipolito M, Hoefels S, Holmans PA, Holsboer F, Hoogendijk WJ, Hottenga JJ, Hultman CM, Hus V, Ingason A, Ising M, Jamain S, Jones EG, Jones I, Jones L, Tzeng JY, Kahler AK, Kahn RS, Kandaswamy R, Keller MC, Kennedy JL, Kenny E, Kent L, Kim Y, Kirov GK, Klauck SM, Klei L, Knowles JA, Kohli MA, Koller DL, Konte B, Korszun A, Krabbendam L, Krasucki R, Kuntsi J, Kwan P, Landen M, Langstrom N, Lathrop M, Lawrence J, Lawson WB, Leboyer M, Ledbetter DH, Lee PH, Lencz T, Lesch KP, Levinson DF, Lewis CM, Li J, Lichtenstein P, Lieberman JA, Lin DY, Linszen DH, Liu CY, Lohoff FW, Loo SK, Lord C, Lowe JK, Lucae S, MacIntyre DJ, Madden PAF, Maestrini E, Magnusson PKE, Mahon PB, Maier W, Malhotra AK, Mane SM, Martin CL, Martin NG, Mattheisen M, Matthews K, Mattingsdal M, McCarroll SA, McGhee KA, McGough JJ, McGrath PJ, McGuffin P, McInnis MG, McIntosh A, McKinney R, McLean AW, McMahon FJ, McMahon WM, McQuillin A, Medeiros H, Medland SE, Meier S, Melle I, Meng F, Meyer J, Middeldorp CM, Middleton L, Milanova V, Miranda A, Monaco AP, Montgomery GW, Moran JL, Moreno‐De‐Luca D, Morken G, Morris DW, Morrow EM, Moskvina V, Muglia P, Muhleisen TW, Muir WJ, Muller‐Myhsok B, Murtha M, Myers RM, Myin‐Germeys I, Neale MC, Nelson SF, Nievergelt CM, Nikolov I, Nimgaonkar V, Nolen WA, Nothen MM, Nurnberger JI, Nwulia EA, Nyholt DR, O'Dushlaine C, Oades RD, Olincy A, Oliveira G, Olsen L, Ophoff RA, Osby U, Owen MJ, Palotie A, Parr JR, Paterson AD, Pato CN, Pato MT, Penninx BW, Pergadia ML, Pericak‐Vance MA, Pickard BS, Pimm J, Piven J, Posthuma D, Potash JB, Poustka F, Propping P, Puri V, Quested DJ, Quinn EM, Ramos‐Quiroga JA, Rasmussen HB, Raychaudhuri S, Rehnstrom K, Reif A, Ribases M, Rice JP, Rietschel M, Roeder K, Roeyers H, Rossin L, Rothenberger A, Rouleau G, Ruderfer D, Rujescu D, Sanders AR, Sanders SJ, Santangelo SL, Sergeant JA, Schachar R, Schalling M, Schatzberg AF, Scheftner WA, Schellenberg GD, Scherer SW, Schork NJ, Schulze TG, Schumacher J, Schwarz M, Scolnick E, Scott LJ, Shi JX, Shilling PD, Shyn SI, Silverman JM, Slager SL, Smalley SL, Smit JH, Smith EN, Sonuga‐Barke EJS, St Clair D, State M, Steffens M, Steinhausen HC, Strauss JS, Strohmaier J, Stroup TS, Sutcliffe JS, Szatmari P, Szelinger S, Thirumalai S, Thompson RC, Todorov AA, Tozzi F, Treutlein J, Uhr M, van den Oord EJCG, Van Grootheest G, Van Os J, Vicente AM, Vieland VJ, Vincent JB, Visscher PM, Walsh CA, Wassink TH, Watson SJ, Weissman MM, Werge T, Wienker TF, Wijsman EM, Willemsen G, Williams N, Willsey AJ, Witt SH, Xu W, Young AH, Yu TW, Zammit S, Zandi PP, Zhang P, Zitman FG, Zollner S, Devlin B, Kelsoe JR, Sklar P, Daly MJ, O'Donovan MC, Craddock N, Sullivan PF, Smoller JW, Kendler KS, Wray NR, Genomi C‐DGP, Genetic IIBD (2013): Genetic relationship between five psychiatric disorders estimated from genome‐wide SNPs. Nat Genet 45:984–994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Lee TW, Girolami M, Sejnowski TJ (1999): Independent component analysis using an extended infomax algorithm for mixed subgaussian and supergaussian sources. Neural Comput 11:417–441. [DOI] [PubMed] [Google Scholar]
  34. Li CY, Mao XZ, Wei LP (2008): Genes and (common) pathways underlying drug addiction. PLoS Comput Biol 4:e2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Liu J, Calhoun VD, Chen J, Claus ED, Hutchison KE (2013): Effect of homozygous deletions at 22q13.1 on alcohol dependence severity and cue‐elicited BOLD response in the precuneus. Addict Biol 18:548–558. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Liu J, Claus ED, Calhoun VD, Hutchison KE (2014): Brain regions affected by impaired control modulate responses to alcohol and smoking cues. J Stud Alcohol Drugs 75:808–816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Liu J, Pearlson G, Windemuth A, Ruano G, Perrone‐Bizzozero NI, Calhoun V (2009): Combining fMRI and SNP data to investigate connections between brain function and genetics using parallel ICA. Hum Brain Mapp 30:241–255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Liu QS, Pu L, Poo MM (2005): Repeated cocaine exposure in vivo facilitates LTP induction in midbrain dopamine neurons. Nature 437:1027–1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Luo XG, Kranzler HR, Zuo LJ, Wang S, Schork NJ, Gelernter J (2007): Multiple ADH genes modulate risk for drug dependence in both African‐ and European‐americans. Hum Mol Genet 16:380–390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Malenka RC, Bear MF (2004): LTP and LTD: An embarrassment of riches. Neuron 44:5–21. [DOI] [PubMed] [Google Scholar]
  41. Mayfield RD, Harris RA, Schuckit MA (2008): Genetic factors influencing alcohol dependence. Br J Pharmacol 154:275–287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Mulligan MK, Ponomarev I, Hitzemann RJ, Belknap JK, Tabakoff B, Harris RA, Crabbe JC, Blednov YA, Grahame NJ, Phillips TJ, Finn DA, Hoffman PL, Iyer VR, Koob GF, Bergeson SE (2006): Toward understanding the genetics of alcohol drinking through transcriptome meta‐analysis. Proc Natl Acad Sci USA 103:6368–6373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Pandey SC (2003): Anxiety and alcohol abuse disorders: A common role for CREB and its target, the neuropeptide Y gene. Trends Pharmacol Sci 24:456–460. [DOI] [PubMed] [Google Scholar]
  44. Park MS, Sohn JH, Suk JA, Kim SH, Sohn S, Sparacio R (2007): Brain substrates of craving to alcohol cues in subjects with alcohol use disorder. Alcohol Alcohol 42:417–422. [DOI] [PubMed] [Google Scholar]
  45. Prescott CA, Kendler KS (1999): Genetic and environmental contributions to alcohol abuse and dependence in a population‐based sample of male twins. Am J Psychiatry 156:34–40. [DOI] [PubMed] [Google Scholar]
  46. Purcell S, Neale B, Todd‐Brown K, Thomas L, Ferreira MA, Bender D, Maller J, Sklar P, de Bakker PI, Daly MJ, Sham PC (2007): PLINK: A tool set for whole‐genome association and population‐based linkage analyses. Am J Hum Genet 81:559–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Radel M, Vallejo RL, Iwata N, Aragon R, Long JC, Virkkunen M, Goldman D (2005): Haplotype‐based localization of an alcohol dependence gene to the 5q34 gamma‐aminobutyric acid type a gene cluster. Arch Gen Psychiatry 62:47–55. [DOI] [PubMed] [Google Scholar]
  48. Ripke S, Sanders AR, Kendler KS, Levinson DF, Sklar P, Holmans PA, Lin DY, Duan J, Ophoff RA, Andreassen OA, Scolnick E, Cichon S, St Clair D, Corvin A, Gurling H, Werge T, Rujescu D, Blackwood DH, Pato CN, Malhotra AK, Purcell S, Dudbridge F, Neale BM, Rossin L, Visscher PM, Posthuma D, Ruderfer DM, Fanous A, Stefansson H, Steinberg S, Mowry BJ, Golimbet V, De Hert M, Jonsson EG, Bitter I, Pietilainen OP, Collier DA, Tosato S, Agartz I, Albus M, Alexander M, Amdur RL, Amin F, Bass N, Bergen SE, Black DW, Borglum AD, Brown MA, Bruggeman R, Buccola NG, Byerley WF, Cahn W, Cantor RM, Carr VJ, Catts SV, Choudhury K, Cloninger CR, Cormican P, Craddock N, Danoy PA, Datta S, de Haan L, Demontis D, Dikeos D, Djurovic S, Donnelly P, Donohoe G, Duong L, Dwyer S, Fink‐Jensen A, Freedman R, Freimer NB, Friedl M, Georgieva L, Giegling I, Gill M, Glenthoj B, Godard S, Hamshere M, Hansen M, Hansen T, Hartmann AM, Henskens FA, Hougaard DM, Hultman CM, Ingason A, Jablensky AV, Jakobsen KD, Jay M, Jurgens G, Kahn RS, Keller MC, Kenis G, Kenny E, Kim Y, Kirov GK, Konnerth H, Konte B, Krabbendam L, Krasucki R, Lasseter VK, Laurent C, Lawrence J, Lencz T, Lerer FB, Liang KY, Lichtenstein P, Lieberman JA, Linszen DH, Lonnqvist J, Loughland CM, Maclean AW, Maher BS, Maier W, Mallet J, Malloy P, Mattheisen M, Mattingsdal M, McGhee KA, McGrath JJ, McIntosh A, McLean DE, McQuillin A, Melle I, Michie PT, Milanova V, Morris DW, Mors O, Mortensen PB, Moskvina V, Muglia P, Myin‐Germeys I, Nertney DA, Nestadt G, Nielsen J, Nikolov I, Nordentoft M, Norton N, Nothen MM, O'Dushlaine CT, Olincy A, Olsen L, O'Neill FA, Orntoft TF, Owen MJ, Pantelis C, Papadimitriou G, Pato MT, Peltonen L, Petursson H, Pickard B, Pimm J, Pulver AE, Puri V, Quested D, Quinn EM, Rasmussen HB, Rethelyi JM, Ribble R, Rietschel M, Riley BP, Ruggeri M, Schall U, Schulze TG, Schwab SG, Scott RJ, Shi J, Sigurdsson E, Silverman JM, Spencer CC, Stefansson K, Strange A, Strengman E, Stroup TS, Suvisaari J, Terenius L, Thirumalai S, Thygesen JH, Timm S, Toncheva D, van den Oord E, van Os J, van Winkel R, Veldink J, Walsh D, Wang AG, Wiersma D, Wildenauer DB, Williams HJ, Williams NM, Wormley B, Zammit S, Sullivan PF, O'Donovan MC, Daly MJ, Gejman PV (2011): Genome‐wide association study identifies five new schizophrenia loci. Nat Genet 43:969–976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Rissanen J (1978): Modeling by shortest data description. Automatica 14:465–471. [Google Scholar]
  50. Robbins TW, Ersche KD, Everitt BJ (2008): Drug addiction and the memory systems of the brain. Addict Rev 2008 1141:1–21. [DOI] [PubMed] [Google Scholar]
  51. Ron D, Jurd R (2005): The “ups and downs” of signaling cascades in addiction. Sci STKE 2005:re14. [DOI] [PubMed] [Google Scholar]
  52. Salvatore JE, Aliev F, Edwards AC, Evans DM, Macleod J, Hickman M, Lewis G, Kendler KS, Loukola A, Korhonen T, Latvala A, Rose RJ, Kaprio J, Dick DM (2014): Polygenic scores predict alcohol problems in an independent sample and show moderation by the environment. Genes (Basel) 5:330–346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Schacht JP, Anton RF, Myrick H (2013): Functional neuroimaging studies of alcohol cue reactivity: A quantitative meta‐analysis and systematic review. Addict Biol 18:121–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Schumann G, Johann M, Frank J, Preuss U, Dahmen N, Laucht M, Rietschel M, Rujescu D, Lourdusamy A, Clarke TK, Krause K, Dyer A, Depner M, Wellek S, Treutlein J, Szegedi A, Giegling I, Cichon S, Blomeyer D, Heinz A, Heath S, Lathrop M, Wodarz N, Soyka M, Spanagel R, Mann K (2008): Systematic analysis of glutamatergic neurotransmission genes in alcohol dependence and adolescent risky drinking behavior. Arch Gen Psychiatry 65:826–838. [DOI] [PubMed] [Google Scholar]
  55. Skinner HA, Horn JL (1984): Alcohol Dependence Scale: Users Guide. Toronto, Canada: Alcohol Research Foundation. [Google Scholar]
  56. Stranger BE, Stahl EA, Raj T (2011): Progress and promise of genome‐wide association studies for human complex trait genetics. Genetics 187:367–383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Sutherland MT, McHugh MJ, Pariyadath V, Stein EA (2012): Resting state functional connectivity in addiction: Lessons learned and a road ahead. Neuroimage 62:2281–2295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Tapert SF, Brown GG, Baratta MV, Brown SA (2004): fMRI BOLD response to alcohol stimuli in alcohol dependent young women. Addict Behav 29:33–50. [DOI] [PubMed] [Google Scholar]
  59. Treutlein J, Cichon S, Ridinger M, Wodarz N, Soyka M, Maier W, Moessner R, Dahmen N, Fehr C, Scherbaum N, Wichmann E, Schreiber S, Dragano N, Sommer WH, Wienker T, Sullivan PF, Nothen M, Kiefer F, Spanagel R, Mann KF, Rietschel M (2010): Genome‐wide association study of alcohol dependence. Alcohol Clin Exp Res 34:43a–43a. [Google Scholar]
  60. Waltereit R, Weller M (2003): Signaling from cAMP/PKA to MAPK and synaptic plasticity. Mol Neurobiol 27:99–106. [DOI] [PubMed] [Google Scholar]
  61. Wand G, Levine M, Zweifel L, Schwindinger W, Abel T (2001): The cAMP‐protein kinase a signal transduction pathway modulates ethanol consumption and sedative effects of ethanol. J Neurosci 21:5297–5303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Wang JC, Grucza R, Cruchaga C, Hinrichs AL, Bertelsen S, Budde JP, Fox L, Goldstein E, Reyes O, Saccone N, Saccone S, Xuei X, Bucholz K, Kuperman S, Nurnberger J, Rice JP, Schuckit M, Tischfield J, Hesselbrock V, Porjesz B, Edenberg HJ, Bierut LJ, Goate AM (2009): Genetic variation in the chrna5 gene affects mRNA levels and is associated with risk for alcohol dependence. Mol Psychiatry 14:501–510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Wojnar M, Brower KJ, Strobbe S, Ilgen M, Matsumoto H, Nowosad I, Sliwerska E, Burmeister M (2009): Association between Val66Met Brain‐derived neurotrophic factor (BDNF) gene polymorphism and Post‐treatment relapse in alcohol dependence. Alcohol Clin Exp Res 33:693–702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Wong CCY, Schumann G (2008): Genetics of addictions: Strategies for addressing heterogeneity and polygenicity of substance use disorders. Philos Trans R Soc B Biol Sci 363:3213–3222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Wrase J, Schlagenhauf F, Kienast T, Wustenberg T, Bermpohl F, Kahnt T, Beek A, Strohle A, Juckel G, Knutson B, Heinz A (2007): Dysfunction of reward processing correlates with alcohol craving in detoxified alcoholics. Neuroimage 35:787–794. [DOI] [PubMed] [Google Scholar]
  66. Zhang H, Kranzler HR, Yang BZ, Luo X, Gelernter J (2008): The oprd1 and oprk1 loci in alcohol or drug dependence: oprd1 variation modulates substance dependence risk. Mol Psychiatry 13:531–543. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Supporting Information Table 1.

Supporting Information Table 4.


Articles from Human Brain Mapping are provided here courtesy of Wiley

RESOURCES