Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

nature reviews neuroscience https://doi.org/10.

1038/s41583-022-00656-8

Perspective Check for updates

The Genetically Informed


Neurobiology of Addiction
(GINA) model
Ryan Bogdan 1
, Alexander S. Hatoum 2
, Emma C. Johnson 2
& Arpana Agrawal 2

Abstract Sections

Addictions are heritable and unfold dynamically across the lifespan. Introduction

One prominent neurobiological theory proposes that substance-induced Brain-based models of


changes in neural circuitry promote the progression of addiction. addiction

Genome-wide association studies have begun to characterize the Genetics of addiction


polygenic architecture undergirding addiction liability and revealed Predispositional and/or
that genetic loci associated with risk can be divided into those causal?

associated with a general broad-spectrum liability to addiction and The GINA model
those associated with drug-specific addiction risk. In this Perspective, Conclusions and perspectives
we integrate these genomic findings with our current understanding
of the neurobiology of addiction to propose a new Genetically
Informed Neurobiology of Addiction (GINA) model.

Department of Psychological and Brain Sciences, Washington University in St. Louis, St. Louis, MO, USA.
1

Department of Psychiatry, Washington University School of Medicine, St. Louis, MO, USA.
2
e-mail: rbogdan@
wustl.edu; arpana@wustl.edu

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 40


Perspective

Introduction whom, where and when one uses a substance) and intrinsic cues (such
The loss of life and socioeconomic costs associated with addictive as mood and physiology) are quickly and strongly paired to drug–
substances burden the world1. There are vast individual differences reward and become conditioned predictors that motivate substance
in patterns of substance use, which range from casual or occasional use17,18. This stage exerts its largest influence on addiction during the
to problematic and disordered. Addiction emerges when, following initial escalating and episodic heavy use of a substance (particularly
chronic regular use, the presence of a substance helps maintain homeo- during adolescence and young adulthood), the re-emergent escalat-
stasis. Behaviourally and psychologically, addiction typically results ing use that follows abstinence, and the use of substance types with
in the attenuation of reward elicited by initial substance use and in the increasing psychoactive impact (higher doses or potency) following
development of compulsive use to ameliorate negative affect as well addiction progression9,19,20.
as the psychological and physiological stress states and withdrawal With continued heavy substance use and progression towards
symptoms that arise when a substance is absent. Increasing quantities, severe SUD, it is proposed that the reinforcing properties of substances
dosage and potencies of substances are often pursued in addiction in an shift to negative reinforcement8,9,21,22. This transition to the withdrawal–
attempt to obtain the increasingly fleeting ‘highs’ experienced during negative affect stage of addiction is marked by distress and anhedonia23–25
initial use. Box 1 provides an outline of the correspondence between as well as by the aversive physiological states (such as blackouts, nau-
this definition of addiction with the definitions of substance use and sea and insomnia) and psychological states (such as anxiety, depres-
substance use disorders (SUDs). sion and heightened stress) that arise in the absence of the effects of
Translational neuroscience research has transformed our under- a drug9. During this stage, substance use is compulsive and functions
standing of addiction and led to its re-characterization as a neuro- to provide relief from these aversive states by returning the body to
biological state rather than a controllable moral failing2. A largely a state of homeostasis. Such homeostasis can now only be achieved
independent line of research has shown that the moderate-to-large when the drug is present10 because a series of neuroadaptations (such
heritability of SUDs is undergirded by a polygenic architecture that is as fewer dopamine receptors and attenuated reward-related dopamine
associated with broad-spectrum liability to addiction as well as dis- release)13 have taken place as natural adjustments to repeated drug
tinct genetic architectures associated with substance-specific risks3–5. exposure. These adaptations also promote anhedonia, through which
This genetic risk is compounded by environmental factors that are other non-drug rewards (such as social interactions, achievement, food
substance specific (such as policy-related access to specific substances) and sex) lose their reinforcing properties. Thus, increasing quantities
and associated with factors related to general psychiatric liability (such of the addictive substance are required to reach homeostasis, with even
as socioeconomic status). greater amounts being required to ‘chase the highs’ that were associ-
In this Perspective, we incorporate knowledge from human genetic ated with the binge–intoxication stage17. The chronic use of addictive
studies of addiction into brain disease models. First, we provide an substances also promotes stress and negative affect (through, for exam-
overview of the three-stage neurobiological model of addiction, which ple, elevations of corticotropin-releasing hormone in the extended
postulates that substance-induced neural changes are the predominant amygdala)26. During substance abstinence, withdrawal and related neg-
contributor to the aetiology of SUDs. We then showcase how contem- ative affect leads to heightened interoceptive salience, through which
porary genetic work has begun to identify the polygenic architecture the physiological arousal associated with withdrawal and negative
underlying addiction risk. Here, we highlight the utility of genetically emotionality potentiates negative reinforcement-related craving10,27.
informed study designs to probe the plausibility of proposed models of In the third stage of this model, the repeated pairing of drug use
addiction. Last, we integrate genetic research into an expanded version with reward and relief results in a cognitive preoccupation–anticipation
of the stage-based neurobiological model of addiction by proposing of the drug in expectation of these effects; this is often characterized
a new model: the Genetically Informed Neurobiology of Addiction by the subjective experience of drug craving. Building on dual systems
(GINA) model. In this model, the interplay between genetic liability models that postulate that self-control arises when deliberative execu-
and substance-induced changes in the neural substrates of positive tive function counteracts more automatic emotionally driven behav-
reinforcement (thought to drive binging, intoxication and escalat- iour28,29, this stage of addiction is proposed to be defined by a loss of
ing use), negative urgency (thought to occur during withdrawal and this regulatory capacity as a result of substance-induced impairment
negative affect), and executive function and/or regulatory capacity of top-down executive function (such as reduced prefrontal control
(important for the preoccupation with and/or anticipation of sub- over striatal and other limbic circuits)8,30,31. Thus, it is proposed that
stance use) along with neural and peripheral substance-specific heavy and sustained substance use causes the prefrontal cortex to
pathways contribute to SUD development. become less efficient at minimizing the direct incentive salience evoked
by substance cues and the emergent negative emotionality, leaving
Brain-based models of addiction individuals with diminished intrinsic ability to combat the throes of
According to the most prominent neurobiological model of addic- addiction, even if there are deep subjective desires to stop.
tion6–10, substance and/or experience-dependent alterations in cor- This model emphasizes how substance-induced neural changes
tiostriatal11 and corticolimbic12 circuitry (Fig. 1) drive three recurring contribute to the development and maintenance of moderate-to-
and non-mutually exclusive stages of addiction: binge–intoxication, severe SUDs. Given that drugs of abuse impact neurotransmitters and
withdrawal–negative affect and preoccupation–anticipation. neuromodulators at a magnitude that does not intrinsically occur17,32,
During the binge–intoxication stage of the neurobiological model it might be assumed that resulting brain changes would be so pen-
of addiction, substance-induced stimulation of neural reward circuitry etrant that addiction would be a foregone conclusion in anyone with
provides positive reinforcement. In support of this idea, all addictive escalating levels of use. In reality, only some individuals exposed to
substances have been shown to directly or indirectly elicit fast and addictive substances develop addiction33 and a significant minority
large increases in dopamine release in the nucleus accumbens (NAc)13,14 of individuals with a SUD recover34. The neurobiological stage-based
that resemble predictive reward signals15,16. Extrinsic cues (such as with model of addiction therefore acknowledges that factors that influence

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 41


Perspective

Box 1

Clinical and genetic distinctions between substance use and


addiction
Not all substance use reflects or results in addiction (nor is provide insight into the extent and nature of the genetic influence
addiction only relevant to psychotropic substances3,232–234, although on the use of and addiction to various substances50,54,46,47,50,81,238–242
here we restrict our discussion to substance addiction). In this article, (see the Table).
we broadly define addiction as the stage at which the pleasurable GWAS also suggest distinctions between substance use
aspects of substance use are attenuated, and compulsive use and SUDs. There is a moderate to high genetic correlation
emerges to ameliorate the negative affect and stress states that between substance use (encompassing ever using a substance
arise in the absence of the substance. and use in daily life) and problematic or disordered use46,50,53,54.
Various diagnostic schemas define substance use disorders However, substance use and SUDs differ in their genetic associa-
(SUDs) in a substance-specific manner as the syndromes that arise tions with other psychosocial factors and psychiatric and medical
from excessive substance or drug involvement accompanied by loss comorbidities. For instance, while problem drinking is genetically
of control over use, use despite deleterious consequences, and im- correlated with negative health outcomes, many studies docu-
pairment. For instance, according to the Diagnostic and Statistical ment that typical alcohol consumption is genetically correlated
Manual of Mental Disorders, Fifth Edition (DSM-5), a specific SUD with higher educational attainment and a lower risk for cardiometa-
(such as opioid use disorder) can be diagnosed when an individual bolic disease and is not significantly related to genetic risk for other
qualifies for 2 or more of 11 criteria that assess physiological, psycho- psychopathologies (reviewed in ref.61). How often someone drinks
logical and interpersonal problems235. The eleventh edition of the is confounded with higher socio-economic advantage, thus biasing
International Classification of Diseases (ICD-11), on the other hand, genetic correlations identified by GWAS that focus on measures of
includes separate entries for substance dependence, harmful pat- drinking in daily life60,62. Likewise, cannabis use and cannabis use
terns of use and hazardous patterns of use236. Harmful or problematic disorder exhibit opposing correlations with educational attain-
substance use is also evaluated in health-care settings using short ment and body mass index but both are genetically related to more
screeners237. Therefore, the definition of addiction that we use here serious psychiatric outcomes such as schizophrenia and depres-
approximates the DSM-5 diagnosis of a moderate or severe SUD or sion46,51. This difference is not as evident for nicotine, where both
the ICD-11 coding of harmful use and dependence. smoking initiation and nicotine dependence appear to be linked
Both propensity for substance use and risk of developing to a greater genetic risk for psychosocial disadvantage and
SUDs are heritable and genome-wide association studies (GWAS) ­psychiatric and somatic illness47.

Substance Use or use disorder Sample size of SNP-heritability Number of independent Genetic correlation Refs.
current largest variants/genes between use and use
GWAS identified disorder

Alcohol Alcohol use disorder/problem 435,563 0.07 29/66 genes 0.77 49,53,54

alcohol use
Drinks/week 941,280 0.04 99/362 genes 47

Tobacco/nicotine Nicotine dependence 58,000 0.09 5/16 genes 0.4–0.5 238

Ever smoked 1,232,091 0.08 378/833 genes 47

Cannabis Cannabis use disorder 384,925 0.12 2/3 genes 0.50 46

Ever used cannabis 184,765 0.11 8/35 genes 51

Opioids Opioid use disorder 639,709a 0.13 10/4 genes NA 48,53,239–241

Cocaine Cocaine use disorder 6,546 b


0.30 1/5 genes NA 81,242

NA, not available; SNP, single nucleotide polymorphism. This GWAS of European and African ancestries has a case N = 20,858 (total N = 639,709), but another recent GWAS reports a
a

larger case N = 31,473. This second GWAS reports a smaller N overall (N = 425,944) but a slightly more diverse sample (including European Americans, African Americans, and Hispanic
Americans)240,241. bThis sample N comes from a trans-ancestral meta-analysis81, while the heritability estimate comes from a separate analysis of individuals of European ancestry242.

the predisposition of an individual to addiction must be considered as emerging evidence reported in a recent preprint highlights substance-
major contributors to the disorder5,9. specific risk factors that take the form of variants in genes within sub-
The neurobiological stage-based model provides a framework of stance-specific metabolic and signalling pathways37. Thus, we can
addiction susceptibility that is not substance specific. While the con- hypothesize that predispositional liability to the addictive properties of
cept of an addiction risk that is shared across substances is supported a specific substance (or a non-substance-related behaviour) may set the
by evidence that SUDs are frequently comorbid with one another35 pace for transitions between the neurobiological stages of addiction
and share similar neural36, genetic37,38 and environmental39 correlates, and intensify the subjective experience of each stage.

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 42


Perspective

Developmental vulnerability and trait-like vulnerability have as CHRNA5 for nicotine and OPRM1 for opioids)46–54 and implicated in
been highlighted by other neurobiological theories of addiction. addiction models (such as CRHR1, encoding corticotropin-releasing
For instance, arising from developmental psychology40, the neuro­ factor receptor 1)34,50,55. Recent comprehensive reviews of these GWAS
developmental model of addiction41 has theorized that adolescence findings are available5,43 (Box 1).
and young adulthood confer broad-spectrum addiction risk owing More generally, GWAS of complex traits (including substance use
to typical patterns of brain maturation that initially prioritize emo- and SUDs) have revealed that, with a few exceptions (such as the effects
tional and social processing over cognitive control and regulation. of the single nucleotide polymorphism (SNP) rs1229984 in ADH1B on
This promotes risk-taking behaviour42 as well as increased impulsive alcohol use49), single genetic variants have small effect sizes and these
attempts to cope with negative emotion, placing adolescents and traits are highly polygenic49. Notably, while twin studies suggest a herit-
young adults at risk for both the pleasurable and negative reinforc- ability of ~50% for a range of addictions, GWAS and whole-exome analy-
ing aspects of SUDs (particularly in the context of underdeveloped ses56,57 can explain, at best, only a quarter of this heritability58, although
physiological tolerance and still-developing regulatory capacity). In the inclusion of less common variants does improve the heritability of
this model, the typical earlier maturation of reward-related and stress some traits57. This discrepancy is typical of most complex traits; it is
and negative affect-related neural circuitry and relatively delayed pre- possible that some of this ‘missing heritability’ resides in rare variants
frontal development are seen as addiction risk factors. It is speculated that will be identified using sequencing technologies59. However, the
that substance use may more profoundly shape neural circuitry, and high polygenicity of addictions also suggests that larger sample sizes
especially prefrontal development, during these periods of extensive may be required to identify additional novel common variants than for
neural maturation. other complex traits5.
The neurodevelopmental and stage-based neurobiological models Three other insights unique to SUDs are notable. First, while
of addiction have unique origins and differentially weight predisposi- genetic correlations between liability to substance use (for example,
tional liability and substance-induced alterations. The neurobiological likelihood of ever using or frequency of use) and problematic or dis-
model arose primarily from data on functional differences in brain ordered use are moderate to high46,50, the genetics of disordered use
activity and receptor densities and emphasizes the role of substance- faithfully reproduce a pattern of correlated medical comorbidities
induced neural plasticity in the aetiology of addiction. On the other (both psychiatric and somatic) and potential indicators of negative life
hand, the neurodevelopmental model relies predominantly on emer- outcomes (such as lower education attainment) whereas the genetics
gent structural changes during adolescence and emphasizes predispo- of substance use have been related to adaptive psychosocial correlates
sitional developmental liability. However, both models highlight the and inconclusively linked with psychopathology (Box 1). Accordingly,
contributions of impulsivity, negative affect and executive function the risk of substance use may represent a mixture of risk for future
(and their neural substrates) in addiction vulnerability. problems and resilience to them46,49,60–62.
Over the past 5 years, we have witnessed immense progress in A second insight recapitulates prior evidence from twin stud-
human genetics (reviewed below) that can shed further light on the ies indicating that genetic liability for SUDs is largely shared across
mechanisms of addiction. With this in mind, it is now important to substances but that there is also important substance-specific liabil-
begin to measure and integrate predispositional genetic risk into brain ity37,63 (Fig. 2). A recent preprint reports that the polygenic architecture
disease models of addiction. underlying the general liability to SUDs includes loci that regulate
dopaminergic signalling37, including signals linked to DRD2 (encod-
Genetics of addiction ing the D2 dopamine receptor) and PDE4B (encoding cAMP-specific
Genetics of substance use and SUDs 3′,5′-cyclic phosphodiesterase 4B)37,50,64. As PDE4B is instrumental
Historically, the search for genetic variation underlying SUDs has in neuroplasticity within prefrontal dopaminergic pathways and is
focused on the genes encoding substance-specific neurotransmit- associated with stress and negative affect in animal models65–67, it rep-
ters or metabolic enzymes (such as opioid or nicotinic receptors, resents an intriguing locus for addiction that is consistent with the
alcohol dehydrogenase, and cytochrome p450) and genes encoding stage-based neurobiological model of addiction. Another recent GWAS
proteins involved in canonical systems that have a widespread effect identified genes contributing to shared liability to substance use,
on psychiatrically relevant behaviours (such as dopaminergic and addictions and other related behaviours (such as a greater number
serotonergic receptors and transporters)43. As in studies of other of sexual partners)64. Genes identified as being associated with traits
complex phenotypes, studies were conducted on relatively modestly with externalizing features included CADM2 (encoding cell adhesion
sized samples and single gene variants, genes or haplotypes were molecule 2), which has also been linked to general liability to SUDs37
examined with exonic variation prioritized. However, the introduc- as well as to substance use, risky sexual behaviour, self-control and
tion of genome-wide association studies (GWAS) to the field led to a obesity68,69. It is thus possible that CADM2 impacts addiction liability
transition from candidate gene validation to genetic exploration, with by influencing early risk-taking and self-control more broadly70 (as
some unexpected consequences. opposed to influencing addiction progression). Intriguingly, despite
For many psychiatric disorders (such as schizophrenia or major the identification of overlapping loci and a high genetic correlation
depressive disorder), the candidate genes that had been hypothesized between the externalizing GWAS64 and a factor identified in the recently
to be involved in disease liability were found by GWAS to be no more reported GWAS representing common addiction liability37, many novel
likely to be associated with disease risk than those selected at random loci are associated with the latter addiction factor, implying that addic-
and, with a few exceptions, novel loci were associated with these disor- tion pathology is partially genetically distinct from general liability to
ders44,45. However, in the case of substance use and addiction, some of externalizing behaviours37.
the strongest significant signals (in addition to novel loci) in GWAS were Third, and finally, these GWAS have shown that, after the common
in candidate genes known to regulate metabolism (such as ADH1B for genetic liability to addiction is considered, residual and substance-
alcohol and CYP2A6 for nicotine), encode receptor binding sites (such specific variation is often conferred by variants in genes encoding

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 43


Perspective

a Corticostriatal circuit Corticolimbic circuit

DMPFC DMPFC
Ventral ACC ACC
striatum Caudate

VMPFC Putamen VMPFC

OFC
Pallidum
VTA Hypothalamus
BNST

Amygdala Amygdala
Hippocampus Hippocampus Autonomic
nuclei

DLPFC DLPFC

VLPFC VLPFC Insula


OFC

b
Prefrontal Prefrontal
cortex cortex

Amygdala Striatum Pallidum Thalamus Thalamus and Amygdala


sensory cortices

Bed nucleus
Hippocampal Autonomic
VTA Hypothalamus Hippocampus of the stria Insula
formation nuclei
terminalis

metabolic factors and substance-specific receptors37. This substance- Correspondence between GWAS and molecular genetics
specific variation is also polygenic; however, the effect sizes of some That addictions originate and induce perturbations in brain-based
individual variants are an order of magnitude larger than those of genetic pathways is supported by sources of genetic data in addition
variants that are common across addictions37,64. to GWAS. Many loci linked to addiction through GWAS have been shown to

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 44


Perspective

Fig. 1 | Corticostriatal and corticolimbic circuits underlying addiction. regulating emotional responses12. Low-resolution and high-resolution sensory
Anatomical locations of (part a) and connections between (part b) the primary information arrives in the basolateral complex of the amygdala from the
nodes within the corticostriatal and corticolimbic circuits that support reward, thalamus and sensory cortices, respectively. Efferent projections from the
emotion, and their regulation and are proposed to influence the binge– centromedial and extended amygdala, including the bed nucleus of the stria
intoxication, withdrawal–negative affect and preoccupation–anticipation stages terminalis (BNST), to autonomic nuclei (such as the parabrachial nucleus),
of addiction. The corticostriatal circuit is critical for reward processing and hypothalamus and hippocampus drive emotional responses, including fear
largely contributes to the binge–intoxication stage of addiction. The striatum conditioning and the generation of stress-related physiological changes.
(comprised of the putamen, caudate and ventral striatum) is the primary node of Direct and indirect connections between the amygdala and insula facilitate
this network. Through its connections with other nodes, the striatum supports interoception (awareness and importance of our physiological states).
learning reward contingencies, hedonic responsiveness, generating motivation Projections from the nucleus basalis of Meynert in the extended amygdala
to pursue rewards and goals, forming and implementing plans to obtain facilitate amygdala-driven arousal and sensitivity of the cortex. Projections
reward, adjusting behaviour and plans according to changing contingencies, from the amygdala to the VMPFC promote subjective awareness and evaluation
and coordinating motor movements in the service of obtaining reward11. of emotion and the integration of affective information (such as motivational
More specifically, dopaminergic projections from the ventral tegmental area information conveyed by the ventral striatum projections shown in the upper
(VTA) to the nucleus accumbens within the ventral striatum support reward panel). Projections from the DLPFC and ventrolateral prefrontal cortex (VLPFC)
prediction and learning in combination with multimodal sensory information to the amygdala through the dorsomedial prefrontal cortex (DMPFC) and
received from the basolateral amygdala and contextual information from the VMPFC promote the regulation of affective responses and physiological arousal.
hippocampal formation. Projections from the striatum to the pallidum support Both the corticostriatal and corticolimbic circuits support executive function
hedonic responsiveness through endogenous opioid stimulation and provide and the regulation of behaviour to influence the preoccupation–anticipation
motivational signals to the ventromedial prefrontal cortex (VMPFC; supporting stage of addiction by contributing to incentive salience (such as the ventral
integration of contextual and interoceptive information, bottom-up drive and striatum projections to the VMPFC and orbitofrontal cortex (OFC) within
top-down regulation) and the dorsolateral prefrontal cortex (DLPFC; supporting the corticostriatal circuit), interoceptive signals associated with withdrawal
goal-directed planning) through thalamic relays. Afferents from the prefrontal physiology and affect (such as the insula within the corticolimbic circuit), and the
cortex to the ventral striatum further serve to facilitate the implementation regulation of behaviour (through the DLPFC, VLPFC and ACC in both circuits).
of plans to obtain reward (DLPFC) as well as flexible behavioural adjustment While there are many additional connections within and between these circuits,
when expected actions do not obtain predicted outcomes (anterior cingulate we present a heuristic model focusing on those most well linked to addiction.
cortex; ACC) and can facilitate or inhibit the motivational significance of These circuits are explained in greater detail in prior publications11,12. Of note,
reward-predictive cues in the environment. The corticolimbic circuit is critical unlike prior depictions of the stage-based neurobiological model, which show
for affective processing and behavioural vigilance; it largely contributes to three circuits corresponding to each stage, we present the corticostriatal and
the withdrawal–negative affect stage of addiction. The amygdala (inclusive corticolimbic circuits, which are hypothesized to predominantly drive the
of the amygdala and the extended amygdala) is the primary node of this binge–intoxication and withdrawal–negative affect stages, respectively.
network; through its connections with other nodes, it supports responses The preoccupation–anticipation stage is undergirded by prefrontal connections
to environmental challenges, including threat and stress, by generating and within and across these circuits in this model.

be expression quantitative trait loci (eQTL) across developmental tegmental area that mapped to well-established addiction pathways
stages in tissue obtained from brain regions implicated in the three- (such as dopaminergic networks)78. These studies suggest that, just
stage neuro­biological model of addiction46,71–75. For instance, variants as the brain responds dynamically to repeated drug exposure, so do
linked to substance use were associated with gene expression and the transcriptome and epigenome. However, this genomic and neural
co-expression network modules in the NAc, anterior cingulate cortex plasticity may not be a consequence of drug exposure alone: novel stud-
(Fig. 1), cerebellum, dorsolateral prefrontal cortex and other brain ies that integrate GWAS signals with multi-omics data have suggested
regions71–73. Gene co-expression patterns were preserved across that a subset of the genes that are differentially expressed also shows
these brain regions, supporting a generalized overall enrichment enriched genetic associations with substance use79. In one study, genes
of these variants in the brain rather than in specific brain regions71. that were differentially expressed in the dorsolateral prefrontal cortex
Beyond eQTL effects, heritability enrichment analyses of GWAS have (DLPFC) of individuals who were alcohol dependent coalesced into
revealed that the genetics of substance use and addiction liability are two co-expression modules that included genes that were enriched in
enriched for tissue-specific and cell-specific regulatory elements spe- alcohol use and addiction GWAS80. In other instances, genes identified
cifically related to chromatin architecture76 (that is, DNA folding) in as having significant association with substance use in GWAS were
primary brain regions specified in the three-stage neurobiological not differentially expressed but genes within their co-expression net-
model of addiction. Genes linked to alcohol and tobacco use and to their works were81. Therefore, GWAS data in combination with multi-omics
problematic use also show higher expression in excitatory neurons in and cross-species findings can provide insights into the neural gene
cortical and midbrain regions as well as the hippocampus, thalamus and networks vulnerable to substance-induced modulation.
amygdala76. Thus, addiction-relevant GWAS have begun to reveal evi-
dence that is convergent with the three-stage neurobiological model of Brain imaging genetics
addiction and highlight the role of predisposition within this framework. The conceptualization of brain structure and function as intermedi-
Drug exposure-induced transcriptomic changes have also been ate phenotypes or endophenotypes82 (which are hypothesized to lie
observed. In one study, human alcohol dependence GWAS-associated between genes and/or environmental experiences and disease pro-
genes showed networks of co-expression in the prefrontal cortex, NAc cesses) generated enthusiasm that genetic research on neural pheno­
and ventral tegmental area of ethanol-exposed mouse brains77. Another types would help characterize the genetic architecture underlying
study found cross-species conservation in gene expression changes complex behaviour, including addiction. For example, the high her-
associated with cocaine exposure within the hippocampus and ventral itability of brain structure83 alongside its objective quantification,

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 45


Perspective

Glossary

Anhedonia Fractional anisotropy Incentive salience Predictive reward signals


The loss of pleasure or lack of reactivity A measure of the degree of anisotropy A cognitive process that motivates Neural signals that demarcate the
to pleasurable stimuli. of a diffusion process ranging from behaviour towards reward. expected delivery of reward following
0 to 1. In the context of diffusion extrinsic and/or intrinsic cues.
Binging tensor imaging, it reflects the uniform Machine learning
Consuming a large amount of a directionality of white-matter fibres in A data-driven approach that iteratively Predispositional liability
substance (typically alcohol) in a short the brain and is often conceptualized as examines a training data set for The aspect of an outcome that is
period of time. an index of white matter integrity and patterns across large numbers and attributable to predispositional (that
structural connectivity. diverse types of variables associated is, genetic variation, prior experiences)
Candidate gene with an outcome and, upon ‘learning’ factors.
A gene posited to be associated with a Gene variants these data patterns, can be used to
phenotype based on prior knowledge. Sections of DNA sequence that differ test whether these patterns accurately Regulatory elements
across groups of individuals. predict the outcome in independent Components of a gene, such as the
Compulsive use data sets. promoter and introns, that regulate its
Drug consumption that is not under Genetic architectures expression.
control and typically functions to Distinct genetic factors that influence Negative reinforcement
achieve drug-present homeostasis and one or more traits. The removal of something unpleasant Resting-state functional
alleviation of negative affect/withdrawal or uncomfortable by a stimulus and/or connectivity
as opposed to drug-induced euphoric Genetic liability behaviour. Correlated signal between brain regions
reward. The contribution of genetic factors to in the absence of any stimulus or task.
the likelihood of observing a phenotype. Negative urgency
Craving A personality facet related to impulsive Single nucleotide
A persistent desire to use a substance. Genetic nurture behaviour in the context of negative polymorphism
The effect of genetically influenced mood or experiences. (SNP). A single base pair in the genome
Developmental vulnerability parent behaviour on offspring behaviour. that varies across individuals.
Vulnerability to a given outcome Pleiotropic effects
that arises in the context of typical Genome-wide association The influences of a variant, gene Trait-like vulnerability
development. studies or groups of variants on multiple Vulnerability to a given trait.
(GWAS). A hypothesis-free analysis phenotypes.
Executive function of the association between common Twin studies
Complex mental processes and genetic variation across the genome Polygenic Comparisons of phenotype correlations
cognition (for example, working and a phenotype. The genetic characteristic of traits that in identical and fraternal twins to parse
memory) that control skills (for example, is due to the aggregated small effects the role of genetic and environmental
organizing, solving) and regulate Genomic structural equation of many genetic variants. effects on a given phenotype or set of
emotion and behaviour. modelling phenotypes.
A statistical genetics method for Positive reinforcement
Expression quantitative trait identifying genetic variants that Reward obtained after a stimulus Withdrawal
loci influence multiple phenotypes as well and/or behaviour. Physical (for example, headaches
(eQTL). Genetic variants that modify the as each individual phenotype. and insomnia) and psychological (for
expression of a gene by acting upon the Positive urgency example, depressed mood) aversive
regulatory elements of the gene. Heritability A facet of personality related to experiences that occur when use of a
The proportion of total variation in a impulsive behaviour in the context substance is discontinued.
phenotype that is due to genetic factors. of anticipated reward.

reliability and proximity to gene function were the basis for presuming genetic overlap, its characterization will be difficult without large sam-
that GWAS of brain structure would yield loci with large effects82. Meta- ples. Notably, these genetic correlations may be constrained by small
analyses of the GWAS of structural brain phenotypes found, instead, effects of brain–behaviour associations91.
that brain imaging phenotypes are highly polygenic and complex Unravelling the genetic architecture of functional neuroimaging
and that their genetic correlation with behavioural outcomes is far phenotypes (such as task-related activity or resting-state functional
more modest than hypothesized84–87. For example, initial genetic corre- connectivity) has been even more challenging. GWAS of task-related
lations estimated between GWAS of substance involvement (including functional MRI (t-fMRI) and resting state functional connectivity
use, problematic use and SUDs) and brain structure were modest; how- pheno­types in the UK Biobank have revealed low heritability and identi-
ever, their magnitude was similar to the genetic correlations between fied few loci86,87,92. This may be partially attributable to the low reliability
these same substance involvement GWAS and other behavioral of traditional t-fMRI93 and the need for large amounts of resting state
traits88–90. These data suggest that, while there is gene–brain–behaviour functional connectivity data to facilitate its reliable measurement94.

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 46


Perspective

Thus, despite the prominence of functional neuroimaging studies of cannabis use was associated with cortical thinning in the superior
of addiction95, the vanishingly low heritability, psychometric chal- and anterior medial prefrontal cortex. This change occurred over a
lenges and practical difficulties in harmonizing the findings of different 5-year period among participants who were cannabis naive at baseline
studies have led to relatively few well-powered genetically informed (age 14) and in a dose-dependent manner that was also associated with
investigations. impulsivity100. However, other longitudinal studies have found evi-
dence supportive of the predispositional model. For instance, reduced
Predispositional and/or causal? DLPFC volume in children who were substance naive was associated
As reviewed above, GWAS and transcriptomic studies of addiction- with an earlier age of drinking initiation when those children reached
related phenotypes highlight the role of predisposition within the adolescence and an attenuation of the typical reduction in drinking
three-stage neurobiological model as well as potential substance- that occurs in young adulthood73. Similarly, two studies found that
induced changes in epigenetic structure and the transcriptomic land- lower orbitofrontal cortex volume in early adolescence preceded and
scape. By contrast, the vast majority of brain-imaging addiction-related was associated with subsequent onset of cannabis use101,102 (but, for
science has interpreted cross-sectional associations between substance evidence of larger volume predicting cannabis use, see ref.103).
involvement and brain phenotypes to putatively reflect causative Two considerations are noteworthy when interpreting longitudi-
substance-induced brain alterations. Below, we review evidence from nal studies. First, adolescence is characterized by dynamic changes in
longitudinal and genetically informed designs that can be used to infer brain development. As such, individual differences in the trajectories
whether substance-related variability in brain structure may plausibly of brain development in youth who initiate substance use may reflect
reflect predispositional risk and/or sequela of substance involvement. substance-induced changes and/or gradations of predispositional
Much like GWAS, these data highlight the need to incorporate genetic influences on neurodevelopment. While experimental evidence
predisposition into neurobiological models of addiction. in non-human animals shows that heavy substance use can reduce
markers of neurogenesis and brain growth104–107, other evidence sug-
Longitudinal studies gests that neural development is significantly genetic in origin108–111.
Longitudinal studies of substance involvement have revealed that Therefore, disentangling the aspects of brain development that are
changes in brain structure and function are associated with escalating attributable to genetic predisposition from those that are sequelae
substance use. For example, the National Consortium on Alcohol and of substance use is challenging, especially when the genes contribut-
Neurodevelopment in Adolescence (NCANDA)96 has found that heavy ing to brain development may exert pleiotropic effects on substance
alcohol use in adolescence is associated with, and precedes, acceler- involvement.
ated cortical grey matter decline, particularly in the medial and dorsal Second, while a dose–response relationship (in which the more
prefrontal cortices97, as well as a decline in white matter integrity98. Simi- one uses a substance, the stronger the association with brain met-
larly, in a cohort examined as part of the IMAGEN study99, the initiation rics) may reflect causal effects112, it is also possible that those with

Fig. 2 | The genomic architecture of SUDs. The


General addiction liability
genetic contribution to individual substance use
disorders (SUDs) is attributable to variants that
influence general addiction liability and substance-
specific variants63. General addiction liability is
Variants for risk-taking/ Variants for negative Variants for executive driven by variants influencing traits that correspond
reward-related impulsivity urgency/negative affect function to the three stages of the neurobiological model of
addiction: reward and risk-taking, negative affect
and urgency, and executive functioning. By contrast,
Traits related to neurobiological risk variants in receptors that respond to the psychoactive
Risk taking Negative affect Executive function components of individual substances or those in genes
metabolizing individual drugs directly influence each
SUD in a substance-specific manner. Furthermore,
genetic variants that influence other psychiatric
Substance use disorders disorders may also independently influence SUDs.
Reciprocally, the genetics underlying general
Alcohol Nicotine Cannabis Opioid Other addiction liability may impact risk of other psychiatric
disorders. Small effects of substance-specific
genetic variants on other psychiatric disorders
are also predicted. In addition to these genetic
Variants in Variants influencing other Subjective well-being and pathways, prolonged substance use and SUDs may
substance-specific genes psychiatric disorders other psychiatric disorders phenotypically influence risk-taking, negative affect,
executive functioning, and psychiatric health and
well-being (double headed dashed red arrows depict
phenotypic associations). Of note, alcohol, nicotine,
Observable traits Latent traits
cannabis, cocaine and opioid use disorders are shown
Effect of genetic liability Effect of genetic liability Reciprocal phenotypic as there are current large genome-wide association
to a trait on the trait itself to a trait on another trait association between traits
studies of these SUDs; however, the genetics of many
(independent of genetics)
other SUDs could be similarly classified.

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 47


Perspective

pre-existing neurobiological liabilities that manifest in altered neuro­ light drinking twins from discordant pairs did not differ in their grey
developmental trajectories in adolescence may be more vulnerable to matter volume nor from twins in pairs where both were heavy drinkers,
escalating and disordered substance use112,113. Indeed, GWAS show that suggesting that the reduced grey matter volume that is associated with
increasing severity of drug use is associated with a polygenic signal alcohol use does not arise as a consequence of use. A recent series of
that is partially distinct from the genes influencing lighter or milder studies of 436 twins 24 years of age from the Minnesota Twin Family
use and is more likely to exert pleiotropic effects on brain develop- Study120 documents support for both predispositional and causal
ment (even when we make simplified assumptions that the effects contributors to associations between substance involvement and
are linear)114. cortical thickness. For instance, a thinner medial orbitofrontal cortex
among those with alcohol, tobacco and/or cannabis use disorders was
Genetic causal modelling attributable to predispositional risk, with some evidence that alcohol
Genomic data can be used to examine the plausibility of hypotheses of or cannabis use disorder may also contribute to these reductions121.
phenotypic causality. Because modest but significant genetic correla- In the same sample, an index of alcohol (but not cannabis) use was
tions have been demonstrated between psychiatric phenotypes and associated with a thinner cortex overall, with evidence that this reflects
brain phenotypes46,49,90, the associations between variants at addiction- both predispositional risk and a potential consequence of alcohol
relevant loci and neural phenotypes have been partially attributed to exposure122. Collectively, these analyses reveal that brain structure
the pleiotropic effects of these variants. However, if addiction is con- correlates of substance use and SUDs may reflect a predispositional
ceptualized as the escalation of drug exposure and the brain phenotype liability to substance involvement as well as a potential consequence
as the target of such prolonged exposure, then any effect of variants of exposure.
in drug-related loci on the brain may represent genetic causality (and
this could be tested via Mendelian randomization115; Fig. 3). Modern Genetic predisposition
genetic causality approaches that account for the polygenic nature of Evidence for correlations between genetic variants identified in
SUDs can estimate the proportion of shared genetic liability or identify GWAS of brain imaging and those identified in GWAS of substance
genetic variants that might be causal116. One such analysis found no use and SUDs89,90 suggest that associations between brain structure
support for a genetically causal effect of problematic alcohol use on and addiction-related behaviour that occur after the onset of exposure
brain structure phenotypes89. By contrast, and consistent with a pre- are confounded by pre-existing pleiotropic liability. Thus, nearly any
dispositional model, it did provide support for the idea that differences evidence for causation would also support predisposition. Ideal sup-
in brain structure (including a lower volume of the basal forebrain and port for predispositional effects arises from studies of brain develop-
a greater volume of the pars opercularis) may plausibly contribute to ment in individuals before their drug exposure. Family studies provide
problematic alcohol use. persuasive support for such pre-existing brain differences in those
genetically enriched for addiction liability123,124. Youth with a family his-
Discordant designs tory of alcohol use disorder have thinner frontal and parietal cortices111
Monozygotic (identical) twins that are discordant for drug exposure and smaller frontal grey matter volume125, larger grey matter volumes
serve as a natural quasi-experiment for the study of causal effects of of the cerebellar lobes126, NAc127 and amygdala110,128, and task-related
drugs in humans117 (Fig. 3). If the correlation between drug use and brain response variability in brain regions related to reward response and
structure is entirely genetic, the brain structure in monozygotic pairs decision-making129–133 than those without such a family history. These
discordant for substance use (or SUDs) would not be different. If, on studies provide compelling support for the association between family
the other hand, there is a difference in brain structure, then contribu- history of addiction and brain structure and function, which (in some
tors beyond shared genetics and prenatal and familial environment instances) was investigated prior to substance use onset, and further
are implicated; these contributors may be exposure specific (that is, support the neurodevelopmental hypothesis because a family his-
causal effects) or due to person-specific third variables (such as early tory of alcohol use disorder was also associated with early behavioural
trauma that motivates substance use and, independently, modifies undercontrol134.
brain structure in that twin). While it is ideal to study identical twins, However, family history is an amalgamation of inherited genetic
dizygotic twins and even non-twin siblings (close in age) can also be risk and genetic nurture135 and can be biased by lack of adequate meas-
used to further parse non-genetic sources of similarity (for example, urement of familial density of risk136. Twin studies disentangle these
dizygotic twins are more closely matched for prenatal exposures than familial effects to some degree by either explicitly modelling genetic
non-twin siblings). nurture or separating genetic and family environmental factors within
Several studies have examined structural brain differences in the offspring137,138. Well-powered neuroimaging studies that also assess
twin pairs who vary in their drug exposure. In data from the Human family history and future substance use, particularly during the devel-
Connectome Project118, it was shown that an association between opmental period prior to onset of substance use, are rare; however, the
cannabis use and reduced volume of subcortical structures was no Adolescent Brain and Cognitive Development (ABCD study) provides
longer apparent when cannabis-using individuals were compared with an opportunity to evaluate the interrelationships between brain and
their co-twins or age-approximate siblings, consistent with the robust substance use development in individuals from ages 9–10 years into
estimates of genetic correlation between cannabis use and subcorti- early adulthood139. In the baseline data from this sample, total brain
cal brain volume119. Similarly, it was discovered that the correlations and regional cortical and subcortical volumes, cortical thickness and
between alcohol consumption and insula and DLPFC volumes were surface area, and fractional anisotropy and mean diffusivity indices
primarily attributable to predispositional factors73. Instead of relying were examined for their association with polygenic liability to alcohol
on discordancy for alcohol use alone, this study also contrasted brain consumption and problem drinking114 (Fig. 3). The polygenic risk score
structure in twin pairs in which both individuals were heavy drinkers as (PRS; the aggregated effects of risk alleles associated with a trait140) for
well as in pairs in which both individuals were light drinkers. Heavy and problematic alcohol use was associated with a lower volume of the left

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 48


Perspective

a b
Genetic 3
variation
Genetic Substance Neurotoxic 1
variation use effects
Changes in brain 2 Substance 4 Neurotoxic
development use effects
Changes in brain
development
5 Changes in brain
development

c d
Population Discovery GWAS Sibling (twins)
of addiction shared genetics

Random segregation
of alleles to gametes
Create a polygenic No substance Substance use
score in an use
independent sample
Risk alleles for
Protective/null
exposure (for
effect alleles
example, addiction)
Brain phenotypes Brain Brain
Substance use
(for example, in phenotype phenotype
and addiction
those who are
yet to use drugs)
Exposure
(for example, Non-chronic
chronic substance use
substance use)
Outcome No outcome
(for example, (for example, no
changes in brain changes in brain
development) development)
Fig. 3 | Using genomics to validate hypotheses of addiction. a, According to Mendelian randomization approaches as well as their limitations, see ref.115.
neurobiological models of addiction, genetic variation influences substance d, Testing the association between polygenic risk140 for addiction and brain
use, which may, in turn, exert neurotoxic effects that alter brain development. imaging phenotypes, including trajectories, in drug-naive individuals
b, According to predispositional models of addiction, genetic risk for substance (left flow chart) is an ideal approach to assess whether pre-existing brain-
use disorders impacts brain development (1) prior to or concurrent with the related differences precede addiction. Here, the effects of genetic variants are
onset of substance use and its escalation and sets the neurobiological stage for taken from a discovery GWAS of addiction and applied to a sample, ideally of
substance use and future addiction (2). Consequent substance involvement (3) individuals without a history of substance use (for example, children), which
(also influenced by genetic risk that is not associated with neural phenotypes) has brain data. A polygenic risk score is created in this new independent sample.
may then causally influence the brain, via neurotoxic mechanisms, to further It is expected that this polygenic risk score will eventually be associated with
potentiate problematic substance use (4). Cyclically, these brain-related substance use and addiction in this sample. However, if it is also associated
changes may further enhance risk for addiction progression (5). c, Mendelian with brain phenotypes prior to the use of substances, then we can infer that
randomization115 and other genetic causal methods can be used to evaluate genetic risk that precedes the onset of substance use contributes to brain
these models. These approaches are based on the fact that parental genotypes development (part b, 1) and later substance use (part b, 3) rather than a causal
conferring risk of exposure (that is, chronic substance use) are equally as effect of substance use on the brain alone (part b, 4). Alternatively, examining
likely to be inherited by the offspring as genotypes that are protective or of no twins (or similarly aged non-twin siblings) that are discordant117 for substance
effect. Individuals inheriting risk alleles or polygenic risk of substance use will involvement can provide information on whether substance-related neural
subsequently be more likely to use drugs; we can then test whether this chronic phenotypes arise from predispositional influences and/or are induced through
use causally alters brain development. In this method, the individual risk alleles substance involvement (right flow chart). If the brains of genetically similar
(or the polygenic risk of drug exposure) are the genetic instrument, and an individuals differ as a function of their substance use, then non-genetic
independent association between this genetic instrument and the outcome mechanisms, including substance-induced changes, might be implicated.
(changes in brain development), as shown in the flow chart, is possible evidence However, if brain phenotypes are similar among those discordant for substance
for causal effects of substance exposure on the brain. The genetic instrument use, this would suggest that predispositional effects, including shared genetic
is assumed to influence the outcome (changes in brain development) solely via variation and environmental exposures, are responsible for their associations
its influence on chronic substance use (dashed line). For a greater discussion of with substance involvement.

frontal pole and greater cortical thickness of the right supramarginal these analyses, data were excluded from the small subset of youths who
gyrus, although nominally significant associations for both typical report substance use141. However, an even earlier epoch of substance
and problematic alcohol use PRS and insula metrics were evident114. exposure — prenatal exposure — also merits consideration. In the ABCD
In another analysis, the PRS for cannabis use disorder but not for can- sample, prenatal exposure to cannabis, particularly beyond the first
nabis use was associated with lower white matter volume46. In each of trimester, is correlated with psychopathology (but not global brain

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 49


Perspective

a
Exposure opportunity Drug expectancies

Onset of use/
experimentation

Occasional,
casual use

Heavy episodic use


Cessation
(binge/intoxication)

Three-stage
addiction
model Sustained
heavy use

Preoccupation– Negative affect–


craving withdrawal

b Heavy use (for example, Drug-specific


episodic or sustained polygenic risk
binge-intoxication)

Striatal/
midbrain

Reward-
related
risk
taking Negative
Negative affect/
Preoccupation Prefrontal Executive urgency Limbic
and craving functioning withdrawal/
and affect stress
response

Polygenic core and drug-specific polygenic risk Environment Brain substrate

structure) outcomes142 and persists as children enter their teens143. influences on behavioural undercontrol and substance use152–156.
Therefore, the study of predisposition that is exclusively related to Similarly, genomic studies are identifying drug-induced epigenetic
genotype requires consideration of family history, genetic nurture and ­alterations in relevant brain regions157.
other third-variable confounders as well as prenatal exposure to sub- Based on these foundational discoveries, we outline an integrated
stance use. However, even studies of prenatal exposure are confounded framework for the development of addiction — the GINA model (Fig. 4).
by intergenerational transmission of genetic predisposition144–151. At the core of the GINA model is polygenic liability. The neural circuitry
underlying reward, negative affect and executive function as well
The GINA model as the drug-specific pathways (such as those featuring drug recep-
Brain imaging studies of addiction have tended to invoke drug-induced tors or enzymes involved in drug metabolism) serve as the substrates
mechanisms of effect whereas genomic studies have mostly relied within which polygenic liability, risk and sequela of addiction unfold.
on predispositional aspects of vulnerability. Each domain implicitly Environmental factors serve as the filter through which gene–brain
considers the relevance of the other to some degree. For instance, associations influence addiction-related behaviour. While the GINA
the neurodevelopmental model references common latent genetic model is presented as a framework for integrating imaging and genetics

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 50


Perspective

Fig. 4 | The GINA model. Addiction may be conceptualized as a developmental environment may be equally important). Each stage aligns most closely with
process or as a syndrome comprised of stages of escalating problem use. While genetic predispositions that act via specific posited brain mechanisms (these
the Genetically Informed Neurobiology of Addiction (GINA) model described are shown ‘stacked’ below that stage of the addiction cycle). All of the addiction
here outlines a testable gene–brain–behaviour mechanism underpinning stages are influenced by a polygenic core, which broadly corresponds to trait
the stages of addiction, it is scalable and can be extended to advance our representations of substance-induced stages of sustained heavy use (binge–
understanding of the process of addiction. a, An illustration of the process intoxication), negative affect (withdrawal–negative affect), and preoccupation or
of addiction, and of those that lead into addiction, serves as a framework for craving (preoccupation–anticipation) and by additional drug-specific polygenic
understanding the GINA model. Exposure opportunity, availability193,194 and risk that influences addiction, partly via the brain (for example, variants in genes
initial expectations surrounding the anticipated subjective effects of substance encoding neurotransmitters) as well as via non-brain mechanisms (such as
use serve as early contributors to drug-seeking behaviours and increase the metabolic variants). Polygenic liability to reward-related risk-taking contributes
likelihood of substance use195–198. Onset of substance use occurs in a subset of to initial phases of binge–intoxication and may promote later escalating use
individuals, with some further entering a phase of casual but repeated substance (shown as heavy — episodic or sustained — use), which plays a role in promoting
use. Depending on the addictive potential of the substance, progression through the reward-related neural response to pleasurable aspects of substance use
periods of heavy episodic use and cessation may then occur (intervening (for example, striatal brain regions6). On the other hand, chronic substance use
aspects of these processes are not depicted). For some substances, periods induces brain-related alterations that culminate in heightened stress states and
of primarily reward-related occasional or casual use, heavy episodic use and negative affect (for instance, those with polygenic liability to negative urgency
cessation may occur (for example, heavy drinking limited to college), during may be more vulnerable to this pathway) via a modified limbic response6.
which time individuals may even meet criteria for milder forms of substance Furthermore, polygenic liability to executive function is likely to be instrumental
use disorders199–201. Not shown are the numerous genetic and environmental in drug craving via changes in prefrontal brain function that result in increasing
influences that promote or deter progression through these substance- difficulties regulating the emotional salience of substance-related stimuli despite
interfacing behaviours. For a further subset of individuals, heavy episodic use the potential of subjective and cognitive desires to stop. Despite the appearance
advances into a phase of sustained heavy use, wherein the pleasurable aspects of in this schematic of a one-to-one correspondence between polygenic liability,
substance use are attenuated and compulsive use emerges to ameliorate negative brain region and addiction stage, the gene–brain–behaviour map is likely to
affect, psychological and/or physiological stress states, and physiological be more interconnected. For instance, sustained heavy use in the context of
withdrawal symptoms164. Withdrawal, and related negative mood, following negative affect may be influenced by polygenic risk to negative urgency and
substance abstinence leads to potentiated interoceptive salience through which affect via limbic pathways and substance-induced alterations in striatal circuits.
physiological arousal associated with withdrawal and negative emotionality The environment provides a filter for genetic liability (that is, the magnitude and
are potentiated21. We propose that this phase reflects moderate to severe nature of genetic effects may be different in differing environmental contexts)
forms of substance use disorders. b, The neurobiological model of addiction and also directly underpins addictions. The brain is depicted as the outer
in the GINA framework. The GINA model places the three stages of addiction substrate from which psychological aspects of addiction emerge. While distinct
(shown around the outside of the image) within the context of a polygenic core brain systems are illustrated, it is likely that networks of brain regions correspond
(grey), environmental filter (blue) and brain substrate (red; width of circles to the three stages of addiction. While not noted here, aspects of addiction arise
does not correspond to any relative magnitude of effect, that is, genes and from and impact other bodily systems as well as the brain.

studies of addiction in both clinical and population cohorts, it can also metabolizing enzymes166 can approach those seen for the apolipopro-
be extended to evaluate the onset of use, occasional use and milder tein ε4 variant and Alzheimer disease risk167. Studying such pronounced
forms of SUD. (but scarce) genetic effects alongside polygenic patterns of common
and drug-specific genetic risk requires novel statistical approaches
Polygenic core that can handle mixtures of distributions of genetic effect sizes and
As a simplifying principle, we posit that four key domains of genetic risk conditional analyses. However, drug-specific loci that encode the
form the polygenic core of addictions. Three are common to all addic- neurotransmitter targets for a drug are rarely so specific. For instance,
tions: genes affecting reward and risk-taking (notably, in the context the rs16969968 variant in CHRNA5, encoding a nicotinic receptor,
of positive urgency), genes affecting negative affect and/or suscepti- was shown to be highly significant for tobacco use phenotypes and
bility to negative urgency, and genes affecting executive functioning also associated with schizophrenia and educational attainment168,169.
and/or regulation. Genetic measurement of these domains, especially Neuroimaging studies of this variant (rs16969968) have linked the
as they pertain to addiction liability, remains incomplete and under- risk allele to greater hippocampal activation in response to smoking
specified. For instance, negative urgency is a hallmark characteristic cues170 and with reduced resting-state connectivity between the dorsal
of SUDs and some comorbid mood disorders26,158. However, current anterior cingulate cortex and the ventral-striatopallidal circuit171 but
GWAS of negative affect rely on heterogeneous constructs (such as with null effects on brain differences in light-smoking adolescents172.
depression or neuroticism)159,160. The genetic disarticulation of the Drug-specific loci also capture some variability in responses to existing
sub-facets of these composites161 (such as negative urgency)70,161,162 as treatments for addiction but findings are mixed173 and specific GWAS of
well as well-powered GWAS of addiction-relevant indices of negative pharmacogenomics response are needed174. Notably, drug-specific loci
affect (such as using substances to cope or stress-responsivity163,164) are evident in GWAS of both substance use and addiction (for example,
using approaches such as genomic structural equation modelling165 will the rs1229984 variant in ADH1B is significant for typical, maximum
be required to fine-tune this polygenic core from an index of general- habitual, and problematic and/or disordered drinking)47,50,175, suggest-
ized risk for psychopathology to an addiction-specific liability factor. ing that their influence on addiction may be routed via their regulation
The fourth source of genetic variability arises from genes encod- of substance consumption and the subjective, and possibly interocep-
ing drug-specific metabolic factors and receptors; while polygenic tive, effects associated with use176–179. Therefore, in the GINA model,
in architecture, some of the drug-specific single loci may exert rela- we place drug-specific polygenic liability in the context of exposure
tively large effects. For instance, the effect sizes of variants in alcohol and, most notably, of heavy episodic and heavy sustained use (Fig. 4),

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 51


Perspective

where it regulates subjective response and sets the pace for entry into The latter represents the form of escalating chronic use that aligns with
and progression within the three-stage addiction model. current conceptualizations of heavy use in the context of addiction.
While behaviourally distinct, these aspects of binge–intoxication may
Brain substrate share genetic and neurobiological contributors.
The GINA model provides a framework for gene–brain–addiction
mechanisms from which we can develop testable hypotheses. For Characterizing substance use in the GINA model
example, polygenic liability to executive function deficits might modify It could be argued that, rather than being classified as disorders, addic-
prefrontal regulatory capacity and, in turn, potentiate preoccupation tions would be better represented as a process or series of interactions
with drugs. However, it would be reductive to assume a one-to-one cor- with psychoactive substances that, in some instances, becomes dis-
respondence between polygenic liability, brain region and behavioural ordered. From this perspective, the addiction process (Fig. 4) begins
manifestation. For instance, it is highly likely that striatal circuitry is with exposure opportunity and expectations regarding the drug use
sensitive to the stages of addiction that evoke positive urgency and that experience. These early stages are strongly motivated by environmental
polygenic liability to risk-taking as well as executive function (that is, factors, although gene–environment correlations influence drug avail-
undercontrol) undergird positive urgency. Polygenic liability to risk- ability and exposure opportunity193,194. Upon onset, initial experiences
taking is also likely to contribute to other stages of substance use and (which may be heritable) and subsequent experiences (which may be
addiction (such as relapse)180 and to affect other brain regions beyond subjective) with individual drugs motivate or deter further use195–198.
striatal regions (Fig. 4). Multi-method studies that integrate polygenic Continued use represents a mixture of pathways — heavy episodic use
risk with multivariate brain and behavioural phenotypes will be neces- may become entrenched and transition to addiction or, for substances
sary to broaden the scope of gene–brain–addiction connections. For with lower addiction potential, settle into patterns of socially accepted or
example, machine learning-based approaches coupled with large-scale intermittent use. Many individuals attempt to quit drug use in their 20s
data181 could be used to perform a systematic, data-driven study of the and 30s, and the GINA model features both early199–201 and later cessation.
complexity underlying the GINA model. The same genetic core that contributes to the stage-based neurobiologi-
cal model is also likely to contribute to these aspects of the addiction pro-
Environmental filter cess. For instance, genetic propensity to risk-taking may motivate early
While not detailed in this Perspective, the GINA model incorporates drug-seeking behaviours64, and some neuroimaging studies suggest pre-
environment as the filter through which addiction emerges. Similar existing brain differences in youth at risk for substance use onset73,101,102.
to genetics, some environmental factors (such as life stress) will gen- Likewise, initial and typical subjective responses to individual substances
eralize across substances and other psychopathology while others may be influenced by variation in drug-specific loci176,196.
(such as policy, taxation and distance to alcohol outlet) are likely to
be more substance specific (although policies do have cross-cutting Relationship with other heuristics
effects182). A proportion of the environmental impact on addiction The GINA model represents our conceptualization of the vast com-
also involves neurobiological mechanisms. For instance, trauma plexities that underlie addictions and is inspired by the extensive
(especially when occurring during early life) is associated with brain output of psychology, psychiatry, neuroscience, genetics and trans-
development and exacerbates addiction risk183–185. However, the envi- lational research generated by international teams of scientists, par-
ronment and genetic susceptibility may be related. Some traumatic ticularly those who forecasted a need to bridge brain and genome
experiences are correlates of genetic risk (such as passive exposure to research120,123,202–206; it is certainly not unique in adopting a multifactorial
early adverse environments that are a product of parental genotype) view. Instead, it represents a conceptual increment that has resulted
while others are modifiers of polygenic liability (such as trauma that from novel study designs, genetic discoveries and rapid increases in
moderates polygenic liability to addiction) and gene expression (such availability of genetically informed neuroimaging data. For instance,
as trauma-induced epigenetic changes)186–189. the Addictions Neuroclinical Assessment (ANA)207 was developed to
guide researchers in designing studies that might test the three-stage
Characterizing addiction in the GINA model neurobiological model of addiction. However, the GINA model provides
The GINA model characterizes addiction as a transition from episodic to a framework from which a series of hypotheses can be tested using
sustained heavy use in the context of emerging negative affect and pre- ANA-derived data. A previous study also integrated evidence from
occupation (Fig. 4). From a diagnostic perspective, this coincides with behavioural, genetic and neuroimaging studies to provide a framework
moderate and severe SUDs, as defined in the Diagnostic and Statistical for a common liability to addictions208 and tested it using multiple
Manual of Mental Disorders, Fifth Edition (DSM-5)190. Recently, individu- sources of data209. These studies were prescient in anticipating the role
als endorsing between two and five DSM-5 criteria (mild or moderate of dopaminergic pathways on common addiction risk, although the
SUDs) were classified as being in a high-risk, sub-threshold state of GINA model has the advantage of leveraging contemporary insights
pre-addiction191 (similar to pre-diabetes) where interventions may be from GWAS to advance a polygenic framework. Other schemas, such
maximally beneficial. While the GINA model is more closely aligned with as the Hierarchical Taxonomy of Psychopathology (HiTOP)210, aim to
addiction per se, those with higher pre-addiction scores, depending on outline the common genetic and neurobiological underpinnings of a
their individual symptomatology, may well be described by the GINA broad range of psychopathologies, including addictions. While the
model. While reminiscent of the three-stage neurobiological model, GINA model includes cross-disorder components (Fig. 2), it is clear
the GINA model separates the broader binge–intoxication stage into that addictions are not merely the product of generalized genetic
heavy use that is either episodic or sustained. The former represents liability to broad-spectrum psychopathology and chronic substance
intermittent reward-motivated accelerations in use (such as alcohol use63. In spirit, the GINA model is aligned with the newly hypothesized
consumption in college students) and may also capture mild SUDs as Aetiologic, Theory-Based, Ontogenetic Hierarchical Framework of
defined in DSM-5 (ref.192) and is thus somewhat distinct from addiction. Alcohol Use Disorder (ETOH)211, which highlights the importance

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 52


Perspective

of distinguishing between common and drug-specific mechanisms of emerging HEALthy Brain and Cognitive Development (HBCD) study,
risk and considering premorbid and drug-acquired pathways of influ- which begins during the prenatal period, will be critical for this process.
ence. However, the ETOH model is finely tailored towards advancing Second, as genomic and imaging data become available in larger,
our understanding of alcohol use disorder, whereas the GINA model population-representative settings that span developmental periods,
focuses on addictions more broadly in the context of common and the GINA model will require retooling to account for lighter and moder-
drug-specific genetic liability. ate levels of substance use and milder forms of addiction. Distinctions
between levels of addiction severity may impact the magnitude of
Predisposition in the context of drug-induced change neurogenetic associations or may map onto qualitatively different
The central aim of the GINA model is to establish the role of genetic polygenic signals and brain regions. From a genomics perspective,
predisposition as a source of individual differences in substance- larger-scale discovery GWAS will profoundly impact the GINA model.
related neural phenotypes. Most neuroimaging studies speculate Currently, far too few GWAS are well-powered to identify genetic sig-
that neuro­biological associations arise as sequelae of drug use, while nals in individuals of ancestries other than Europeans, which limits
widespread evidence of pleiotropy212 motivates geneticists to be partial the equitable application of PRS218. In addition, while smaller studies
to correlational mechanisms — the GINA model presents one possible of copy number variants exist219,220, large-scale meta-analyses of struc-
reconciliation of these disparate perspectives. Simply put, while the tural variants221 are needed. Transcriptomic analyses provided some of
consequence of chronic drug use may be evident in the brain, genetic the earliest validation of GWAS discoveries; however, most have been
factors that predate onset of substance use may influence whether or conducted in bulk tissue and greater specificity is needed. Results for
not an individual will initiate drug use, escalate in use and progress to other psychiatric disorders have highlighted the importance of study-
addiction. Furthermore, trait correlates of substance-induced states ing the enrichment of genetic signals in single cells or single nuclei, thus
serve as additional sources of individual differences (Fig. 4). For exam- partitioning heritability (even within a single brain region) into specific
ple, we theorize that limbic adaptation to chronic drug use is likely to cell types222,223. Beyond eQTL, other cell type-specific annotations, such
induce a negative affect state in most individuals, but that those with as enhancer effects224, could also be incorporated into PRS develop-
a genetic predisposition to negative affect may be more susceptible ment. Such data are essential to providing a neuro-cellular perspective
to this transition. Thus, individual differences in the developmental to the GINA model; however, at this time, larger GWAS are likely to be
tempo and severity of the individual stages are likely to be due to poly- the most influential source of data225. Nonetheless, multi-omics data
genic liability, either via direct effects of genomic variation or via early also provide valuable data matrices for emerging methods to develop
influences on brain variability, similar to environmental moderators functional PRS226–228, which appear to be more portable across ancestral
of risk. populations than traditional effect size-based PRS229.

Conclusions and perspectives A consideration


Future directions Brain neuroimaging and genetics have substantially advanced our
Heuristic models, such as the GINA model, are intended to be dynamic understanding of the biological bases of addiction. Both approaches
and riddled with chasms that anticipated advances in genetics, have opponents who have argued that the conceptualization of addic-
neuro­science and psychiatry can bridge. Recent criticism of genetic tion as a ‘brain disease’ discounts the role of socio-political factors,
approaches surrounds the questionable prognostic utility of addiction which may be more modifiable than one’s biology230,231. Yet, influential
PRS213. Neuroimaging data are also associated with small effects91,214,215, environmental provocations may, in fact, not be predictable nor malle-
and researchers in this field have grappled with methodological chal- able. There are also deeper ethical issues surrounding the use of either
lenges such as the reliability and heritability of t-fMRI studies86,93. genetics or neurobiology — or for that matter, environmental factors —
The following emerge as priority areas for improvement. to prospectively categorize an individual by their future addiction
First, there is a need to incorporate evidence from studies of the diagnosis. It is therefore worth considering the trade-off between using
trajectories of brain development and from studies of brain function. genetics and neuroscience to prognosticate addiction risk as a source
Most addiction neuroscience work has focused on brain function95. of stigma versus the potential for predispositional mechanisms to
However, typical measurements of t-fMRI and resting-state functional unburden individuals and societies of ill-construed notions regarding
connectivity are characterized by low heritability87, which may arise the moral valence of persons using substances.
(at least in part) from reliability challenges93,94. Going forward, out-
lining the genetic architecture of brain function will require better Summary
phenotyping. In addition to improving the reliability of univariate Weaving genetics and neuroscience together, especially in
metrics (such as dense-sampling94), multivariate approaches may the context of environmental considerations, can provide
have great utility. Indeed, polyneural phenotypes (that is, those with appreciable insights into the origins of addiction. It may even
differentially weighted regional associations, much like PRS) may be provide clues for prevention of the critical transition from non-
required to meaningfully predict individual differences in complex problematic to maladaptive use and illuminate efficacious thera-
behaviour. While such ‘lumping’ approaches may disappoint those peutic pathways. By proposing the integrative GINA model, we
hoping to link complex behaviour and genetic variation to readily encourage the adoption of a multifactorial perspective of addiction —
interpretable brain regions according to their known role in behaviour, a process that represents a dynamic cascade of genetic predispo-
they may be replicable and reliable, similar to polygenic approaches216. sition and environmental risk and resilience that is enacted via
Characterizing the longitudinal trajectories of brain development that developmentally relevant brain maturation and substance-induced
contribute to periods of addiction vulnerability is equally important. brain alterations.
Efforts targeting brain development217, especially large longitudinal
samples, such as ABCD of mid-childhood to young adulthood139 and the Published online: 29 November 2022

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 53


Perspective

References 36. Mackey, S. et al. Mega-analysis of gray matter volume in substance dependence: general
1. Peacock, A. et al. Global statistics on alcohol, tobacco and illicit drug use: 2017 status and substance-specific regional effects. Am. J. Psychiatry 176, 119–128 (2019).
report. Addiction 113, 1905–1926 (2018). 37. Hatoum, A. S. et al. Multivariate genome-wide association meta-analysis of over 1 million
2. Volkow, N. D., Gordon, J. A. & Koob, G. F. Choosing appropriate language to reduce the subjects identifies loci underlying multiple substance use disorders. medRxiv https://
stigma around mental illness and substance use disorders. Neuropsychopharmacology doi.org/10.1101/2022.01.06.22268753v1 (2022).
46, 2230–2232 (2021). 38. Schoeler, T. et al. Novel biological insights into the common heritable liability to
3. Agrawal, A. et al. The genetics of addiction-a translational perspective. Transl. Psychiatry substance involvement: a multivariate genome-wide association study. Biol. Psychiatry
2, e140 (2012). https://doi.org/10.1016/j.biopsych.2022.07.027 (2022).
4. Kendler, K. S. et al. Recent advances in the genetic epidemiology and molecular genetics 39. Zucker, R. A. et al. Assessment of culture and environment in the Adolescent Brain
of substance use disorders. Nat. Neurosci. 15, 181–189 (2012). and Cognitive Development Study: Rationale, description of measures, and early data.
5. Gelernter, J. & Polimanti, R. Genetics of substance use disorders in the era of big data. Dev. Cogn. Neurosci. 32, 107–120 (2018).
Nat. Rev. Genet. 22, 712–729 (2021). 40. Chassin, L., Sher, K. J., Hussong, A. & Curran, P. The developmental psychopathology
6. Koob, G. F. & Volkow, N. D. Neurocircuitry of addiction. Neuropsychopharmacology 35, of alcohol use and alcohol disorders: research achievements and future directions.
217–238 (2010). Dev. Psychopathol. 25, 1567–1584 (2013).
7. Volkow, N. D., Koob, G. F. & McLellan, A. T. Neurobiologic advances from the brain disease 41. Casey, B. J. & Jones, R. M. Neurobiology of the adolescent brain and behavior:
model of addiction. N. Engl. J. Med. 374, 363–371 (2016). implications for substance use disorders. J. Am. Acad. Child Adolesc. Psychiatry 49,
8. Koob, G. F. & Volkow, N. D. Neurobiology of addiction: a neurocircuitry analysis. 1189–1201 (2010).
Lancet Psychiatry 3, 760–773 (2016). 42. Rapuano, K. M. et al. Behavioral and brain signatures of substance use vulnerability in
9. Volkow, N. D. & Boyle, M. Neuroscience of addiction: relevance to prevention and childhood. Dev. Cogn. Neurosci. 46, 100878 (2020).
treatment. Am. J. Psychiatry 175, 729–740 (2018). 43. Deak, J. D. & Johnson, E. C. Genetics of substance use disorders: a review. Psychol. Med.
10. Koob, G. F. Drug addiction: hyperkatifeia/negative reinforcement as a framework for 51, 2189–2200 (2021).
medications development. Pharmacol. Rev. 73, 163–201 (2021). 44. Johnson, E. C. et al. No evidence that schizophrenia candidate genes are more associated
11. Haber, S. N. Corticostriatal circuitry. Dialogues Clin. Neurosci. 18, 7–21 (2016). with schizophrenia than noncandidate genes. Biol. Psychiatry 82, 702–708 (2017).
12. Janak, P. H. & Tye, K. M. From circuits to behaviour in the amygdala. Nature 517, 284–292 45. Border, R. et al. No support for historical candidate gene or candidate gene-by-interaction
(2015). hypotheses for major depression across multiple large samples. Am. J. Psychiatry 176,
13. Volkow, N. D., Fowler, J. S., Wang, G. J., Baler, R. & Telang, F. Imaging dopamine’s role in 376–387 (2019).
drug abuse and addiction. Neuropharmacology 56, 3–8 (2009). 46. Johnson, E. C. et al. A large-scale genome-wide association study meta-analysis
14. Volkow, N. D., Michaelides, M. & Baler, R. The neuroscience of drug reward and addiction. of cannabis use disorder. Lancet Psychiatry 7, 1032–1045 (2020).
Physiol. Rev. 99, 2115–2140 (2019). 47. Liu, M. et al. Association studies of up to 1.2 million individuals yield new insights into the
15. Schultz, W. Dopamine reward prediction error coding. Dialogues Clin. Neurosci. 18, genetic etiology of tobacco and alcohol use. Nat. Genet. 51, 237–244 (2019).
23–32 (2016). 48. Polimanti, R. et al. Leveraging genome-wide data to investigate differences between
16. Keiflin, R. & Janak, P. H. Dopamine prediction errors in reward learning and addiction: opioid use vs. opioid dependence in 41,176 individuals from the Psychiatric Genomics
from theory to neural circuitry. Neuron 88, 247–263 (2015). Consortium. Mol. Psychiatry 25, 1673–1687 (2020).
17. Volkow, N. D. & Morales, M. The brain on drugs: from reward to addiction. Cell 162, 49. Walters, R. K. et al. Transancestral GWAS of alcohol dependence reveals common
712–725 (2015). genetic underpinnings with psychiatric disorders. Nat. Neurosci. 21, 1656–1669 (2018).
18. Filbey, F. M. et al. fMRI study of neural sensitization to hedonic stimuli in long-term, daily 50. Zhou, H. et al. Genome-wide meta-analysis of problematic alcohol use in 435,563
cannabis users. Hum. Brain Mapp. 37, 3431–3443 (2016). individuals yields insights into biology and relationships with other traits. Nat. Neurosci.
19. Perry, J. L. & Carroll, M. E. The role of impulsive behavior in drug abuse. 23, 809–818 (2020).
Psychopharmacology 200, 1–26 (2008). 51. Pasman, J. A. et al. GWAS of lifetime cannabis use reveals new risk loci, genetic overlap
20. Willuhn, I., Burgeno, L. M., Groblewski, P. A. & Phillips, P. E. Excessive cocaine use results with psychiatric traits, and a causal influence of schizophrenia. Nat. Neurosci. 21,
from decreased phasic dopamine signaling in the striatum. Nat. Neurosci. 17, 704–709 1161–1170 (2018).
(2014). 52. Xu, K. et al. Genome-wide association study of smoking trajectory and meta-analysis
21. Koob, G. F. Negative reinforcement in drug addiction: the darkness within. Curr. Opin. of smoking status in 842,000 individuals. Nat. Commun. 11, 5302 (2020).
Neurobiol. 23, 559–563 (2013). 53. Zhou, H. et al. Association of OPRM1 functional coding variant with opioid use disorder:
22. Koob, G. F. & Le Moal, M. Plasticity of reward neurocircuitry and the ‘dark side’ of drug a genome-wide association study. JAMA Psychiatry 77, 1072–1080 (2020).
addiction. Nat. Neurosci. 8, 1442–1444 (2005). 54. Kranzler, H. R. et al. Genome-wide association study of alcohol consumption and use
23. Ahmed, S. H. & Koob, G. F. Transition from moderate to excessive drug intake: change in disorder in 274,424 individuals from multiple populations. Nat. Commun. 10, 1499 (2019).
hedonic set point. Science 282, 298–300 (1998). 55. Gelernter, J. et al. Genome-wide association study of maximum habitual alcohol
24. Martz, M. E. et al. Association of marijuana use with blunted nucleus accumbens intake in>140,000 U.S. European and African American veterans yields novel risk loci.
response to reward anticipation. JAMA Psychiatry 73, 838–844 (2016). Biol. Psychiatry 86, 365–376 (2019).
25. Balodis, I. M. & Potenza, M. N. Anticipatory reward processing in addicted 56. Brazel, D. M. et al. Exome Chip meta-analysis fine maps causal variants and elucidates the
populations: a focus on the monetary incentive delay task. Biol. Psychiatry 77, 434–444 genetic architecture of rare coding variants in smoking and alcohol use. Biol. Psychiatry
(2015). 85, 946–955 (2019).
26. Zorrilla, E. P., Logrip, M. L. & Koob, G. F. Corticotropin releasing factor: a key role in the 57. Jang, S. K. et al. Rare genetic variants explain missing heritability in smoking. Nat. Hum.
neurobiology of addiction. Front. Neuroendocrinol. 35, 234–244 (2014). Behav. https://doi.org/10.1038/s41562-022-01408-5 (2022).
27. Naqvi, N. H., Gaznick, N., Tranel, D. & Bechara, A. The insula: a critical neural substrate 58. Sullivan, P. F. & Geschwind, D. H. Defining the genetic, genomic, cellular, and diagnostic
for craving and drug seeking under conflict and risk. Ann. N. Y. Acad. Sci. 1316, 53–70 architectures of psychiatric disorders. Cell 177, 162–183 (2019).
(2014). 59. Yang, J. et al. Genetic variance estimation with imputed variants finds negligible missing
28. McClure, S. M. & Bickel, W. K. A dual-systems perspective on addiction: contributions heritability for human height and body mass index. Nat. Genet. 47, 1114–1120 (2015).
from neuroimaging and cognitive training. Ann. N. Y. Acad. Sci. 1327, 62–78 (2014). 60. Mallard, T. T. et al. Item-level genome-wide association study of the alcohol use disorders
29. Heather, N. The concept of akrasia as the foundation for a dual systems theory of identification test in three population-based cohorts. Am. J. Psychiatry 179, 58–70 (2022).
addiction. Behav. Brain Res. 390, 112666 (2020). 61. Sanchez-Roige, S., Palmer, A. A. & Clarke, T. K. Recent efforts to dissect the genetic basis
30. Cadet, J. L. & Bisagno, V. Neuropsychological consequences of chronic drug use: of alcohol use and abuse. Biol. Psychiatry 87, 609–618 (2020).
relevance to treatment approaches. Front. Psychiatry 6, 189 (2015). 62. Marees, A. T. et al. Genetic correlates of socio-economic status influence the pattern
31. Brockett, A. T., Pribut, H. J., Vazquez, D. & Roesch, M. R. The impact of drugs of shared heritability across mental health traits. Nat. Hum. Behav. 5, 1065–1073 (2021).
of abuse on executive function: characterizing long-term changes in neural 63. Hatoum, A. S. et al. The addiction risk factor: a unitary genetic vulnerability
correlates following chronic drug exposure and withdrawal in rats. Learn. Mem. 25, characterizes substance use disorders and their associations with common correlates.
461–473 (2018). Neuropsychopharmacology 47, 1739–1745 (2022).
32. Gerdeman, G. L., Partridge, J. G., Lupica, C. R. & Lovinger, D. M. It could be habit 64. Karlsson Linner, R. et al. Multivariate analysis of 1.5 million people identifies genetic
forming: drugs of abuse and striatal synaptic plasticity. Trends Neurosci. 26, 184–192 associations with traits related to self-regulation and addiction. Nat. Neurosci. 24,
(2003). 1367–1376 (2021).
33. Grant, B. F. et al. Epidemiology of DSM-5 drug use disorder: results from the national 65. Liu, X., Zhong, P., Vickstrom, C., Li, Y. & Liu, Q. S. PDE4 inhibition restores the balance
epidemiologic survey on alcohol and related conditions-III. JAMA Psychiatry 73, 39–47 between excitation and inhibition in VTA dopamine neurons disrupted by repeated in
(2016). vivo cocaine exposure. Neuropsychopharmacology 42, 1991–1999 (2017).
34. Sinha, R. New findings on biological factors predicting addiction relapse vulnerability. 66. Tibbo, A. J. & Baillie, G. S. Phosphodiesterase 4B: master regulator of brain signaling.
Curr. Psychiatry Rep. 13, 398–405 (2011). Cells 9, 1254 (2020).
35. Stinson, F. S. et al. Comorbidity between DSM-IV alcohol and specific drug use disorders 67. Kuroiwa, M. et al. Phosphodiesterase 4 inhibition enhances the dopamine D1 receptor/
in the United States: results from the National Epidemiologic Survey on Alcohol and PKA/DARPP-32 signaling cascade in frontal cortex. Psychopharmacology 219, 1065–1079
Related Conditions. Drug Alcohol Depend. 80, 105–116 (2005). (2012).

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 54


Perspective

68. Arends, R. M. et al. Associations between the CADM2 gene, substance use, risky sexual 101. Luby, J. L. et al. Developmental trajectories of the orbitofrontal cortex and anhedonia
behavior, and self-control: A phenome-wide association study. Addict. Biol. 26, e13015 in middle childhood and risk for substance use in adolescence in a longitudinal sample
(2021). of depressed and healthy preschoolers. Am. J. Psychiatry 175, 1010–1021 (2018).
69. Morris, J. et al. Genetic variation in CADM2 as a link between psychological traits and 102. Cheetham, A. et al. Orbitofrontal cortex volume and effortful control as prospective
obesity. Sci. Rep. 9, 7339 (2019). risk factors for substance use disorder in adolescence. Eur. Addict. Res. 23, 37–44 (2017).
70. Sanchez-Roige, S. et al. CADM2 is implicated in impulsive personality traits by genome- 103. Wade, N. E. et al. Orbitofrontal cortex volume prospectively predicts cannabis and other
and phenome-wide association studies in humans, with further support from studies substance use onset in adolescents. J. Psychopharmacol. 33, 1124–1131 (2019).
of Cadm2 mutant mice. medRxiv https://doi.org/10.1101/2022.01.29.22270095v3 (2022). 104. Taffe, M. A. et al. Long-lasting reduction in hippocampal neurogenesis by alcohol
71. Gerring, Z. F., Vargas, A. M., Gamazon, E. R. & Derks, E. M. An integrative systems-based consumption in adolescent nonhuman primates. Proc. Natl Acad. Sci. USA 107,
analysis of substance use: eQTL-informed gene-based tests, gene networks, and 11104–11109 (2010).
biological mechanisms. Am. J. Med. Genet. B Neuropsychiatr. Genet. 186, 162–172 (2021). 105. Shnitko, T. A., Liu, Z., Wang, X., Grant, K. A. & Kroenke, C. D. Chronic alcohol drinking
72. Kapoor, M. et al. Multi-omics integration analysis identifies novel genes for alcoholism slows brain development in adolescent and young adult nonhuman primates. eNeuro 6,
with potential overlap with neurodegenerative diseases. Nat. Commun. 12, 5071 (2021). ENEURO.0044-19.2019 (2019).
73. Baranger, D. A. A. et al. Convergent evidence for predispositional effects of brain gray 106. Kibaly, C., Xu, C., Cahill, C. M., Evans, C. J. & Law, P. Y. Non-nociceptive roles of opioids
matter volume on alcohol consumption. Biol. Psychiatry 87, 645–655 (2020). in the CNS: opioids’ effects on neurogenesis, learning, memory and affect. Nat. Rev.
74. Marees, A. T. et al. Post-GWAS analysis of six substance use traits improves the Neurosci. 20, 5–18 (2019).
identification and functional interpretation of genetic risk loci. Drug Alcohol Depend. 107. Coleman, L. G. Jr., He, J., Lee, J., Styner, M. & Crews, F. T. Adolescent binge drinking alters
206, 107703 (2020). adult brain neurotransmitter gene expression, behavior, brain regional volumes, and
75. Lin, H. et al. Prefrontal cortex eQTLs/mQTLs enriched in genetic variants associated with neurochemistry in mice. Alcohol. Clin. Exp. Res. 35, 671–688 (2011).
alcohol use disorder and other diseases. Epigenomics 12, 789–800 (2020). 108. Douet, V., Chang, L., Cloak, C. & Ernst, T. Genetic influences on brain developmental
76. Sey, N. Y. A. et al. Chromatin architecture in addiction circuitry identifies risk genes and trajectories on neuroimaging studies: from infancy to young adulthood. Brain Imaging
potential biological mechanisms underlying cigarette smoking and alcohol use traits. Behav. 8, 234–250 (2014).
Mol. Psychiatry 27, 3085–3094 (2022). 109. Brouwer, R. M. et al. The speed of development of adolescent brain age depends on sex
77. Mignogna, K. M., Bacanu, S. A., Riley, B. P., Wolen, A. R. & Miles, M. F. Cross-species and is genetically determined. Cereb. Cortex 31, 1296–1306 (2021).
alcohol dependence-associated gene networks: Co-analysis of mouse brain gene 110. Dager, A. D. et al. Shared genetic factors influence amygdala volumes and risk for
expression and human genome-wide association data. PLoS One 14, e0202063 (2019). alcoholism. Neuropsychopharmacology 40, 412–420 (2015).
78. Huggett, S. B., Bubier, J. A., Chesler, E. J. & Palmer, R. H. C. Do gene expression findings 111. Henderson, K. E. et al. Cortical thickness in adolescents with a family history of alcohol
from mouse models of cocaine use recapitulate human cocaine use disorder in reward use disorder. Alcohol. Clin. Exp. Res. 42, 89–99 (2018).
circuitry? Genes Brain Behav. 20, e12689 (2021). 112. Baranger, D. A. A. & Bogdan, R. Editorial: causal, predispositional, or correlate? Group
79. Palmer, R. H. C. et al. Multi-omic and multi-species meta-analyses of nicotine differences in cognitive control-related brain function in cannabis-using youth raise new
consumption. Transl. Psychiatry 11, 98 (2021). questions. J. Am. Acad. Child Adolesc. Psychiatry 58, 665–667 (2019).
80. Kapoor, M. et al. Analysis of whole genome-transcriptomic organization in brain to 113. Gage, S. H., Munafo, M. R., MacLeod, J., Hickman, M. & Smith, G. D. Cannabis and
identify genes associated with alcoholism. Transl. Psychiatry 9, 89 (2019). psychosis. Lancet Psychiatry 2, 380 (2015).
81. Huggett, S. B. & Stallings, M. C. Genetic architecture and molecular neuropathology 114. Hatoum, A. S. et al. Polygenic risk scores for alcohol involvement relate to brain structure
of human cocaine addiction. J. Neurosci. 40, 5300–5313 (2020). in substance-naive children: Results from the ABCD study. Genes Brain Behav. 20, e12756
82. Gottesman, I. I. & Gould, T. D. The endophenotype concept in psychiatry: etymology (2021).
and strategic intentions. Am. J. Psychiatry 160, 636–645 (2003). 115. Munafo, M. R., Higgins, J. P. T. & Smith, G. D. Triangulating evidence through the inclusion
83. Jansen, A. G., Mous, S. E., White, T., Posthuma, D. & Polderman, T. J. What twin studies of genetically informed designs. Cold Spring Harb. Perspect. Med. 11, a040659 (2021).
tell us about the heritability of brain development, morphology, and function: a review. 116. O’Connor, L. J. & Price, A. L. Distinguishing genetic correlation from causation across 52
Neuropsychol. Rev. 25, 27–46 (2015). diseases and complex traits. Nat. Genet. 50, 1728–1734 (2018).
84. Grasby, K. L. et al. The genetic architecture of the human cerebral cortex. Science 367, 117. Gillespie, N. A. & Kendler, K. S. Use of genetically informed methods to clarify the nature
eaay6690 (2020). of the association between cannabis use and risk for schizophrenia. JAMA Psychiatry 78,
85. Hofer, E. et al. Genetic correlations and genome-wide associations of cortical structure in 467–468 (2021).
general population samples of 22,824 adults. Nat. Commun. 11, 4796 (2020). 118. Elam, J. S. et al. The human connectome project: a retrospective. Neuroimage 244,
86. Elliott, L. T. et al. Genome-wide association studies of brain imaging phenotypes 118543 (2021).
in UK Biobank. Nature 562, 210–216 (2018). 119. Pagliaccio, D. et al. Shared predisposition in the association between cannabis use and
87. Smith, S. M. et al. An expanded set of genome-wide association studies of brain imaging subcortical brain structure. JAMA Psychiatry 72, 994–1001 (2015).
phenotypes in UK Biobank. Nat. Neurosci. 24, 737–745 (2021). 120. Iacono, W. G. & McGue, M. Minnesota twin family study. Twin Res. 5, 482–487 (2002).
88. Bulik-Sullivan, B. et al. An atlas of genetic correlations across human diseases and traits. 121. Harper, J. et al. Orbitofrontal cortex thickness and substance use disorders in emerging
Nat. Genet. 47, 1236–1241 (2015). adulthood: causal inferences from a co-twin control/discordant twin study. Addiction
89. Hatoum, A. S., Johnson, E. C., Agrawal, A. & Bogdan, R. Brain structure and problematic 116, 2548–2558 (2021).
alcohol use: a test of plausible causation using latent causal variable analysis. 122. Harper, J. et al. The effects of alcohol and cannabis use on the cortical thickness of
Brain Imaging Behav. 15, 2741–2745 (2021). cognitive control and salience brain networks in emerging adulthood: a co-twin control
90. Rabinowitz, J. A. et al. Shared genetic etiology between cortical brain morphology study. Biol. Psychiatry 89, 1012–1022 (2021).
and tobacco, alcohol, and cannabis use. Cereb. Cortex 32, 796–807 (2022). 123. Hill, S. Y. Trajectories of alcohol use and electrophysiological and morphological indices
91. Marek, S. et al. Reproducible brain-wide association studies require thousands of brain development: distinguishing causes from consequences. Ann. N. Y. Acad. Sci.
of individuals. Nature 603, 654–660 (2022). 1021, 245–259 (2004).
92. Guimaraes, J. P. O. F. T., Sprooten, E., Beckmann, C. F., Franke, B. & Bralten, J. Shared 124. Rangaswamy, M. & Porjesz, B. Uncovering genes for cognitive (dys)function and
genetic influences on resting-state functional networks of the brain. Hum. Brain Mapp. predisposition for alcoholism spectrum disorders: a review of human brain oscillations
43, 1787–1803 (2022). as effective endophenotypes. Brain Res. 1235, 153–171 (2008).
93. Elliott, M. L. et al. What is the test-retest reliability of common task-functional MRI 125. Weiland, B. J. et al. Substance abuse risk in emerging adults associated with smaller
measures? New empirical evidence and a meta-analysis. Psychol. Sci. 31, 792–806 frontal gray matter volumes and higher externalizing behaviors. Drug Alcohol Depend.
(2020). 137, 68–75 (2014).
94. Gordon, E. M. et al. Precision functional mapping of individual human brains. Neuron 95, 126. Hill, S. Y., Lichenstein, S. D., Wang, S. & O’Brien, J. Volumetric differences in cerebellar
791–807.e7 (2017). lobes in individuals from multiplex alcohol dependence families and controls:
95. Hayes, A., Herlinger, K., Paterson, L. & Lingford-Hughes, A. The neurobiology of their relationship to externalizing and internalizing disorders and working memory.
substance use and addiction: evidence from neuroimaging and relevance to treatment. Cerebellum 15, 744–754 (2016).
BJPsych Adv. 26, 367–378 (2020). 127. Cservenka, A., Gillespie, A. J., Michael, P. G. & Nagel, B. J. Family history density of
96. Brown, S. A. et al. The National Consortium on Alcohol and NeuroDevelopment in alcoholism relates to left nucleus accumbens volume in adolescent girls. J. Stud. Alcohol
Adolescence (NCANDA): A multisite study of adolescent development and substance Drugs 76, 47–56 (2015).
use. J. Stud. Alcohol Drugs 76, 895–908 (2015). 128. Hill, S. Y. et al. Right amygdala volume in adolescent and young adult offspring
97. Pfefferbaum, A. et al. Altered brain developmental trajectories in adolescents after from families at high risk for developing alcoholism. Biol. Psychiatry 49, 894–905
initiating drinking. Am. J. Psychiatry 175, 370–380 (2018). (2001).
98. Zhao, Q. et al. Association of heavy drinking with deviant fiber tract development 129. Acheson, A. et al. Increased forebrain activations in youths with family histories of
in frontal brain systems in adolescents. JAMA Psychiatry 78, 407–415 (2021). alcohol and other substance use disorders performing a Go/NoGo task. Alcohol. Clin.
99. Mascarell Maricic, L. et al. The IMAGEN study: a decade of imaging genetics Exp. Res. 38, 2944–2951 (2014).
in adolescents. Mol. Psychiatry 25, 2648–2671 (2020). 130. Cservenka, A., Herting, M. M. & Nagel, B. J. Atypical frontal lobe activity during verbal
100. Albaugh, M. D. et al. Association of cannabis use during adolescence with working memory in youth with a family history of alcoholism. Drug Alcohol Depend. 123,
neurodevelopment. JAMA Psychiatry 78, 1–11 (2021). 98–104 (2012).

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 55


Perspective

131. Cservenka, A. & Nagel, B. J. Risky decision-making: an FMRI study of youth at high risk for 161. Nagel, M., Watanabe, K., Stringer, S., Posthuma, D. & van der Sluis, S. Item-level analyses
alcoholism. Alcohol. Clin. Exp. Res. 36, 604–615 (2012). reveal genetic heterogeneity in neuroticism. Nat. Commun. 9, 905 (2018).
132. Yau, W. Y. et al. Nucleus accumbens response to incentive stimuli anticipation in 162. Sanchez-Roige, S. et al. Genome-wide association studies of impulsive personality traits
children of alcoholics: relationships with precursive behavioral risk and lifetime alcohol (BIS-11 and UPPS-P) and drug experimentation in up to 22,861 adult research participants
use. J. Neurosci. 32, 2544–2551 (2012). identify loci in the CACNA1I and CADM2 genes. J. Neurosci. 39, 2562–2572 (2019).
133. Heitzeg, M. M., Nigg, J. T., Yau, W. Y., Zucker, R. A. & Zubieta, J. K. Striatal dysfunction 163. Zorrilla, E. P. & Koob, G. F. Impulsivity derived from the dark side: neurocircuits that
marks preexisting risk and medial prefrontal dysfunction is related to problem drinking in contribute to negative urgency. Front. Behav. Neurosci. 13, 136 (2019).
children of alcoholics. Biol. Psychiatry 68, 287–295 (2010). 164. Koob, G. F. Anhedonia, hyperkatifeia, and negative reinforcement in substance use
134. Zucker, R. A., Heitzeg, M. M. & Nigg, J. T. Parsing the undercontrol/disinhibition pathway disorders. Curr. Top. Behav. Neurosci. 58, 147–165 (2022).
to substance use disorders: a multilevel developmental problem. Child Dev. Perspect. 5, 165. Grotzinger, A. D. et al. Genomic structural equation modelling provides insights into
248–255 (2011). the multivariate genetic architecture of complex traits. Nat. Hum. Behav. 3, 513–525
135. Kong, A. et al. The nature of nurture: Effects of parental genotypes. Science 359, (2019).
424–428 (2018). 166. Edenberg, H. J. & McClintick, J. N. Alcohol dehydrogenases, aldehyde dehydrogenases,
136. Molina, B. S., Donovan, J. E. & Belendiuk, K. A. Familial loading for alcoholism and and alcohol use disorders: a critical review. Alcohol. Clin. Exp. Res. 42, 2281–2297 (2018).
offspring behavior: mediating and moderating influences. Alcohol. Clin. Exp. Res. 34, 167. Andrews, S. J., Fulton-Howard, B. & Goate, A. Interpretation of risk loci from genome-
1972–1984 (2010). wide association studies of Alzheimer’s disease. Lancet Neurol. 19, 326–335 (2020).
137. Balbona, J. V., Kim, Y. & Keller, M. C. Estimation of parental effects using polygenic scores. 168. Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological
Behav. Genet. 51, 264–278 (2021). insights from 108 schizophrenia-associated genetic loci. Nature 511, 421–427 (2014).
138. Maes, H. H. et al. Cross-cultural comparison of genetic and cultural transmission of 169. Okbay, A. et al. Genome-wide association study identifies 74 loci associated with
smoking initiation using an extended twin kinship model. Twin Res. Hum. Genet. 21, educational attainment. Nature 533, 539–542 (2016).
179–190 (2018). 170. Janes, A. C. et al. Association between CHRNA5 genetic variation at rs16969968 and
139. Volkow, N. D. et al. The conception of the ABCD study: From substance use to a broad brain reactivity to smoking images in nicotine dependent women. Drug Alcohol Depend.
NIH collaboration. Dev. Cogn. Neurosci. 32, 4–7 (2018). 120, 7–13 (2012).
140. Wray, N. R. et al. From basic science to clinical application of polygenic risk scores: 171. Hong, L. E. et al. A genetically modulated, intrinsic cingulate circuit supports human
a primer. JAMA Psychiatry 78, 101–109 (2021). nicotine addiction. Proc. Natl Acad. Sci. USA 107, 13509–13514 (2010).
141. Lisdahl, K. M. et al. Substance use patterns in 9-10 year olds: Baseline findings from 172. Chaarani, B. et al. Low smoking exposure, the adolescent brain, and the modulating role
the adolescent brain cognitive development (ABCD) study. Drug Alcohol Depend. 227, of CHRNA5 polymorphisms. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 4, 672–679
108946 (2021). (2019).
142. Paul, S. E. et al. Associations between prenatal cannabis exposure and childhood 173. Hartwell, E. E. et al. Systematic review and meta-analysis of the moderating effect
outcomes: results from the ABCD study. JAMA Psychiatry 78, 64–76 (2021). of rs1799971 in OPRM1, the mu-opioid receptor gene, on response to naltrexone
143. Baranger, D. A. A. et al. Association of mental health burden with prenatal cannabis treatment of alcohol use disorder. Addiction 115, 1426–1437 (2020).
exposure from childhood to early adolescence: longitudinal findings from the 174. Biernacka, J. M. et al. Genetic contributions to alcohol use disorder treatment outcomes:
Adolescent Brain Cognitive Development (ABCD) study. JAMA Pediatrics https:// a genome-wide pharmacogenomics study. Neuropsychopharmacology 46, 2132–2139
doi.org/10.1001/jamapediatrics.2022.3191 (2022). (2021).
144. Roffman, J. L. et al. Association of adverse prenatal exposure burden with child 175. Deak, J. D. et al. Million Veteran Program. Genome-wide investigation of maximum
psychopathology in the Adolescent Brain Cognitive Development (ABCD) study. habitual alcohol intake in US veterans in relation to alcohol consumption traits and
PLoS One 16, e0250235 (2021). alcohol use disorder. JAMA Netw. Open 5, e2238880 (2022).
145. El Marroun, H. et al. Prenatal tobacco exposure and brain morphology: a prospective 176. Sherva, R. et al. Association of a single nucleotide polymorphism in neuronal
study in young children. Neuropsychopharmacology 39, 792–800 (2014). acetylcholine receptor subunit alpha 5 (CHRNA5) with smoking status and with
146. Ross, E. J., Graham, D. L., Money, K. M. & Stanwood, G. D. Developmental ‘pleasurable buzz’ during early experimentation with smoking. Addiction 103, 1544–1552
consequences of fetal exposure to drugs: what we know and what we still must learn. (2008).
Neuropsychopharmacology 40, 61–87 (2015). 177. McCarthy, D. M., Pedersen, S. L., Lobos, E. A., Todd, R. D. & Wall, T. L. ADH1B*3 and
147. D’Onofrio, B. M., Sjolander, A., Lahey, B. B., Lichtenstein, P. & Oberg, A. S. Accounting response to alcohol in African-Americans. Alcohol. Clin. Exp. Res. 34, 1274–1281 (2010).
for confounding in observational studies. Annu. Rev. Clin. Psychol. 16, 25–48 (2020). 178. Duranceaux, N. C. et al. Associations of variations in alcohol dehydrogenase genes with
148. Quinn, P. D. et al. Association between maternal smoking during pregnancy and severe the level of response to alcohol in non-Asians. Alcohol. Clin. Exp. Res. 30, 1470–1478
mental illness in offspring. JAMA Psychiatry 74, 589–596 (2017). (2006).
149. Micalizzi, L. et al. A sibling-comparison study of smoking during pregnancy and risk for 179. Jensen, K. P. et al. A CHRNA5 smoking risk variant decreases the aversive effects
reading-related problems. Neurotoxicol Teratol. 84, 106961 (2021). of nicotine in humans. Neuropsychopharmacology 40, 2813–2821 (2015).
150. Donald, K. A. et al. Neuroimaging effects of prenatal alcohol exposure on the developing 180. Kreek, M. J., Nielsen, D. A., Butelman, E. R. & LaForge, K. S. Genetic influences on
human brain: a magnetic resonance imaging review. Acta Neuropsychiatr. 27, 251–269 impulsivity, risk taking, stress responsivity and vulnerability to drug abuse and addiction.
(2015). Nat. Neurosci. 8, 1450–1457 (2005).
151. Easey, K. E. et al. Characterization of alcohol polygenic risk scores in the context of 181. Wright, S. N. & Little, A. R. NIDA vision for big data science to understand the biological
mental health outcomes: Within-individual and intergenerational analyses in the Avon underpinnings of substance use disorders. Neuropsychopharmacology 46, 262 (2021).
Longitudinal Study of Parents and Children. Drug Alcohol Depend. 221, 108654 (2021). 182. Veligati, S. et al. Changes in alcohol and cigarette consumption in response to medical
152. Vrieze, S. I., McGue, M., Miller, M. B., Hicks, B. M. & Iacono, W. G. Three mutually and recreational cannabis legalization: Evidence from U.S. state tax receipt data.
informative ways to understand the genetic relationships among behavioral disinhibition, Int. J. Drug Policy 75, 102585 (2020).
alcohol use, drug use, nicotine use/dependence, and their co-occurrence: twin 183. Kendler, K. S. et al. Childhood sexual abuse and adult psychiatric and substance
biometry, GCTA, and genome-wide scoring. Behav. Genet. 43, 97–107 (2013). use disorders in women: an epidemiological and cotwin control analysis. Arch. Gen.
153. Dick, D. M., Adkins, A. E. & Kuo, S. I. Genetic influences on adolescent behavior. Psychiatry 57, 953–959 (2000).
Neurosci. Biobehav. Rev. 70, 198–205 (2016). 184. Sartor, C. E., Agrawal, A., McCutcheon, V. V., Duncan, A. E. & Lynskey, M. T. Disentangling
154. Kendler, K. S., Prescott, C. A., Myers, J. & Neale, M. C. The structure of genetic and the complex association between childhood sexual abuse and alcohol-related problems:
environmental risk factors for common psychiatric and substance use disorders in men a review of methodological issues and approaches. J. Stud. Alcohol Drugs 69, 718–727
and women. Arch. Gen. Psychiatry 60, 929–937 (2003). (2008).
155. Iacono, W. G., Malone, S. M. & McGue, M. Behavioral disinhibition and the development 185. Tomasi, D. & Volkow, N. D. Associations of family income with cognition and brain
of early-onset addiction: common and specific influences. Annu. Rev. Clin. Psychol. 4, structure in USA children: prevention implications. Mol. Psychiatry 26, 6619–6629 (2021).
325–348 (2008). 186. Yehuda, R. & Lehrner, A. Intergenerational transmission of trauma effects: putative role
156. McGue, M., Irons, D. & Iacono, W. G. The adolescent origins of substance use disorders: of epigenetic mechanisms. World Psychiatry 17, 243–257 (2018).
a behavioral genetic perspective. Nebr Symp. Motiv. 61, 31–50 (2014). 187. Meyers, J. et al. Early sexual trauma exposure and neural response inhibition
157. Hamilton, P. J. & Nestler, E. J. Epigenetics and addiction. Curr. Opin. Neurobiol. 59, in adolescence and young adults: trajectories of frontal theta oscillations during
128–136 (2019). a Go/No-Go Task. J. Am. Acad. Child Adolesc. Psychiatry 58, 242–255.e2 (2019).
158. Pang, R. D., Farrahi, L., Glazier, S., Sussman, S. & Leventhal, A. M. Depressive symptoms, 188. Meyers, J. L. et al. Psychosocial moderation of polygenic risk for cannabis involvement:
negative urgency and substance use initiation in adolescents. Drug Alcohol Depend. 144, the role of trauma exposure and frequency of religious service attendance. Transl.
225–230 (2014). Psychiatry 9, 269 (2019).
159. Levey, D. F. et al. Bi-ancestral depression GWAS in the Million Veteran Program and meta- 189. Szutorisz, H. & Hurd, Y. L. Feeding the developing brain: the persistent epigenetic effects
analysis in >1.2 million individuals highlight new therapeutic directions. Nat. Neurosci. 24, of early life malnutrition. Biol. Psychiatry 80, 730–732 (2016).
954–963 (2021). 190. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders,
160. Nagel, M. et al. Meta-analysis of genome-wide association studies for neuroticism in Fifth Edition (American Psychiatric Association, 2013).
449,484 individuals identifies novel genetic loci and pathways. Nat. Genet. 50, 920–927 191. McLellan, A. T., Koob, G. F. & Volkow, N. D. Preaddiction-a missing concept for treating
(2018). substance use disorders. JAMA Psychiatry 79, 749–751 (2022).

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 56


Perspective

192. Lane, S. P., Steinley, D. & Sher, K. J. Meta-analysis of DSM alcohol use disorder criteria 226. Liang, Y. et al. Polygenic transcriptome risk scores (PTRS) can improve portability
severities: structural consistency is only ‘skin deep’. Psychol. Med. 46, 1769–1784 (2016). of polygenic risk scores across ancestries. Genome Biol. 23, 23 (2022).
193. Hines, L. A. et al. Overlap of heritable influences between cannabis use disorder, 227. Marquez-Luna, C. et al. Incorporating functional priors improves polygenic prediction
frequency of use and opportunity to use cannabis: trivariate twin modelling and accuracy in UK Biobank and 23andMe data sets. Nat. Commun. 12, 6052 (2021).
implications for genetic design. Psychol. Med. 48, 2786–2793 (2018). 228. Hu, Y. et al. Leveraging functional annotations in genetic risk prediction for human
194. Gillespie, N. A. et al. Longitudinal modeling of genetic and environmental influences complex diseases. PLoS Comput. Biol. 13, e1005589 (2017).
on self-reported availability of psychoactive substances: alcohol, cigarettes, marijuana, 229. Amariuta, T. et al. Improving the trans-ancestry portability of polygenic risk scores by
cocaine and stimulants. Psychol. Med. 37, 947–959 (2007). prioritizing variants in predicted cell-type-specific regulatory elements. Nat. Genet. 52,
195. Agrawal, A., Madden, P. A., Bucholz, K. K., Heath, A. C. & Lynskey, M. T. Initial reactions to 1346–1354 (2020).
tobacco and cannabis smoking: a twin study. Addiction 109, 663–671 (2014). 230. Hall, W., Carter, A. & Forlini, C. Brain disease model of addiction: misplaced priorities?
196. Edwards, A. C. et al. Meta-analysis of genetic influences on initial alcohol sensitivity. Lancet Psychiatry 2, 867 (2015).
Alcohol. Clin. Exp. Res. 42, 2349–2359 (2018). 231. Risch, N. & Merikangas, K. The future of genetic studies of complex human diseases.
197. Grant, J. D. et al. Subjective reactions to cocaine and marijuana are associated with abuse Science 273, 1516–1517 (1996).
and dependence. Addict. Behav. 30, 1574–1586 (2005). 232. Potenza, M. N. et al. Gambling disorder. Nat. Rev. Dis. Prim. 5, 51 (2019).
198. Lyons, M. J. et al. How do genes influence marijuana use? The role of subjective effects. 233. Slutske, W. S. et al. Common genetic vulnerability for pathological gambling and alcohol
Addiction 92, 409–417 (1997). dependence in men. Arch. Gen. Psychiatry 57, 666–673 (2000).
199. Windle, M. A multilevel developmental contextual approach to substance use and 234. Kaye, W. H. et al. Does a shared neurobiology for foods and drugs of abuse contribute to
addiction. Biosocieties 5, 124–136 (2010). extremes of food ingestion in anorexia and bulimia nervosa? Biol. Psychiatry 73, 836–842
200. Lee, M. R. & Sher, K. J. “Maturing out” of binge and problem drinking. Alcohol Res. 39, (2013).
31–42 (2018). 235. Hasin, D. S. et al. DSM-5 criteria for substance use disorders: recommendations and
201. Verges, A. et al. Refining the notion of maturing out: results from the national epidemiologic rationale. Am. J. Psychiatry 170, 834–851 (2013).
survey on alcohol and related conditions. Am. J. Public. Health 103, e67–e73 (2013). 236. Saunders, J. B., Degenhardt, L., Reed, G. M. & Poznyak, V. Alcohol use disorders in ICD-11:
202. Iacono, W. G. et al. The utility of twins in developmental cognitive neuroscience research: past, present, and future. Alcohol. Clin. Exp. Res. 43, 1617–1631 (2019).
How twins strengthen the ABCD research design. Dev. Cogn. Neurosci. 32, 30–42 (2018). 237. Saunders, J. B., Aasland, O. G., Babor, T. F., de la Fuente, J. R. & Grant, M. Development
203. Porjesz, B., Jones, K. & Begleiter, H. The genetics of oscillations in the human brain. of the alcohol use disorders identification test (AUDIT): WHO collaborative project on
Suppl. Clin. Neurophysiol. 57, 441–449 (2004). early detection of persons with harmful alcohol consumption–II. Addiction 88, 791–804
204. Anokhin, A. P. Genetic psychophysiology: advances, problems, and future directions. (1993).
Int. J. Psychophysiol. 93, 173–197 (2014). 238. Quach, B. C. et al. Expanding the genetic architecture of nicotine dependence and its
205. Mackey, S. et al. Genetic imaging consortium for addiction medicine: from neuroimaging shared genetics with multiple traits. Nat. Commun. 11, 5562 (2020).
to genes. Prog. Brain Res. 224, 203–223 (2016). 239. Gaddis, N. et al. Multi-trait genome-wide association study of opioid addiction: OPRM1
206. Van Essen, D. C. et al. The WU-Minn human connectome project: an overview. and beyond. Sci. Rep. 12, 16873 (2022).
Neuroimage 80, 62–79 (2013). 240. Kember, R. L. et al. Cross-ancestry meta-analysis of opioid use disorder uncovers novel
207. Kwako, L. E., Momenan, R., Litten, R. Z., Koob, G. F. & Goldman, D. Addictions loci with predominant effects in brain regions associated with addiction. Nat. Neurosci.
neuroclinical assessment: a neuroscience-based framework for addictive disorders. https://doi.org/10.1038/s41593-022-01160-z (2022).
Biol. Psychiatry 80, 179–189 (2016). 241. Deak, J. D. et al. Genome-wide association study in individuals of European and
208. Vanyukov, M. M. et al. Liability to substance use disorders: 1. Common mechanisms and African ancestry and multi-trait analysis of opioid use disorder identifies 19 independent
manifestations. Neurosci. Biobehav. Rev. 27, 507–515 (2003). genome-wide significant risk loci. Mol. Psychiatry https://doi.org/10.1038/
209. Vanyukov, M. M. et al. Liability to substance use disorders: 2. A measurement approach. s41380-022-01709-1 (2022).
Neurosci. Biobehav. Rev. 27, 517–526 (2003). 242. Cabana-Dominguez, J., Shivalikanjli, A., Fernandez-Castillo, N. & Cormand, B. Genome-
210. Latzman, R. D. & DeYoung, C. G., Hitop Neurobiological Foundations Workgroup. wide association meta-analysis of cocaine dependence: Shared genetics with comorbid
Using empirically-derived dimensional phenotypes to accelerate clinical conditions. Prog. Neuropsychopharmacol. Biol. Psychiatry 94, 109667 (2019).
neuroscience: the Hierarchical Taxonomy of Psychopathology (HiTOP) framework.
Neuropsychopharmacology 45, 1083–1085 (2020). Acknowledgements
211. Boness, C. L., Watts, A. L., Moeller, K. N. & Sher, K. J. The etiologic, theory-based, The authors acknowledge the following funding from the United States National Institutes
ontogenetic hierarchical framework of alcohol use disorder: a translational systematic of Health: R.B. (R01DA54750; R21AA27827, U01DA055367), A.S.H. (T32DA007261,
review of reviews. Psychol. Bull. 147, 1075–1123 (2021). K01AA030083), E.C.J. (K01DA51759), A.A. (K02DA32573, R01DA54750). Funders were not
212. Watanabe, K. et al. A global overview of pleiotropy and genetic architecture in complex involved in the preparation of this manuscript in any way.
traits. Nat. Genet. 51, 1339–1348 (2019).
213. Barr, P. B. et al. Using polygenic scores for identifying individuals at increased risk of
Author contributions
substance use disorders in clinical and population samples. Transl. Psychiatry 10, 196
All authors researched data for the article and wrote the article. A.A., A.S.H. and R.B.
(2020).
contributed substantially to discussion of the content. A.A., E.C.J. and R.B. reviewed and/or
214. Dick, A. S. et al. Meaningful associations in the adolescent brain cognitive development
edited the manuscript before submission.
study. Neuroimage 239, 118262 (2021).
215. Baranger, D. A. A. et al. Borderline personality traits are not correlated with brain
structure in two large samples. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 5, Competing interests
669–677 (2020). The authors declare no competing interests.
216. Bogdan, R., Baranger, D. A. A. & Agrawal, A. Polygenic risk scores in clinical psychology:
bridging genomic risk to individual differences. Annu. Rev. Clin. Psychol. 14, 119–157 (2018). Additional information
217. Frangou, S. et al. Cortical thickness across the lifespan: Data from 17,075 healthy Correspondence should be addressed to Ryan Bogdan or Arpana Agrawal.
individuals aged 3-90 years. Hum. Brain Mapp. 43, 431–451 (2022).
218. Martin, A. R. et al. Clinical use of current polygenic risk scores may exacerbate health Peer review information Nature Reviews Neuroscience thanks the anonymous reviewers for
disparities. Nat. Genet. 51, 584–591 (2019). their contribution to the peer review of this work.
219. Li, D. et al. Genome-wide association study of copy number variations (CNVs) with opioid
dependence. Neuropsychopharmacology 40, 1016–1026 (2015). Reprints and permissions information is available at www.nature.com/reprints.
220. Sulovari, A., Liu, Z., Zhu, Z. & Li, D. Genome-wide meta-analysis of copy number
variations with alcohol dependence. Pharmacogenomics J. 18, 398–405 (2018). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
221. Mukamel, R. E. et al. Protein-coding repeat polymorphisms strongly shape diverse published maps and institutional affiliations.
human phenotypes. Science 373, 1499–1505 (2021).
222. Skene, N. G. et al. Genetic identification of brain cell types underlying schizophrenia. Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
Nat. Genet. 50, 825–833 (2018). article under a publishing agreement with the author(s) or other rightsholder(s); author self-
223. Ruzicka, W. B. et al. Single-cell dissection of schizophrenia reveals neurodevelopmental- archiving of the accepted manuscript version of this article is solely governed by the terms of
synaptic axis and transcriptional resilience. medRxiv https://doi.org/10.1101/ such publishing agreement and applicable law.
2020.11.06.20225342v1 (2020).
224. Nasser, J. et al. Genome-wide enhancer maps link risk variants to disease genes. Nature Related links
593, 238–243 (2021). HEALthy Brain and Cognitive Development (HBCD) study: https://heal.nih.gov/research/
225. Baranger, D. A. A. et al. Multi-omics analyses cannot identify true-positive novel infants-and-children/healthy-brain
associations from underpowered genome-wide association studies of four brain-related
traits. bioRxiv https://doi.org/10.1101/2022.04.13.487655 (2022). © Springer Nature Limited 2022

Nature Reviews Neuroscience | Volume 24 | January 2023 | 40–57 57

You might also like