Sunteți pe pagina 1din 18

Neuron

Review

Neuroscience of Behavioral and Pharmacological


Treatments for Addictions
Marc N. Potenza,1,2,3,* Mehmet Sofuoglu,1,3 Kathleen M. Carroll,1,4 and Bruce J. Rounsaville1,4
1Department of Psychiatry
2Department of Neurobiology and Child Study Center
Yale University School of Medicine, New Haven, CT 06519, USA
3These authors contributed equally to this work
4These authors contributed equally to this work

*Correspondence: marc.potenza@yale.edu
DOI 10.1016/j.neuron.2011.02.009

Although substantial advances have been made in behavioral and pharmacological treatments for addic-
tions, moving treatment development to the next stage may require novel ways of approaching addictions,
particularly ways based on new findings regarding the neurobiological underpinnings of addictions that also
assimilate and incorporate relevant information from earlier approaches. In this review, we first briefly review
theoretical and biological models of addiction and then describe existing behavioral and pharmacologic ther-
apies for the addictions within this framework. We then propose new directions for treatment development
and targets that are informed by recent evidence regarding the heterogeneity of addictions and the neurobi-
ological contributions to these disorders.

Overview ioral engagement representing core elements (Holden, 2001;


Despite intensive research and significant advances, drug addic- Potenza, 2006; Shaffer, 1999). If these are considered the central
tions remain a substantial public health problem. Drug addictions elements of addictions, then behaviors like gambling may be
cost U.S. society hundreds of billions of dollars annually and considered from an addictions perspective. Consistent with
impact not only the addicted individuals, but also their spouses, this notion and existing clinical, preclinical, and neurobiological
children, employers, and others (Uhl and Grow, 2004; Volkow data, pathological gambling is being considered for reclassifica-
et al., 2011). Furthermore, costs may be even higher as nondrug tion with substance use disorders into an addictions category in
disorders (e.g., related to food and gambling) have recently been DSM-V (Holden, 2010).
conceptualized within an addiction framework, with neurobiolog- In addition to similarities across addictive disorders, there are
ical data supporting similarities across substance dependences, also differences relevant to individual addictions that are related
obesity, and pathological gambling (Frascella et al., 2010; Grant to features like the sites of action of the drugs being abused and
et al., 2010b; Kenny, 2011; Potenza, 2008). Given the additional the social acceptability and availability of the behavior or
health burdens of these conditions (e.g., obesity and tobacco substance, and these represent important considerations with
consumption represent two top causes of preventable death respect to the neurobiologies and treatments of addictions. For
[Danaei et al., 2009; Kenny, 2011]), addictions arguably represent example, while compulsivity may cut across addictions, aspects
our nation’s (and the world’s) main health problem. Thus, the of tolerance and withdrawal may differ and reflect specific neuro-
development of improved prevention and treatment strategies adaptations related to individual substances or behaviors (Dalley
is of paramount importance. et al., 2011; Kenny, 2011; Sulzer, 2011). Thus, considering
In order to best prevent and treat addictions, it is important to the mechanisms underlying addictions in general as well as
have a clear understanding of which disorders constitute addic- features unique to individual disorders is important in treatment
tions, and this point has been debated considerably over time. development.
The term addiction, derived from a Latin word meaning ‘‘bound Multiple, non-mutually exclusive models (e.g., incentive
to’’ or ‘‘enslaved by,’’ was initially not linked to substance use salience [Robinson and Berridge, 2001], allostasis [Edwards
(Maddux and Desmond, 2000). However, over the past several and Koob, 2010; Koob and Le Moal, 2001], reward deficiency
hundred years, addiction became associated with excessive [Blum et al., 1996]) have been proposed for addictions. While
alcohol and then drug use such by the 1980s it was largely they each have unique features, they also include common
synonymous with compulsive drug use (O’Brien et al., 2006). features related to the proposed core elements of addiction
However, observations that individuals with gambling problems described above. Across these models, motivational neurocircui-
share clinical, phenomenological, genetic, and other biological try functions to favor drug use (or behavioral engagement) over
similarities with people with drug dependences has prompted other aspects of life (e.g., studying for tests, going to work, or
reconsideration of the core features of addiction, with continued caring for one’s family). Consistently, addiction has been termed
performance of the behavior despite adverse consequences, a condition of motivated behavior going awry (Volkow and Li,
compulsive engagement, or diminished control over the 2004) and neurobiological models of motivational circuitry have
behavior, and an appetitive urge or craving state prior to behav- been proposed for addictions and addiction vulnerability

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 695


Neuron

Review

(Chambers et al., 2003; Everitt and Robbins, 2005; George and necessitates executive control, which may be mediated via
Koob, 2010). In these models, cortico-striato-pallido-thalamo- top-down control of the prefrontal cortex over subcortical
cortical loops form a central feature underlying motivated behav- processes promoting motivations to engage in the addictive
iors (Alexander et al., 1990; Alexander et al., 1986). Other brain behavior (Chambers et al., 2003; Everitt and Robbins, 2005).
regions and circuits contribute importantly to motivated behav- Both positive reinforcement processes (e.g., seeking a drug
iors, with regions like the amygdala providing important affective high) and negative reinforcement processes (e.g., seeking
information, the hippocampus important contextual memory relief from stressful or negative mood states) may be linked to
information, the hypothalamus and septum important homeo- environmental or internal cues in fashions that are behaviorally
static information, and the insula important information related engrained, resistant to change, and linked to powerful motiva-
to interoceptive processing (Chambers et al., 2003; Everitt and tional craving states (Chambers et al., 2007). Thus, therapies
Robbins, 2005; George and Koob, 2010; Naqvi and Bechara, may be needed to target strong learned associations between
2009). Additionally, cingulate cortices provide important contribu- drug use and immediate positive or negative reinforcements.
tions, with the anterior and posterior components contributing to
emotional regulation, cognitive control, and stress responsive- Phases of Treatment
ness (Botvinick et al., 2001; Bush et al., 2000; Sinha, 2008). While The treatment for addictions can be divided into three phases:
often relatively simplistic, such models, particularly when consid- detoxification, initial recovery, and relapse prevention. The first
ered from a systems perspective (i.e., these brain regions function phase, detoxification, has the goal of achieving abstinence that
in circuits rather than in isolation), provide a neurobiological basis is sufficiently sustained to yield a safe reduction in immediate
for developing new treatments for addictions and investigating withdrawal symptoms. The second phase, initial recovery, has
the mechanisms by which effective therapies for addictions work. goals of developing sustained motivation to avoid relapse,
Aspects of the development of addictions can be understood learning strategies for tolerating craving induced by external or
on the basis of both positive and negative reinforcements linked internal cues, and developing new patterns of behavior that
to drug use. Drug experimentation typically begins during adoles- entail replacement of drug-induced reinforcement with alterna-
cence, initially resulting in hedonic experiences that generate tive rewards. The third phase, relapse prevention, takes place
relatively immediate positive reinforcement for use with little or after a period of sustained abstinence and requires subjects to
no negative consequences (Rutherford et al., 2010; Wagner develop long-term strategies that will allow them to replace
and Anthony, 2002). Yet as drug use continues, neuroadapta- past drug behaviors with new, healthy behaviors.
tions occur relating to the development of drug tolerance, result- As discussed above, the drug addiction cycle is maintained
ing in a reduction in the pleasurable sensations achieved from through repeated use and alterations to motivational neurocircui-
a similar initial level of drug use. Although the precise adaptations try, including dopaminergic systems. Given the need to disengage
remain a topic of current investigation, motivational from sustained patterns of use and related neuroadaptations,
neurocircuitry and multiple neurotransmitter systems, particu- detoxification frequently requires pharmacological intervention.
larly dopamine, are implicated (Rutherford et al., 2010; Sulzer, These initial drug treatments may involve choosing a replacement
2011). As this cycle continues, subjects increase the frequency substance that has a similar mode of action on the neurobiological
and amount of drug use to gain the same rewarding experience. substrate, while having a slower and more sustained effect.
For many drugs, increased use also leads to withdrawal symp- Behaviorally speaking, this results in withdrawal symptoms that
toms when drug use is curtailed or cut down. As withdrawal are made less acute but more prolonged and gradual. For
symptoms can at least be temporarily relieved by continued, example, drugs with longer half-lives than herion (e.g., metha-
and escalating, drug use, a vicious cycle is established. Over done) can be used in addicted individuals during detoxification.
time, hedonic motivations for substance use diminish while nega- Successful resolution of detoxification requires sustained
tive reinforcement motivations increase, with drug-taking behav- motivation to tolerate withdrawal symptoms. The second phase,
iors becoming less rewarding and more compulsive or habitual initial recovery, is often aided by external, structural controls
over time. This shift has been proposed to involve a progression (e.g., hospitalization) that limit access to drugs once withdrawal
of involvement of ventral to dorsal cortico-striato-pallido-tha- symptoms have been alleviated. Yet, ultimately, the initial
lamo-cortical circuitry (Brewer and Potenza, 2008; Everitt and recovery phase must teach the addicted individual ways to
Robbins, 2005; Fineberg et al., 2010; Haber and Knutson, sustain motivations to avoid relapse, learn strategies for toler-
2010). From a molecular level, dopamine function, particularly ating and resisting drug cravings induced by external or internal
striatal D2/D3 receptor function, appears relevant to this process cues, and develop new patterns of behavior that entail replace-
and has been implicated across addictive disorders (Kenny, ment of drug-induced reinforcement with alternative rewards.
2011; Steeves et al., 2009; Wang et al., 2009). However, multiple Learning these new behavioral strategies can also be aided by
other neurotransmitter systems contribute and may represent the longer-term administration of medications such as those
better treatment targets, particularly as D2/D3 receptor antago- used in the detoxification process (e.g., drugs that block or
nists have not demonstrated clinical efficacy for addictions. reduce drug rewards, reduce craving by substituting for drug
From a cognitive perspective, attempts to control or eliminate effects) or by the additional augmentation with drugs that help
addictive behaviors are usually motivated by the delayed nega- to reduce mental and physical symptoms not necessarily related
tive consequences of use. The individual’s cognitive recognition to drug use (e.g., independent depression or anxiety disorders).
of these negative consequences may lead to attempts to develop The third phase, relapse prevention, is perhaps the most
changed attitudes and drug-using behaviors. This process difficult to achieve given the long-term brain adaptations

696 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review
Figure 1. The Brain Regions Proposed
to Mediate the Behavioral and
Pharmacological Treatments of Addictions
For simplicity, only a few key brain regions are
included in the figure. See text for details. Abbrevi-
ations: AD, adrenergic receptors; CBT, cognitive
behavioral therapy; CM, contingency manage-
ment; DA, dopamine; DAT, dopamine transporter;
GABA, gamma-aminobutyric acid; Glu, glutamate;
l.cer, locus ceruleus; MI, motivational interviewing;
Nac, nucleus accumbens; nAChR, nicotinic cholin-
ergic receptor; NE, norepinephrine; NET, norepi-
nephrine transporter; VTA, ventral tegmental area.

Behavioral Therapies Strategies


and Targets
Compulsive drug use despite negative
consequences and despite the desire to
quit can be understood as entailing two
processes that are targets for behavioral
therapies: (1) the excessive desire to
use or craving for substances; and (2)
insufficient impulse control associated
resulting from sustained drug abuse. Indeed, relapses often with neurocognitive impairment. In the sections below we briefly
occur and many relapse prevention programs involve review three broad categories of behavioral interventions that
a continued support system (e.g., Alcoholics Anonymous) to have achieved consistent empirical support for substance use
aid in maintaining new behaviors developed during initial problems through randomized controlled trials. These are (1)
recovery. Threats to recovery involve both external and internal brief and motivational models, (2) contingency management
cues that lead to waning motivation, attenuation of external or models, and (3) cognitive behavioral models.
internal controls, and revival of associative learning cues linking Brief Motivational Models
drug use to hedonic experiences and can be triggered by both A surprising revelation of the past 20 years of treatment research
external cues and internal cues. External cues include exposure in the addictions has been the efficacy and durability of brief
to drugs or to people, places, or things associated with drug use. behavioral therapies for many individuals with substance use
Internal cues include hedonic experiences that may be problems (Burke et al., 2003; Miller and Rollnick, 2002). Rela-
enhanced by resumed drug use or dysphoric experiences that tively brief, focused interventions consisting of as little of a single
may be mediated by such factors as stress, interpersonal session have not only been demonstrated to be effective, but in
conflict, or symptoms of comorbid mental disorders such as several studies have also been shown to be as effective as
depression. lengthier, more intensive approaches. The efficacy of brief
At all three phases, social processes can improve executive motivational approaches appears to extend to addictions that
functioning through a variety of mechanisms, including do not involve ingested substances (Burke et al., 2003), including
enhancing motivation, reducing friction and stress, providing pathological gambling and eating disorders (Brewer et al.,
alternative rewards associated with avoiding drug use, and 2008b; Frascella et al., 2010), suggesting that similar neural
providing external constraints. These factors can be conceived mechanisms may underlie therapeutic effects across addictions.
of as enhancing cortically mediated executive control over Although the precise neural mechanisms mediating the effects of
addictive behaviors (Volkow et al., 2011). brief motivational interventions are not known, processes
In the next sections, we will briefly review the major behavioral involving the receipt of health-related information and recom-
and pharmacological treatments for addictions and describe the mendations from a professional may prompt individuals to alter
targets of these treatments. In a simplified description, the neural their decision-making processes to focus on more future-
processes targeted by treating addictions can be characterized oriented goals. Thus, brain motivational circuitry in general and
as ‘‘top-down’’ or ‘‘bottom-up.’’ Top-down interventions attempt specific regions implicated in risk-reward decision making
to change cognitions and behaviors mediated by enhanced (e.g., ventromedial prefrontal cortex), cognitive control (e.g.,
prefrontal cortical function and altered executive control. anterior cingulate cortex), and planning and executive func-
Bottom-up interventions target the subcortical processes, tioning (e.g., dorsolateral prefrontal cortex) in particular may
including the striatal reward pathways, that mediate dysphoric represent important brain regions for consideration (Bechara,
symptoms and learned association pathways that do not require 2003; Bush et al., 2002; Dalley et al., 2011).
or necessarily involve cortical activity. As a broad generalization, Contingency Management Models
current behavioral treatments appear strongest at changing Another major development in the treatment of substance
top-down activity while pharmacological treatments tend to use problems has involved findings regarding the efficacy of
target bottom-up processes (Figure 1). contingency management interventions (Dutra et al., 2008;

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 697


Neuron

Review

Lussier et al., 2006). Based on principles of behavioral pharma- decrease even after CBT treatment concludes, so-called
cology and operant conditioning, contingency management ‘‘sleeper effects’’ (Carroll et al., 1994).
approaches recognize that abused substances are powerful As with other behavioral therapies, the neural mechanisms
reinforcers and are implemented with the idea that immediate underlying CBT remain poorly understood and, in comparison
reinforcement of abstinence (or other behaviors incompatible to brief motivational interventions and contingency manage-
with substance use) can reliably, and comparatively easily, ment, may be particularly complex given the multifaceted nature
interrupt substance use for a large number of individuals. In the of CBT. For example, CBT typically consists of multiple sessions
case of substance dependence, individuals are provided or modules, with each having a specific focus (Carroll et al.,
concrete rewards, often cash, that generally escalate in value 1998; Petry, 2005). Accordingly, different modules may prefer-
and are contingent on submitting drug-free urine specimens entially induce changes in specific neural circuits. For example,
(Higgins et al., 1991). Beyond producing some of the largest modules that teach coping strategies for managing cravings
and most consistent effect sizes in substance abuse treatment may specifically influence or involve brain regions implicated in
(Dutra et al., 2008), these approaches have broad utility cue-induced drug craving (e.g., medial prefrontal and anterior
and can be targeted to improve treatment adherence, including cingulate cortices in cocaine dependence [Childress et al.,
medication compliance that often undercuts the efficacy of 1999; Wexler et al., 2001]) and/or work through altering func-
available pharmacotherapies (Carroll et al., 2004). tional connectivity within brain circuits related to craving.
Contingency management for addictions can be conceptual- Consistent with this notion, an fMRI study investigating cue-
ized within a behavioral neuroeconomic framework (Glimcher induced craving and using instructions based on CBT cognitive
and Rustichini, 2004). Individuals with addictions as compared strategies to focus on long-term consequences of tobacco use
to those without typically place comparably greater values on rather than short-term pleasurable tobacco associations found
immediate rewards, and future rewards are more rapidly deval- that dorsolateral prefrontal cortical regions exerted control
ued, a process termed delay or temporal discounting. This rapid over ventral striatal activation in the regulation of craving (Kober
discounting has been observed across groups of individuals with et al., 2010). These findings are reminiscent of a study of
different addictions, both substance and nonsubstance, in tobacco smokers who were exposed to tobacco cues in an
active and remitted addictions, and with respect to both drugs emotional Stroop task and received a combination of behavioral
and money (Johnson et al., 2007; Johnson et al., 2010; Petry, therapy and nicotine replacement (Janes et al., 2010). Individ-
2001a, 2001b; Ross et al., 2009). From a neurobiological uals who showed greater functional connectivity between
perspective, the selection of small immediate rewards typically prefrontal cortical regions and brain areas involved in craving
activates ‘‘reward’’ regions like the ventral striatum and ventro- and interoceptive processing (anterior cingulate cortex and
medial prefrontal cortex whereas the selection of larger, delayed insula) demonstrated greater success in treatment (Janes
rewards activates more dorsal cortical regions (Kable and et al., 2010).
Glimcher, 2007; McClure et al., 2004). Steep temporal discount- Other aspects of CBT may involve the recruitment and
ing has been associated with poor treatment outcome for strengthening of other circuits. For example, consider the
addictions (Krishnan-Sarin et al., 2007), may be amenable to learning of alternate coping strategies. Training on a visual
treatment (Bickel et al., 2011), and may involve cortical and perception learning task led to strengthened connectivity of
subcortical systems involved in decision making (Bickel and Yi, circuitry involved in spatial attention, and these changes were
2008) (see also Balleine et al. (2007) and related articles in the observed in brain activity during rest (Lewis et al., 2009). Restful
volume). waking brain activity has been termed the default mode network,
Cognitive Behavioral Models and although changes in default mode processing have been
Another set of approaches that has emerged with strong support proposed to underlie both effective behavioral and pharmaco-
from randomized trials includes cognitive behavioral therapies logical treatment of nicotine dependence (Costello et al.,
(CBTs), which seek to help the individual recognize behavioral 2010), the relationship between default mode processing and
patterns and cognitions that maintain substance use and to learning changes in CBT has not been examined.
learn and then implement skills and strategies to change Other CBT modules (for example, those relating to financial
those patterns and interrupt substance use (Dutra et al., 2008; management in pathological gambling) may more closely involve
Irvin et al., 1999; Magill and Ray, 2009; Tolin, 2010). These neurocircuitry implicated in the processing of monetary rewards
approaches are based on principles of operant as well as clas- or financial decision making (Kable and Glimcher, 2007; Knutson
sical conditioning, for example, seeking to heighten the individ- and Greer, 2008). As individuals with addictions typically differ
uals’ awareness of cues previously paired with substance use, from control subjects in the function of such circuitry (Tanabe
reduction of exposure to such cues, and implementation of skills et al., 2007; Wrase et al., 2007), it is tempting to speculate that
to be aware of and tolerate cue-induced craving. CBT effective CBT might ‘‘normalize’’ these circuits and that such
approaches emphasize the development of cognitive strategies normalization would be related to completion of the correspond-
to countervail the strong drives for drugs associated with condi- ing CBT module. CBT-related changes over a longer time period,
tioned cravings, as well as to fortify behavioral controls through including ‘‘sleeper effects,’’ may involve circuitry underlying
learning to employ alternative coping mechanisms or to seek and cognitive function and affective control, as has been observed
value alternative, socially sanctioned rewards that are incompat- with CBT in other disorders like depression and obsessive-
ible with drug abuse. CBT approaches appear to have particu- compulsive disorder (Goldapple et al., 2004; Ritchey et al.,
larly durable effects in that substance use often continues to 2010; Saxena et al., 2009; Siegle et al., 2006).

698 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review

The Future of Behavioral Therapies has been reported to induce white matter integrity changes in
Several novel approaches to achieving recovery from addictions the corona radiata, a tract connecting the anterior cingulate
are receiving empirical support, and in some cases these may cortex to other brain structures (Tang et al., 2010).
complement existing strategies through more efficient targeting Additional behavioral therapeutic advances might be gleaned
of cognitive, emotional, and behavioral domains or deficits, as from considering approaches to nonsubstance addictions. For
well as their neural correlates. Novel cognitive remediation strat- example, imaginal desensitization has shown some efficacy in
egies, aimed at strengthening brain function, may have potential the treatment of pathological gambling (Brewer et al., 2008b;
in addiction treatment. Cognitive remediation strategies involve Grant et al., 2009), and this approach of controlled exposure to
repeated intensive exposure to computerized exercises in- gambling-related cues may help uncouple cues from engage-
tended to strengthen memory, attention, planning, and other ment in addictive behaviors and thus might be anticipated to
aspects of executive functioning. Given its novelty, this influence prefrontal control over motivation (George and Koob,
approach has promise and is consistent with adult neuroplastic- 2010). Participation in 12-step programs (e.g., Alcoholics
ity (Ersche and Sahakian, 2007) and findings in nonaddicted Anonymous) may also induce specific neuroadaptations. Like
populations. For example, cognitive remediation strategies CBT, 12-step programs have multiple components (steps) (Alco-
improve not only neurocognitive functioning in individuals with holics Anonymous, 1986), and these may be differentially linked
schizophrenia, but also their general social and occupational to specific brain circuits. For example, step eight involves
functioning (Bell et al., 2001). Specifically, measures of neuro- a willingness to make amends to those harmed, and performing
cognition (assessing attention, memory, and problem solving) such behaviors may involve changes neurocircuitry implicated in
and measures of social cognition and adjustment were improved social reciprocity and moral decision making (Moll et al., 2005;
over a two-year period (Hogarty et al., 2004). There is preliminary Potenza, 2009a). Although a 12-step program is not a behavioral
evidence that computerized cognitive remediation improves therapy per se, many individuals receiving formal treatment for
cognitive functioning in substance users with neuropsycho- addictions also attend 12-step programs. Thus, considering
logical deficits and also improves treatment engagement the contributions of 12-step participation to treatment outcome
and outcome (Bickel et al., 2011; Grohman et al., 2006; Wexler, and corresponding changes in neurocircuitry is important.
2011). For example, working memory training was found to
reduce impulsive choice measures of temporal discounting Pharmacological Treatments and Targets
in substance abusers (Bickel et al., 2011). The extent to Multiple pharmacological targets have been identified for the
which such changes reflect increased top-down control treatment of addictive disorders. ‘‘Classic’’ approaches tend to
through enhanced prefrontal cortical function requires direct target the drug ‘‘reward’’ system, such as normalization of func-
investigation. tion through agonist approaches and negative reinforcement
Other approaches receiving empirical support in addictions strategies. These approaches are informed by study of neuro-
treatment are mindfulness-based therapies. Based in part on transmitters affected by substances of abuse (Koob and Volkow,
Buddhist tenets and practices, mindfulness-based therapies 2010; Reissner and Kalivas, 2010; Sulzer, 2011), with recent
have been developed to target stress and negative mood states approaches emphasizing the targeting of individual vulnerabil-
in depression and examined in preliminary studies of addictions ities and cognitive function (George and Koob, 2010).
(Brewer et al., 2010). In a pilot study, mindfulness training, Medications Targeting Positive Reinforcement
compared with CBT, demonstrated comparable efficacy on or Drug Reward
measures of retention and abstinence and was more effective Positive reinforcement is defined as any stimulus that increases
in diminishing subjective and biological stress responsiveness the probability of the preceding behavior and typically involves
(self-reported anxiety following personalized stress exposure a hedonic reward. Self-administration is the primary measure
and sympathetic/vagal ratios, respectively) (Brewer et al., for drug reinforcement, and almost all reinforcing drugs induce
2009). These findings suggest that mindfulness-based therapies subjective drug reward or ‘‘liking’’ in humans. The exact function
may be particularly helpful in targeting negative reinforcement of dopamine in addictive behavior continues to be debated
processes, like stress-induced cravings, in addictions, and this (Dalley and Everitt, 2009; Kenny, 2011; Lajtha and Sershen,
may be of particular relevance to relapse prevention as stress- 2010; Schultz, 2010, 2011). According to Robinson and Berridge,
induced cravings measures predict relapse (Sinha et al., 2006). dopamine mainly mediates incentive-salience or ‘‘wanting’’ while
Given the neurobiology of stress responsiveness and increased drug pleasure or ‘‘liking’’ is mediated by other neurotransmitters
cortico-striato-limbic activations to stress in addicted individuals including endogenous opioids, gamma-aminobutyric acid
(Koob and Zorrilla, 2010; Sinha, 2008), it is tempting to speculate (GABA), and endocannabinoids (Berridge et al., 2009; Horder
that mindfulness-based therapies may normalize stress-related et al., 2010; Robinson and Berridge, 1993). The hypothesis is sup-
responses in addicted individuals. Mindfulness-based therapies ported by a human PET imaging study in which dopamine release
may be particularly applicable to women, as cocaine-dependent by amphetamine was correlated with drug ‘‘wanting’’ but not with
women as compared to cocaine dependent men demonstrate mood elevation (Leyton et al., 2002). In addition, acute phenylal-
relatively increased cortico-striato-limbic activations to stress anine-tyrosine depletion, which reduces the precursor levels for
cues (Potenza et al., 2007). Changes related to mindfulness- dopamine, resulted in attenuated cue and cocaine-induced
based therapies may involve white matter changes in brain drug craving but not euphoria or self-administration of cocaine
regions implicated in emotional regulation and cognitive control (Leyton et al., 2005). Further, dopamine receptor antagonists
as meditation, a component of mindfulness-based therapies, do not consistently block cocaine-induced ‘‘high’’ in humans

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 699


Neuron

Review

Table 1. Pharmacotherapies Used for the Treatment of Addictive Disorders


Medication Delivery System Mechanism of action Type of Addiction Efficacy
Methadone Oral solution or tablet m opioid agonist Opioid Effective in retaining patients
in treatment and reducing
heroin use (Mattick et al.,
2003)
Buprenorphine Sublingual tablet alone or Partial m opioid agonist and k Opioid Effective in retaining patients
with naltrexone opioid antagonist and reducing heroin use
(Mattick et al., 2004)
Naltrexone Oral tablet, Opioid antagonist Opioid Better opioid use outcomes
Extended-release injectable in high retention groups
suspension (Johansson et al., 2006)
Alcohol Significantly reduces
relapses but not to drinking
(Srisurapanont and
Jarusuraisin, 2005)
Disulfiram Oral tablet Increases acetaldehyde by Alcohol May reduce drinking in
inhibition of aldehyde compliant patients.
dehydrogenase Overall efficacy is
questionable (Mann, 2004)
Acamprosate Oral tablet NMDA receptor modulator Alcohol Two-fold increase in
abstinence rates at 1 year
(Mason and Heyser, 2010)
Nicotine Replacement Transdermal patch, gum, nAChR agonist Nicotine Two-fold increase in the
Treatments lozenge, nasal spray, and odds of quitting smoking
oral inhaler (Fiore et al., 2008)
Bupropion Oral tablet DA and NE reuptake blocker, Nicotine Two-fold increase in the
and nAChR antagonist odds of quitting smoking
(Eisenberg et al., 2008)
Varenicline Oral tablet Partial agonist for the a4b2 Nicotine 2- to 3-fold increase in the
and full agonist for the a7 odds of quitting smoking at
nAChR 6 months (Cahill et al., 2010)
Abbreviations: DA, dopamine; nAChR, nicotinic cholinergic receptor; NE, norepinephrine; NMDA, n-methyl-d-aspartate.

(Brauer and De Wit, 1997; Haney et al., 2001). Additional support the storage of dopamine in intracellular vesicles (Partilla et al.,
also comes from the food literature where differences in dopa- 2006; Sulzer, 2011). The long-term safety and abuse liability of
mine-related neural responses to highly versus less palatable amphetamines as a treatment for cocaine addiction remain to
foods are observed (Kenny, 2011). These clinical as well as other be determined.
preclinical findings (Berridge et al., 2009) provide indirect Another example of an agonist approach for cocaine depen-
evidence for a limited role of dopamine for drug ‘‘liking.’’ Identi- dence is modafinil, which has stimulant-like effects. Modafinil
fying the neurotransmitter mechanisms that mediate drug is a weak dopamine transporter inhibitor and increases synaptic
‘‘wanting’’ and ‘‘liking’’ responses may facilitate development of dopamine levels (Volkow et al., 2009). It also stimulates hypotha-
new pharmacotherapy targets for addictive disorders. lamic orexin neurons, reduces GABA release, and increases
1. Agonist approaches. Agonist medications have their main glutamate release (Martı́nez-Raga et al., 2008). While initial
impact on the same types of neurotransmitter receptors as those randomized clinical trials with modafinil were promising for
stimulated by abused substances. The general strategy of cocaine addiction (Dackis et al., 2005), a multisite clinical trial
agonist treatments is to substitute a safer, longer-acting drug was negative (Anderson et al., 2009). However, modafinil may
for a more risky, short-acting one. Examples of agonist treatment act as a cognitive enhancing agent in stimulant-dependent indi-
include methadone for opioid dependence and nicotine replace- viduals, improving learning through neural regions (insula and
ment treatment for smoking cessation (Table 1). Agonist treat- ventromedial prefrontal and anterior cingulate cortices) impli-
ment approaches have also been examined for cocaine depen- cated in learning and cognitive control (Ghahremani et al., 2011).
dence (Herin et al., 2010). Most notably, dextroamphetamine 2. Antagonist approaches. Antagonists block the effects of
has reduced drug use in short-term clinical trials in cocaine drugs by either pharmacological or pharmacokinetic mecha-
(Grabowski et al., 2004; Shearer et al., 2003) and methamphet- nisms. Antagonist treatment approaches have been especially
amine users (Longo et al., 2010; Shearer et al., 2001). Amphet- useful for opioid drugs. An example of pharmacological antago-
amines, similar to cocaine, increase synaptic dopamine levels nism is blockage of opioid effects by the m opioid antagonist
by inhibiting monoamine transporters and also by disrupting naltrexone or by buprenorphine, a partial m opioid agonist and

700 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review

k opioid antagonist. Buprenorphine and naltrexone block the transmitter systems in the brain including dopamine, norepi-
rewarding effects of opioids and are effective for the treatment nephrine, corticotrophin releasing hormone (CRH), GABA, and
for opioid addiction. Naltrexone also attenuates the rewarding glutamate (Chen et al., 2010; Koob and Le Moal, 2005). These
effects of alcohol by presumably blocking m opioid receptors adaptive changes are thought to underlie the negative reinforc-
(Ray et al., 2008), and this mechanism probably contributes to ing effects of abstinence from drug use that are clinically
naltrexone’e efficacy for the treatment of alcohol addiction observed as withdrawal symptoms, craving for drug use, and
(Sulzer, 2011). Similarly, varenicline, a partial agonist for the negative mood states like anhedonia and anxiety. Increased
alpha4beta2 nicotinic receptor, attenuates the rewarding effects norepinephrine activity is associated especially with opioid and
of nicotine (Patterson et al., 2009; Sofuoglu et al., 2009; West alcohol withdrawal states. Development of sensitization to
et al., 2008) and is effective for the treatment of nicotine depen- drug-related cues, perceived as craving induced by drug cues,
dence (Table 1). probably involve adaptive changes in the dopamine, GABA,
More recently, immunotherapies have been developed for the and glutamate systems (Schmidt and Pierce, 2010). Reduction
treatment of cocaine, methamphetamine, and nicotine addic- in dopamine levels in the ‘‘reward’’ circuit is thought to mediate
tions (Orson et al., 2008). Immunotherapies antagonize drug anhedonia commonly observed following abstinence from drugs
effects via pharmacokinetic mechanisms (LeSage et al., 2006). (Treadway and Zald, 2011). Examples of medications targeting
The antibodies produced by immunotherapies sequester the negative reinforcement of drugs include methadone or bupre-
drug in the circulation and reduce the amount of drug and the norphine, drugs that relieve opioid withdrawal symptoms.
speed at which it reaches the brain. This results in attenuated Nicotine replacement products, bupropion, and the partial nico-
rewarding effects of the drug of abuse (Haney et al., 2010). While tinic agonist varenicline relieve nicotine withdrawal symptoms
initial clinical trials suggest some promise (Martell et al., 2005, and attenuate the negative mood states after smoking cessation
2009), to date the efficacy of vaccines has been undercut by (Patterson et al., 2009; Sofuoglu et al., 2009). Acamprosate, an
a substantial induction period required to achieve clinically approved medication for the treatment of alcohol dependence,
significant levels of circulating antibodies and only partial attenuates withdrawal symptoms and craving for alcohol (Gual
blockade of drug effects even when antibody levels are maxi- and Lehert, 2001).
mized. An important limitation of vaccines is that the antibodies Medications targeting the noradrenergic system have shown
produced are specific for a given drug of abuse, a characteristic promising results for treatments targeting withdrawal or relapse.
that will limit their clinical efficacy in polydrug abusers. The most Preclinical and human laboratory studies suggest that lofexidine,
promising use of vaccine may be to prevent relapse in individual an alpha2-adrenergic agonist, may attenuate stress-induced
whose drug use is limited to a single agent. relapse in cocaine and opioid users (Highfield et al., 2001; Sinha
A potentially promising target for agonist and antagonist treat- et al., 2007). Cocaine users with more severe withdrawal symp-
ment of cocaine addiction is the D3 dopamine receptor (Heid- toms respond more favorably to propranolol, a beta-adrenergic
breder and Newman, 2010). Like the D2 dopamine receptor, antagonist (Kampman et al., 2006). Clinical trials are underway to
the D3 dopamine receptor is expressed at high levels in the stria- test the efficacies of carvedilol, an alpha and beta-adrenergic
tum, but compared to the D2 dopamine receptor, it is particularly antagonist, and guanfacine, an alpha2-adrenergic agonist, in
highly expressed in the ventral striatum. While D3 agonists treating cocaine or methamphetamine addiction.
partially reproduce cocaine reinforcement, D3 antagonists or Several agents targeting glutamate system are also under
partial agonists attenuate cocaine reinforcement (Achat-Mendes investigation as potential treatment medications. Memantine,
et al., 2010). D3 partial agonists (CJB090, BP 897, and others) a noncompetitive n-methyl-d-aspartate (NMDA) glutamate
can act like agonists and stimulate dopamine receptors when receptor antagonist, has also shown efficacy in reducing cue-
endogenous levels of dopamine are low, as in cocaine with- induced craving for alcohol in alcohol dependent patients (Kru-
drawal. In contrast, when dopamine receptors are stimulated pitsky et al., 2007). In pathological gambling, memantine may
after cocaine use, D3 partial agonists can act like antagonists be efficacious and operate by reducing cognitive measures of
in blocking the effects of cocaine (Martelle et al., 2007). However, compulsivity (Grant et al., 2010). However, clinical trials with
drugs with D2 and D3 antagonistic properties have not demon- memantine have demonstrated negative findings for alcohol
strated clinical efficacy for drug or nonsubstance addictions (Evans et al., 2007) and cocaine dependence (Bisaga et al.,
(Fong et al., 2008), D2/D3 antagonists have been associated 2010). A neutraceutical that targets the glutamate system is
with promoting of gambling-related motivations in pathological N-acetyl cysteine, a natural compound used for the treatment
gambling (Zack and Poulos, 2007), and dopamine agonists of acetaminophen overdose. N-acetyl cysteine’s proposed anti-
(including D3-preferring drugs) have been associated with addictive effects include normalization of reduced extracellular
nonsubstance addictions like pathological gambling in the treat- glutamate levels in the nucleus accumbens by stimulating the
ment of Parkinson’s disease (Weintraub et al., 2010). As such, cystine-glutamate antiporter (Baker et al., 2003). N-acetyl
the efficacies and tolerabilities of D3 partial agonists need careful cysteine has shown some positive results in small clinical trials
examination in people with addictions. Additionally, drugs that for cocaine and nicotine addiction and pathological gambling
target striatal dopamine function through indirect manners (Grant et al., 2007; Knackstedt et al., 2009; Mardikian et al.,
(e.g., through serotonin 1B receptors) also warrant consideration 2007). Larger studies are underway to test its efficacy in these
for treatment development (Hu et al., 2010). disorders. In addition, compounds targeting metabotropic gluta-
3. Medications targeting negative reinforcement of drugs. Drug mate receptors have shown efficacy in blocking reinstatement of
addiction is associated with adaptive changes in multiple neuro- drug use behavior in animal models for relapse. For example,

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 701


Neuron

Review

LY379268, an agonist of the group II metabotropic glutamate coping strategies. The range of deficits that is found in addicted
receptors, reduces self-administration and reinstatement of individuals includes attention, working memory, and response
drug-seeking behavior for nicotine (Liechti et al., 2007), alcohol inhibition, functions that are attributed to the prefrontal cortex.
(Sidhpura et al., 2010; Zhao et al., 2006), and cocaine (Adewale Cognitive functioning in the prefrontal cortex is modulated
et al., 2006; Baptista et al., 2004). Several metabotropic gluta- by many neurotransmitters, including glutamate, GABA, acetyl-
mate agonists are available for human use and should be evalu- choline, and monoamines: dopamine, serotonin, and norepi-
ated for the treatment of addictive disorders. nephrine (Robbins and Arnsten, 2009). Many cognitive
Medications Targeting Individual Vulnerability Factors enhancers targeting these neurotransmitters are in different
to Addiction stages of development.
Individuals vary in their vulnerability to addiction. For example, In a recent proof-of-concept study, we examined the efficacy
among those who had tried cocaine, only about 17% become of galantamine, a cholinesterase inhibitor, as a cognitive
addicted (Wagner and Anthony, 2002). For alcohol, about 15% enhancer in abstinent cocaine users (M.S., J. Poling, and
of those who drink eventually become dependent, while 30% K.M.C., unpublished data). Cholinesterase inhibitors, including
of those who try smoking become addicted smokers. These tacrine, rivastigmine, donepezil, and galantamine, have been
proportions are similar to those observed in preclinical models used for the treatment of dementia and other disorders charac-
of addiction (Belin et al., 2008). The individual factors contrib- terized by cognitive impairment, including Parkinson’s disease,
uting to vulnerability to addiction are complex and have not yet traumatic brain injury, and schizophrenia (Giacobini, 2004).
been fully elucidated (George and Koob, 2010; Kreek et al., Cholinesterase inhibitors increase the synaptic concentrations
2005; Le Moal, 2009; Sinha, 2008; Uhl et al., 2009). Comorbid of acetylcholine (ACh), which lead to increased stimulation of
psychiatric conditions and cognitive deficits are two examples both nicotinic and muscarinic cholinergic receptors. Galant-
of individual vulnerability factors that could be targeted by phar- amine is also an allosteric modulator of the a7 and a4b2 nicotinic
macotherapies. ACh receptor (nAChR) subtypes (Schilström et al., 2007). In our
1. Treatments targeting comorbid psychiatric conditions. study, 10 day treatment with galantamine, compared to placebo,
Comorbidity exists between drug addiction and primary psychi- improved the attention and working memory functions in absti-
atric disorders including schizophrenia, mood and anxiety disor- nent cocaine users (M.S., J. Poling, and K.M.C., unpublished
ders, and attention-deficit hyperactivity disorder (Hasin et al., data). These findings support the promise of galantamine as
2007; Kessler et al., 2005). For example, among individuals a cognitive enhancer among cocaine users. This study did not
with schizophrenia, 40% to 60% abuse drugs or alcohol and examine treatment effect on cocaine use because participants
over 90% smoke cigarettes (George et al., 2002). Addicted indi- had to be abstinent of drug use to allow accurate assessment
viduals with comorbid psychiatric disorders tend to have poorer of galantamine on cognitive performance. Additional clinical
outcomes than those without comorbidity (Brady and Sinha, trials are underway to test the efficacy of galantamine in the
2005; Havassy et al., 2004; Potenza, 2007). One of the possible treatment of cocaine-addicted individuals.
mechanisms underlying this high comorbidity is self-medication, Another promising medication for cognitive enhancement is
which posits that individual with primary psychiatric disorders atomoxetine, a selective norepinephrine transporter (NET) inhib-
use drugs or alcohol to relieve specific symptoms (e.g., negative itor used for the treatment of attention deficit hyperactivity
affect) or side effects of their treatment medications (e.g., seda- disorder (ADHD). In prefrontal cortex, the NET is responsible
tion). Alternatively, common genetic and other neurobiolgical for the reuptake of norepinephrine as well as dopamine into
factors may lead to high comorbidity between addictions and presynaptic nerve terminals (Kim et al., 2006). As a result, atom-
other psychiatric disorders (Chambers et al., 2001; Potenza oxetine increases both NE and dopamine levels in the PFC, and
et al., 2005). Common vulnerability factors may include both actions may contribute to the cognitive-enhancing effects
increased impulsivity, reward sensitivity, and cognitive deficits. of atomoxetine (Bymaster et al., 2002). Consistent with preclin-
One implication of comorbidity is that effective treatment of ical studies (Jentsch et al., 2009; Seu et al., 2009), atomoxetine
psychiatric disorders may also reduce substance use, although improves attention and response inhibition functions in healthy
existing clinical trials indicate mixed results in this regard (Nunes controls and patients with ADHD (Chamberlain et al., 2007,
and Levin, 2004). 2009; Faraone et al., 2005). Attention and response inhibition
2. Medications targeting cognitive deficits. A large body of functions are essential in optimum cognitive control needed to
evidence has documented cognitive deficits in chronic alcohol, prevent drug use, and atomoxetine in preclinical models dimin-
cocaine, methamphetamine, and cannabis users (Ersche et al., ished drug-seeking behaviors (Economidou et al., 2011). Both
2006; Ersche and Sahakian, 2007; Goldstein and Volkow, attention and response inhibition are impaired in cocaine users
2002). Cognitive deficits may represent a particular challenge (Li et al., 2006; Monterosso et al., 2005). Whether these cognitive
for treatment-seeking users who require intact cognitive func- functions can be improved and drug use curtailed with atomox-
tioning in order to engage in treatment and learn new behavioral etine remains to be determined in cocaine users.
strategies in order to stop their drug use. As demonstrated In addition to cholinesterase inhibitors and atomoxetine, there
previously, cognitive deficits are associated with higher rates are many other potential cognitive enhancers include modafinil,
of attrition and poor treatment outcome (Aharonovich et al., amphetamines, partial nAChR agonists, like varenicline, and
2006; Bates et al., 2006). Cognitive enhancement strategies metabotropic glutamate agonists (Olive, 2010). The safety and
may be especially important early in the treatment by improving efficacy of these medications remain to be tested in clinical
their ability to learn, remember, and implement new skills and studies with addicted individuals.

702 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review

Combined Behavioral and Pharmacological Treatment involving pre- and posttreatment imaging are needed to assess
Approaches more directly the relationships between specific treatments,
While great progress has been made in identification of effective outcome measures, and neural functions.
pharmacotherapies and behavioral therapies for the addictions, Although it is tempting to speculate that specific combinations
no existing treatment, delivered alone, is completely effective of treatments (e.g., behavioral and pharmacological therapies
(Carroll and Onken, 2005; Vocci et al., 2005). Thus, an important that theoretically engage top-down processes and bottom-up
strategy to enhance the efficacy of monotherapies is to combine processes, respectively; Figure 1) may have complementary
them with one or more alternative treatments (National Institute mechanisms of action, the precise mechanisms for synergism
on Drug Abuse, 2007). The results of combined treatments can between behavioral and pharmacotherapies are not well under-
be additive, interactive, nonadditive (adding a second treatment stood and require direct investigation. Existing data offer some
neither adds nor subtracts), or subtractive. Strategies for insight. For example, consistent with the notion of pharmaco-
choosing treatments to combine include (1) use of complemen- therapies working in a bottom-up fashion, bupropion treatment
tary efficacious treatments that address weakness in either of tobacco smokers was associated with less craving and dimin-
therapy alone, (2) use of efficacious treatments that target the ished limbic activation to smoking cues when attempting to
same processes in different ways, and (3) use of treatments resist craving, whereas placebo treatment did not demonstrate
that are not efficacious alone but catalyze each other. changes in limbic activations (Culbertson et al., 2011). However,
Frequently, these strategies involve combining a top-down in a study of tobacco smokers receiving treatment with bupro-
approach with a bottom-up intervention, such as combinations pion, practical group counseling, or pill placebo, individuals
of behavioral and pharmacotherapies (Figure 1). receiving either active treatment differed from those receiving
There are multiple examples of behavioral and pharmacolog- placebo by showing greater reduction in glucose metabolism
ical treatments having complementary effects. A classic posttreatment in the posterior cingulate cortex (Costello et al.,
example is the combination of methadone maintenance with 2010). The decreased metabolism was not related to cigarette
behavioral therapies (McLellan et al., 1993; Peirce et al., 2006). use measures and appeared largely similar across the behavioral
Without behavioral treatments, provision of methadone was and pharmacological therapies. As the posterior cingulate is an
associated with early treatment failure and dropout (Ball and integral component of the default mode network, the authors
Ross, 1991). Another example of this strategy involves antide- speculated that effective treatments for nicotine dependence
pressant medications and cognitive behavioral therapy, each may improve default mode network functioning, moving individ-
of which has been demonstrated to reduce depression in uals toward better goal-oriented states (Costello et al., 2010).
depressed smokers (Hall et al., 2002). Antidepressants are tar- As children with ADHD show suppression of default mode pro-
geted at neurotransmitter systems thought to underlie depres- cessing in response to stimulant treatment (Peterson et al.,
sion symptoms while CBT attempts to change behaviors and 2009), the findings suggest that improved default mode process-
cognitions associated with maintaining depression (DeRubeis ing function may represent an important treatment target across
et al., 1999; DeRubeis et al., 2008). An example of catalytic, or disorders characterized by impaired impulse control. Moreover,
synergistic, treatment effects is provided by studies that as posterior cingulate activation during drug craving has been
combine contingency management with tricyclic antidepres- associated with treatment outcome for cocaine dependence
sants for cocaine abuse in methadone-maintained patients (Kos- (Kosten et al., 2006), the findings also suggest an important
ten et al., 2003; Poling et al., 2006). In both of these trials, neither role for posterior cingulate function for treatment outcome
tricyclics nor contingency management was efficacious alone across addictions and one that may also relate to the involve-
but the combination yielded superior results compared to a stan- ment of the posterior cingulate in circuits related to emotional
dard treatment condition. Behavioral therapies may also work in and motivational processing (Sinha, 2008). Such possibilities
a complementary fashion, particularly in different stages of treat- warrant direct examination.
ment. For example, motivational interventions may help engage
individuals in treatment, contingency management may help Using Neuroscience to Investigate Treatment
maintain individuals in treatment, and CBT may help with long- Mechanisms
term abstinence through relapse prevention and ‘‘sleeper As reviewed above, traditional pharmacologic approaches to
effects.’’ Although not linked to a specific therapy, there are addiction have focused on exploiting our understanding of the
data to suggest that these different aspects of treatment specific actions of various neurotransmitters in the brain (e.g.,
outcome are differentially associated with specific neural dopamine for reward, opioids for pleasure, and adrenergic
circuits. For example, in cocaine-dependent individuals, neurochemicals for excitement) (Potenza, 2008). While
pretreatment fMRI measures of cognitive control were differen- continuing to increase our understanding of the neurochemical
tially associated with outcome measures of retention and absti- underpinnings of addictions remains important (particularly for
nence, with retention correlating with activation in the dorsolat- pharmacotherapy development), approaches to understanding
eral prefrontal cortex (implicated in executive functioning) and brain function related to addictions are increasingly focusing
abstinence with activation in striatal and ventromedial prefrontal on neural systems in the pathophysiologies of addictions.
cortical regions (implicated in reward processing and decision Thus, incorporating pre- and posttreatment neuroimaging
making) (Brewer et al., 2008a). As this study involved a small measures into randomized clinical trials for addictions is partic-
sample of subjects receiving combinations of behavioral and ularly important if we are to identify neural predictors and
pharmacological therapies, additional larger controlled studies correlates of effective treatments for these disorders.

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 703


Neuron

Review

There exist multiple considerations when integrating neuroi- represent important targets for treatment development (e.g.,
maging and clinical trials for addictions. While some are practical potential endophenotypes like impulsivity or compulsivity [Dalley
(e.g., a relatively short time frame between evaluation/randomi- et al., 2011]), whereas others (e.g., developmental stages, sex
zation and scanning requiring coordination between an interdis- differences, stage of the addiction process) may represent
ciplinary research team, questions as to how best to manage and important considerations when targeting or matching specific
consider recency of drug use—and potentially intoxication or treatments to specific individuals.
withdrawal—with respect to scanning), others are theoretical Endophenotypes represent particularly attractive therapeutic
(e.g., selecting measures that are theoretically related to the targets as they may associate more closely to biological mecha-
therapies’ proposed mechanisms of action, a notion consistent nisms than do heterogeneous psychiatric disorders like addic-
with selecting evaluative measures in clinical trials in general tions (Fineberg et al., 2010; Gottesman and Gould, 2003). One
[Walker et al., 2006]). An important advantage of fMRI in this potential endophenotype relevant to addiction treatment is
respect is the ability to monitor brain activity (via blood oxygen impulsivity (Dalley et al., 2011). Preclinical data indicate that
level dependent [BOLD] signal) during task performance. As impulsive tendencies prior to drug exposure both are linked to
such, specific fMRI paradigms may offer particular insights into ventral striatal dopamine function and predict the development
the mechanisms of action of particular therapies. For example, of addictive behaviors (Belin et al., 2008; Dalley et al., 2007).
effective contingency management, involving the delivery of Studies also link midbrain to ventral striatal dopamine pathways
small immediate rewards based on positive short-term behav- to impulsivity in people (Buckholtz et al., 2010). Clinical data
iors (e.g., drug abstinence) may be expected to involve changes suggest that impulsivity is associated with addiction severity
in reward processing that can be assessed through fMRI and that changes in addiction severity during treatment correlate
paradigms like the monetary incentive delay task (Andrews with changes in impulsivity (Blanco et al., 2009). Thus, targeting
et al., 2010). Alternatively, specific aspects of CBT, such as impulsivity through behavioral or pharmacological mechanisms
developing skills to cope with drug cues or triggers, might that promote self-control warrants consideration. As elevated
involve changes in brain circuitry underlying regulation of craving impulsivity may predate addictive problems, such interventions
or cognitive control that may be assessed through different may be considered at early points in either the addictive process
fMRI paradigms (Brewer et al., 2008a; Janes et al., 2010; or in development. This latter point seems particularly salient as
Kober et al., 2010). Other fMRI paradigms (e.g., those probing individual differences in self-control during childhood predict
stress responsiveness) may be particularly well suited for important measures of functioning during adolescence and
investigating mechanisms underlying mindfulness-based into adulthood (Lehrer, 2009; Mischel et al., 1989). Furthermore,
therapies (Brewer et al., 2009; Sinha et al., 2005). Additionally, as substance exposure during adolescence may lead to greater
advances in fMRI technology that facilitate real-time feedback impulsivity in adulthood (Nasrallah et al., 2009), early intervention
of regional brain activation may be used to investigate features appears particularly important.
relevant to specific therapies (e.g., control of craving in Targeting of specific factors may be complicated by the
CBT and meditational states in mindfulness-based therapies) complexities of the constructs. For example, impulsivity is
(deCharms, 2008). a multifaceted construct that factors into two or more domains
Conversely, novel methods of treatment delivery, such as (Meda et al., 2009; Moeller et al., 2001). Two domains repeatedly
computer-assisted delivery of CBT (Carroll et al., 2008; Carroll identified include those related to choice/decision making and
et al., 2009), may facilitate understanding of treatment mecha- response disinhibition, and each appears relevant to addiction
nisms through neuroimaging studies. Given the consistency (de Wit, 2009; Perry and Carroll, 2008; Potenza and de Wit,
with which it is delivered, computerized treatment offers 2010; Reynolds et al., 2006; Verdejo-Garcı́a et al., 2008). The
a more robust and standardized form of treatment. The conse- specific domains of impulsivity may relate differentially to other
quent reduction in variance in the treatment variable may relevant psychobiological processes (e.g., reward processing
increase the power of fMRI paradigms to detect processes and cognitive control appear theoretically and biologically linked
that are specific to this form of treatment, offering an advantage to choice and response impulsivity, respectively) and thus
in small-sample fMRI studies (Frewen et al., 2008). Also, compo- combinations of therapies that preferentially target each domain
nents of computer-delivered treatments could conceivably be may be needed to optimize treatments.
studied directly using fMRI. As self-report and behavioral measures of impulsivity have
been found to factor separately (Meda et al., 2009) and behav-
Future Directions: Individual Differences, ioral and self-reported measures of the same constructs (e.g.,
Endophenotypes, and Treatment Matching temporal discounting) may not correlate with one another and
One current focus in optimizing treatment involves identifying be differentially related to treatment outcome (Krishnan-Sarin
individual differences related to addiction treatment outcome et al., 2007), a broad range of self-report, behavioral, and biolog-
to guide the selection of therapies. While the consideration of ical assessments (including neurocognitive ones) may provide
individual differences is not new (e.g., Project MATCH investi- the deep phenotyping that will be vital to treatment develop-
gated individual differences and treatment specificity with argu- ments for addictions. Additionally, as brain circuits underlying
ably limited success [Cutler and Fishbein, 2005]), recent motivation, reward responsiveness, decision making, and
approaches have considered individual differences from behavioral control are undergoing significant changes during
a different perspective (e.g., as possible endophenotypes [Got- periods of increased addiction vulnerability such as adolescence
tesman and Gould, 2003]). Some individual differences may (Casey et al., 2010; Chambers et al., 2003; Rutherford et al.,

704 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review

2010; Somerville et al., 2010), developmental considerations are a sex- or culture-specific fashion, thorough assessments and
important in this process. large samples involving targeted recruitment may be necessary
Potential endophenotypes may underlie multiple kinds of to optimize treatment strategies for individuals.
addictions (Frascella et al., 2010). However, specific drugs are Drug-Related Brain Changes: Consideration
also associated with unique short- and long-term effects, of Nonsubstance Addictions
including potential neurotoxicities. Drug exposure may have Given the potential neurotoxic and neuroadaptation effects of
specific influences on brain structure and function, and such abused substances, understanding the neuroscience of addic-
changes warrant particular attention as they relate to treatment tive processes may be enhanced by focusing on addictions
development. For example, cocaine use has been associated that do not necessarily involve use of psychoactive substances.
with metabolic impairments, with increasing chronicity of use For example, obesity shares similarities with drug addictions
progressively influencing cortical regions from more ventral at neurobiological levels (e.g., with respect to striatal D2/D3
and medial regions to more dorsal and lateral ones (Beveridge dopamine receptor function), and these similarities may inform
et al., 2008). These findings are consistent with a broad range treatment and policy strategies (A.N. Gearhardt, C.M. Grilo,
of cognitive deficits observed in cocaine dependent individuals, R.J. DeLeone, K.D. Brownell, and M.N.P., unpublished data;
including on tasks associated with ventromedial prefrontal Vanbuskirk and Potenza, 2010). Pathological gambling also
cortical function (Bechara, 2003) as well as ones linked to dorso- demonstrates clinical and biological similarities with drug addic-
lateral prefrontal cortical function and associated with treatment tions (Holden, 2010; Potenza, 2006, 2008). Consistently, treat-
outcome measures of retention (Brewer et al., 2008a; Streeter ments, particularly those with proposed mechanisms of action
et al., 2008). Other brain differences, such as white matter integ- (e.g., modulation of neurotransmission in the mesolimbic dopa-
rity (Lim et al., 2002, 2008; Moeller et al., 2005; Moeller et al., mine pathway by opioid receptor antagonists like naltrexone or
2007), have been observed in cocaine dependence and associ- nalmefene or enhancing cognitive function via glutamatergic
ated with disadvantageous decision making (Lane et al., 2010) agents like memantine) that target features observed across
and treatment outcome (Xu et al., 2010). Both pharmacological addictions, appear efficacious for both substance and gambling
(Harsan et al., 2008; Schlaug et al., 2009) and behavioral (Tang addictions (Brewer et al., 2008b; Cheon et al., 2008; Grant et al.,
et al., 2010) approaches may alter white matter integrity. Thus, 2010; Potenza, 2008). Furthermore, among individuals with
white matter integrity may represent an underexamined thera- pathological gambling, response to an opioid receptor antago-
peutic target in addictions. Additionally, investigating means nist appears strongly related to a family history of alcoholism
for altering synaptic connections, including rapid mechanisms (Grant et al., 2008), suggesting a possible endophenotype
related to brief exposure to antiglutamatergic drugs (Li et al., common to pathological gambling and alcoholism. However,
2010), may aid addiction treatment development efforts, partic- other features, such as executive processes involving dorsal
ularly as related to stress or other negative reinforcement prefrontal cortical function, appear more impaired in individuals
processes. These considerations underscore the promise of with alcoholism than in those with gambling problems, consis-
developing and testing (both singly and in combination) pharma- tent with neurotoxic influences of alcohol (Lawrence et al.,
cological and behavioral treatments aimed at improving cogni- 2009; Potenza, 2009b). As pathological gambling is unhindered
tive functions such as attention, working memory, decision by drug-on-brain-substrate effects that may complicate the
making, and self-control. Relating the results of these treatments treatment of substance addictions, it represents an important
to measures of impulsivity and brain function can provide disorder for better understanding substance addictions and their
evidence for mechanisms of these treatments. treatments.
Endophenotypes may track closely with genetic factors, and
individual differences related to addictions and their treatments Conclusions
may be influenced by genetic, environmental, or interactive influ- Although significant advances have been made over the past
ences (Goldman et al., 2005; Renthal and Nestler, 2008). As several decades in the development of effective treatments for
commonly occurring allelic variants have been variably linked addictions, they remain a substantial public health problem.
to treatment outcomes for addictions (e.g., a functional variant The development of neuroscience methodologies for assessing
of the gene encoding the m opioid receptor has been associated brain structure and function provides an exciting opportunity for
with opioid antagonist treatment outcome in some (Oslin et al., applying these tools to understand and improve treatments.
2003) but not other (Arias et al., 2008) studies of alcohol depen- Additional research efforts should define novel targets for treat-
dence or heavy drinking) and specific environmental exposures ment (e.g., cognitive function, control of craving, impulsivity,
in conjunction with commonly occurring allelic variants may shift compulsivity, and/or self-control), implement tools for assessing
the risk for developing and treating addictions (e.g., stress expo- these targets over time (including self-report, behavioral, neuro-
sure and serotonin-transporter-encoding genetic variants cognitive/neural measures), and identify clinically relevant indi-
interact to influence alcohol intake in young adults and may be vidual differences that may be used to guide the selections of
linked to ondansetron response in alcohol dependence [Johnson therapies, including combinations of therapies that may operate
et al., 2008; Laucht et al., 2009; Sinha, 2009]), it will be important in complementary or synergistic fashions. As effects of drug use
to carefully assess multiple environmental and genetic measures on brain and brain function may be a major factor underlying
as related to treatment outcome. Furthermore, as timing of envi- ability to benefit from treatment, direct investigation of drug-
ronmental exposures may differentially impact individuals (e.g., related influences on brain structure and function are warranted
influences of trauma early versus later in life) and do so in in translational and longitudinal studies. Concurrent investigation

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 705


Neuron

Review

of substance and nonsubstance addictions should be especially Baker, D.A., McFarland, K., Lake, R.W., Shen, H., Toda, S., and Kalivas, P.W.
(2003). N-acetyl cysteine-induced blockade of cocaine-induced reinstate-
informative. ment. Ann. N. Y. Acad. Sci. 1003, 349–351.

Ball, J.C., and Ross, A. (1991). The Effectiveness of Methadone Maintenance


ACKNOWLEDGMENTS Treatment (New York, NY: Springer Verlag).

The authors report that they have no financial conflicts of interest with respect Balleine, B.W., Doya, K., O’Doherty, J., and Sakagami, M. (2007). Current
to the content of this manuscript. Dr. Potenza has received financial support or trends in decision making. Ann. N. Y. Acad. Sci. 1104, xi–xv.
compensation for the following: Dr. Potenza has consulted for and advised
Baptista, M.A., Martin-Fardon, R., and Weiss, F. (2004). Preferential effects of
Boehringer Ingelheim; has consulted for and has financial interests in
the metabotropic glutamate 2/3 receptor agonist LY379268 on conditioned
Somaxon; has received research support from the National Institutes of reinstatement versus primary reinforcement: Comparison between cocaine
Health, Veteran’s Administration, Mohegan Sun Casino, the National Center and a potent conventional reinforcer. J. Neurosci. 24, 4723–4727.
for Responsible Gaming and its affiliated Institute for Research on Gambling
Disorders, and Forest Laboratories, Ortho-McNeil, Oy-Control/Biotie and Bates, M.E., Pawlak, A.P., Tonigan, J.S., and Buckman, J.F. (2006). Cognitive
Glaxo-SmithKline pharmaceuticals; has participated in surveys, mailings or impairment influences drinking outcome by altering therapeutic mechanisms
telephone consultations related to drug addiction, impulse control disorders of change. Psychol. Addict. Behav. 20, 241–253.
or other health topics; has consulted for law offices and the federal public
defender’s office in issues related to impulse control disorders; provides Bechara, A. (2003). Risky business: Emotion, decision-making, and addiction.
clinical care in the Connecticut Department of Mental Health and Addiction J. Gambl. Stud. 19, 23–51.
Services Problem Gambling Services Program; has performed grant reviews
Belin, D., Mar, A.C., Dalley, J.W., Robbins, T.W., and Everitt, B.J. (2008). High
for the National Institutes of Health and other agencies; has guest-edited jour- impulsivity predicts the switch to compulsive cocaine-taking. Science 320,
nal sections; has given academic lectures in grand rounds, CME events and 1352–1355.
other clinical or scientific venues; and has generated books or book chapters
for publishers of mental health texts. The content of this manuscript is solely Bell, M., Bryson, G., Greig, T., Corcoran, C., and Wexler, B.E. (2001). Neuro-
the responsibility of the authors and does not necessarily represent the official cognitive enhancement therapy with work therapy: Effects on neuropsycho-
views of the funding agencies. Support was provided by National Institute logical test performance. Arch. Gen. Psychiatry 58, 763–768.
on Drug Abuse grants P50-DA09241, P50-DA016556, P20-DA027844,
R37-DA015969, K02-DA021304, R01-DA020908, RC1-DA028279, Berridge, K.C., Robinson, T.E., and Aldridge, J.W. (2009). Dissecting
R01-DA018647, R21 DA029445, and R01-DA019039, the National Institute components of reward: ‘Liking’, ‘wanting’, and learning. Curr. Opin. Pharma-
on Alcoholism and Alcohol Abuse grant RL1-AA017539, the VISN 1 Mental col. 9, 65–73.
Illness Educational, Research, and Clinical Center (MIRECC), and a Center
Beveridge, T.J.R., Gill, K.E., Hanlon, C.A., and Porrino, L.J. (2008). Review.
of Excellence in Gambling Research from the National Center for Responsible Parallel studies of cocaine-related neural and cognitive impairment in humans
Gaming. and monkeys. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, 3257–3266.

Bickel, W.K., and Yi, R. (2008). Temporal discounting as a measure of execu-


REFERENCES tive function: Insights from the competing neuro-behavioral decision system
hypothesis of addiction. Adv. Health Econ. Health Serv. Res. 20, 289–309.
Achat-Mendes, C., Grundt, P., Cao, J., Platt, D.M., Newman, A.H., and Speal-
man, R.D. (2010). Dopamine D3 and D2 receptor mechanisms in the abuse- Bickel, W.K., Yi, R., Landes, R.D., Hill, P.F., and Baxter, C. (2011). Remember
related behavioral effects of cocaine: Studies with preferential antagonists in the future: Working memory training decreases delay discounting among stim-
squirrel monkeys. J. Pharmacol. Exp. Ther. 334, 556–565. ulant addicts. Biol. Psychiatry 69, 260–265.

Adewale, A.S., Platt, D.M., and Spealman, R.D. (2006). Pharmacological Bisaga, A., Aharonovich, E., Cheng, W.Y., Levin, F.R., Mariani, J.J., Raby,
stimulation of group ii metabotropic glutamate receptors reduces cocaine W.N., and Nunes, E.V. (2010). A placebo-controlled trial of memantine for
self-administration and cocaine-induced reinstatement of drug seeking in cocaine dependence with high-value voucher incentives during a pre-random-
squirrel monkeys. J. Pharmacol. Exp. Ther. 318, 922–931. ization lead-in period. Drug Alcohol Depend. 111, 97–104.

Aharonovich, E., Hasin, D.S., Brooks, A.C., Liu, X., Bisaga, A., and Nunes, E.V. Blanco, C., Potenza, M.N., Kim, S.W., Ibáñez, A., Zaninelli, R., Saiz-Ruiz, J.,
(2006). Cognitive deficits predict low treatment retention in cocaine dependent and Grant, J.E. (2009). A pilot study of impulsivity and compulsivity in patho-
patients. Drug Alcohol Depend. 81, 313–322. logical gambling. Psychiatry Res. 167, 161–168.

Alexander, G.E., Crutcher, M.D., and DeLong, M.R. (1990). Basal ganglia-tha- Blum, K., Cull, J.G., Braverman, E.R., and Comings, D.E. (1996). Reward defi-
lamocortical circuits: Parallel substrates for motor, oculomotor, ‘‘prefrontal’’ ciency syndrome. Am. Sci. 84, 132–145.
and ‘‘limbic’’ functions. Prog. Brain Res. 85, 119–146.
Botvinick, M.M., Braver, T.S., Barch, D.M., Carter, C.S., and Cohen, J.D.
(2001). Conflict monitoring and cognitive control. Psychol. Rev. 108, 624–652.
Alexander, G.E., DeLong, M.R., and Strick, P.L. (1986). Parallel organization of
functionally segregated circuits linking basal ganglia and cortex. Annu. Rev. Brady, K.T., and Sinha, R. (2005). Co-occurring mental and substance use
Neurosci. 9, 357–381. disorders: The neurobiological effects of chronic stress. Am. J. Psychiatry
162, 1483–1493.
Anderson, A.L., Reid, M.S., Li, S.H., Holmes, T., Shemanski, L., Slee, A., Smith,
E.V., Kahn, R., Chiang, N., Vocci, F., et al. (2009). Modafinil for the treatment of Brauer, L.H., and De Wit, H. (1997). High dose pimozide does not block
cocaine dependence. Drug Alcohol Depend. 104, 133–139. amphetamine-induced euphoria in normal volunteers. Pharmacol. Biochem.
Behav. 56, 265–272.
Andrews, M.M., Meda, S.A., Thomas, A.D., Potenza, M.N., Krystal, J.H.,
Worhunsky, P., Stevens, M.C., O’Malley, S.S., Book, G.A., and Pearlson, Brewer, J.A., Worhunsky, P.D., Carroll, K.M., Rounsaville, B.J., and Potenza,
G.D. (2010). Individuals family history positive for alcoholism show fMRI M.N. (2008a). Pretreatment brain activation during stroop task is associated
abnormalities in reward sensitivity that are related to impulsivity factors. with outcomes in cocaine-dependent patients. Biol. Psychiatry 64, 998–1004.
Biol. Psychiatry, in press. Published online November 30, 2010. 10.1016/
j.biopsych.2010.09.049. Brewer, J.A., Grant, J.E., and Potenza, M.N. (2008b). The treatment of patho-
logic gambling. Addict. Disorders Treatment 7, 1–14.
Alcoholics Anonymous. (1986). Twelve Steps and Twelve Traditions, Vol. 32nd
Printing (New York, NY: Alcoholics Anonymous World Services, Inc.). Brewer, J.A., and Potenza, M.N. (2008). The neurobiology and genetics of
impulse control disorders: Relationships to drug addictions. Biochem. Phar-
Arias, A.J., Armeli, S., Gelernter, J., Covault, J., Kallio, A., Karhuvaara, S., macol. 75, 63–75.
Koivisto, T., Mäkelä, R., and Kranzler, H.R. (2008). Effects of opioid receptor
gene variation on targeted nalmefene treatment in heavy drinkers. Alcohol. Brewer, J.A., Sinha, R., Chen, J.A., Michalsen, R.N., Babuscio, T.A., Nich, C.,
Clin. Exp. Res. 32, 1159–1166. Grier, A., Bergquist, K.L., Reis, D.L., Potenza, M.N., et al. (2009). Mindfulness

706 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review
training reduces stress reactivity and relapse in substance abuse: Results from Chambers, R.A., Taylor, J.R., and Potenza, M.N. (2003). Developmental
a randomized, controlled pilot study. Subst. Abus. 30, 306–317. neurocircuitry of motivation in adolescence: A critical period of addiction
vulnerability. Am. J. Psychiatry 160, 1041–1052.
Brewer, J.A., Bowen, S., Smith, J.T., Marlatt, G.A., and Potenza, M.N. (2010).
Mindfulness-based treatments for co-occurring depression and substance Chen, B.T., Hopf, F.W., and Bonci, A. (2010). Synaptic plasticity in the meso-
use disorders: What can we learn from the brain? Addiction 105, 1698–1706. limbic system: Therapeutic implications for substance abuse. Ann. N Y Acad.
Sci. 1187, 129–139.
Buckholtz, J.W., Treadway, M.T., Cowan, R.L., Woodward, N.D., Li, R., Ansari,
M.S., Baldwin, R.M., Schwartzman, A.N., Shelby, E.S., Smith, C.E., et al. Cheon, Y., Park, J., Joe, K.-H., and Kim, D.-J. (2008). The effect of 12-week
(2010). Dopaminergic network differences in human impulsivity. Science open-label memantine treatment on cognitive function improvement in
329, 532. patients with alcohol-related dementia. Int. J. Neuropsychopharmacol. 11,
971–983.
Burke, B.L., Arkowitz, H., and Menchola, M. (2003). The efficacy of motiva-
tional interviewing: A meta-analysis of controlled clinical trials. J. Consult. Childress, A.R., Mozley, P.D., McElgin, W., Fitzgerald, J., Reivich, M., and
Clin. Psychol. 71, 843–861. O’Brien, C.P. (1999). Limbic activation during cue-induced cocaine craving.
Am. J. Psychiatry 156, 11–18.
Bush, G., Luu, P., and Posner, M.I. (2000). Cognitive and emotional influences
in anterior cingulate cortex. Trends Cogn. Sci. (Regul. Ed.) 4, 215–222. Costello, M.R., Mandelkern, M.A., Shoptaw, S., Shulenberger, S., Baker, S.K.,
Abrams, A.L., Xia, C., London, E.D., and Brody, A.L. (2010). Effects of treat-
ment for tobacco dependence on resting cerebral glucose metabolism.
Bush, G., Vogt, B.A., Holmes, J., Dale, A.M., Greve, D., Jenike, M.A., and
Neuropsychopharmacology 35, 605–612.
Rosen, B.R. (2002). Dorsal anterior cingulate cortex: A role in reward-based
decision making. Proc. Natl. Acad. Sci. USA 99, 523–528.
Culbertson, C.S., Bramen, J., Cohen, M.S., London, E.D., Olmstead, R.E.,
Gan, J.J., Costello, M.R., Shulenberger, S., Mandelkern, M.A., and Brody,
Bymaster, F.P., Katner, J.S., Nelson, D.L., Hemrick-Luecke, S.K., Threlkeld, A.L. (2011). Effect of bupropion treatment on brain activation induced by ciga-
P.G., Heiligenstein, J.H., Morin, S.M., Gehlert, D.R., and Perry, K.W. (2002).
rette-related cues in smokers. Arch. Gen. Psychiatry, in press. Published
Atomoxetine increases extracellular levels of norepinephrine and dopamine online January 3, 2011. 10.1001/archgenpsychiatry.2010.193.
in prefrontal cortex of rat: A potential mechanism for efficacy in attention
deficit/hyperactivity disorder. Neuropsychopharmacology 27, 699–711. Cutler, R.B., and Fishbein, D.A. (2005). Are alcoholism treatments effective?
The Project MATCH data. BMC Public Health 5, 75.
Cahill, K.M., Stead, L.F., and Lancaster, T. (2010). Nicotine receptor partial
agonists for smoking cessation. Cochrane Database Syst. Rev. 12, CD006103. Dackis, C.A., Kampman, K.M., Lynch, K.G., Pettinati, H.M., and O’Brien, C.P.
(2005). A double-blind, placebo-controlled trial of modafinil for cocaine depen-
Carroll, K.M., Connors, G.J., Cooney, N.L., DiClemente, C.C., Donovan, D.M., dence. Neuropsychopharmacology 30, 205–211.
Kadden, R.R., Longabaugh, R.L., Rounsaville, B.J., Wirtz, P.W., and Zweben,
A. (1998). A Cognitive Behavioral Approach: Treating Cocaine Addiction Dalley, J.W., and Everitt, B.J. (2009). Dopamine receptors in the learning,
(Rockville, MD: NIDA). memory and drug reward circuitry. Semin. Cell Dev. Biol. 20, 403–410.

Carroll, K.M., Kosten, T.R., and Rounsaville, B.J. (2004). Choosing a behavioral Dalley, J.W., Everitt, B.J., and Robbins, T.W. (2011). Impulsivity, compulsivity,
therapy platform for pharmacotherapy of substance users. Drug Alcohol and top-down cognitive control. Neuron 69, this issue, 680–694.
Depend. 75, 123–134.
Dalley, J.W., Fryer, T.D., Brichard, L., Robinson, E.S.J., Theobald, D.E.H.,
Carroll, K.M., Ball, S.A., Martino, S., Nich, C., Babuscio, T.A., Nuro, K.F., Lääne, K., Peña, Y., Murphy, E.R., Shah, Y., Probst, K., et al. (2007). Nucleus
Gordon, M.A., Portnoy, G.A., and Rounsaville, B.J. (2008). Computer-assisted accumbens D2/3 receptors predict trait impulsivity and cocaine reinforce-
delivery of cognitive-behavioral therapy for addiction: A randomized trial of ment. Science 315, 1267–1270.
CBT4CBT. Am. J. Psychiatry 165, 881–888.
Danaei, G., Ding, E.L., Mozaffarian, D., Taylor, B., Rehm, J., Murray, C.J.L.,
Carroll, K.M., Ball, S.A., Martino, S., Nich, C., Babuscio, T.A., and Rounsaville, and Ezzati, M. (2009). The preventable causes of death in the United States:
B.J. (2009). Enduring effects of a computer-assisted training program for Comparative risk assessment of dietary, lifestyle, and metabolic risk factors.
cognitive behavioral therapy: A 6-month follow-up of CBT4CBT. Drug Alcohol PLoS Med. 6, e1000058.
Depend. 100, 178–181.
de Wit, H. (2009). Impulsivity as a determinant and consequence of drug use:
Carroll, K.M., and Onken, L.S. (2005). Behavioral therapies for drug abuse. Am. A review of underlying processes. Addict. Biol. 14, 22–31.
J. Psychiatry 162, 1452–1460.
deCharms, R.C. (2008). Applications of real-time fMRI. Nat. Rev. Neurosci. 9,
Carroll, K.M., Rounsaville, B.J., Nich, C., Gordon, L.T., Wirtz, P.W., and Gawin, 720–729.
F.H. (1994). One-year follow-up of psychotherapy and pharmacotherapy for
DeRubeis, R.J., Gelfand, L.A., Tang, T.Z., and Simons, A.D. (1999). Medica-
cocaine dependence. Delayed emergence of psychotherapy effects. Arch.
tions versus cognitive behavior therapy for severely depressed outpatients:
Gen. Psychiatry 51, 989–997.
Mega-analysis of four randomized comparisons. Am. J. Psychiatry 156,
1007–1013.
Casey, B.J., Duhoux, S., and Malter Cohen, M. (2010). Adolescence: What
do transmission, transition, and translation have to do with it? Neuron 67, DeRubeis, R.J., Siegle, G.J., and Hollon, S.D. (2008). Cognitive therapy versus
749–760. medication for depression: Treatment outcomes and neural mechanisms.
Nat. Rev. Neurosci. 9, 788–796.
Chamberlain, S.R., Del Campo, N., Dowson, J., Müller, U., Clark, L., Robbins,
T.W., and Sahakian, B.J. (2007). Atomoxetine improved response inhibition Dutra, L., Stathopoulou, G., Basden, S.L., Leyro, T.M., Powers, M.B., and Otto,
in adults with attention deficit/hyperactivity disorder. Biol. Psychiatry 62, M.W. (2008). A meta-analytic review of psychosocial interventions for
977–984. substance use disorders. Am. J. Psychiatry 165, 179–187.
Chamberlain, S.R., Hampshire, A., Müller, U., Rubia, K., Del Campo, N., Craig, Economidou, D., Dalley, J.W., and Everitt, B.J. (2011). Selective norepineph-
K., Regenthal, R., Suckling, J., Roiser, J.P., Grant, J.E., et al. (2009). rine reuptake inhibition by atomoxetine prevents cue-induced heroin and
Atomoxetine modulates right inferior frontal activation during inhibitory cocaine seeking. Biol. Psychiatry 69, 266–274.
control: A pharmacological functional magnetic resonance imaging study.
Biol. Psychiatry 65, 550–555. Edwards, S., and Koob, G.F. (2010). Neurobiology of dysregulated motiva-
tional systems in drug addiction. Future Neurol. 5, 393–401.
Chambers, R.A., Krystal, J.H., and Self, D.W. (2001). A neurobiological basis
for substance abuse comorbidity in schizophrenia. Biol. Psychiatry 50, 71–83. Eisenberg, M.J., Filion, K.B., Yavin, D., Bélisle, P., Mottillo, S., Joseph, L.,
Gervais, A., O’Loughlin, J., Paradis, G., Rinfret, S., and Pilote, L. (2008).
Chambers, R.A., Bickel, W.K., and Potenza, M.N. (2007). A scale-free systems Pharmacotherapies for smoking cessation: A meta-analysis of randomized
theory of motivation and addiction. Neurosci. Biobehav. Rev. 31, 1017–1045. controlled trials. CMAJ 179, 135–144.

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 707


Neuron

Review
Ersche, K.D., Fletcher, P.C., Roiser, J.P., Fryer, T.D., London, M., Robbins, Grabowski, J., Rhoades, H., Stotts, A., Cowan, K., Kopecky, C., Dougherty, A.,
T.W., and Sahakian, B.J. (2006). Differences in orbitofrontal activation during Moeller, F.G., Hassan, S., and Schmitz, J. (2004). Agonist-like or antagonist-
decision-making between methadone-maintained opiate users, heroin users like treatment for cocaine dependence with methadone for heroin depen-
and healthy volunteers. Psychopharmacology (Berl.) 188, 364–373. dence: Two double-blind randomized clinical trials. Neuropsychopharmacol-
ogy 29, 969–981.
Ersche, K.D., and Sahakian, B.J. (2007). The neuropsychology of amphet-
amine and opiate dependence: Implications for treatment. Neuropsychol. Grant, J.E., Kim, S.W., Hollander, E., and Potenza, M.N. (2008). Predicting
Rev. 17, 317–336. response to opiate antagonists and placebo in the treatment of pathological
gambling. Psychopharmacology (Berl.) 200, 521–527.
Evans, S.M., Levin, F.R., Brooks, D.J., and Garawi, F. (2007). A pilot double-
blind treatment trial of memantine for alcohol dependence. Alcohol. Clin. Grant, J.E., Donahue, C.B., Odlaug, B.L., Kim, S.W., Miller, M.J., and Petry,
Exp. Res. 31, 775–782. N.M. (2009). Imaginal desensitisation plus motivational interviewing for patho-
logical gambling: A randomised controlled trial. Br. J. Psychol. 195, 266–267.
Everitt, B.J., and Robbins, T.W. (2005). Neural systems of reinforcement for
drug addiction: From actions to habits to compulsion. Nat. Neurosci. 8, Grant, J.E., Kim, S.W., and Odlaug, B.L. (2007). N-acetyl cysteine, a gluta-
1481–1489. mate-modulating agent, in the treatment of pathological gambling: A pilot
study. Biol. Psychiatry 62, 652–657.
Faraone, S.V., Biederman, J., Spencer, T., Michelson, D., Adler, L., Reimherr,
F., and Seidman, L. (2005). Atomoxetine and stroop task performance in adult Grant, J.E., Chamberlain, S.R., Odlaug, B.L., Potenza, M.N., and Kim, S.W.
attention-deficit/hyperactivity disorder. J. Child Adolesc. Psychopharmacol. (2010). Memantine shows promise in reducing gambling severity and cognitive
15, 664–670. inflexibility in pathological gambling: A pilot study. Psychopharmacology
(Berl.) 212, 603–612.
Fineberg, N.A., Potenza, M.N., Chamberlain, S.R., Berlin, H.A., Menzies, L.,
Bechara, A., Sahakian, B.J., Robbins, T.W., Bullmore, E.T., and Hollander, Grant, J.E., Potenza, M.N., Weinstein, A., and Gorelick, D.A. (2010b). Introduc-
E. (2010). Probing compulsive and impulsive behaviors, from animal models tion to behavioral addictions. Am. J. Drug Alcohol Abuse 36, 233–241.
to endophenotypes: A narrative review. Neuropsychopharmacology 35,
591–604. Grohman, K., Fals-Stewart, W., and Donnelly, K. (2006). Improving treatment
response of cognitively impaired veterans with neuropsychological rehabilita-
Fiore, M.C., Jaen, C.R., Baker, T.B., Bailey, W.C., Benowitz, N.L., Curry, S.J., tion. Brain Cogn. 60, 203–204.
Dorfman, S.F., Froelicher, E.S., Goldstein, M.G., Healton, C.G., et al. (2008).
Treating Tobacco Use and Dependence: 2008 Update. Clinical Practice Gual, A., and Lehert, P. (2001). Acamprosate during and after acute alcohol
Guidelines (Rockville, MD: U.S. Department of Health and Human Services. withdrawal: A double-blind placebo-controlled study in Spain. Alcohol
Public Health Service). Alcohol. 36, 413–418.
Fong, T., Kalechstein, A., Bernhard, B., Rosenthal, R., and Rugle, L. (2008). A Haber, S.N., and Knutson, B. (2010). The reward circuit: Linking primate
double-blind, placebo-controlled trial of olanzapine for the treatment of video anatomy and human imaging. Neuropsychopharmacology 35, 4–26.
poker pathological gamblers. Pharmacol. Biochem. Behav. 89, 298–303.
Hall, S.M., Humfleet, G.L., Reus, V.I., Muñoz, R.F., Hartz, D.T., and Maude-
Frascella, J., Potenza, M.N., Brown, L.L., and Childress, A.R. (2010). Shared Griffin, R. (2002). Psychological intervention and antidepressant treatment in
brain vulnerabilities open the way for nonsubstance addictions: Carving addic- smoking cessation. Arch. Gen. Psychiatry 59, 930–936.
tion at a new joint? Ann. N. Y. Acad. Sci. 1187, 294–315.
Haney, M., Gunderson, E.W., Jiang, H., Collins, E.D., and Foltin, R.W. (2010).
Frewen, P.A., Dozois, D.J., and Lanius, R.A. (2008). Neuroimaging studies of
Cocaine-specific antibodies blunt the subjective effects of smoked cocaine in
psychological interventions for mood and anxiety disorders: Empirical and
humans. Biol. Psychiatry 67, 59–65.
methodological review. Clin. Psychol. Rev. 28, 228–246.
Haney, M., Ward, A.S., Foltin, R.W., and Fischman, M.W. (2001). Effects of
George, O., and Koob, G.F. (2010). Individual differences in prefrontal cortex
ecopipam, a selective dopamine D1 antagonist, on smoked cocaine self-
function and the transition from drug use to drug dependence. Neurosci.
administration by humans. Psychopharmacology (Berl.) 155, 330–337.
Biobehav. Rev. 35, 232–247.

George, T.P., Vessicchio, J.C., Termine, A., Bregartner, T.A., Feingold, A., Harsan, L.A., Steibel, J., Zaremba, A., Agin, A., Sapin, R., Poulet, P., Guignard,
Rounsaville, B.J., and Kosten, T.R. (2002). A placebo controlled trial of bupro- B., Parizel, N., Grucker, D., Boehm, N., et al. (2008). Recovery from chronic
pion for smoking cessation in schizophrenia. Biol. Psychiatry 52, 53–61. demyelination by thyroid hormone therapy: Myelinogenesis induction and
assessment by diffusion tensor magnetic resonance imaging. J. Neurosci.
Ghahremani, D.G., Tabibnia, G., Monterosso, J., Hellemann, G., Poldrack, 28, 14189–14201.
R.A., and London, E.D. (2011). Effect of modafinil on learning and task-related
brain activity in methamphetamine-dependent and healthy individuals. Hasin, D.S., Stinson, F.S., Ogburn, E., and Grant, B.F. (2007). Prevalence,
Neuropsychopharmacology, in press. Published online February 2, 2011. correlates, disability, and comorbidity of DSM-IV alcohol abuse and depen-
10.1038/npp.2010.233. dence in the United States: Results from the National Epidemiologic Survey
on Alcohol and Related Conditions. Arch. Gen. Psychiatry 64, 830–842.
Giacobini, E. (2004). Cholinesterase inhibitors: New roles and therapeutic
alternatives. Pharmacol. Res. 50, 433–440. Havassy, B.E., Alvidrez, J., and Owen, K.K. (2004). Comparisons of patients
with comorbid psychiatric and substance use disorders: Implications for treat-
Glimcher, P.W., and Rustichini, A. (2004). Neuroeconomics: The consilience of ment and service delivery. Am. J. Psychiatry 161, 139–145.
brain and decision. Science 306, 447–452.
Heidbreder, C.A., and Newman, A.H. (2010). Current perspectives on selective
Goldapple, K., Segal, Z., Garson, C., Lau, M., Bieling, P., Kennedy, S., and dopamine D(3) receptor antagonists as pharmacotherapeutics for addictions
Mayberg, H. (2004). Modulation of cortical-limbic pathways in major depres- and related disorders. Ann. N. Y. Acad. Sci. 1187, 4–34.
sion: Treatment specific effects of cognitive behavioral therapy compared to
paroxetine. Arch. Gen. Psychiatry 61, 34–41. Herin, D.V., Rush, C.R., and Grabowski, J. (2010). Agonist-like pharmaco-
therapy for stimulant dependence: Preclinical, human laboratory, and clinical
Goldman, D., Oroszi, G., and Ducci, F. (2005). The genetics of addictions: studies. Ann. N. Y. Acad. Sci. 1187, 76–100.
Uncovering the genes. Nat. Rev. Genet. 6, 521–532.
Higgins, S.T., Delaney, D.D., Budney, A.J., Bickel, W.K., Hughes, J.R., Foerg,
Goldstein, R.Z., and Volkow, N.D. (2002). Drug addiction and its underlying F., and Fenwick, J.W. (1991). A behavioral approach to achieving initial
neurobiological basis: Neuroimaging evidence for the involvement of the cocaine abstinence. Am. J. Psychiatry 148, 1218–1224.
frontal cortex. Am. J. Psychiatry 159, 1642–1652.
Highfield, D., Yap, J., Grimm, J.W., Shalev, U., and Shaham, Y. (2001).
Gottesman, I.I., and Gould, T.D. (2003). The endophenotype concept Repeated lofexidine treatment attenuates stress-induced, but not drug
in psychiatry: Etymology and strategic intentions. Am. J. Psychiatry 160, cues-induced reinstatement of a heroin-cocaine mixture (speedball) seeking
636–645. in rats. Neuropsychopharmacology 25, 320–331.

708 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review
Hogarty, G.E., Flesher, S., Ulrich, R., Carter, M., Greenwald, D., Pogue-Geile, Kober, H., Mende-Siedlecki, P., Kross, E.F., Weber, J., Mischel, W., Hart, C.L.,
M., Kechavan, M., Cooley, S., DiBarry, A.L., Garrett, A., et al. (2004). Cognitive and Ochsner, K.N. (2010). Prefrontal-striatal pathway underlies cognitive regu-
enhancement therapy for schizophrenia: Effects of a 2-year randomized trial lation of craving. Proc. Natl. Acad. Sci. USA 107, 14811–14816.
on cognition and behavior. Arch. Gen. Psychiatry 61, 866–876.
Koob, G.F., and Le Moal, M. (2001). Drug addiction, dysregulation of reward,
Holden, C. (2001). ‘Behavioral’ addictions: Do they exist? Science 294, and allostasis. Neuropsychopharmacology 24, 97–129.
980–982.
Koob, G.F., and Le Moal, M. (2005). Plasticity of reward neurocircuitry and the
Holden, C. (2010). Psychiatry. Behavioral addictions debut in proposed ‘dark side’ of drug addiction. Nat. Neurosci. 8, 1442–1444.
DSM-V. Science 327, 935.
Koob, G.F., and Volkow, N.D. (2010). Neurocircuitry of addiction. Neuropsy-
Horder, J., Harmer, C.J., Cowen, P.J., and McCabe, C. (2010). Reduced neural chopharmacology 35, 217–238.
response to reward following 7 days treatment with the cannabinoid CB(1)
antagonist rimonabant in healthy volunteers. Int. J. Neuropsychopharmacol. Koob, G.F., and Zorrilla, E.P. (2010). Neurobiological mechanisms of addic-
13, 1103–1113. tion: Focus on corticotropin-releasing factor. Curr. Opin. Investig. Drugs 11,
63–71.
Hu, J., Henry, S., Gallezot, J.-D., Ropchan, J., Neumaier, J.F., Potenza, M.N.,
Sinha, R., Krystal, J.H., Huang, Y., Ding, Y.-S., et al. (2010). Serotonin 1B Kosten, T., Oliveto, A., Feingold, A., Poling, J., Sevarino, K., McCance-Katz, E.,
receptor imaging in alcohol dependence. Biol. Psychiatry 67, 800–803. Stine, S., Gonzalez, G., and Gonsai, K. (2003). Desipramine and contingency
management for cocaine and opiate dependence in buprenorphine main-
Irvin, J.E., Bowers, C.A., Dunn, M.E., and Wang, M.C. (1999). Efficacy of tained patients. Drug Alcohol Depend. 70, 315–325.
relapse prevention: A meta-analytic review. J. Consult. Clin. Psychol. 67,
563–570. Kosten, T.R., Scanley, B.E., Tucker, K.A., Oliveto, A., Prince, C., Sinha, R.,
Potenza, M.N., Skudlarski, P., and Wexler, B.E. (2006). Cue-induced brain
Janes, A.C., Pizzagalli, D.A., Richardt, S., deB Frederick, B., Chuzi, S., Pachas, activity changes and relapse in cocaine-dependent patients. Neuropsycho-
G., Culhane, M.A., Holmes, A.J., Fava, M., Evins, A.E., and Kaufman, M.J. pharmacology 31, 644–650.
(2010). Brain reactivity to smoking cues prior to smoking cessation predicts
ability to maintain tobacco abstinence. Biol. Psychiatry 67, 722–729. Kreek, M.J., Nielsen, D.A., Butelman, E.R., and LaForge, K.S. (2005). Genetic
influences on impulsivity, risk taking, stress responsivity and vulnerability to
Jentsch, J.D., Aarde, S.M., and Seu, E. (2009). Effects of atomoxetine and drug abuse and addiction. Nat. Neurosci. 8, 1450–1457.
methylphenidate on performance of a lateralized reaction time task in rats.
Psychopharmacology (Berl.) 202, 497–504. Krishnan-Sarin, S., Reynolds, B., Duhig, A.M., Smith, A., Liss, T., McFetridge,
A., Cavallo, D.A., Carroll, K.M., and Potenza, M.N. (2007). Behavioral impul-
Johansson, B.A., Berglund, M., and Lindgren, A. (2006). Efficacy of mainte- sivity predicts treatment outcome in a smoking cessation program for adoles-
nance treatment with naltrexone for opioid dependence: A meta-analytical cent smokers. Drug Alcohol Depend. 88, 79–82.
review. Addiction 101, 491–503.
Krupitsky, E.M., Neznanova, O., Masalov, D., Burakov, A.M., Didenko, T.,
Johnson, B.A., Javors, M.A., Roache, J.D., Seneviratne, C., Bergeson, S.E., Romanova, T., Tsoy, M., Bespalov, A., Slavina, T.Y., Grinenko, A.A., et al.
Ait-Daoud, N., Dawes, M.A., and Ma, J.Z. (2008). Can serotonin transporter (2007). Effect of memantine on cue-induced alcohol craving in recovering
genotype predict serotonergic function, chronicity, and severity of drinking? alcohol-dependent patients. Am. J. Psychiatry 164, 519–523.
Prog. Neuropsychopharmacol. Biol. Psychiatry 32, 209–216.
Lajtha, A., and Sershen, H. (2010). Heterogeneity of reward mechanisms.
Johnson, M.W., Bickel, W.K., and Baker, F. (2007). Moderate drug use and Neurochem. Res. 35, 851–867.
delay discounting: A comparison of heavy, light, and never smokers. Exp.
Clin. Psychopharmacol. 15, 187–194. Lane, S.D., Steinberg, J.L., Ma, L., Hasan, K.M., Kramer, L.A., Zuniga, E.A.,
Narayana, P.A., and Moeller, F.G. (2010). Diffusion tensor imaging and deci-
Johnson, M.W., Bickel, W.K., Baker, F., Moore, B.A., Badger, G.J., and Bud- sion making in cocaine dependence. PLoS ONE 5, e11591.
ney, A.J. (2010). Delay discounting in current and former marijuana-dependent
individuals. Exp. Clin. Psychopharmacol. 18, 99–107. Laucht, M., Treutlein, J., Schmid, B., Blomeyer, D., Becker, K., Buchmann,
A.F., Schmidt, M.H., Esser, G., Jennen-Steinmetz, C., Rietschel, M., et al.
Kable, J.W., and Glimcher, P.W. (2007). The neural correlates of subjective (2009). Impact of psychosocial adversity on alcohol intake in young adults:
value during intertemporal choice. Nat. Neurosci. 10, 1625–1633. Moderation by the LL genotype of the serotonin transporter polymorphism.
Biol. Psychiatry 66, 102–109.
Kampman, K.M., Dackis, C., Lynch, K.G., Pettinati, H., Tirado, C., Gariti, P.,
Sparkman, T., Atzram, M., and O’Brien, C.P. (2006). A double-blind, Lawrence, A.J., Luty, J., Bogdan, N.A., Sahakian, B.J., and Clark, L. (2009).
placebo-controlled trial of amantadine, propranolol, and their combination Problem gamblers share deficits in impulsive decision-making with alcohol-
for the treatment of cocaine dependence in patients with severe cocaine with- dependent individuals. Addiction 104, 1006–1015.
drawal symptoms. Drug Alcohol Depend. 85, 129–137.
Le Moal, M. (2009). Drug abuse: Vulnerability and transition to addiction. Phar-
Kenny, P.J. (2011). Reward mechanisms in obesity: New insights and future macopsychiatry 42 (Suppl 1 ), S42–S55.
directions. Neuron 69, this issue, 664–679.
Lehrer, J. (2009). DON’T! The secret of self-control. The New Yorker, pp. 26-
Kessler, R.C., Chiu, W.T., Demler, O., Merikangas, K.R., and Walters, E.E. 32. http://www.newyorker.com/reporting/2009/05/18/090518fa_fact_lehrer.
(2005). Prevalence, severity, and comorbidity of 12-month DSM-IV disorders
in the National Comorbidity Survey Replication. Arch. Gen. Psychiatry 62, LeSage, M.G., Keyler, D.E., and Pentel, P.R. (2006). Current status of immuno-
617–627. logic approaches to treating tobacco dependence: Vaccines and nicotine-
specific antibodies. AAPS J. 8, E65–E75.
Kim, C.H., Hahn, M.K., Joung, Y., Anderson, S.L., Steele, A.H., Mazei-Robin-
son, M.S., Gizer, I., Teicher, M.H., Cohen, B.M., Robertson, D., et al. (2006). A Lewis, C.M., Baldassarre, A., Committeri, G., Romani, G.L., and Corbetta, M.
polymorphism in the norepinephrine transporter gene alters promoter activity (2009). Learning sculpts the spontaneous activity of the resting human brain.
and is associated with attention-deficit hyperactivity disorder. Proc. Natl. Proc. Natl. Acad. Sci. USA 106, 17558–17563.
Acad. Sci. USA 103, 19164–19169.
Leyton, M., Boileau, I., Benkelfat, C., Diksic, M., Baker, G., and Dagher, A.
Knackstedt, L.A., LaRowe, S., Mardikian, P., Malcolm, R., Upadhyaya, H., (2002). Amphetamine-induced increases in extracellular dopamine, drug
Hedden, S., Markou, A., and Kalivas, P.W. (2009). The role of cystine-gluta- wanting, and novelty seeking: A PET/[11C]raclopride study in healthy men.
mate exchange in nicotine dependence in rats and humans. Biol. Psychiatry Neuropsychopharmacology 27, 1027–1035.
65, 841–845.
Leyton, M., Casey, K.F., Delaney, J.S., Kolivakis, T., and Benkelfat, C. (2005).
Knutson, B., and Greer, S.M. (2008). Anticipatory affect: Neural correlates Cocaine craving, euphoria, and self-administration: A preliminary study of
and consequences for choice. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, the effect of catecholamine precursor depletion. Behav. Neurosci. 119,
3771–3786. 1619–1627.

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 709


Neuron

Review
Li, C.S.R., Milivojevic, V., Kemp, K.A., Hong, K., and Sinha, R. (2006). McClure, S.M., Laibson, D.I., Loewenstein, G., and Cohen, J.D. (2004).
Performance monitoring and stop signal inhibition in abstinent patients with Separate neural systems value immediate and delayed monetary rewards.
cocaine dependence. Drug Alcohol Depend. 85, 205–212. Science 306, 503–507.

Li, N., Lee, B., Liu, R.J., Banasr, M., Dwyer, J.M., Iwata, M., Li, X.Y., Aghaja- McLellan, A.T., Arndt, I.O., Metzger, D.S., Woody, G.E., and O’Brien, C.P.
nian, G., and Duman, R.S. (2010). mTOR-dependent synapse formation (1993). The effects of psychosocial services in substance abuse treatment.
underlies the rapid antidepressant effects of NMDA antagonists. Science JAMA 269, 1953–1959.
329, 959–964.
Meda, S.A., Stevens, M.C., Potenza, M.N., Pittman, B., Gueorguieva, R.,
Liechti, M.E., Lhuillier, L., Kaupmann, K., and Markou, A. (2007). Metabotropic Andrews, M.M., Thomas, A.D., Muska, C., Hylton, J.L., and Pearlson, G.D.
glutamate 2/3 receptors in the ventral tegmental area and the nucleus (2009). Investigating the behavioral and self-report constructs of impulsivity
accumbens shell are involved in behaviors relating to nicotine dependence. domains using principal component analysis. Behav. Pharmacol. 20, 390–399.
J. Neurosci. 27, 9077–9085.
Miller, W.R., and Rollnick, S. (2002). Motivational Interviewing: Preparing
People for Change, Second Edition (New York, NY: Guilford Press).
Lim, K.O., Choi, S.J., Pomara, N., Wolkin, A., and Rotrosen, J.P. (2002).
Reduced frontal white matter integrity in cocaine dependence: A controlled Mischel, W., Shoda, Y., and Rodriguez, M.I. (1989). Delay of gratification in
diffusion tensor imaging study. Biol. Psychiatry 51, 890–895. children. Science 244, 933–938.

Lim, K.O., Wozniak, J.R., Mueller, B.A., Franc, D.T., Specker, S.M., Rodriguez, Moeller, F.G., Barratt, E.S., Dougherty, D.M., Schmitz, J.M., and Swann, A.C.
C.P., Silverman, A.B., and Rotrosen, J.P. (2008). Brain macrostructural and (2001). Psychiatric aspects of impulsivity. Am. J. Psychiatry 158, 1783–1793.
microstructural abnormalities in cocaine dependence. Drug Alcohol Depend.
92, 164–172. Moeller, F.G., Hasan, K.M., Steinberg, J.L., Kramer, L.A., Dougherty, D.M.,
Santos, R.M., Valdes, I., Swann, A.C., Barratt, E.S., and Narayana, P.A.
Longo, M., Wickes, W., Smout, M., Harrison, S., Cahill, S., and White, J.M. (2005). Reduced anterior corpus callosum white matter integrity is related to
(2010). Randomized controlled trial of dexamphetamine maintenance for the increased impulsivity and reduced discriminability in cocaine-dependent
treatment of methamphetamine dependence. Addiction 105, 146–154. subjects: Diffusion tensor imaging. Neuropsychopharmacology 30, 610–617.

Lussier, J.P., Heil, S.H., Mongeon, J.A., Badger, G.J., and Higgins, S.T. (2006). Moeller, F.G., Hasan, K.M., Steinberg, J.L., Kramer, L.A., Valdes, I., Lai, L.Y.,
A meta-analysis of voucher-based reinforcement therapy for substance use Swann, A.C., and Narayana, P.A. (2007). Diffusion tensor imaging eigenvalues:
disorders. Addiction 101, 192–203. Preliminary evidence for altered myelin in cocaine dependence. Psychiatry
Res. 154, 253–258.
Maddux, J.F., and Desmond, D.P. (2000). Addiction or dependence? Addic- Moll, J., Zahn, R., de Oliveira-Souza, R., Krueger, F., and Grafman, J. (2005).
tion 95, 661–665. Opinion: The neural basis of human moral cognition. Nat. Rev. Neurosci. 6,
799–809.
Magill, M., and Ray, L.A. (2009). Cognitive-behavioral treatment with adult
alcohol and illicit drug users: A meta-analysis of randomized controlled trials. Monterosso, J.R., Aron, A.R., Cordova, X., Xu, J., and London, E.D. (2005).
J. Stud. Alcohol Drugs 70, 516–527. Deficits in response inhibition associated with chronic methamphetamine
abuse. Drug Alcohol Depend. 79, 273–277.
Mann, K. (2004). Pharmacotherapy of alcohol dependence: A review of the
clinical data. CNS Drugs 18, 485–504. Naqvi, N.H., and Bechara, A. (2009). The hidden island of addiction: The insula.
Trends Neurosci. 32, 56–67.
Mardikian, P.N., LaRowe, S.D., Hedden, S., Kalivas, P.W., and Malcolm, R.J.
(2007). An open-label trial of N-acetylcysteine for the treatment of cocaine Nasrallah, N.A., Yang, T.W.H., and Bernstein, I.L. (2009). Long-term risk pref-
dependence: A pilot study. Prog. Neuropsychopharmacol. Biol. Psychiatry erence and suboptimal decision making following adolescent alcohol use.
31, 389–394. Proc. Natl. Acad. Sci. USA 106, 17600–17604.

National Institute on Drug Abuse. (2007). Principles of Drug Addiction Treat-


Martell, B.A., Mitchell, E., Poling, J., Gonsai, K., and Kosten, T.R. (2005).
ment: A Research Based Guide (Bethesda, MD: NIDA).
Vaccine pharmacotherapy for the treatment of cocaine dependence. Biol.
Psychiatry 58, 158–164. Nunes, E.V., and Levin, F.R. (2004). Treatment of depression in patients with
alcohol or other drug dependence: A meta-analysis. JAMA 291, 1887–1896.
Martell, B.A., Orson, F.M., Poling, J., Mitchell, E., Rossen, R.D., Gardner, T.,
and Kosten, T.R. (2009). Cocaine vaccine for the treatment of cocaine depen- O’Brien, C.P., Volkow, N., and Li, T.K. (2006). What’s in a word? Addiction
dence in methadone-maintained patients: A randomized, double-blind, versus dependence in DSM-V. Am. J. Psychiatry 163, 764–765.
placebo-controlled efficacy trial. Arch. Gen. Psychiatry 66, 1116–1123.
Olive, M.F. (2010). Cognitive effects of Group I metabotropic glutamate
Martelle, J.L., Claytor, R., Ross, J.T., Reboussin, B.A., Newman, A.H., and receptor ligands in the context of drug addiction. Eur. J. Pharmacol. 639,
Nader, M.A. (2007). Effects of two novel D3-selective compounds, 47–58.
NGB 2904 [N-(4-(4-(2,3-dichlorophenyl)piperazin-1-yl)butyl)-9H-fluorene-2-
carboxamide] and CJB 090 [N-(4-(4-(2,3-dichlorophenyl)piperazin-1-yl) Orson, F.M., Kinsey, B.M., Singh, R.A., Wu, Y., Gardner, T., and Kosten, T.R.
butyl)-4-(pyridin-2-yl)benzamide], on the reinforcing and discriminative stim- (2008). Substance abuse vaccines. Ann. N Y Acad. Sci. 1141, 257–269.
ulus effects of cocaine in rhesus monkeys. J. Pharmacol. Exp. Ther. 321,
573–582. Oslin, D.W., Berrettini, W., Kranzler, H.R., Pettinati, H., Gelernter, J., Volpicelli,
J.R., and O’Brien, C.P. (2003). A functional polymorphism of the mu-opioid
Martı́nez-Raga, J., Knecht, C., and Cepeda, S. (2008). Modafinil: A useful receptor gene is associated with naltrexone response in alcohol-dependent
medication for cocaine addiction? Review of the evidence from neuropharma- patients. Neuropsychopharmacology 28, 1546–1552.
cological, experimental and clinical studies. Curr Drug Abuse Rev 1, 213–221.
Partilla, J.S., Dempsey, A.G., Nagpal, A.S., Blough, B.E., Baumann, M.H., and
Rothman, R.B. (2006). Interaction of amphetamines and related compounds at
Mason, B.J., and Heyser, C.J. (2010). The neurobiology, clinical efficacy and the vesicular monoamine transporter. J. Pharmacol. Exp. Ther. 319, 237–246.
safety of acamprosate in the treatment of alcohol dependence. Expert Opin.
Drug Saf. 9, 177–188. Patterson, F., Jepson, C., Strasser, A.A., Loughead, J., Perkins, K.A., Gur,
R.C., Frey, J.M., Siegel, S., and Lerman, C. (2009). Varenicline improves
Mattick, R.P., Breen, C., Kimber, J., and Davoli, M. (2003). Methadone main- mood and cognition during smoking abstinence. Biol. Psychiatry 65, 144–149.
tenance therapy versus no opioid replacement therapy for opioid dependence.
Cochrane Database Syst. Rev., CD002209. Peirce, J.M., Petry, N.M., Stitzer, M.L., Blaine, J.D., Kellogg, S., Satterfield, F.,
Schwartz, M., Krasnansky, J., Pencer, E., Silva-Vazquez, L., et al. (2006).
Mattick, R.P., Kimber, J., Breen, C., and Davoli, M. (2004). Buprenorphine Effects of lower-cost incentives on stimulant abstinence in methadone main-
maintenance versus placebo or methadone maintenance for opioid depen- tenance treatment: A National Drug Abuse Treatment Clinical Trials Network
dence. Cochrane Database Syst. Rev., CD002207. study. Arch. Gen. Psychiatry 63, 201–208.

710 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.


Neuron

Review
Perry, J.L., and Carroll, M.E. (2008). The role of impulsive behavior in drug Ross D., Kincaid H., Spurrett D., and Collins P., eds. (2009). What is addiction?
abuse. Psychopharmacology (Berl.) 200, 1–26. (Cambridge, MA: MIT Press).

Peterson, B.S., Potenza, M.N., Wang, Z., Zhu, H., Martin, A., Marsh, R., Ples- Rutherford, H.J.V., Mayes, L.C., and Potenza, M.N. (2010). Neurobiology of
sen, K.J., and Yu, S. (2009). An FMRI study of the effects of psychostimulants adolescent substance use disorders: Implications for prevention and treat-
on default-mode processing during Stroop task performance in youths with ment. Child Adolesc. Psychiatr. Clin. N. Am. 19, 479–492.
ADHD. Am. J. Psychiatry 166, 1286–1294.
Saxena, S., Gorbis, E., O’Neill, J., Baker, S.K., Mandelkern, M.A., Maidment,
Petry, N.M. (2001a). Delay discounting of money and alcohol in actively using K.M., Chang, S., Salamon, N., Brody, A.L., Schwartz, J.M., and London,
alcoholics, currently abstinent alcoholics, and controls. Psychopharmacology E.D. (2009). Rapid effects of brief intensive cognitive-behavioral therapy on
(Berl.) 154, 243–250. brain glucose metabolism in obsessive-compulsive disorder. Mol. Psychiatry
14, 197–205.
Petry, N.M. (2001b). Pathological gamblers, with and without substance use
disorders, discount delayed rewards at high rates. J. Abnorm. Psychol. 110, Schilström, B., Ivanov, V.B., Wiker, C., and Svensson, T.H. (2007). Galant-
482–487.
amine enhances dopaminergic neurotransmission in vivo via allosteric poten-
tiation of nicotinic acetylcholine receptors. Neuropsychopharmacology 32,
Petry, N.M. (2005). Pathological Gambling: Etiology, Comorbidity and Treat-
43–53.
ment (Washington, DC: American Psychological Association).

Poling, J., Oliveto, A., Petry, N.M., Sofuoglu, M., Gonsai, K., Gonzalez, G., Schlaug, G., Marchina, S., and Norton, A. (2009). Evidence for plasticity in
Martell, B., and Kosten, T.R. (2006). Six-month trial of bupropion with contin- white-matter tracts of patients with chronic Broca’s aphasia undergoing
gency management for cocaine dependence in a methadone-maintained intense intonation-based speech therapy. Ann. N. Y. Acad. Sci. 1169, 385–394.
population. Arch. Gen. Psychiatry 63, 219–228.
Schmidt, H.D., and Pierce, R.C. (2010). Cocaine-induced neuroadaptations in
Potenza, M.N., and de Wit, H. (2010). Control yourself: Alcohol and impulsivity. glutamate transmission: Potential therapeutic targets for craving and addic-
Alcohol. Clin. Exp. Res. 34, 1303–1305. tion. Ann. N. Y. Acad. Sci. 1187, 35–75.

Potenza, M.N. (2006). Should addictive disorders include non-substance- Schultz, W. (2010). Dopamine signals for reward value and risk: Basic and
related conditions? Addiction 101 (Suppl 1 ), 142–151. recent data. Behav. Brain Funct. 6, 24.

Potenza, M.N. (2007). Impulse control disorders and co-occurring disorders: Schultz, W. (2011). Potential effects of addictive drugs on neuronal reward, risk
Dual diagnosis considerations. J. Dual Diagn. 3, 47–57. and decision mechanisms. Neuron 69, this issue, 603–617.
Potenza, M.N. (2008). Review. The neurobiology of pathological gambling and Seu, E., Lang, A., Rivera, R.J., and Jentsch, J.D. (2009). Inhibition of the norepi-
drug addiction: An overview and new findings. Philos. Trans. R. Soc. Lond. B nephrine transporter improves behavioral flexibility in rats and monkeys.
Biol. Sci. 363, 3181–3189. Psychopharmacology (Berl.) 202, 505–519.
Potenza, M.N. (2009a). Gambling and morality: A neuropsychiatric perspec- Shaffer, H.J. (1999). Strange bedfellows: A critical view of pathological
tive. In Gambling: Mapping the American Moral Landscape, A. Wolfe and gambling and addiction. Addiction 94, 1445–1448.
E.C. Owens, eds. (Waco, TX: Baylor University Press), pp. 175–192.

Potenza, M.N. (2009b). Non-substance and substance addictions. Addiction Shearer, J., Wodak, A., Mattick, R.P., Van Beek, I., Lewis, J., Hall, W., and
104, 1016–1017. Dolan, K. (2001). Pilot randomized controlled study of dexamphetamine
substitution for amphetamine dependence. Addiction 96, 1289–1296.
Potenza, M.N., Bergquist, K.L., Milovojevic, V., Lacadie, C.M., Rajeevan, N.,
Li, C.S.R., and Sinha, R. (2007). Sex differences in the neural correlates of Shearer, J., Wodak, A., van Beek, I., Mattick, R.P., and Lewis, J. (2003). Pilot
stress and craving in cocaine dependence. J. Womens Health (Larchmt) 16, randomized double blind placebo-controlled study of dexamphetamine for
1108. cocaine dependence. Addiction 98, 1137–1141.

Potenza, M.N., Xian, H., Shah, K., Scherrer, J.F., and Eisen, S.A. (2005). Sidhpura, N., Weiss, F., and Martin-Fardon, R. (2010). Effects of the mGlu2/3
Shared genetic contributions to pathological gambling and major depression agonist LY379268 and the mGlu5 antagonist MTEP on ethanol seeking and
in men. Arch. Gen. Psychiatry 62, 1015–1021. reinforcement are differentially altered in rats with a history of ethanol depen-
dence. Biol. Psychiatry 67, 804–811.
Ray, L.A., Hutchison, K.E., MacKillop, J., Miranda, R., Jr., Audette, A., Swift,
R., and Monti, P.M. (2008). Effects of naltrexone during the descending limb Siegle, G.J., Carter, C.S., and Thase, M.E. (2006). Use of FMRI to predict
of the blood alcohol curve. Am. J. Addict. 17, 257–264. recovery from unipolar depression with cognitive behavior therapy. Am.
J. Psychiatry 163, 735–738.
Reissner, K.J., and Kalivas, P.W. (2010). Using glutamate homeostasis as
a target for treating addictive disorders. Behav. Pharmacol. 21, 514–522. Sinha, R. (2008). Chronic stress, drug use, and vulnerability to addiction. Ann.
N. Y. Acad. Sci. 1141, 105–130.
Renthal, W., and Nestler, E.J. (2008). Epigenetic mechanisms in drug addic-
tion. Trends Mol. Med. 14, 341–350. Sinha, R. (2009). Stress and addiction: A dynamic interplay of genes, environ-
ment, and drug intake. Biol. Psychiatry 66, 100–101.
Reynolds, B., Ortengren, A., Richards, J.B., and de Wit, H. (2006). Dimensions
of impulsive behavior: Personality and behavioral measures. Personality and
Sinha, R., Garcia, M., Paliwal, P., Kreek, M.J., and Rounsaville, B.J. (2006).
Individual Differences 40, 305–315.
Stress-induced cocaine craving and hypothalamic-pituitary-adrenal
Ritchey, M., Dolcos, F., Eddington, K.M., Strauman, T.J., and Cabeza, R. responses are predictive of cocaine relapse outcomes. Arch. Gen. Psychiatry
(2010). Neural correlates of emotional processing in depression: Changes 63, 324–331.
with cognitive behavioral therapy and predictors of treatment response.
J. Psychiatr. Res. Sinha, R., Kimmerling, A., Doebrick, C., and Kosten, T.R. (2007). Effects of
lofexidine on stress-induced and cue-induced opioid craving and opioid absti-
Robbins, T.W., and Arnsten, A.F. (2009). The neuropsychopharmacology of nence rates: Preliminary findings. Psychopharmacology (Berl.) 190, 569–574.
fronto-executive function: Monoaminergic modulation. Annu. Rev. Neurosci.
32, 267–287. Sinha, R., Lacadie, C., Skudlarski, P., Fulbright, R.K., Rounsaville, B.J., Kos-
ten, T.R., and Wexler, B.E. (2005). Neural activity associated with stress-
Robinson, T.E., and Berridge, K.C. (1993). The neural basis of drug craving: induced cocaine craving: A functional magnetic resonance imaging study.
An incentive-sensitization theory of addiction. Brain Res. Brain Res. Rev. 18, Psychopharmacology (Berl.) 183, 171–180.
247–291.
Sofuoglu, M., Herman, A.I., Mooney, M., and Waters, A.J. (2009). Varenicline
Robinson, T.E., and Berridge, K.C. (2001). Incentive-sensitization and addic- attenuates some of the subjective and physiological effects of intravenous
tion. Addiction 96, 103–114. nicotine in humans. Psychopharmacology (Berl.) 207, 153–162.

Neuron 69, February 24, 2011 ª2011 Elsevier Inc. 711


Neuron

Review
Somerville, L.H., Jones, R.M., and Casey, B.J. (2010). A time of change: Volkow, N.D., Baler, R.D., and Goldstein, R.Z. (2011). Addiction: Pulling at the
Behavioral and neural correlates of adolescent sensitivity to appetitive and neural threads of society. Neuron 69, this issue, 599–602.
aversive environmental cues. Brain Cogn. 72, 124–133.
Volkow, N.D., Fowler, J.S., Logan, J., Alexoff, D., Zhu, W., Telang, F., Wang,
Srisurapanont, M., and Jarusuraisin, N. (2005). Naltrexone for the treatment of G.J., Jayne, M., Hooker, J.M., Wong, C., et al. (2009). Effects of modafinil on
alcoholism: A meta-analysis of randomized controlled trials. Int. J. Neuropsy- dopamine and dopamine transporters in the male human brain: Clinical impli-
chopharmacol. 8, 267–280. cations. JAMA 301, 1148–1154.

Steeves, T.D.L., Miyasaki, J., Zurowski, M., Lang, A.E., Pellecchia, G., Wagner, F.A., and Anthony, J.C. (2002). From first drug use to drug depen-
Van Eimeren, T., Rusjan, P., Houle, S., and Strafella, A.P. (2009). Increased dence; developmental periods of risk for dependence upon marijuana,
striatal dopamine release in Parkinsonian patients with pathological gambling: cocaine, and alcohol. Neuropsychopharmacology 26, 479–488.
A [11C] raclopride PET study. Brain 132, 1376–1385.
Walker, M., Toneatto, T., Potenza, M.N., Petry, N., Ladouceur, R., Hodgins,
Streeter, C.C., Terhune, D.B., Whitfield, T.H., Gruber, S., Sarid-Segal, O., D.C., el-Guebaly, N., Echeburua, E., and Blaszczynski, A. (2006). A framework
Silveri, M.M., Tzilos, G., Afshar, M., Rouse, E.D., Tian, H., et al. (2008). Perfor- for reporting outcomes in problem gambling treatment research: The Banff,
mance on the Stroop predicts treatment compliance in cocaine-dependent Alberta Consensus. Addiction 101, 504–511.
individuals. Neuropsychopharmacology 33, 827–836.
Wang, G.J., Volkow, N.D., Thanos, P.K., and Fowler, J.S. (2009). Imaging of
Sulzer, D. (2011). How addictive drugs disrupt dopamine neurotransmission. brain dopamine pathways: Implications for understanding obesity. J. Addict.
Neuron 69, this issue, 628–649. Med. 3, 8–18.
Tanabe, J., Thompson, L., Claus, E., Dalwani, M., Hutchison, K., and Banich, Weintraub, D., Koester, J., Potenza, M.N., Siderowf, A.D., Stacy, M.A., Voon,
M.T. (2007). Prefrontal cortex activity is reduced in gambling and nongambling V., Whetteckey, J., Wunderlich, G.R., and Lang, A.E. (2010). Impulse control
substance users during decision-making. Hum. Brain Mapp. 28, 1276–1286. disorders in Parkinson disease: A cross-sectional study of 3090 patients.
Arch. Neurol. 67, 589–595.
Tang, Y.Y., Lu, Q., Geng, X., Stein, E.A., Yang, Y., and Posner, M.I. (2010).
Short-term meditation induces white matter changes in the anterior cingulate. West, R., Baker, C.L., Cappelleri, J.C., and Bushmakin, A.G. (2008). Effect of
Proc. Natl. Acad. Sci. USA 107, 15649–15652. varenicline and bupropion SR on craving, nicotine withdrawal symptoms,
and rewarding effects of smoking during a quit attempt. Psychopharmacology
Tolin, D.F. (2010). Is cognitive-behavioral therapy more effective than other
(Berl.) 197, 371–377.
therapies? A meta-analytic review. Clin. Psychol. Rev. 30, 710–720.

Treadway, M.T., and Zald, D.H. (2011). Reconsidering anhedonia in depres- Wexler, B.E., Gottschalk, C.H., Fulbright, R.F., Prohovnik, I., Lacadie, C.M.,
Rounsaville, B.J., and Gore, J.C. (2001). Functional magnetic resonance
sion: Lessons from translational neuroscience. Neurosci. Biobehav. Rev. 35,
537–555. imaging of cocaine craving. Am. J. Psychiatry 158, 86–95.

Uhl, G.R., Drgon, T., Johnson, C., and Liu, Q.R. (2009). Addiction genetics and Wexler, B.E. (2011). Computerized cognitive remediation treatment for
pleiotropic effects of common haplotypes that make polygenic contributions substance abuse disorders. Biol. Psychiatry 69, 197–198.
to vulnerability to substance dependence. J. Neurogenet. 23, 272–282.
Wrase, J., Schlagenhauf, F., Kienast, T., Wüstenberg, T., Bermpohl, F., Kahnt,
Uhl, G.R., and Grow, R.W. (2004). The burden of complex genetics in brain T., Beck, A., Ströhle, A., Juckel, G., Knutson, B., and Heinz, A. (2007). Dysfunc-
disorders. Arch. Gen. Psychiatry 61, 223–229. tion of reward processing correlates with alcohol craving in detoxified alco-
holics. Neuroimage 35, 787–794.
Vanbuskirk, K.A., and Potenza, M.N. (2010). The treatment of obesity and its
co-occurrence with substance use disorders. J. Addict. Med. 4, 1–10. Xu, J., DeVito, E.E., Worhunsky, P.D., Carroll, K.M., Rounsaville, B.J., and
Potenza, M.N. (2010). White matter integrity is associated with treatment
Verdejo-Garcı́a, A., Lawrence, A.J., and Clark, L. (2008). Impulsivity as a vulner- outcome measures in cocaine dependence. Neuropsychopharmacology 35,
ability marker for substance-use disorders: Review of findings from high-risk 1541–1549.
research, problem gamblers and genetic association studies. Neurosci. Bio-
behav. Rev. 32, 777–810. Zack, M., and Poulos, C.X. (2007). A D2 antagonist enhances the rewarding
and priming effects of a gambling episode in pathological gamblers. Neuro-
Vocci, F.J., Acri, J., and Elkashef, A. (2005). Medication development psychopharmacology 32, 1678–1686.
for addictive disorders: The state of the science. Am. J. Psychiatry 162,
1432–1440. Zhao, Y., Dayas, C.V., Aujla, H., Baptista, M.A., Martin-Fardon, R., and Weiss,
F. (2006). Activation of group II metabotropic glutamate receptors attenuates
Volkow, N.D., and Li, T.K. (2004). Drug addiction: The neurobiology of behav- both stress and cue-induced ethanol-seeking and modulates c-fos expression
iour gone awry. Nat. Rev. Neurosci. 5, 963–970. in the hippocampus and amygdala. J. Neurosci. 26, 9967–9974.

712 Neuron 69, February 24, 2011 ª2011 Elsevier Inc.

S-ar putea să vă placă și