Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 May 1.
Published in final edited form as: Nat Rev Neurosci. 2011 Oct 12;12(11):623–637. doi: 10.1038/nrn3111

Transcriptional and Epigenetic Mechanisms of Addiction

Alfred J Robison 1, Eric J Nestler 1
PMCID: PMC3272277  NIHMSID: NIHMS352346  PMID: 21989194

Preface

Investigations of long-term changes in brain structure and function that accompany chronic exposure to drugs of abuse suggest that alterations in gene regulation contribute importantly to the addictive phenotype. We review multiple mechanisms by which drugs alter the transcriptional potential of genes, from the mobilization or repression of the transcriptional machinery to epigenetics — including alterations in the accessibility of genes within their native chromatin structure and the regulation of gene expression by non-coding RNAs. Increasing evidence implicates these various mechanisms of gene regulation in the lasting changes that drugs of abuse induce in brain, and offer novel inroads for addiction therapy.

Introduction

Drug addiction exacts an enormous medical, financial, and emotional toll on society in the form of overdose and health complications, family disintegration, loss of employment, and crime. NIH’s National Institute on Drug Abuse (NIDA) estimates that the total cost of drug abuse in the U.S. exceeds $600 billion annually, and it is particularly alarming to note a sharp increase in abuse of prescription drugs and in teenage drug abuse in general (www.nida.nih.gov/). These data substantiate the need for increased study of the neuronal effects of drugs of abuse and the mechanisms of addiction in the expectation of uncovering novel targets for treating and preventing addictive disorders.

Although most individuals are exposed to abused drugs, only a subset experience the loss of control over drug use and compulsion for drug seeking and taking that defines the addicted state. Entrance into this state is strongly influenced by both an individual’s genetic constitution and the psychological and social context in which drug exposure occurs13. Although the genetic risk for addiction is roughly 50%, the specific genes involved remain almost completely unknown. The addictive phenotype can persist for the length of an individual’s life, with drug craving and relapse occurring even after decades of abstinence. This persistence suggests that drugs induce long-lasting changes in the brain that underlie addiction behaviors.

The many cells of an individual organism, though they contain essentially identical complements of DNA, differentiate to form distinct tissues and organs through regulated changes in the transcriptional potential of each gene based on environmental cues, cell-to-cell signals, and probably random factors4. It is becoming apparent that many of the same processes of gene regulation involved in this normal differentiation of cells and tissues during development are also engaged in the adult organism to mediate cellular adaptation to environmental stimuli5,6. The processes involved in the regulation of transcriptional potential are varied and highly complex, and include activation and inhibition of transcription factors, modification of chromatin and DNA structure, and induction of non-coding RNAs. Increasing evidence supports the hypothesis that each of these mechanisms of epigenetic regulation is directly affected by drugs of abuse, and that such adaptations are one of the main processes by which drugs induce highly stable changes in the brain that mediate the addicted phenotype. This Review summarizes the findings that support this hypothesis, and highlights areas where future research will extend this fundamental knowledge of addiction and exploit it for new therapeutics.

Drug Action and Gene Transcription

A seemingly equivalent syndrome of addiction can occur with exposure to a wide variety of chemical substances or even rewarding activities, from cocaine to gambling to sex. One common mechanism across these various forms of addiction is thought to be activation of the brain’s reward circuitry, which centers on dopaminergic neurons in the ventral tegmental area (VTA) of the midbrain and their projections to the limbic system, in particular, the nucleus accumbens (NAc; also known as ventral striatum), dorsal striatum, amygdala, hippocampus, and regions of prefrontal cortex (Figure 1)79. This reward circuitry is activated by stimuli or pursuits that promote evolutionary fitness of the organism, like nutrient-rich foods, sex, and social stimulation. As drugs of abuse activate this circuitry far more strongly and persistently than natural rewards, and without association with productive behavioral outcomes, chronic exposure to drugs modulates brain reward regions in part through a homeostatic desensitization that renders the individual unable to attain sufficient feelings of reward in the absence of drug. An alternate, but not mutually exclusive, hypothesis focuses on sensitization, whereby drugs alter the reward circuitry to cause increased assignment of incentive salience to drug cues, effectively making drug-associated environmental stimuli more difficult to ignore and leading to intense drug craving and relapse10. Pathological drug-induced changes in the reward circuitry further impair behavioral control.

Figure 1. Brain reward circuitry.

Figure 1

The brain on the left depicts dopaminergic afferents (light blue arrows) which originate in the ventral tegmental area (blue) and release dopamine in the nucleus accumbens (red) and many other limbic targets. Also shown are other monoaminergic nuclei — the noradrenergic locus coeruleus (green) and serotonergic dorsal raphe (yellow) — which modulate drug reward and other actions. The brain on the right highlights glutamatergic regions that are important for reward: medial prefrontal cortex (green), orbitofrontal cortex (yellow), anterior cingulate cortex (dark blue), thalamus (purple), hippocampus (orange), and amygdala (aqua), all of which send excitatory projections to the nucleus accumbens (red). Drugs of abuse alter this reward circuitry in complex ways, which lead to addiction.

Virtually all rewarding drugs or activities increase dopaminergic transmission from the VTA to the NAc and other target limbic regions, though they each employ partly distinct mechanisms and in some cases involve other neurotransmitter systems as well79. The actions of drugs on the NAc are further complicated by the cellular heterogeneity of this brain region (BOX 1). Although drugs differ in their acute mechanisms of action, the common syndrome of addiction suggests that chronic exposure to these distinct acute mechanisms induces some shared molecular adaptations in brain reward regions that mediate the lasting nature of the addictive phenotype.

Box 1. Cellular Organizaton of Nucleus Accumbens.

The nucleus accumbens (NAc) is composed of multiple neuronal cell types (see the figure), with each cell type apparently exhibiting different transcriptional responses to drugs of abuse and mediating distinct aspects of drug reward and addiction. Glutamatergic afferents from the hippocampus, prefrontal cortex, and amygdala, among other regions, excite all subtypes of NAc neurons36, with such excitation differentially regulating drug reward and motivation, as shown by recent optogenetic experiments130,131. These excitatory inputs are modulated by dopamine afferents from the VTA (ventral tegmental area), and psychostimulant drugs such as cocaine and amphetamine act by directly prolonging the effects of these dopamine signals. Excitatory inputs to the NAc are also modulated by endogenous opioid peptides that are both expressed locally and released by input neurons. Opiate drugs thus act directly on NAc neurons that express opioid receptors; they also promote dopamine release in the NAc indirectly by inhibiting VTA GABAergic interneurons. Cannabinoids also have a role in regulating NAc neurons. They act primarily by locally repressing the function of glutamatergic synapses.

Much work is needed to further parse the cellular specificity of drug action in the NAc. 95% of NAc neurons are GABAergic MSNs (medium spiny neurons), which can be further differentiated into those MSNs that express the D1 dopamine receptor (D1-type MSNs) and express dynorphin and substance P and those that express the D2 dopamine receptor (D2-type MSNs) and express enkephalin132. Drug induction of ΔFosB133,134, and the effects of ΔFosB and G9a on cell morphology and behavior, differ between D1-type and D2-type MSNs135, and neuronal activity of these two cell types causes opposing effects on the rewarding properties of cocaine131. In addition, acute cocaine causes extracellular signal-regulated kinase (ERK)-dependent phosphorylation of MSK1 (mitogen- and stress-activated kinase-1) and of histone 3 specifically in D1-type MSNs75, although the functional consequences of this histone modification are not yet known. In contrast, the effects of cannabinoids seem to predominate at glutamatergic synapses on D2-type MSNs136. About 1–2% of NAc neurons are aspiny large cholinergic interneurons, which have been shown to play an important role in cocaine reward130, and a similar number are GABAergic interneurons, the function of which are less well understood.

Although these studies are important, to date they have barely scratched the surface of what promises to be an important new focus in addiction research: to overlay the alterations in transcriptional potential of genes induced by chronic exposure to drugs onto the map of cellular subtypes in the NAc.

Box 1

We and others have long hypothesized that changes in the transcriptional potential of genes, through the actions of transcription factors, chromatin modifications, and noncoding RNAs, contribute importantly to many of the neuroadaptations that result from chronic exposure to drugs of abuse (Figure 2)11. We know that many mRNAs display altered expression in brain reward regions after chronic drug exposure, which suggests that transcription of individual genes is differentially regulated under these conditions. Over the past ~5 years, studies at the chromatin level have confirmed the involvement of such transcriptional mechanisms in vivo. Moreover, beyond stable changes in steady-state mRNA levels, this work has demonstrated that the “inducibility” of a gene—its ability to be induced or repressed in response to the next drug exposure or some other environmental stimulus—is also altered by chronic drug exposure and that such gene “priming” or “desensitization” is mediated by stable drug-induced changes in the chromatin state around individual genes (Figure 3).

Figure 2. Mechanisms of transcriptional and epigenetic regulation by drugs of abuse.

Figure 2

In eukaryotic cells, DNA is organized by wrapping around histone octomers to form nucleosomes, which are then further organized and condensed to form chromosomes (right). Only by temporarily unraveling compacted chromatin can the DNA of a specific gene be made accessible to the transcriptional machinery. Drugs (left) act through synaptic targets to alter intracellular signaling cascades, which leads to the activation or inhibition of transcription factors and of many other nuclear targets including chromatin regulatory proteins; the detailed mechanisms involved in the latter remain poorly understood. This leads to the induction or repression of particular genes, including those for noncoding RNAs; altered expression of some of these genes can in turn further regulate gene transcription. It is hypothesized that some of these drug-induced changes at the chromatin level are extremely stable and thereby underlie the long-lasting behaviors that define addiction. CREB, cAMP response element binding protein; DNMTs, DNA methyltransferases; HATs, histone acetyltransferases; HDACs, histone deacetylases; HDMs, histone demethylases; HMTs, histone methyltransferases; MEF2, myocyte enhancing factor-2; NFκB, nuclear factor κB; pol II, polymerase II.

Figure 3. Gene priming and desensitization.

Figure 3

In addition to regulating the steady-state expression levels of certain genes, cocaine induces latent effects at many other genes, which alter their inducibility in response to a subsequent stimulus. A. Analysis of mRNA expression after acute or chronic cocaine. Heat maps marked with an asterisk (*) show all genes that are upregulated in the NAc 1 hr after a cocaine challenge in naive animals (acute), in animals treated repeatedly with cocaine (repeated + acute), or in animals after 1 wk of withdrawal from repeated cocaine (repeated wd + acute). Associated heat maps show how the same genes were affected under the other two conditions. Desensitized transcriptional responses after repeated cocaine are indicated (***). B. Early evidence suggests that epigenetic mechanisms are important in mediating such gene priming and desensitization and that many such changes are latent, meaning that they are not reflected by stable changes in steady-state mRNA levels. Rather, such changes alter chromatin structure such that later drug challenge induces a given gene to a greater (primed) or lesser (desensitized) extent based on the epigenetic modifications induced by previous chronic drug exposure. A major goal of current research is to identify the chromatin signatures that underlie gene priming and desensitization.

This transcriptional and epigenetic model of chronic drug action provides a plausible mechanism for how environmental influences during development can increase (or decrease) the risk for addiction later in life. For example, there is mounting evidence that stress during adolescence increases the risk of addiction, and that exposure to drugs in utero increases the risk in adolescence and adulthood12,13. Long-lasting changes in gene transcription or in the potential for transcription that result from early-life stress or drug exposure — mediated at the chromatin level in the absence of genetic differences in primary DNA sequence — might render an adult brain more vulnerable to the addictive process. As alterations in transcriptional potential can last for many years, this model also explains how relapse can occur despite decades of abstinence.

Recent studies of rodent models of addiction have provided considerable support for this hypothesis and have contributed importantly to our understanding of in vivo transcriptional and epigenetic regulation in the brain. Here, we highlight key examples of transcriptional and epigenetic mechanisms of drug action, and identify some of the novel potential targets for therapeutic intervention during the addiction process.

Transcription Factors in Addiction

The classic mechanism for regulation of gene expression is through the actions of transcription factors: proteins that, in response to cell signaling pathways, bind to specific sequences of DNA, generally in the promoter or enhancer regions of target genes, and increase or repress their expression by promoting or blocking the recruitment, respectively, of the RNA polymerase-II transcriptional complex. Transcription factors operate as part of large protein complexes, with their mechanisms of action eventually involving alterations in chromatin structure (see below). Although neurons contain hundreds of transcription factors, studies of adaptations induced by drugs of abuse have focused primarily on a small subset.

ΔFosB

ΔFosB14 is encoded by the fosB gene and shares homology with other Fos family transcription factors. It heterodimerizes with Jun family proteins to form activator protein 1 (AP-1) complexes that bind to AP-1 sites in responsive genes to regulate transcription. There is some evidence from in vitro studies that ΔFosB may also homodimerize15. Although all Fos family proteins are induced transiently by acute drug exposure, chronic administration of virtually any drug of abuse induces the long-lasting expression specifically of ΔFosB14,16,17, a process most robust in the NAc and dorsal striatum, but also seen in several other reward-related brain regions including prefrontal cortex16. ΔFosB induction in the NAc and dorsal striatum by drugs of abuse, whether investigator-administered or self-administered, occurs only in the subtype of medium spiny neuron (MSN) that expresses D1 dopamine receptors (D1-type MSNs)14. ΔFosB — a C-terminal truncation of full-length FosB that is generated by alternative splicing — lacks the two degron domains that are present in the full-length protein and conserved among all other Fos family proteins. This absence results in a four-fold increase in protein stability18. In addition, ΔFosB is phosphorylated in vivo at serine 27, and at several other sites, and this phosphorylation further stabilizes the protein by roughly 10-fold, both in vitro and in vivo19,20. This intrinsic and regulated protein stability is a particularly interesting feature of the molecule, as it provides a molecular mechanism by which drug-induced changes in gene expression can persist for weeks after drug intake stops.

ΔFosB has been linked directly to several addiction-related behaviors. In adult bitransgenic mice where removal of doxycycline induces ΔFosB overexpression specifically in D1-type MSNs of the NAc and dorsal striatum, such induction causes increased locomotor sensitivity to cocaine21, increased conditioned place-preference to cocaine and morphine21,22, and increased cocaine self-administration23. Meanwhile, viral-mediated overexpression studies show that cocaine induction of ΔFosB in orbitofrontal cortex (OFC), a subregion of prefrontal cortex, mediates the ability of chronic cocaine to induce tolerance to the cognition-disrupting effects of acute drug exposure24. Such overexpression also enhances impulsivity during drug withdrawal, and both of these effects further promote drug self-administration24,25. Importantly, genetic or viral overexpression of ΔJunD, a dominant negative mutant of JunD which antagonizes ΔFosB- and other AP-1-mediated transcriptional activity, in the NAc or OFC blocks these key effects of drug exposure14,2224. This indicates that ΔFosB is both necessary and sufficient for many of the changes wrought in the brain by chronic drug exposure. ΔFosB is also induced in D1-type NAc MSNs by chronic consumption of several natural rewards, including sucrose, high fat food, sex, wheel running, where it promotes that consumption14,2630. This implicates ΔFosB in the regulation of natural rewards under normal conditions and perhaps during pathological addictive-like states.

Progress has been made in identifying the broad range of transcriptional targets (some activated, some repressed) through which ΔFosB these produces various behavioral phenotypes in response to drug exposure31,32. By regulating numerous genes related to dendritic spine architecture, including synaptotagmin, microtubule associated proteins, activity-regulated cytoskeleton-associated protein (ARC), actin-related proteins, cyclin-dependent kinase-5 (CDK5), and kinesin3133, among others, ΔFosB is important for the structural plasticity that is induced in NAc by cocaine3436: it is both necessary and sufficient for cocaine-induced increases in the dendritic spine number of NAc MSNs37 (BOX 2). As will be seen below, ΔFosB controls the activity of several other transcriptional and epigenetic regulatory proteins, which then further influence NAc dendritic arborizations, suggesting that ΔFosB serves as one of the master control proteins governing this structural plasticity. ΔFosB also regulates proteins that are important for glutamatergic synaptic function and plasticity, including AMPA receptor subunits21,38 and Ca2+-calmodulin-dependent kinase II (CaMKII)31,39, consistent with the hypothesis that it mediates key aspects of the synaptic plasticity exhibited by MSNs after drug exposure34,40.

Box 2. Epigenetic Regulation and Dendritic Spine Plasticity.

In order for changes in gene transcription and chromatin modifications to affect complex behaviors such as addiction, they must result in some functional output, such as a change in neuronal excitability (intrinsic membrane properties) or connectivity (synapse number or strength). Indeed, it is clear that nearly all drugs of abuse alter the structural connectivity of neurons in the reward circuitry, an effect most evident in changes in the number, shape, and size of dendritic spines on MSNs (medium spiny neurons, as depicted on the left) in the nucleus accumbens (NAc)3436. These changes seem to be behaviorally relevant, as they correlate with behavioral sensitization137, however, certain conditions that increase spine density cause the opposite behavioral effects60,107. Moreover, the nature of these changes varies with the abused substance, time of withdrawal, and method of intake, even within a single brain region. For instance, experimenter-administered cocaine increases the number of thin spines on NAc MSNs during and shortly after chronic exposure, but increases mushroom spines and dendritic complexity during withdrawal34,36. Moreover, opiates and psychostimulants both induce locomotor activity acutely and locomotor and reward sensitization chronically138, whereas morphine consistently reduces NAc MSN spine density and complexity34,35; resolving this discrepancy is an important future research goal. It is also likely that structural plasticity of the NAc plays a role in volition and decision-making, as self-administered drugs generally cause larger changes in spine density than the same doses administered by experimenters35,36. Although the molecular underpinnings of these structural changes remain incompletely understood, several factors that control gene transcription and chromatin regulation have been implicated (as depicted in the example dendrites on the right). These include ΔFosB37, cAMP response element binding protein (CREB)44, myocyte enhancing factor-2 (MEF2)60, G9a37, and DNA methyltransferase-3a (DNMT3a)71, each of which has been linked directly to cocaine regulation of NAc MSN spine density. A key goal is to now identify how these epigenetic factors control cytoskeletal and cytoskeleton-altering genes to regulate spine morphology and consequently changes in neuronal circuitry and addiction-related behaviors.

Box 2

Though ΔFosB is far more stable than all other transcription factors linked to addiction to date, drug relapse can occur after decades of abstinence, a timescale dwarfing even phosphorylated ΔFosB’s prolonged turnover rate. The question of whether ΔFosB can remain stably linked to individual gene promoters for longer periods of time, or whether ΔFosB can induce longer-lasting changes to the chromatin structure of individual genes (see Gene Priming and Desensitization below), to influence relapse behavior long after total cellular levels of the protein return to baseline remains a major focus for the field.

CREB (cAMP response element binding protein)

CREB forms homodimers that can bind to genes at cAMP response elements (CREs), but primarily activates transcription after it has been phosphorylated at Ser133 (by any of several protein kinases), which allows recruitment of CREB-binding protein (CBP) that then promotes transcription (see below)41,42. The mechanism by which CREB activation represses the expression of certain genes is less well understood. Both psychostimulants (cocaine and amphetamine) and opiates increase CREB activity, acutely and chronically — as measured by increased phospho-CREB (pCREB) or reporter gene activity in CRE-LacZ transgenic mice — in multiple brain regions, including the NAc and dorsal striatum4143. Experiments involving the inducible overexpression of CREB or a dominant negative mutant in bitransgenic mice or with viral vectors have shown that CREB induction in the NAc, which occurs in both D1- and D2-type MSNs41, decreases the rewarding effects of cocaine and of opiates44,45, an effect that promotes drug self-administration presumably via negative reinforcement46. In contrast to cocaine and opiates, CREB shows more complicated and varied responses to other drugs of abuse or rewards. For example, chronic nicotine47 or ethanol48,49 administration reduces pCREB levels in the NAc. On the other hand, CREB activity appears necessary for nicotine to establish a place preference50, exposure to Δ9-tetrahydrocannabinol (the active compound in marijuana) increases pCREB in the prefrontal cortex and hippocampus51, and stimuli associated with natural reward increase pCREB in the NAc52. In addition, other CREB family proteins, such as ICER (inducible cAMP repressor) and ATFs (activating transcription factors), have been implicated in the long-term actions of drugs of abuse and require further study53.

CREB activity has been directly linked to the functional activity of NAc MSNs. CREB overexpression increases, whereas dominant-negative CREB decreases, the electrical excitability of MSNs54. Possible differences between D1- and D2-type MSNs have not yet been explored. The observation that viral-mediated overexpression of a K+ channel subunit in the NAc, which decreases MSN excitability, enhances locomotor responses to cocaine suggests that CREB acts as a break on behavioral sensitization to cocaine by upregulating MSN excitability54. Numerous target genes for CREB have been identified that mediate these and other effects on NAc MSNs31,32,42,44,55. Prominent examples include the opioid peptide dynorphin which feeds back and suppresses dopaminergic signaling to the NAc41,44, as well as certain ion channels and glutamate receptor subunits which control NAc excitability54,55. It is interesting to compare these effects of CREB in the NAc to similar data from the locus coeruleus, where CREB has also been found to increase neuronal excitability and thereby mediate aspects of drug tolerance and dependence (BOX 3).

Box 3. Morphine Action in Locus Coeruleus: From Gene Transcription to Neuronal Excitability.

The locus coeruleus (LC) is the major noradrenergic nucleus in the brain, and it has served as a useful model of opiate action11, 139. Acute morphine decreases the firing rate of LC neurons, whereas chronic exposure to the drug allows the rate to return to baseline (a phenomenon known as tolerance), and withdrawal from morphine causes firing rates to increase dramatically over baseline (a phenomenon that is characteristic of dependence and withdrawal) (see traces on left of figure)140,141. Chronic morphine exerts these effects on firing rate in part by upregulating the cAMP–cAMP response element binding protein (CREB) pathway – including induction of adenylyl cyclase type 8 (AC8) and CREB itself, (upward bold green arrows in the figure). As this pathway is acutely inhibited by the drug, cAMP-CREB upregulation can be seen as a classic negative feedback mechanism. These cellular and molecular effects of chronic morphine are independent of synaptic inputs and can be induced by direct activation of opioid receptors on LC neurons in brain slices141. Moreover, the proposed role for CREB in LC, which was based originally on overexpression systems, has been validated more recently by the local knockout of endogenous CREB from LC neurons141. Activation of the cAMP–CREB pathway in LC neurons is behaviorally relevant, in that it contributes to symptoms of physical opiate dependence and withdrawal, which are mediated in part by LC activation. These studies establish the molecular details of a transcriptional mechanism of intrinsic homeostatic plasticity involved in the development of opiate tolerance and dependence, which have provided key insight into the chronic actions of opiates and of other drugs of abuse in several other CNS regions, including those directly related to reward, such as the nucleus accumbens and ventral tegmental area11. PKA, protein kinase A; Rp-cAMP, a competitive inhibitor of cAMP-dependent processes; dnCREB, dominant negative CREB; AC8 KO, AC8 knock-out mouse.

Box 3

NFκB (nuclear factor κB)

NFκB, a transcription factor that is rapidly activated by diverse stimuli, was studied initially for its role in inflammation and immune responses, and linked more recently to synaptic plasticity and memory56. NFκB has been demonstrated to be induced in the NAc by repeated cocaine administration, where it is required for cocaine’s induction of NAc MSN dendritic spines (Box 2) and sensitization to the rewarding effects of the drug57. It has also been associated with nicotine dependence in human populations58. A major goal of current research is to identify the target genes through which NFκB causes this cellular and behavioral plasticity. Interestingly, cocaine induction of NFκB is mediated via ΔFosB14, illustrating complex transcriptional cascades involved in drug action. The role of NFκB in MSN spinogenesis has recently been extended to stress and depression models59, a finding of particular importance considering the comorbidity of depression and addiction, and the well-studied phenomenon of stress-induced relapse to drug abuse.

MEF2 (myocyte enhancing factor-2)

Multiple MEF2 proteins are expressed in brain, including NAc MSNs, where they form homo- and heterodimers that can activate or repress gene transcription depending on the nature of the proteins they recruit (i.e., p300, a coactivator, vs. class II histone deacetylases [HDACs], co-repressors [see below]). Recent work outlines a possible mechanism by which chronic cocaine suppresses striatal MEF2 activity in part through a D1 receptor–cAMP-dependent inhibition of calcineurin, a Ca2+-dependent protein phosphatase60. Cocaine regulation of CDK5, which is also a target for cocaine and ΔFosB as stated earlier33, may be involved as well. This reduction in MEF2 activity is required for cocaine induction of MSN dendritic spine number, but seems to inhibit behavioral sensitization to cocaine60. Although these data suggest that MEF2 plays an important role in the structural and behavioral changes resulting from repeated cocaine administration, they also demonstrate an apparent inconsistency between MSN spine increases and behavioral sensitization to cocaine that merits further study34. Although ethanol has been shown to decrease MEF2 expression in rat cardiomyocytes61, little is known about the effects of other drugs of abuse on MEF2 function in the brain.

Additional Transcription Factors

The transcription factors listed here are the ones most extensively studied in addiction models, however, increasing evidence links several others to drug exposure, including the glucocorticoid receptor, nucleus accumbens 1 transcription factor (NAC1), early growth response factors (EGRs), and signal transducers and activators of transcription (STATs)11,14. For example, glucocorticoid receptor expression is required in dopaminoceptive neurons for cocaine seeking62 but not morphine responses63, and this gene may be associated with initiation of alcohol abuse in teenagers64.

Epigenetics of Addiction

Over the past decade, research into the regulation of transcriptional potential through modification of DNA and chromatin structure has exploded. As it became clear that epigenetic change underlies adaptations in the adult organism, investigations of epigenetic mechanisms have proven fruitful in numerous fields, including drug addiction65,66. Here, we introduce three major mechanisms of epigenetic regulation — histone tail modification, DNA methylation, and microRNAs — and summarize the major findings that have linked each of these mechanisms to addiction.

Histone Tail Modification

Most DNA in eukaryotic cells is densely packed in chromatin, where 147 base pairs (bp) are wrapped around a nucleosome core in ~1.7 superhelical turns67. Nucleosomes are composed of octamers that contain four histone dimers, one each of histones H2A, H2B, H3, and H4, with H1 binding to spans of non-nucleosomal DNA. Numerous types of posttranslational modifications of the N-terminal tails of histones alter chromatin compaction to create more “open” (euchromatin, which is transcriptionally permissive) vs. “closed” (heterochromatin, which is transcriptionally repressive) states68 (Figure 3).

Many residues in the tails of histones are covalently modified in numerous ways, resulting in a complex “code” that is thought to control the accessibility of individual genes to the transcriptional machinery69. Histone acetylation, which negates the positive charge of lysine residues in the histone tail, is associated with transcriptional activation. This process is controlled by histone acetyltransferases (HATs) and HDACs, each of which comprises multiple enzyme classes whose expression and activity are exquisitely regulated67. Histone methylation has been associated with both transcriptional activation and repression depending on the particular residue and the extent of methylation70,71: both lysine and arginine residues can be methylated by several families of histone methyltransferases (HMTs), and this reaction can be reversed by equally diverse histone demethylases. Histone tail modifications also include phosphorylation, ubiquitination, sumoylation, ADP ribosylation, among many others67. The prospect of deciphering the histone code is daunting, given the seemingly infinite number of possible patterns of histone modifications, and the possibility that a particular pattern may have varying meaning depending on the individual gene involved. Nevertheless, new tools are accelerating progress in mapping the epigenetic state of individual gene promoters and the genome as a whole, and future research will determine the feasibility of identifying functionally meaningful chromatin codes72.

Multiple drugs of abuse induce changes in histone acetylation in brain, and evidence has begun to accumulate that these modifications underlie some of the functional abnormalities found in addiction models66,70. First, global (i.e., total cellular) levels of H3 and H4 acetylation are increased in the NAc after acute or chronic exposure to cocaine65,73, and gene promoters that show increased H3 vs. H4 acetylation have been mapped genome-wide32. Despite these global increases, many genes show decreased histone acetylation after chronic cocaine, raising a key question as to what governs gene-specific acetylation changes in the face of global modifications. Another key question concerns the precise intracellular signaling cascades through which cocaine induces changes in histone acetylation — there is some information that such changes may be specific to D1-type MSNs and involve regulation of growth factor-associated kinases74,75. Second, alcohol withdrawal has been demonstrated to increase HDAC activity and reduce histone acetylation in the mouse amygdala76, and the commonly abused inhalant benzyl alcohol regulates potassium channels that are tied to alcohol tolerance via H4 acetylation in Drosophila77. Third, exposure to Δ9-THC, the active ingredient in marijuana, increases HDAC3 in trophoblast cells78. However, this alteration was absent in a genome-wide screen of brain tissue from Δ9-THC-treated mice79, demonstrating that experiments on cell lines can yield effects that are very different from those found in a complex heterogeneous tissue like the brain. These data highlight the need for further research to define the effects of drugs of abuse on histone acetylation in brain in a region- and cell type-specific manner and to identify the specific HAT and HDAC subtypes and intracellular signaling pathways that mediate this regulation in vivo.

Experimental alterations in histone acetylation potently affect addiction-related behaviors. Short-term administration of non-specific HDAC inhibitors, either systemic or intra-NAc, potentiates place conditioning and locomotor responses to psychostimulants and to opiates65,73,80. More prolonged HDAC inhibition has been reported to induce changes in the opposite direction81,82, perhaps through adaptations that oppose initial enzyme inhibition. Studies of specific HDAC isoforms have yielded interesting information: overexpression of HDAC4 or HDAC5 decreases behavioral responses to cocaine73,80, whereas genetic deletion of HDAC5 hypersensitizes mice to the chronic (but not acute) effects of the drug80. Likewise, mutant mice with reduced expression of CBP, a major HAT in brain, exhibit decreased sensitivity to chronic cocaine83. Much additional work is needed to define the influence of specific HAT and HDAC subtypes on addiction-related phenomena.

The potential complexity involved is indicated by recent findings on sirtuins, which are considered Class III HDACs but in reality influence many non-histone proteins. Genome-wide studies of chromatin alterations in the NAc after chronic cocaine revealed upregulation of two sirtuins, SIRT1 and SIRT2. Pharmacological inhibition of sirtuins decreases cocaine place preference and self-administration, whereas activation increases rewarding responses to cocaine32. SIRT1 and SIRT2 induction is associated with increased H3 acetylation and increased ΔFosB binding at their gene promoters32, suggesting that sirtuins are downstream targets of ΔFosB. Work is now needed to identify the proteins that are affected by cocaine-induced regulation of these sirtuins. For example, sirtuins deacetylate several transcription factors such as forkhead box (FoxO) proteins, and serve scaffolding functions by contributing to transcriptional repressive complexes84,85, processes which now warrant study in cocaine models. These findings illustrate the ability of genome-wide efforts to identify fundamentally new mechanisms involved in drug action.

Histone methylation is directly regulated by drugs of abuse as well: global levels of histone 3 lysine 9 dimethylation (H3K9me2) are reduced in the NAc after chronic cocaine37 and a genome-wide screen revealed alterations in H3K9me2 binding on the promoters of numerous genes in this brain region32; both increases and decreases were observed, indicating again that epigenetic modifications at individual genes often defy global changes. The global decrease in H3K9me2 in the NAc is likely mediated by cocaine-induced downregulation of two HMTs, G9a and G9a-like protein (GLP), which catalyze H3K9me237. These adaptations mediate enhanced responsiveness to cocaine, as selective knockout or pharmacological inhibition of G9a in the NAc promotes cocaine-induced behaviors, whereas G9a overexpression has the opposite effect. G9a likewise mediates the ability of cocaine to increase the spine density of NAc MSNs37 (Box 2). Interestingly, there is a functional feedback loop between G9a and ΔFosB: ΔFosB seems to be responsible for cocaine-induced suppression of G9a, and G9a binds to and represses the fosb promoter, such that G9a downregulation may promote the accumulation of ΔFosB observed after chronic cocaine37. In addition, G9a and ΔFosB share many of the same target genes.

Chronic cocaine also downregulates H3K9me3, a mark of heterochromatin, specifically in the NAc and this change is associated with a decrease in the total amount of heterochromatin in NAc MSN nuclei and an increase in the volume of these nuclei86. Genome-wide mapping of H3K9me3 after chronic cocaine indicates that most of the cocaine regulation of this mark occurs at non-genic regions, including at repetitive line elements, which are consequently induced by cocaine86. Although the functional implications of this regulation are not yet known, these findings highlight the profound effects that cocaine exerts on the genome within NAc neurons.

Studies are now needed to examine the actions of other drugs of abuse on these histone endpoints, as well as the effect of drugs on many other types of histone modifications known to regulate eukaryotic gene expression in other systems, in addiction models. Examples include recent, preliminary observations of chronic cocaine regulation of histone arginine methylation and poly-ADP ribosylation, of several families of chromatin remodeling proteins, and of histone variant subunits in the NAc, all of which illustrate the complexity of epigenetic changes associated with drug exposure8790.

Moreover, it will be important to relate drug-induced modifications of histones, occurring at specific drug-regulated genes, with the recruitment of numerous additional proteins that ultimately constitute the transcriptional activation or repression complexes that mediate such regulation. For example, early studies have demonstrated that cocaine induction of CDK5 in the NAc involves a cascade of events which include binding of ΔFosB to the Cdk5 gene promoter, followed by the recruitment of CBP, increased H3 acetylation, and the recruitment of specific chromatin remodeling factors, such as transcription activator BRG173 (Figure 4). Such activation also involves reduced repressive histone methylation at this promoter, which is mediated via cocaine suppression of G9a. In contrast, a very different cascade mediates chronic amphetamine repression of the c-fos gene. Here, ΔFosB binds to the c-fos promoter and recruits HDAC1 and SIRT1, and presumably numerous other proteins91. Also, chronic amphetamine induces increased repressive histone methylation at the c-Fos promoter, perhaps mediated via increased G9a binding37. It is interesting that such increased G9a binding occurs despite the global decrease in G9a expression, once again highlighting gene-specific changes that occur on top of global modifications. Understanding the molecular basis of such gene-specific modifications — e.g., why ΔFosB triggers a cascade of transcriptional activation when it binds to one promoter, but a cascade of transcriptional repression when it binds to another — is a crucial goal of current research. To date, these efforts have been pursued on a protein-by-protein basis, which is experimentally painstaking. A major need in the field is to develop tools to analyze the complete protein complexes recruited to individual genes in concert with drug exposure.

Figure 4. Epigenetic basis of drug regulation of gene expression.

Figure 4

The figure is based on the mechanisms by which chronic cocaine, through ΔFosB, activates the cdk5 gene (top) and represses the c-fos gene (bottom). Top: ΔFosB binds to the cdk5 gene and recruits several co-activators, including CBP (CREB binding protein) — a type of histone acetyltransferase (HAT) leading to increased histone acetylation, transcription factor BRG1 (also known as brahma-related gene 1) — a type of chromatin remodeling factor — and SUG1 (proteasome 26S ATPase subunit 5), another type of chromatin regulatory protein. ΔFosB also represses G9a expression, leading to reduced repressive histone methylation at the cdk5 gene. The net result is gene activation and increased CDK5 expression. Bottom: In contrast, ΔFosB binds to the c-fos gene and recruits several co-repressors, including HDAC1 (histone deacetylase 1) and SIRT 1 (sirtuin 1). The gene also shows increased G9a binding and repressive histone methylation (despite global decreases in these marks). The net result is c-fos gene repression. As transcriptional regulatory complexes contain dozens or hundreds of proteins, much further work is needed to further define the activational and repressive complexes that cocaine recruits to particular genes to mediate their transcriptional regulation and to explore the range of distinct activational and repressive complexes involved in cocaine action.

DNA Methylation

Methylation of DNA occurs at the 5′ position of cytosine nucleotides, with the resulting methyl group projecting into the major groove of the DNA double helix92. In mammals, this occurs almost exclusively in 5′-CpG-3′ sequences and methylation is common throughout the genome — ~3% of all cytosines in human DNA are methylated93 — with proper cytosine methylation required for normal development, genetic imprinting, and X-chromosomal inactivation94. CpG sequences are not evenly dispersed throughout the genome, but rather concentrated in regions termed CpG islands. These are CG-rich regions that overlap with the promoters of 50–60% of human genes and are typically methylated to a much lower extent than CpG dinucleotides found outside of islands95. CpG methylation is catalyzed by a family of enzymes termed DNA methyltransferases (DNMTs), some of which are responsible for maintenance of DNA methyl states whereas others perform de novo CpG methylation92,93. The process of demethylation is less well understood, and may utilize DNA repair mechanisms, such as growth arrest and DNA damage-inducible protein GADD45 (Gadd45)93 and methylcytosine dioxygenase Tet19698. A variant of DNA methylation, 5-hydroxycytosine methylation, also seems to be important in gene regulation99,100 but has not yet been investigated in addiction models.

DNA methylation is generally considered to repress gene transcription through recruitment of corepressor complexes (e.g., HDACs, HMTs) that can sterically hinder the transcriptional machinery or modify nucleosome structure. Such complexes involve several DNA methyl-binding domain proteins (MBDs)93, which are required for normal cell growth and development. Indeed, mutations in methyl CpG binding protein 2 (MeCP2), a prominent MBD, cause the majority of Rett Syndrome cases and are found in a small number of patients with other autism spectrum disorders94.

There are multiple known links between DNA methylation and addiction. Cocaine self-administration increases MeCP2 expression in the NAc101 and dorsal striatum102, and lentiviral knockdown of MeCP2 in the dorsal striatum (but not the NAc) decreases drug intake under extended but not limited access conditions66. Hypomorphic Mecp2 mutant mice show reduced locomotor sensitization and place conditioning with chronic amphetamine103, however, the same study reported that viral knockdown of MeCP2 in the NAc increases amphetamine place conditioning whereas local overexpression decreases this behavioral response104. The reasons for this discrepancy are unclear, but it seems likely that developmental abnormalities in the mutant mice, or the effects of reduced Mecp2 expression in other brain regions, explain these differences. These findings therefore emphasize the importance of utilizing inducible and brain region-specific tools.

Two possible mechanisms for the actions of MeCP2 in drug reward have been proposed. First, a reduction in MeCP2 prevents amphetamine-mediated increases in NAc dendritic spine density while increasing the number of GABAergic synapses103. This is complemented by a GABAergic interneuron-specific increase in MeCP2 phosphorylation in the NAc, which regulates its transcriptional activity and correlates strongly with behavioral sensitization to amphetamine103. An alternative model suggests that MeCP2 represses the transcription of specific microRNAs (see below), resulting in reduced repression of brain-derived neurotrophic factor (BDNF)105, which is also a target for CREB. BDNF has previously been described to promote cocaine self-administration106, consistent with the MeCP2 data. Though these models are not mutually exclusive, further work is necessary to integrate them with our growing understanding of the multiple brain regions and cell types involved in reward behaviors.

A direct link between CpG methylation and addiction involves DNMT3a. Repeated cocaine administration dynamically regulates DNMT3a expression in the mouse NAc, with decreases seen during early phases of withdrawal and sustained increases seen at later time points82,107. Experimental reduction of DNMT3a activity in the adult NAc, achieved either via viral-mediated local knockout in floxed Dnmt3a mice or via local infusion of a DNMT inhibitor, increases behavioral responses to cocaine, whereas DNMT3a overexpression in this region decreases these responses, but also has the paradoxical effect of increasing NAc MSN spine density107, similar to the effects of MEF2 manipulation in this brain region60. A major goal of current research is to identify the specific genes whose methylation status changes in response to chronic cocaine and consequently regulates cellular and behavioral adaptations to the drug.

These observations that chronic cocaine alters DNMT3a and MBDs in the NAc and dorsal striatum raise the possibility that drug-induced changes in DNA methylation might also occur in germ cells and be passed onto to subsequent generations to regulate the propensity of the offspring for addictive behaviors. Such trans-generational transmission of DNA methylation changes and resulting behavioral plasticity remains highly speculative, although recent research has demonstrated robust effects of adult cocaine exposure in rats on cocaine responses in their progeny108.

Gene Priming and Desensitization

Ongoing studies of chromatin regulation in addiction models support the view that epigenetic modifications at individual genes, in addition to underlying stable changes in the steady-state levels of mRNA expression of certain genes, alter the inducibility of many additional genes in response to some subsequent stimulus in the absence of changes in baseline expression levels. Although such studies are still in relatively early stages of development, these types of latent epigenetic changes can be viewed as “molecular scars” that dramatically alter an individual’s adaptability and contribute importantly to the addicted state.

Such priming and desensitization of genes is evident in a recently published microarray study37. Numerous desensitized genes were identified: ~10% of genes whose transcription is induced acutely in the NAc by cocaine are no longer induced by a cocaine challenge after prior chronic exposure to the drug (Figure 3A). Conversely, numerous genes are primed: genes that are not affected by acute cocaine become inducible after a chronic course of cocaine, with ~3-fold more genes being induced in cocaine-experienced animals. The mechanisms underlying such gene desensitization and priming remain incompletely understood; our hypothesis is that epigenetic mechanisms are crucial (Figure 3B). A subset of primed genes show reduced binding of G9a and H3K9me2 at their promoters in the NAc, suggesting the involvement of this epigenetic mark37. Desensitization of the c-fos gene in the NAc, discussed above and depicted in Figure 4, involves stable increases in the binding of ΔFosB, G9a, and related co-repressors, which—although not affecting steady-state levels of c-Fos mRNA—dramatically repress its inducibility to subsequent drug exposure91.

A major need for the field is to now investigate many additional chromatin mechanisms that are recruited by drug exposure to mediate gene priming and desensitization and to understand the detailed mechanisms that target those particular genes. The goal of such studies would be to identify “chromatin signatures” that underlie such long-lasting regulation. The prominence of gene priming and desensitization indicates that studies of steady-state mRNA levels per se would miss important aspects of drug regulation that are not captured at the particular time point examined. For example, the aforementioned microarray study37 measured mRNA levels 1 hr after a cocaine challenge, and preliminary evidence suggests that a partly distinct set of genes show evidence of priming and desensitization at 4 hr. These observations highlight the unique utility of genome-wide assays of chromatin regulation, as such assays would reveal priming and desensitization more globally32.

MicroRNAs

Increasing attention has focused on a variety of non-coding RNAs that are important in biological regulation109. These include microRNAs, which are generally around 22 bp long, are found in all mammalian cells, and are post-translational regulators that bind to complementary sequences on target mRNAs to repress translation and thus silence gene expression. Like histone modifications and DNA methylation, expression of microRNAs can alter the transcriptional potential of a gene in the absence of any change to the DNA sequence, and thus can be considered an epigenetic phenomenon. Several recent studies have implicated microRNAs in addiction behaviors, and miRNAs altered by drugs of abuse have been shown to regulate the expression of many proteins strongly linked to addiction110.

Cocaine self-administration in rats reportedly increases expression of the microRNA miR-212 in striatum, and experimentally increasing miR-212 levels in this region decreases cocaine reward111. The actions of miR-212 depend on upregulation of CREB, which is known to decrease the rewarding effects of cocaine (see above), and more recent work demonstrates that MeCP2 may interact homeostatically with miR-212 to control BDNF expression and cocaine intake105. It has been proposed that this CREB–miR-212–MeCP2–BDNF mechanism is at least partially responsible for cocaine tolerance and escalating intake. miR-124 and miR-181a are also regulated in brain by chronic cocaine, where they are decreased and increased, respectively112. miR-124 overexpression in the NAc reduces cocaine place conditioning, whereas overexpression of miR-181a has the opposite effect113, suggesting that drug regulation of these microRNAs may also act as mechanisms of tolerance and escalating intake. Like miR-212, miR-124 and miR-181a may operate through the CREB–BDNF pathway, as miR-124 overexpression downregulates both of these genes. However, these microRNAs have also been shown to affect the expression of the dopamine transporter, so their mechanisms of action are likely to be complex114. Finally, arginine exporter protein ARGO2 — which is important in microRNA-mediated gene silencing — along with several specific microRNAs have recently been implicated in cocaine regulation of gene expression selectively in the D2 subclass of striatal MSNs115.

Other drugs of abuse have been linked to microRNAs as well. Opioid receptor activation downregulates miR-190 in cultured rat hippocampal neurons in a beta-arrestin2-dependent manner116, and the let-7 family of microRNA precursors is upregulated by chronic morphine exposure in mice117. Interestingly, the μ opioid receptor is itself a direct target for let-7, and the resulting repression of the receptor has been suggested as a novel mechanism for opiate tolerance117. In zebrafish and in cultured immature rat neurons, morphine decreases miR-133b expression, and this might influence dopamine neuron differentiation114. Additionally, both acute and chronic alcohol exposure upregulates miR-9 in cultured striatal neurons, and this may contribute to alcohol tolerance through regulation of large-conductance Ca2+ activated K+ (BK) channels118. miR-9 seems to preferentially downregulate BK channel isoforms that are sensitive to alcohol potentiation, perhaps shifting BK channel expression toward more tolerant subytpes119. miR-9 also targets the D2 dopamine receptor119, and so probably influences alcohol reward.

In the future, next-generation sequencing of microRNAs in several brain regions after exposure to drugs of abuse will be essential to uncover regulation of specific microRNAs and eventually the genes they regulate. Indeed, this process has already begun, as such screens are revealing numerous mcicroRNAs regulated in the NAc after chronic cocaine115,120. For example, cocaine regulation of the miR-8 family suggests novel mechanisms for drug-induced alterations in the neuronal cytoskeletal and synaptic structure120. Exploring this mechanism in drug-induced regulation of NAc dendritic morphology is an important line of future investigation.

Future Directions

This Review has summarized the increasing array of findings that support a role for regulation of the transcriptional potential of myriad genes in the brain’s maladaptations to drugs of abuse. The mechanisms of transcriptional and epigenetic regulation are themselves varied and highly complex, and future studies are needed to catalogue the vast number of regulatory events that occur as well as to understand the precise underlying mechanisms involved. Key questions include: What controls the recruitment or expulsion of individual transcriptional regulatory proteins to a particular target gene? Our hypothesis is that the underlying epigenetic state of that gene is a crucial determining factor, but then what controls the formation and maintenance of distinct epigenetic states at particular genes? Also, what are the intracellular signaling cascades that transduce the initial drug action at the neurotransmitter-receptor level to the neuronal nucleus to regulate the epigenetic state of specific subsets of genes?

The existing literature on transcriptional and epigenetic mechanisms of addiction is limited in several key ways. Most studies to date have employed conditioned place preference and locomotor sensitization paradigms. While these behavioral assays provide useful insight into an animal’s sensitivity to the actions of drugs of abuse on the brain’s reward circuitry, they do not provide direct measures of drug reinforcement or addiction per se. Rather, the field needs to make greater use of drug self-administration and relapse assays, which are considered the best available animal models of addiction121123. Likewise, most studies have utilized experimenter-administered drugs of abuse, even though we know that drugs exert some distinct actions when self-administered or given within a particular environmental context. Work is also needed to move beyond the relatively short time frames of most current experiments to examine transcriptional and epigenetic endpoints after much longer periods of drug exposure and longer periods of withdrawal from drug exposure, as well as to extend what has largely been studies of cocaine action in NAc to studies of several other drugs and several other reward-related regions. Future studies of gene regulation will better inform drug discovery efforts as they increasingly incorporate experimental paradigms that better model human addiction.

Another limitation of the existing literature is the reliance of many studies on overexpression systems, viral or transgenic, which often induce levels of expression far greater than those seen under normal conditions or even after drug treatment. Such overexpression of transcription factors, chromatin regulatory proteins, or their dominant negative mutants can lead to artifactual changes in gene expression and subsequent alterations in cell morphology, physiology, or behavior. It is reassuring that many of the phenomena described here that utilize overexpression systems have been validated with other methods: e.g., those genes regulated by overexpression of ΔFosB in the NAc of inducible bitransgenic mice31 overlap extensively with genes that show enrichment of endogenous ΔFosB binding after cocaine32. Similar caveats exist for the use of constitutive knockout animals, where loss of a gene in early development and in all tissues makes it difficult to interpret any changes observed in drug regulation involving a single brain region of an adult animal. Ultimately, a truly accurate understanding of the transcriptional and epigenetic regulation of the addiction process will require the generation of novel tools that control protein expression with greater spatial, temporal, and accumulation precision.

Methodological advances in epigenetics are needed as well. Current levels of experimental proof of epigenetic mechanisms of drug action have to date involved the overexpression or deletion of a given epigenetic protein (HAT, HDAC, HMT, DNMT, etc.) within a brain region of interest. However, such manipulations affect the epigenetic states of perhaps thousands of genes without targeting those genes that are specifically altered by drug exposure. Being able to experimentally manipulate the epigenetic state of an individual gene within a discrete brain region of an adult animal would represent a major advance for the field. Tools such as artificially designed zinc-finger proteins124 or sequence-specific transcription activator-like effectors (TALEs)125, designed to bind specific DNA sequences in vivo, would offer exciting possibilities for future studies. Similarly, all genome-wide studies of drug-induced epigenetic changes in brain have thus far utilized total extracts of brain regions, even though we know that drugs produce very different effects on distinct neuronal and non-neuronal cell types within a given region. Genome-wide epigenetic analyses in a cell type-specific manner are a critical need in addiction research126.

Advances in bioinformatics are also needed. Genome-wide studies of transcription factor binding and chromatin modifications generate enormous datasets, which require the development of more optimal tools to effectively mine the resulting data. For example, it will be crucial moving forward to overlay such epigenetic analyses with genome-wide changes in RNA expression and to compare data obtained in animal models with those from human postmortem brain tissue. In a similar vein, the studies reviewed here on drug regulation of gene expression must be integrated over several other levels of analysis. How do individual differences in genome sequences relate to individual differences in epigenetic regulation? Do drug-induced epigenetic modifications occur in peripheral tissues such as blood and do any such changes reflect addiction-relevant phenomena? Recent studies, for example, have found altered levels of methylation of the monoamine oxidase-A (MAOA) and MAOB gene promoters in blood of smokers127,128. Additionally, altered methylation of the MAOA gene is associated with nicotine and alcohol dependence in women but not men129, emphasizing the need for studies of sex differences in epigenetic regulation in addiction models, which heretofore have focused almost exclusively on male animals (BOX 4).

Box 4. Sex Differences in Drug Addiction: Epigenetic Mechanisms?

Addiction research has historically neglected female subjects, particularly at the animal level, although both human and animal studies have found robust sex differences in drug responses143,144. In self-administration studies with various drugs, female rats are more responsive in general, and exhibit particularly enhanced responses in the transition phases of acquisition or relapse compared to the maintenance phase145,146. In addition, the locomotor effects of many psychostimulants are greater in female rats147,148. Although, in general, ovariectomy reduces these differences and estrogen administration increases them, this is not true of all drugs of abuse, and some contradictory results have been reported143. These data suggest that drugs of abuse have differential effects on the two sexes, and that the reward system may be different between men and women; clinical evidence bears out these hypotheses. For the most part, women have a later age of onset for substance abuse, although they progress to addiction more rapidly than men149. In the specific case of cocaine, women report shorter periods of abstinence, have greater drug intake, and respond more strongly to cue-induced craving143. These differences may be directly related to the brain’s reward circuitry, as men have been reported to show greater striatal dopamine release than women in response to psychostimulant challenges150. Interestingly, stress upregulates the expression of DNMTs (DNA methyltransferases) and MBDs (DNA methyl-binding domain proteins) in the NAc107; these effects predominate in females and inhibition of DNMT3a in the NAc of female rats increases natural reward151, suggesting that the sexes may undergo differential epigenetic regulation of the reward circuitry. Furthermore, as we know that activation of the reward circuitry by sexual behavior induces ΔFosB27,29,30 and other regulators of transcription, there is little doubt that future studies will reveal further sexual dimorphism in the regulation of transcriptional and epigenetic mechanisms by drugs of abuse, findings which may have important consequences for treatment.

As information on transcriptional and epigenetic mechanisms of addiction accumulates, it is essential to integrate it with equally important information regarding posttranscriptional (translational and posttranslational) regulation to obtain a complete understanding of how chronic exposure to a drug of abuse changes the brain to cause addiction. The ultimate goal of this research is to understand basic principles of neuronal and behavioral adaptation and, ultimately, to identify new targets for the treatment of addictive disorders and new methods for their prevention.

“At-a-Glance” Summary.

  • We hypothesize that changes in the transcriptional potential of genes, through the actions of drug-regulated transcription factors, chromatin modifications, and noncoding RNAs, contribute importantly to the neuroadaptations that underlie addiction. This review highlights key examples of such transcriptional and epigenetic mechanisms of addiction, and identifies some of the novel potential targets for therapeutic intervention during the addiction process.

  • The nucleus accumbens, a region central to the processing of reward and the addicting actions of virtually all drugs of abuse, contains a complex milieu of cell types and receives input from and sends signals to a variety of brain regions. Chronic exposure to drugs of abuse alters gene expression patterns, the morphology, and ultimately the functional activity of nucleus accumbens neurons, which neuroadaptations which contribute importantly to the addiction process.

  • Chronic exposure to drugs of abuse alters the expression or activity of numerous transcription factors, including ΔFosB, CREB (cAMP response element binding protein), NFκB (nuclear factor κB), and MEF2 (myocyte-enhancing factor-2). Manipulation of these factors, specifically in the nucleus accumbens or other parts of the brain’s reward circuitry, alters specific molecular, cellular, and behavioral responses in rodent models of addiction, which defines the functional role of these factors and their target genes in addiction.

  • Epigenetic regulation underlies many adaptations of an adult organism to environmental stimuli, such as seen during drug addiction. Posttranslational modification of histone tails and direct modification of DNA, as well as altered levels or activity of a host of other chromatin remodeling proteins, mediate the ability of drugs of abuse, after chronic exposure, to alter the expression of specific genes in the brain’s reward circuitry.

  • Ongoing studies of chromatin regulation in addiction models support the view that epigenetic changes at individual genes alter not only the steady-state levels of their expression but also their inducibility in response to some subsequent stimulus. We propose that these latent epigenetic changes, termed “gene priming or desensitization,” alter an individual’s adaptability and contribute importantly to the addicted state.

  • Several recent studies have implicated microRNAs in addiction-related behaviors in animal models, and several specific microRNAs, whose expression is altered by drugs of abuse in brain reward regions, have been shown to regulate the expression of several proteins strongly linked to addiction.

  • Key questions include: What controls the recruitment or expulsion of individual transcriptional and chromatin regulatory proteins to a particular target gene? What controls the formation and maintenance of distinct epigenetic states at particular genes? How are the actions of drugs of abuse, all of which initially target the synapse, transduced to the neuronal nucleus to regulate the epigenetic state and transcriptional potential of individual genes?

Glossary

Conditioned place preference

A behavioral test where animals learn to prefer an environment associated with rewarding drug administration. It provides an indirect measure of drug reward

Degron domain

A specific amino acid sequence that indicates the start site for degradation of a protein via proteasomal or other proteolytic processes

Dendritic spine

A small protrusion from a dendrite that is typically associated with synaptic input from a glutamatergic axon at its tip, but which may receive other inputs along its sides or neck

Dependence

Altered physiological state that develops to compensate for persistent drug exposure and that gives rise to a withdrawal syndrome after cessation of drug exposure

DNA methyltransferases (DNMTs)

Enzymes that methylate CpG residues in DNA

Dominant negative mutant

A mutant molecule that forms heteromeric complexes with the wild type to yield a non-functional complex

Histone acetyltransferases (HATs)

Enzymes that catalyze the acetylation of histone N-terminal tails

Histone deacetylases (HDACs)

Enzymes that catalyze the deacetylation of histone N-terminal tails

Histone demethylases (HDMs)

Enzymes that catalyze the demethylation of histone N-terminal tails

Histone methyltransferases (HMTs)

Enzymes that catalyze the methylation of histone N-terminal tails

Hypomorphic

a mutation that causes a wild-type gene product to be produced at a reduced level

Limbic system

A collection of cortical and subcortical structures important for processing memory and emotional information. Prominent structures include the hippocampus and amygdala

Medium spiny neurons (MSNs)

The main cell population of the ventral and dorsal striatum; these GABAergic projection neurons form the two main outputs of these structures, called the direct (D1-type MSNs) and indirect (D2-type MSNs) pathways

Nucleosome

The basic building block of chromatin in which 147 base pairs of DNA are wrapped (~1.65 turns) around a core histone octamer

Self-administration

A form of operant conditioning using a drug as a reward, generally by administration through an intravenous line that is controlled directly by the animal’s actions

Sensitization

Enhanced drug responsiveness with repeated exposure to a constant dose

Sirtuins

Categorized as Class III histone deacetylases, but serve as protein deacetylases for many non-histone proteins as well as part of transcription repressive complexes apparently independent of catalytic activity

Tolerance

Reduced drug responsiveness with repeated exposure to a constant dose

Biographies

Eric J. Nestler, MD, PhD: Eric Nestler is the Nash Family Professor of Neuroscience, Chair of the Department of Neuroscience, and Director of the Friedman Brain Institute at the Mount Sinai School of Medicine. He previously served on the faculties of Yale University (New Haven, Connecticut, USA) and the University of Texas Southwestern Medical Center at Dallas (Texas, USA). He received his BA, PhD, and MD from Yale University, and completed his residency training in psychiatry there. He was elected to the Institute of Medicine in 1998 and to the American Academy of Arts and Sciences in 2005. His research focuses on molecular mechanisms of drug addiction and depression in animal models.

Alfred J. Robison, PhD: A.J. Robison is a postdoctoral fellow in the Department of Neuroscience at the Mount Sinai School of Medicine. He received a BS with Honors from Rhodes College in 1999. In 2005, he received a PhD in Molecular Physiology and Biophysics from Vanderbilt University, where his studies focused on the molecular composition of the synapse mentored by Dr. Roger Colbran. He joined the Nestler lab in 2008; his research there focuses on the molecular mechanisms of brain plasticity as they pertain to learning, memory, and addiction in rodent models.

References

  • 1.Kendler KS, Myers J, Prescott CA. Specificity of genetic and environmental risk factors for symptoms of cannabis, cocaine, alcohol, caffeine, and nicotine dependence. Arch Gen Psychiatry. 2007;64:1313–1320. doi: 10.1001/archpsyc.64.11.1313. 64/11/1313 [pii] [DOI] [PubMed] [Google Scholar]
  • 2.Volkow N, Rutter J, Pollock JD, Shurtleff D, Baler R. One SNP linked to two diseases-addiction and cancer: a double whammy? Nicotine addiction and lung cancer susceptibility. Mol Psychiatry. 2008;13:990–992. doi: 10.1038/mp.2008.71. mp200871 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Goldman D, Oroszi G, Ducci F. The genetics of addictions: uncovering the genes. Nat Rev Genet. 2005;6:521–532. doi: 10.1038/nrg1635. nrg1635 [pii] [DOI] [PubMed] [Google Scholar]
  • 4.Henikoff S, Matzke MA. Exploring and explaining epigenetic effects. Trends Genet. 1997;13:293–295. doi: 10.1016/s0168-9525(97)01219-5. S0168952597012195 [pii] [DOI] [PubMed] [Google Scholar]
  • 5.Sutherland JE, Costa M. Epigenetics and the environment. Ann N Y Acad Sci. 2003;983:151–160. doi: 10.1111/j.1749-6632.2003.tb05970.x. [DOI] [PubMed] [Google Scholar]
  • 6.Fraga MF, et al. Epigenetic differences arise during the lifetime of monozygotic twins. Proc Natl Acad Sci U S A. 2005;102:10604–10609. doi: 10.1073/pnas.0500398102. 0500398102 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Koob GF, Le Moal M. Neurobiology of Addiction. Academic Press; 2005. [Google Scholar]
  • 8.Kalivas PW, Volkow ND. The neural basis of addiction: a pathology of motivation and choice. Am J Psychiatry. 2005;162:1403–1413. doi: 10.1176/appi.ajp.162.8.1403. 162/8/1403 [pii] [DOI] [PubMed] [Google Scholar]
  • 9.Hyman SE, Malenka RC, Nestler EJ. Neural mechanisms of addiction: the role of reward-related learning and memory. Annu Rev Neurosci. 2006;29:565–598. doi: 10.1146/annurev.neuro.29.051605.113009. [DOI] [PubMed] [Google Scholar]
  • 10.Robinson TE, Berridge KC. The neural basis of drug craving: an incentive-sensitization theory of addiction. Brain Res Brain Res Rev. 1993;18:247–291. doi: 10.1016/0165-0173(93)90013-p. [DOI] [PubMed] [Google Scholar]
  • 11.Nestler EJ. Molecular basis of long-term plasticity underlying addiction. Nat Rev Neurosci. 2001;2:119–128. doi: 10.1038/35053570. [DOI] [PubMed] [Google Scholar]
  • 12.Andersen SL, Teicher MH. Desperately driven and no brakes: developmental stress exposure and subsequent risk for substance abuse. Neurosci Biobehav Rev. 2009;33:516–524. doi: 10.1016/j.neubiorev.2008.09.009. S0149-7634(08)00166-8 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Malanga CJ, Kosofsky BE. Does drug abuse beget drug abuse? Behavioral analysis of addiction liability in animal models of prenatal drug exposure. Brain Res Dev Brain Res. 2003;147:47–57. doi: 10.1016/j.devbrainres.2003.09.019. S0165380603003079 [pii] [DOI] [PubMed] [Google Scholar]
  • 14.Nestler EJ. Transcriptional mechanisms of addiction: role of DeltaFosB. Philos Trans R Soc Lond B Biol Sci. 2008;363:3245–3255. doi: 10.1098/rstb.2008.0067. Review. 925643J51371U6N7 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Jorissen HJ, et al. Dimerization and DNA-binding properties of the transcription factor DeltaFosB. Biochemistry. 2007;46:8360–8372. doi: 10.1021/bi700494v. [DOI] [PubMed] [Google Scholar]
  • 16.Perrotti LI, et al. Distinct patterns of DeltaFosB induction in brain by drugs of abuse. Synapse. 2008;62:358–369. doi: 10.1002/syn.20500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Hiroi N, et al. FosB mutant mice: loss of chronic cocaine induction of Fos-related proteins and heightened sensitivity to cocaine’s psychomotor and rewarding effects. Proc Natl Acad Sci U S A. 1997;94:10397–10402. doi: 10.1073/pnas.94.19.10397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Carle TL, et al. Proteasome-dependent and -independent mechanisms for FosB destabilization: identification of FosB degron domains and implications for DeltaFosB stability. Eur J Neurosci. 2007;25:3009–3019. doi: 10.1111/j.1460-9568.2007.05575.x. EJN5575 [pii] [DOI] [PubMed] [Google Scholar]
  • 19.Ulery PG, Rudenko G, Nestler EJ. Regulation of DeltaFosB stability by phosphorylation. J Neurosci. 2006;26:5131–5142. doi: 10.1523/JNEUROSCI.4970-05.2006. 26/19/5131 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ulery-Reynolds PG, Castillo MA, Vialou V, Russo SJ, Nestler EJ. Phosphorylation of DeltaFosB mediates its stability in vivo. Neuroscience. 2009;158:369–372. doi: 10.1016/j.neuroscience.2008.10.059. S0306-4522(08)01596-0 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kelz MB, et al. Expression of the transcription factor deltaFosB in the brain controls sensitivity to cocaine. Nature. 1999;401:272–276. doi: 10.1038/45790. [DOI] [PubMed] [Google Scholar]
  • 22.Zachariou V, et al. An essential role for DeltaFosB in the nucleus accumbens in morphine action. Nat Neurosci. 2006;9:205–211. doi: 10.1038/nn1636. nn1636 [pii] [DOI] [PubMed] [Google Scholar]
  • 23.Colby CR, Whisler K, Steffen C, Nestler EJ, Self DW. Striatal cell type-specific overexpression of DeltaFosB enhances incentive for cocaine. J Neurosci. 2003;23:2488–2493. doi: 10.1523/JNEUROSCI.23-06-02488.2003. 23/6/2488 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Winstanley CA, et al. DeltaFosB induction in orbitofrontal cortex mediates tolerance to cocaine-induced cognitive dysfunction. J Neurosci. 2007;27:10497–10507. doi: 10.1523/JNEUROSCI.2566-07.2007. 27/39/10497 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Winstanley CA, et al. DeltaFosB induction in orbitofrontal cortex potentiates locomotor sensitization despite attenuating the cognitive dysfunction caused by cocaine. Pharmacol Biochem Behav. 2009;93:278–284. doi: 10.1016/j.pbb.2008.12.007. S0091-3057(08)00402-4 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Werme M, et al. Delta FosB regulates wheel running. J Neurosci. 2002;22:8133–8138. doi: 10.1523/JNEUROSCI.22-18-08133.2002. 22/18/8133 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wallace DL, et al. The influence of DeltaFosB in the nucleus accumbens on natural reward-related behavior. J Neurosci. 2008;28:10272–10277. doi: 10.1523/JNEUROSCI.1531-08.2008. 28/41/10272 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Teegarden SL, Nestler EJ, Bale TL. Delta FosB-mediated alterations in dopamine signaling are normalized by a palatable high-fat diet. Biol Psychiatry. 2008;64:941–950. doi: 10.1016/j.biopsych.2008.06.007. S0006-3223(08)00704-X [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hedges VL, Chakravarty S, Nestler EJ, Meisel RL. Delta FosB overexpression in the nucleus accumbens enhances sexual reward in female Syrian hamsters. Genes Brain Behav. 2009;8:442–449. doi: 10.1111/j.1601-183X.2009.00491.x. GBB491 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Pitchers KK, et al. DeltaFosB in the nucleus accumbens is critical for reinforcing effects of sexual reward. Genes Brain Behav. 2010;9:831–840. doi: 10.1111/j.1601-183X.2010.00621.xGBB621. [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.McClung CA, Nestler EJ. Regulation of gene expression and cocaine reward by CREB and DeltaFosB. Nat Neurosci. 2003;6:1208–1215. doi: 10.1038/nn1143nn1143. [pii] [DOI] [PubMed] [Google Scholar]
  • 32.Renthal W, et al. Genome-wide analysis of chromatin regulation by cocaine reveals a role for sirtuins. Neuron. 2009;62:335–348. doi: 10.1016/j.neuron.2009.03.026. S0896-6273(09)00241-4 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Bibb JA, et al. Effects of chronic exposure to cocaine are regulated by the neuronal protein Cdk5. Nature. 2001;410:376–380. doi: 10.1038/3506659135066591. [pii] [DOI] [PubMed] [Google Scholar]
  • 34.Russo SJ, et al. The addicted synapse: mechanisms of synaptic and structural plasticity in nucleus accumbens. Trends Neurosci. 2010;33:267–276. doi: 10.1016/j.tins.2010.02.002. S0166-2236(10)00020-2 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Robinson TE, Kolb B. Structural plasticity associated with exposure to drugs of abuse. Neuropharmacology. 2004;47(Suppl 1):33–46. doi: 10.1016/j.neuropharm.2004.06.025. S0028390804001959 [pii] [DOI] [PubMed] [Google Scholar]
  • 36.Kalivas PW. The glutamate homeostasis hypothesis of addiction. Nat Rev Neurosci. 2009;10:561–572. doi: 10.1038/nrn2515. nrn2515 [pii] [DOI] [PubMed] [Google Scholar]
  • 37.Maze I, et al. Essential role of the histone methyltransferase G9a in cocaine-induced plasticity. Science. 2010;327:213–216. doi: 10.1126/science.1179438. 327/5962/213 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Vialou V, et al. DeltaFosB in brain reward circuits mediates resilience to stress and antidepressant responses. Nat Neurosci. 2010;13:745–752. doi: 10.1038/nn.2551. nn.2551 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Robison AJ, et al. Chronic cocaine engages a feedback loop involving ΔFosB and CaMKII in the nucleus accumbens. Society for Neuroscience Annual Meeting. 2011;909.23 [Google Scholar]
  • 40.Wolf ME, Ferrario CR. AMPA receptor plasticity in the nucleus accumbens after repeated exposure to cocaine. Neurosci Biobehav Rev. 2010;35:185–211. doi: 10.1016/j.neubiorev.2010.01.013. S0149-7634(10)00014-X [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Carlezon WA, Jr, Duman RS, Nestler EJ. The many faces of CREB. Trends Neurosci. 2005;28:436–445. doi: 10.1016/j.tins.2005.06.005. S0166-2236(05)00158-X [pii] [DOI] [PubMed] [Google Scholar]
  • 42.Briand LA, Blendy JA. Molecular and genetic substrates linking stress and addiction. Brain Res. 2010;1314:219–234. doi: 10.1016/j.brainres.2009.11.002. S0006-8993(09)02394-4 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Edwards S, Graham DL, Bachtell RK, Self DW. Region-specific tolerance to cocaine-regulated cAMP-dependent protein phosphorylation following chronic self-administration. Eur J Neurosci. 2007;25:2201–2213. doi: 10.1111/j.1460-9568.2007.05473.x. EJN5473 [pii] [DOI] [PubMed] [Google Scholar]
  • 44.Carlezon WA, Jr, et al. Regulation of cocaine reward by CREB. Science. 1998;282:2272–2275. doi: 10.1126/science.282.5397.2272. [DOI] [PubMed] [Google Scholar]
  • 45.Barrot M, et al. CREB activity in the nucleus accumbens shell controls gating of behavioral responses to emotional stimuli. Proc Natl Acad Sci U S A. 2002;99:11435–11440. doi: 10.1073/pnas.172091899172091899. [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Larson EB, et al. Over-expression of CREB in the nucleus accumbens shell increases cocaine reinforcement in self-administering rats. J Neurosci. 2011 doi: 10.1523/JNEUROSCI.3070-11.2011. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Pluzarev O, Pandey SC. Modulation of CREB expression and phosphorylation in the rat nucleus accumbens during nicotine exposure and withdrawal. J Neurosci Res. 2004;77:884–891. doi: 10.1002/jnr.20216. [DOI] [PubMed] [Google Scholar]
  • 48.Misra K, Roy A, Pandey SC. Effects of voluntary ethanol intake on the expression of Ca(2+)/calmodulin-dependent protein kinase IV and on CREB expression and phosphorylation in the rat nucleus accumbens. Neuroreport. 2001;12:4133–4137. doi: 10.1097/00001756-200112210-00054. [DOI] [PubMed] [Google Scholar]
  • 49.Li J, Li YH, Yuan XR. Changes of phosphorylation of cAMP response element binding protein in rat nucleus accumbens after chronic ethanol intake: naloxone reversal. Acta Pharmacol Sin. 2003;24:930–936. [PubMed] [Google Scholar]
  • 50.Brunzell DH, Mineur YS, Neve RL, Picciotto MR. Nucleus accumbens CREB activity is necessary for nicotine conditioned place preference. Neuropsychopharmacology. 2009;34:1993–2001. doi: 10.1038/npp.2009.11. npp200911 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Rubino T, et al. Cellular mechanisms underlying the anxiolytic effect of low doses of peripheral Delta9-tetrahydrocannabinol in rats. Neuropsychopharmacology. 2007;32:2036–2045. doi: 10.1038/sj.npp.1301330. 1301330 [pii] [DOI] [PubMed] [Google Scholar]
  • 52.Shiflett MW, Mauna JC, Chipman AM, Peet E, Thiels E. Appetitive Pavlovian conditioned stimuli increase CREB phosphorylation in the nucleus accumbens. Neurobiol Learn Mem. 2009;92:451–454. doi: 10.1016/j.nlm.2009.02.010. S1074-7427(09)00061-6 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Green TA, et al. Environmental enrichment produces a behavioral phenotype mediated by low cyclic adenosine monophosphate response element binding (CREB) activity in the nucleus accumbens. Biol Psychiatry. 2010;67:28–35. doi: 10.1016/j.biopsych.2009.06.022. S0006-3223(09)00830-0 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Dong Y, et al. CREB modulates excitability of nucleus accumbens neurons. Nat Neurosci. 2006;9:475–477. doi: 10.1038/nn1661. nn1661 [pii] [DOI] [PubMed] [Google Scholar]
  • 55.Huang YH, et al. CREB modulates the functional output of nucleus accumbens neurons: a critical role of N-methyl-D-aspartate glutamate receptor (NMDAR) receptors. J Biol Chem. 2008;283:2751–2760. doi: 10.1074/jbc.M706578200. M706578200 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Meffert MK, Chang JM, Wiltgen BJ, Fanselow MS, Baltimore D. NF-kappa B functions in synaptic signaling and behavior. Nat Neurosci. 2003;6:1072–1078. doi: 10.1038/nn1110nn1110. [pii] [DOI] [PubMed] [Google Scholar]
  • 57.Russo SJ, et al. Nuclear factor kappa B signaling regulates neuronal morphology and cocaine reward. J Neurosci. 2009;29:3529–3537. doi: 10.1523/JNEUROSCI.6173-08.2009. 29/11/3529 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Sullivan PF, et al. Candidate genes for nicotine dependence via linkage, epistasis, and bioinformatics. Am J Med Genet B Neuropsychiatr Genet. 2004;126B:23–36. doi: 10.1002/ajmg.b.20138. [DOI] [PubMed] [Google Scholar]
  • 59.Christoffel DJ, et al. IkappaB kinase regulates social defeat stress-induced synaptic and behavioral plasticity. J Neurosci. 2011;31:314–321. doi: 10.1523/JNEUROSCI.4763-10.2011. 31/1/314 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Pulipparacharuvil S, et al. Cocaine regulates MEF2 to control synaptic and behavioral plasticity. Neuron. 2008;59:621–633. doi: 10.1016/j.neuron.2008.06.020. S0896-6273(08)00538-2 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Chen L, et al. Chronic ethanol feeding impairs AMPK and MEF2 expression and is associated with GLUT4 decrease in rat myocardium. Exp Mol Med. 2010;42:205–215. doi: 10.3858/emm.2010.42.3.021. emm.2010.42.021 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Ambroggi F, et al. Stress and addiction: glucocorticoid receptor in dopaminoceptive neurons facilitates cocaine seeking. Nat Neurosci. 2009;12:247–249. doi: 10.1038/nn.2282. nn.2282 [pii] [DOI] [PubMed] [Google Scholar]
  • 63.Barik J, et al. Glucocorticoid receptors in dopaminoceptive neurons, key for cocaine, are dispensable for molecular and behavioral morphine responses. Biol Psychiatry. 2010;68:231–239. doi: 10.1016/j.biopsych.2010.03.037. S0006-3223(10)00400-2 [pii] [DOI] [PubMed] [Google Scholar]
  • 64.Desrivieres S, et al. Glucocorticoid receptor (NR3C1) gene polymorphisms and onset of alcohol abuse in adolescents. Addict Biol. 2011;16:510–513. doi: 10.1111/j.1369-1600.2010.00239.xADB239. [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.McQuown SC, Wood MA. Epigenetic regulation in substance use disorders. Curr Psychiatry Rep. 2010;12:145–153. doi: 10.1007/s11920-010-0099-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Renthal W, Nestler EJ. Epigenetic mechanisms in drug addiction. Trends Mol Med. 2008;14:341–350. doi: 10.1016/j.molmed.2008.06.004. S1471-4914(08)00135-4 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Borrelli E, Nestler EJ, Allis CD, Sassone-Corsi P. Decoding the epigenetic language of neuronal plasticity. Neuron. 2008;60:961–974. doi: 10.1016/j.neuron.2008.10.012. S0896-6273(08)00885-4 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Berger SL. The complex language of chromatin regulation during transcription. Nature. 2007;447:407–412. doi: 10.1038/nature05915. nature05915 [pii] [DOI] [PubMed] [Google Scholar]
  • 69.Jenuwein T, Allis CD. Translating the histone code. Science. 2001;293:1074–1080. doi: 10.1126/science.1063127293/5532/1074. [pii] [DOI] [PubMed] [Google Scholar]
  • 70.Maze I, Nestler EJ. The epigenetic landscape of addiction. Ann N Y Acad Sci. 2011;1216:99–113. doi: 10.1111/j.1749-6632.2010.05893.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Su IH, Tarakhovsky A. Lysine methylation and ‘signaling memory’. Curr Opin Immunol. 2006;18:152–157. doi: 10.1016/j.coi.2006.01.012. S0952-7915(06)00010-0 [pii] [DOI] [PubMed] [Google Scholar]
  • 72.Rumbaugh G, Miller CA. Epigenetic changes in the brain: measuring global histone modifications. Methods Mol Biol. 2011;670:263–274. doi: 10.1007/978-1-60761-744-0_18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Kumar A, et al. Chromatin remodeling is a key mechanism underlying cocaine-induced plasticity in striatum. Neuron. 2005;48:303–314. doi: 10.1016/j.neuron.2005.09.023. S0896-6273(05)00790-7 [pii] [DOI] [PubMed] [Google Scholar]
  • 74.Schroeder FA, et al. Drug-induced activation of dopamine D(1) receptor signaling and inhibition of class I/II histone deacetylase induce chromatin remodeling in reward circuitry and modulate cocaine-related behaviors. Neuropsychopharmacology. 2008;33:2981–2992. doi: 10.1038/npp.2008.15. npp200815 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Bertran-Gonzalez J, et al. Opposing patterns of signaling activation in dopamine D1 and D2 receptor-expressing striatal neurons in response to cocaine and haloperidol. J Neurosci. 2008;28:5671–5685. doi: 10.1523/JNEUROSCI.1039-08.2008. 28/22/5671 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Pandey SC, Ugale R, Zhang H, Tang L, Prakash A. Brain chromatin remodeling: a novel mechanism of alcoholism. J Neurosci. 2008;28:3729–3737. doi: 10.1523/JNEUROSCI.5731-07.2008. 28/14/3729 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Wang Y, Krishnan HR, Ghezzi A, Yin JC, Atkinson NS. Drug-induced epigenetic changes produce drug tolerance. PLoS Biol. 2007;5:e265. doi: 10.1371/journal.pbio.0050265. 07-PLBI-RA-0430 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Khare M, Taylor AH, Konje JC, Bell SC. Delta9-tetrahydrocannabinol inhibits cytotrophoblast cell proliferation and modulates gene transcription. Mol Hum Reprod. 2006;12:321–333. doi: 10.1093/molehr/gal036. gal036 [pii] [DOI] [PubMed] [Google Scholar]
  • 79.Parmentier-Batteur S, Jin K, Xie L, Mao XO, Greenberg DA. DNA microarray analysis of cannabinoid signaling in mouse brain in vivo. Mol Pharmacol. 2002;62:828–835. doi: 10.1124/mol.62.4.828. [DOI] [PubMed] [Google Scholar]
  • 80.Renthal W, et al. Histone deacetylase 5 epigenetically controls behavioral adaptations to chronic emotional stimuli. Neuron. 2007;56:517–529. doi: 10.1016/j.neuron.2007.09.032. S0896-6273(07)00766-0 [pii] [DOI] [PubMed] [Google Scholar]
  • 81.Romieu P, et al. Histone deacetylase inhibitors decrease cocaine but not sucrose self-administration in rats. J Neurosci. 2008;28:9342–9348. doi: 10.1523/JNEUROSCI.0379-08.2008. 28/38/9342 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Kim WY, Kim S, Kim JH. Chronic microinjection of valproic acid into the nucleus accumbens attenuates amphetamine-induced locomotor activity. Neurosci Lett. 2008;432:54–57. doi: 10.1016/j.neulet.2007.12.005. S0304-3940(07)01261-X [pii] [DOI] [PubMed] [Google Scholar]
  • 83.Levine AA, et al. CREB-binding protein controls response to cocaine by acetylating histones at the fosB promoter in the mouse striatum. Proc Natl Acad Sci U S A. 2005;102:19186–19191. doi: 10.1073/pnas.0509735102. 0509735102 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Finkel T, Deng CX, Mostoslavsky R. Recent progress in the biology and physiology of sirtuins. Nature. 2009;460:587–591. doi: 10.1038/nature08197. nature08197 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Wang F, Tong Q. SIRT2 suppresses adipocyte differentiation by deacetylating FOXO1 and enhancing FOXO1’s repressive interaction with PPARgamma. Mol Biol Cell. 2009;20:801–808. doi: 10.1091/mbc.E08-06-0647. E08-06-0647 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Maze I, et al. Cocaine dynamically regulates heterochromatin and repetitive element unsilencing in nucleus accumbens. Proc Natl Acad Sci U S A. 2011;108:3035–3040. doi: 10.1073/pnas.1015483108. 1015483108 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Sun H, et al. Cocaine and stress regulates ATPase-containing chromatin remodelers. Society for Neuroscience Annual Meeting. 2011;909.14 [Google Scholar]
  • 88.Damez-Werno D, et al. Histone arginine methylation in the nucleus accumbens in response to chronic cocaine and social stress. Society for Neuroscience Annual Meeting. 2011;909.16 [Google Scholar]
  • 89.Kennedy PJ, et al. Differential Histone H2A variant expression in the Nucleus Accumbens Following Repeated Exposure to Cocaine or Morphine. Society for Neuroscience Annual Meeting. 2011;909.15 [Google Scholar]
  • 90.Scobie K, Damez-Werno D, Sun H, Kennedy PJ, Nestler EJ. Role of poly(ADP-ribosyl)ation in drug-seeking behavior andresiliency to stress. Society for Neuroscience Annual Meeting. 2011;909.18 [Google Scholar]
  • 91.Renthal W, et al. Delta FosB mediates epigenetic desensitization of the c-fos gene after chronic amphetamine exposure. J Neurosci. 2008;28:7344–7349. doi: 10.1523/JNEUROSCI.1043-08.2008. 28/29/7344 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Newell-Price J, Clark AJ, King P. DNA methylation and silencing of gene expression. Trends Endocrinol Metab. 2000;11:142–148. doi: 10.1016/s1043-2760(00)00248-4. S1043-2760(00)00248-4 [pii] [DOI] [PubMed] [Google Scholar]
  • 93.Kim JK, Samaranayake M, Pradhan S. Epigenetic mechanisms in mammals. Cell Mol Life Sci. 2009;66:596–612. doi: 10.1007/s00018-008-8432-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Bird A. The methyl-CpG-binding protein MeCP2 and neurological disease. Biochem Soc Trans. 2008;36:575–583. doi: 10.1042/BST0360575. BST0360575 [pii] [DOI] [PubMed] [Google Scholar]
  • 95.Wang Y, Leung FC. An evaluation of new criteria for CpG islands in the human genome as gene markers. Bioinformatics. 2004;20:1170–1177. doi: 10.1093/bioinformatics/bth059bth059. [pii] [DOI] [PubMed] [Google Scholar]
  • 96.Guo JU, Su Y, Zhong C, Ming GL, Song H. Hydroxylationof 5-methylcytosine by TET1 promotes active DNA demethylation in the adult brain. Cell. 2011;145:423–434. doi: 10.1016/j.cell.2011.03.022. S0092-8674(11)00299-6 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Williams K, et al. TET1 and hydroxymethylcytosine in transcription and DNA methylation fidelity. Nature. 2011;473:343–348. doi: 10.1038/nature10066. nature10066 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Wu H, et al. Dual functions of Tet1 in transcriptional regulation in mouse embryonic stem cells. Nature. 2011;473:389–393. doi: 10.1038/nature09934. nature09934 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Pastor WA, et al. Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature. 2011;473:394–397. doi: 10.1038/nature10102. nature10102 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Ficz G, et al. Dynamic regulation of 5-hydroxymethylcytosine in mouse ES cells and during differentiation. Nature. 2011;473:398–402. doi: 10.1038/nature10008. nature10008 [pii] [DOI] [PubMed] [Google Scholar]
  • 101.Host L, Dietrich JB, Carouge D, Aunis D, Zwiller J. Cocaine self-administration alters the expression of chromatin-remodelling proteins; modulation byhistone deacetylase inhibition. J Psychopharmacol. 2011;25:222–229. doi: 10.1177/0269881109348173. 0269881109348173 [pii] [DOI] [PubMed] [Google Scholar]
  • 102.Cassel S, et al. Fluoxetine and cocaine induce the epigenetic factors MeCP2 and MBD1 in adult rat brain. Mol Pharmacol. 2006;70:487–492. doi: 10.1124/mol.106.022301. mol.106.022301 [pii] [DOI] [PubMed] [Google Scholar]
  • 103.Deng JV, et al. MeCP2 in the nucleus accumbens contributes to neural and behavioral responses to psychostimulants. Nat Neurosci. 2010;13:1128–1136. doi: 10.1038/nn.2614. nn.2614 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Grigoryev YA, et al. Deconvoluting post-transplant immunity: cell subset-specific mapping reveals pathways for activation and expansion of memory T, monocytes and B cells. PLoS One. 2010;5:e13358. doi: 10.1371/journal.pone.0013358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Im HI, Hollander JA, Bali P, Kenny PJ. MeCP2 controls BDNF expression and cocaine intake through homeostatic interactions with microRNA-212. Nat Neurosci. 2010;13:1120–1127. doi: 10.1038/nn.2615. nn.2615 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Graham DL, et al. Dynamic BDNF activity in nucleus accumbens with cocaine use increases self-administration and relapse. Nat Neurosci. 2007;10:1029–1037. doi: 10.1038/nn1929. nn1929 [pii] [DOI] [PubMed] [Google Scholar]
  • 107.LaPlant Q, et al. Dnmt3a regulates emotional behavior and spine plasticity in the nucleus accumbens. Nat Neurosci. 2010;13:1137–1143. doi: 10.1038/nn.2619. nn.2619 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Vassoler FM, White SL, Ortinski PI, Sadri-Vakili G, Pierce RC. Society for Neuroscience Annual Meeting [Google Scholar]
  • 109.Taft RJ, Pang KC, Mercer TR, Dinger M, Mattick JS. Non-coding RNAs: regulators of disease. J Pathol. 2010;220:126–139. doi: 10.1002/path.2638. [DOI] [PubMed] [Google Scholar]
  • 110.Li MD, van der Vaart AD. MicroRNAs in addiction: adaptation’s middlemen? Mol Psychiatry. 2011 doi: 10.1038/mp.2011.58. mp201158 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Hollander JA, et al. Striatal microRNA controls cocaine intake through CREB signalling. Nature. 2010;466:197–202. doi: 10.1038/nature09202. nature09202 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Chandrasekar V, Dreyer JL. microRNAs miR-124, let-7d and miR-181a regulate cocaine-induced plasticity. Mol Cell Neurosci. 2009;42:350–362. doi: 10.1016/j.mcn.2009.08.009. S1044-7431(09)00190-0 [pii] [DOI] [PubMed] [Google Scholar]
  • 113.Chandrasekar V, Dreyer JL. Regulation of MiR-124, Let-7d, and MiR-181a in the accumbens affects the expression, extinction, and reinstatement of cocaine-induced conditioned place preference. Neuropsychopharmacology. 2011;36:1149–1164. doi: 10.1038/npp.2010.250. npp2010250 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Sanchez-Simon FM, Zhang XX, Loh HH, Law PY, Rodriguez RE. Morphine regulates dopaminergic neuron differentiation via miR-133b. Mol Pharmacol. 2010;78:935–942. doi: 10.1124/mol.110.066837. mol.110.066837 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Schaefer A, et al. Argonaute 2 in dopamine 2 receptor-expressing neurons regulates cocaine addiction. J Exp Med. 2010;207:1843–1851. doi: 10.1084/jem.20100451. jem.20100451 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Zheng H, et al. mu-Opioid receptor agonists differentially regulate the expression of miR-190 and NeuroD. Mol Pharmacol. 2010;77:102–109. doi: 10.1124/mol.109.060848. mol.109.060848 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.He Y, Yang C, Kirkmire CM, Wang ZJ. Regulation of opioid tolerance by let-7 family microRNA targeting the mu opioid receptor. J Neurosci. 2010;30:10251–10258. doi: 10.1523/JNEUROSCI.2419-10.2010. 30/30/10251 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Pietrzykowski AZ, et al. Posttranscriptional regulation of BK channel splice variant stability by miR-9 underlies neuroadaptation to alcohol. Neuron. 2008;59:274–287. doi: 10.1016/j.neuron.2008.05.032. S0896-6273(08)00529-1 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Pietrzykowski AZ. The role of microRNAs in drug addiction: a big lesson from tiny molecules. Int Rev Neurobiol. 2010;91:1–24. doi: 10.1016/S0074-7742(10)91001-5. S0074-7742(10)91001-5 [pii] [DOI] [PubMed] [Google Scholar]
  • 120.Eipper-Mains JE, et al. microRNA-Seq reveals cocaine-regulated expression of striatal microRNAs. RNA. 2011 doi: 10.1261/rna.2775511. rna.2775511 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Pelloux Y, Everitt BJ, Dickinson A. Compulsive drug seeking by rats under punishment: effects of drug taking history. Psychopharmacology (Berl) 2007;194:127–137. doi: 10.1007/s00213-007-0805-0. [DOI] [PubMed] [Google Scholar]
  • 122.Pickens CL, et al. Neurobiology of the incubation of drug craving. Trends Neurosci. 2011;34:411–420. doi: 10.1016/j.tins.2011.06.001. S0166-2236(11)00089-0 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.O’Connor EC, Chapman K, Butler P, Mead AN. The predictive validity of the rat self-administration model for abuse liability. Neurosci Biobehav Rev. 2011;35:912–938. doi: 10.1016/j.neubiorev.2010.10.012. S0149-7634(10)00176-4 [pii] [DOI] [PubMed] [Google Scholar]
  • 124.Laganiere J, et al. An engineered zinc finger protein activator of the endogenous glial cell line-derived neurotrophic factor gene provides functional neuroprotection in a rat model of Parkinson’s disease. J Neurosci. 2010;30:16469–16474. doi: 10.1523/JNEUROSCI.2440-10.2010. 30/49/16469 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Zhang F, et al. Efficient construction of sequence-specific TAL effectors for modulating mammalian transcription. Nat Biotechnol. 2011;29:149–153. doi: 10.1038/nbt.1775. nbt.1775 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Cheung I, et al. Developmental regulation and individual differences of neuronal H3K4me3 epigenomes in the prefrontal cortex. Proc Natl Acad Sci U S A. 2010;107:8824–8829. doi: 10.1073/pnas.1001702107. 1001702107 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Philibert RA, et al. The effect of smoking on MAOA promoter methylation in DNA prepared from lymphoblasts and whole blood. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:619–628. doi: 10.1002/ajmg.b.31031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Launay JM, et al. Smoking induces long-lasting effects through a monoamine-oxidase epigenetic regulation. PLoS One. 2009;4:e7959. doi: 10.1371/journal.pone.0007959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Philibert RA, Gunter TD, Beach SR, Brody GH, Madan A. MAOA methylation is associated with nicotine and alcohol dependence in women. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:565–570. doi: 10.1002/ajmg.b.30778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Witten IB, et al. Cholinergic interneurons control local circuit activity and cocaine conditioning. Science. 2010;330:1677–1681. doi: 10.1126/science.1193771. 330/6011/1677 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Lobo MK, et al. Cell type-specific loss of BDNF signaling mimics optogenetic control of cocaine reward. Science. 2010;330:385–390. doi: 10.1126/science.1188472. 330/6002/385 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Self DW. In: The Dopamine Receptors. Neve KA, editor. Vol. 2010. Humana Press; 2010. pp. 479–524. [Google Scholar]
  • 133.Lee KW, et al. Cocaine-induced dendritic spine formation in D1 and D2 dopamine receptor-containing medium spiny neurons in nucleus accumbens. Proc Natl Acad Sci U S A. 2006;103:3399–3404. doi: 10.1073/pnas.0511244103. 0511244103 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Nye HE, Hope BT, Kelz MB, Iadarola M, Nestler EJ. Pharmacological studiesof the regulation of chronic FOS-related antigen induction by cocaine in the striatum and nucleus accumbens. J Pharmacol Exp Ther. 1995;275:1671–1680. [PubMed] [Google Scholar]
  • 135.Maze I, et al. G9a regulates cocaine-induced behavioral and transcriptional plasticity in a cell-type specific manner. Society for Neuroscience. 2010;574.7 [Google Scholar]
  • 136.Singla S, Kreitzer AC, Malenka RC. Mechanisms for synapse specificity during striatal long-term depression. J Neurosci. 2007;27:5260–5264. doi: 10.1523/JNEUROSCI.0018-07.2007. 27/19/5260 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Li Y, Acerbo MJ, Robinson TE. The induction of behavioural sensitization is associated with cocaine-induced structural plasticity in the core (but not shell) of the nucleus accumbens. Eur J Neurosci. 2004;20:1647–1654. doi: 10.1111/j.1460-9568.2004.03612.xEJN3612. [pii] [DOI] [PubMed] [Google Scholar]
  • 138.Russo SJ, Mazei-Robison MS, Ables JL, Nestler EJ. Neurotrophic factors and structural plasticity in addiction. Neuropharmacology. 2009;56(Suppl 1):73–82. doi: 10.1016/j.neuropharm.2008.06.059. S0028-3908(08)00214-1 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Van Bockstaele EJ, Reyes BA, Valentino RJ. The locus coeruleus: A key nucleus where stress and opioids intersect to mediate vulnerability to opiate abuse. Brain Res. 2010;1314:162–174. doi: 10.1016/j.brainres.2009.09.036. S0006-8993(09)01939-8 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Han MH, et al. Role of cAMP response element-binding protein in the rat locus ceruleus: regulation of neuronal activity and opiate withdrawal behaviors. J Neurosci. 2006;26:4624–4629. doi: 10.1523/JNEUROSCI.4701-05.2006. 26/17/4624 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Cao JL, et al. Essential role of the cAMP-cAMP response-element binding protein pathway in opiate-induced homeostatic adaptations of locus coeruleus neurons. Proc Natl Acad Sci U S A. 2010;107:17011–17016. doi: 10.1073/pnas.1010077107. 1010077107 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Nestler EJ, Aghajanian GK. Molecular and cellular basis of addiction. Science. 1997;278:58–63. doi: 10.1126/science.278.5335.58. [DOI] [PubMed] [Google Scholar]
  • 143.Fattore L, Altea S, Fratta W. Sex differences in drug addiction: a review of animal and human studies. Womens Health (Lond Engl) 2008;4:51–65. doi: 10.2217/17455057.4.1.51. [DOI] [PubMed] [Google Scholar]
  • 144.Carroll ME, Anker JJ. Sex differences and ovarian hormones in animal models of drug dependence. Horm Behav. 2010;58:44–56. doi: 10.1016/j.yhbeh.2009.10.001. S0018-506X(09)00221-9 [pii] [DOI] [PubMed] [Google Scholar]
  • 145.Lynch WJ, Carroll ME. Sex differences in the acquisition of intravenously self-administered cocaine and heroin in rats. Psychopharmacology (Berl) 1999;144:77–82. doi: 10.1007/s002130050979. [DOI] [PubMed] [Google Scholar]
  • 146.Roth ME, Carroll ME. Sex differences in the escalation of intravenous cocaine intake following long-or short-access to cocaine self-administration. Pharmacol Biochem Behav. 2004;78:199–207. doi: 10.1016/j.pbb.2004.03.018S0091305704001029. [pii] [DOI] [PubMed] [Google Scholar]
  • 147.Cailhol S, Mormede P. Strain and sex differences in the locomotor response and behavioral sensitization to cocaine in hyperactive rats. Brain Res. 1999;842:200–205. doi: 10.1016/s0006-8993(99)01742-4. S0006-8993(99)01742-4 [pii] [DOI] [PubMed] [Google Scholar]
  • 148.Robinson TE, Becker JB, Presty SK. Long-term facilitation of amphetamine-induced rotational behaviorand striatal dopamine release produced by a single exposure to amphetamine: sex differences. Brain Res. 1982;253:231–241. doi: 10.1016/0006-8993(82)90690-4. 0006-8993(82)90690-4 [pii] [DOI] [PubMed] [Google Scholar]
  • 149.Hernandez-Avila CA, Rounsaville BJ, Kranzler HR. Opioid-, cannabis-and alcohol-dependent women show more rapid progression to substance abuse treatment. Drug Alcohol Depend. 2004;74:265–272. doi: 10.1016/j.drugalcdep.2004.02.001S0376871604000304. [pii] [DOI] [PubMed] [Google Scholar]
  • 150.Munro CA, et al. Sex differences in striatal dopamine release in healthy adults. Biol Psychiatry. 2006;59:966–974. doi: 10.1016/j.biopsych.2006.01.008. S0006-3223(06)00133-8 [pii] [DOI] [PubMed] [Google Scholar]
  • 151.Hodes GE, Christoffel DJ, Golden SA, Ahn HF, Russo SJ. Sex differences in epigenetic regulation of stress-related disorders. Society for Neuroscience Annual Meeting. 2011;219 [Google Scholar]

RESOURCES