Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Prog Neurobiol. Author manuscript; available in PMC 2014 Jan 1.
Published in final edited form as:
PMCID: PMC3525776
NIHMSID: NIHMS419755
PMID: 23085425

Addiction-Related Gene Regulation: Risks of Exposure to Cognitive Enhancers vs. Other Psychostimulants

Abstract

The psychostimulants methylphenidate (Ritalin, Concerta), amphetamine (Adderall), and modafinil (Provigil) are widely used in the treatment of medical conditions such as attention-deficit hyperactivity disorder and narcolepsy and, increasingly, as “cognitive enhancers” by healthy people. The long-term neuronal effects of these drugs, however, are poorly understood. A substantial amount of research over the past 2 decades has investigated the effects of psychostimulants such as cocaine and amphetamines on gene regulation in the brain because these molecular changes are considered critical for psychostimulant addiction. This work has determined in some detail the neurochemical and cellular mechanisms that mediate psychostimulant-induced gene regulation and has also identified the neuronal systems altered by these drugs. Among the most affected brain systems are corticostriatal circuits, which are part of cortico-basal ganglia-cortical loops that mediate motivated behavior. The neurotransmitters critical for such gene regulation are dopamine in interaction with glutamate, while other neurotransmitters (e.g., serotonin) play modulatory roles. This review presents (1) an overview of the main findings on cocaine- and amphetamine-induced gene regulation in corticostriatal circuits in an effort to provide a cellular framework for (2) an assessment of the molecular changes produced by methylphenidate, medical amphetamine (Adderall), and modafinil. The findings lead to the conclusion that protracted exposure to these cognitive enhancers can induce gene regulation effects in corticostriatal circuits that are qualitatively similar to those of cocaine and other amphetamines. These neuronal changes may contribute to the addiction liability of the psychostimulant cognitive enhancers.

Keywords: amphetamine, cocaine, cognitive enhancer, cortex, dopamine, gene regulation, methylphenidate, psychostimulant, striatum

1. Introduction

Cognitive enhancers, sometimes called “smart drugs” or “memory enhancers,” are substances taken with the expectation that they increase mental functions such as attention, concentration, alertness, memory, motivation, planning, and decision making (Svetlov et al., 2007; Lanni et al., 2008; Husain and Mehta, 2011). The most widely used cognitive enhancers include the psychostimulant medications methylphenidate (Ritalin, Concerta), amphetamine (Adderall), and modafinil (Provigil). The oldest of these drugs is amphetamine, which was first synthesized in 1887 and has been used in the clinic since the 1930s (Berman et al., 2009). Methylphenidate, first produced in 1944, has also been used as a medication for many decades (Leonard et al., 2004), whereas modafinil was introduced only in the early 1990s (Minzenberg and Carter, 2008).

1.1. Medical and nonprescription uses of psychostimulants

These psychostimulant medications are valued and widely prescribed for their efficacy in controlling symptoms of attention-deficit hyperactivity disorder (ADHD) (methylphenidate, amphetamine) or excessive daytime sleepiness associated with narcolepsy and other sleep disorders (modafinil, amphetamine, methylphenidate) (Leonard et al., 2004; Minzenberg and Carter, 2008; Berman et al., 2009). Furthermore, there is a rationale for the use of methylphenidate and modafinil in the treatment of behavioral deficits associated with psychostimulant addiction (e.g., Goldstein et al., 2010; Volkow et al., 2010; Loland et al., 2012; Reichel and See, 2012; for reviews, see Brady et al., 2011; Sofuoglu et al., 2012). But these medications are also recognized by the US Drug Enforcement Administration (DEA) for their abuse potential and are therefore classified as Schedule II (amphetamine, methylphenidate) or Schedule IV (modafinil) controlled substances.

ADHD is among the most common neurobehavioral disorders and, in the United States, affects approximately 7.8% of children aged 4 to 17 and 4.4% of adults (Kollins, 2008). It is arguably the dramatic increase in diagnosis and pharmacological treatment of ADHD over the past 2 decades that has led to a parallel increase in production of these psychostimulants (Biederman et al., 2007; Kollins, 2008; Swanson and Volkow, 2008; Berman et al., 2009). For example, the number of prescriptions for amphetamine increased 16-fold during the 1990s, and in 2000 the US annual production of amphetamine reached 30,000 kg (Berman et al., 2009). Likewise, the DEA aggregate production quotas for methylphenidate increased from 5,000 kg in 1993 to 15,000 kg in 2000 to 50,000 kg in 2009.1

Not surprisingly, with increasing availability came increasing diversion and use of psychostimulant medications as cognitive enhancers or party drugs (Kollins et al., 2001; Svetlov et al., 2007; Kollins, 2008; Swanson and Volkow, 2008; Wilens et al., 2008; Mache et al., 2012). It is difficult to accurately estimate the amounts and extent of their use as cognitive enhancers, but surveys indicate that students are frequent consumers of these drugs (Svetlov et al., 2007). Thus, studies of the misuse and diversion of prescription ADHD medications found that rates of self-reported past-year use range from 4% to 30% among college students (Kollins et al., 2001; Kollins, 2008; Wilens et al., 2008; Berman et al., 2009) and 5% to 9% in grade school– and high school–aged children (Wilens et al., 2008). The most often reported motives for illicit use among college students are to increase attention, concentration, or alertness (to help study), and to a lesser extent to get “high” (Babcock and Byrne, 2000; Teter et al., 2006; White et al., 2006). Furthermore, the trend for using cognitive enhancers is growing not only among students (Greely et al., 2008): in a recent poll of 1,400 academics (Nature readers), 20% indicated that they had used methylphenidate (62%) or modafinil (44%) to combat jet lag, to improve general concentration, or to assist them in a particular task (Maher, 2008).

1.2. Neurobehavioral and molecular impacts of psychostimulant use

It remains controversial whether the medical use of psychostimulants is completely safe (Kollins, 2008; Wilens et al., 2008; see Section 8), especially in children and adolescents (Carlezon and Konradi, 2004; Andersen, 2005; Berman et al., 2009). Even less clear are the potential long-term effects of cognitive enhancer use in the healthy, in part because widespread use is a relatively new phenomenon and adverse effects of early drug exposure may appear only late in life (e.g., Bolanos et al., 2003; Tropea et al., 2008; Warren et al., 2011). There is concern especially that long-term exposure to psychostimulants during the sensitive period of brain development may increase the risk for maladaptive neurobehavioral changes that may facilitate drug addiction and other neuropsychiatric disorders (Bolanos et al., 2003; Warren et al., 2011; for reviews, see Carlezon and Konradi, 2004; Andersen, 2005; Carrey and Wilkinson, 2011; Marco et al., 2011).

There is little doubt that changes in gene regulation produced by illicit psychostimulants such as cocaine play a critical role in addiction (Hyman and Nestler, 1996; Nestler, 2001) because only these molecular changes endure long enough to mediate behavioral pathologies that can last a lifetime such as addiction (Renthal and Nestler, 2008). Therefore, the addiction liability of medical psychostimulants most likely also rests on their propensity to induce altered gene regulation.

In this article, we review changes in gene regulation produced by medical amphetamine (Adderall), methylphenidate, and modafinil as determined in animal models. We discuss these findings in the context of the known molecular changes that are produced by illicit psychostimulants (e.g., cocaine) and are considered part of the molecular basis for drug addiction.

2. Neurochemical effects of psychostimulant cognitive enhancers

2.1. Changes in monoamine transmission

Psychostimulants cause, among other effects, amplification of monoamine neurotransmission by promoting release and/or blocking reuptake of monoamines and thus prolonging their actions (Natarajan and Yamamoto, 2011; Sulzer, 2011). Psychostimulant-induced potentiation of the dopamine transmission (Di Chiara and Imperato, 1988) is considered critical to the addiction process, whereas serotonin and norepinephrine play modulatory roles (Berke and Hyman, 2000; Nestler, 2001).

Adderall is a mixture of D- and L-amphetamine salts (Berman et al., 2009) and, like cocaine, amphetamines produce elevated extracellular levels of the monoamines dopamine, norepinephrine, and serotonin (Di Chiara and Imperato, 1988; Hurd and Ungerstedt, 1989; Ritz et al., 1990; Kuczenski and Segal, 1997, 2001). Modafinil seems to primarily inhibit dopamine and norepinephrine reuptake (Madras et al., 2006; Volkow et al., 2009; Schmitt and Reith, 2011; Loland et al., 2012) but, probably indirectly, affects other neurotransmitters as well (e.g., histamine, orexin and serotonin; see Minzenberg and Carter, 2008). Better established are the effects of methylphenidate. Methylphenidate binds to and blocks the dopamine and norepinephrine transporters (Schweri et al., 1985; Gatley et al., 1996) and thus produces overflow of these two monoamines (Hurd and Ungerstedt, 1989; Kuczenski and Segal, 1997; Volkow et al., 1998; Gerasimov et al., 2000; Bymaster et al., 2002; Berridge et al., 2006). In contrast, methylphenidate has low affinity for the serotonin transporter (Pan et al., 1994; Wall et al., 1995; Gatley et al., 1996; Bymaster et al., 2002) and produces minimal or no effects on serotonin levels, even with high doses (30 mg/kg, i.p.) (Kuczenski and Segal, 1997; Segal and Kuczenski, 1999; Kankaanpaa et al., 2002).

In vivo microdialysis studies demonstrate that these psychostimulant-induced neurochemical effects are very robust in the prefrontal cortex and in parts of the basal ganglia (Figure 1), especially the striatum (dorsal striatum/caudate-putamen, ventral striatum/nucleus accumbens) (Di Chiara and Imperato, 1988; Hurd and Ungerstedt, 1989; Kuczenski and Segal, 1997, 2001; Gerasimov et al., 2000; Berridge et al., 2006). Low, clinically relevant doses of cognitive enhancers seem to preferentially boost extracellular levels of dopamine and norepinephrine in the prefrontal cortex (Berridge and Devilbiss, 2011, but see Volkow et al., 2001). Higher doses (presumably associated with abuse) predominantly affect dopamine in the striatum due to orders of magnitude higher levels of dopamine tissue content in the striatum.

An external file that holds a picture, illustration, etc.
Object name is nihms419755f1.jpg

Schematic illustrations of cortico-basal ganglia-thalamocortical circuits. (A) Striatal sectors used for mapping gene expression and their main cortical inputs (arrows) are shown for frontal, rostral, middle, and caudal levels of the rat forebrain (for details, see Willuhn et al., 2003; Yano and Steiner, 2005a,b). Psychostimulant-induced gene regulation is maximal in the dorsal sensorimotor sectors of the middle and caudal striatum (darkest shading), which receive inputs from the medial agranular (M2), primary motor (M1), and somatosensory (SS) cortex (see Section 3.1.2). Limbic (white), associative (light grey), and sensorimotor sectors (darker grey) are indicated. The scatterplot (inset upper right) displays the association between methylphenidate-induced Zif268 expression in individual striatal sectors and Zif268 expression in their indicated cortical input regions (values averaged if more than one input). Values are differences in gene expression between animals sacrificed 40 minutes after methylphenidate administration (5 mg/kg, i.p.) and controls sacrificed immediately after drug injection, and are expressed as the percentage of maximal increase in the striatum (see Yano and Steiner, 2005a). CG, cingulate; I, insular; IL, infralimbic; I/LO, insular/lateral orbital; P, piriform; PL, prelimbic. *** p < 0.001. (B) Direct and indirect striatal output pathways in the cortico-basal ganglia-thalamocortical circuits. Direct pathway (striatonigral) neurons contain mainly D1 dopamine receptors and the neuropeptides substance P (SP) and dynorphin (DYN), whereas neurons that give rise to the indirect pathway (striatopallidal neurons) express mostly D2 receptors and the peptide enkephalin (ENK), in addition to their main neurotransmitter γ-aminobutyric acid (GABA). + and − denote facilitatory and inhibitory, respectively. GLU, glutamate; GPe, globus pallidus external segment; GPi, globus pallidus internal segment; SNr, substantia nigra pars reticulata; STN, subthalamic nucleus

2.2. Other effects

In addition to the direct neurochemical effects described above, all of these drugs have a number of other acute effects that can, directly or indirectly, further modify monoamine (and other) transmission (see Yano and Steiner, 2007). The following examples pertain to methylphenidate: (1) Recent studies showed that acute methylphenidate administration alters the distribution and function of the vesicular monoamine transporter-2 (VMAT-2) in the striatum (Sandoval et al., 2002, 2003), similar to cocaine (Fleckenstein et al., 2009). (2) Methylphenidate produces enhanced phosphorylation of glutamate receptors (GluR1) in the prefrontal cortex, similar to amphetamine (Pascoli et al., 2005). (3) Methylphenidate affects second messenger cascades that mediate dopamine signaling. Thus, acute methylphenidate was found to increase and decrease phosphorylation of DARPP-32 at Thr34 and Thr75, respectively, in striatal slices from adult mice, an effect that was dependent on D1 dopamine receptor stimulation (Fukui et al., 2003). These findings demonstrate that there are several independent mechanisms by which cognitive enhancers can affect addiction-related neurotransmission.

Chronic perturbation of neurotransmission by psychostimulants often elicits compensatory (homeostatic) neuroadaptations, which are considered critical for addiction and dependence (Hyman and Nestler, 1996). Thus, repeated treatment with such drugs produces neuronal changes ranging from altered cell signaling (Yano and Steiner, 2007; McGinty et al., 2008) to structural modifications (e.g., in dendritic spine density; Robinson and Kolb, 1997; Jedynak et al., 2007; Kim et al., 2009), and the longevity of these alterations likely requires adaptations in gene expression (Renthal and Nestler, 2008).

Many excellent reviews have surveyed the effects of psychostimulants (primarily amphetamines and cocaine) on gene regulation and their role in addiction in general (e.g., Hyman and Nestler, 1996; Harlan and Garcia, 1998; Torres and Horowitz, 1999; Berke and Hyman, 2000; Nestler, 2001; Kelley, 2004; Hyman, 2005; McGinty et al., 2008; Renthal and Nestler, 2008). In this review we first summarize the molecular changes produced by amphetamine2 and cocaine to provide context for a discussion of findings on the effects of medical amphetamine (Adderall), methylphenidate, and the little that is known about modafinil. Most of these findings were obtained in rat and mouse models.

3. Gene regulation by amphetamine and cocaine in corticostriatal circuits

Most studies on the molecular effects of psychostimulants have focused on gene regulation in dopamine target areas, especially the striatum, which displays particularly robust changes in gene regulation after treatments with amphetamine, cocaine, and other abused drugs (Harlan and Garcia, 1998; Berke and Hyman, 2000).

The striatum, the main input nucleus of the basal ganglia, is an important component of cortico-basal ganglia-cortical circuits (Gerfen and Bolam, 2010; Figure 1), which play a critical role in motivational, executive, and motor aspects of all goal-directed behavior and thus in addiction (Steiner, 2010). Psychostimulant-induced molecular changes in these circuits through the striatum are important for various aspects of addiction, including abnormal reward processing, habit formation, and compulsive behavior (Robbins and Everitt, 1999; Berke and Hyman, 2000; Hyman and Malenka, 2001; Gerdeman et al., 2003; Everitt and Robbins, 2005; Belin and Everitt, 2010). However, some of the most affected (dorsal) striatal circuits (Willuhn et al., 2003; Yano and Steiner, 2005a; 2005b; Unal et al., 2009) also participate in frontostriatal attentional networks and may thus be a therapeutic target in ADHD (Robbins et al., 1998; Solanto, 2002). We therefore focus on psychostimulant-induced gene regulation in corticostriatal circuits.

Microarray investigations indicate that hundreds of genes are affected by dopamine and psychostimulants in these circuits (Berke et al., 1998; McClung and Nestler, 2003; Konradi et al., 2004; Yuferov et al., 2005; Adriani et al., 2006a; Adriani et al., 2006b; Black et al., 2006; Yano and Steiner, 2007; Heiman et al., 2008). However, the vast majority of studies have assessed effects on the expression of neuropeptide transmitters and immediate-early genes (IEGs). Neuropeptides are often selectively contained in specific neuronal subtypes and thus serve as cell type markers (see Section 3.1.3), but they also modulate basal ganglia functions on several levels (e.g., Steiner, 2010). IEGs are useful as markers for cell activation due to their rapid and transient induction by neuronal activity and drug treatments (Sharp et al., 1993; Chaudhuri, 1997; Harlan and Garcia, 1998). They are thus frequently used to map drug effects in the brain.

Immediate-early genes are also of interest because of their direct involvement in neuroplasticity. Many IEGs encode transcription factors that regulate the expression of other genes (e.g., c-Fos, Zif268; Knapska and Kaczmarek, 2004). Others (e.g., Homer 1a) code for members of a family of scaffolding proteins that anchor receptors to the postsynaptic density and play a role in receptor trafficking, dendritic spine formation, and other processes of synaptic plasticity (Xiao et al., 2000; Thomas, 2002). These latter processes may be involved in the abnormal spine formation in striatal neurons produced by psychostimulant treatment (Robinson and Kolb, 1997; Ferrario et al., 2005; Jedynak et al., 2007; Kim et al., 2009).

In the following sections, we first describe the functional domains of the striatum and then present studies that illustrate psychostimulant-induced effects on gene regulation in these domains. The findings reveal which cortico-basal ganglia-cortical circuits (functional domains, cell types) are affected by these drugs and establish a cellular framework for evaluating and understanding the effects of cognitive enhancers such as methylphenidate and modafinil. We then provide examples of molecular changes induced by repeated psychostimulant treatment and discuss their potential functional significance.

3.1. Corticostriatal circuits affected

3.1.1. Functional domains of the striatum

The functional domains of the striatum are defined by their cortical inputs (Figure 1). According to current models of basal ganglia function, the basal ganglia and cortex are interconnected by several parallel anatomical circuits/loops that arise in the cortex and project in a topographical manner to the striatum and from there via the basal ganglia output nuclei and thalamus back to the cortex (Alexander et al., 1986; Albin et al., 1989; Alexander et al., 1990; Groenewegen et al., 1990; Haber, 2003; Joel and Weiner, 1994; Redgrave et al., 2010). Functionally, these circuits can be roughly categorized as limbic, associative, and sensorimotor, arising, respectively, in the limbic, associative, and sensory and motor regions of the cortex and projecting to their associated domains in all basal ganglia nuclei. The behavioral consequences of psychostimulant-induced molecular changes (or indeed of any pathological changes) in the basal ganglia are therefore dependent on the particular circuits affected. Thus there is interest in determining which circuits/functional domains in the striatum are altered by these drugs.

3.1.2. Topography of psychostimulant-induced gene regulation

Early descriptions of the regional distribution of psychostimulant-induced gene regulation in the striatum were in relatively vague terms anatomically (e.g., dorsolateral quadrant, dorsal vs. ventral). Nevertheless, the findings were clear and consistent between laboratories that these effects differ considerably between the different striatal regions. This variability is characteristic of the effects of amphetamine, cocaine (Figure 2A), and methylphenidate (Figure 2B).

An external file that holds a picture, illustration, etc.
Object name is nihms419755f2.jpg

Psychostimulant-induced immediate-early gene expression in corticostriatal circuits as determined by in situ hybridization histochemistry. (A) Cocaine-induced c-Fos expression. Top: Film autoradiograms depict c-Fos expression in coronal sections from rostral (top), middle (center), and caudal striatal levels (bottom) in rats that received a vehicle injection (control, left halfbrain) or a cocaine injection (25 mg/kg; right halfbrain) and were sacrificed 30 minutes later (Brandon and Steiner, 2003). Striatum (S) and nucleus accumbens (NAc) are outlined in the cocaine-treated animals. Note the considerable regional differences in the c-Fos response. Bottom: Time course of cocaine (30 mg/kg)-induced c-Fos and Zif268 expression in the dorsal striatum on the middle level (mean density, expressed as percentage of maximal induction) (Steiner and Gerfen, 1998). Basal expression is indicated by broken lines. (B) Methylphenidate-induced gene expression. Top: Film autoradiograms show c-Fos (left), Zif268 (middle), and Homer 1a expression (right) at 0 minutes (control, left halfbrain) and 40 minutes (c-Fos) or 1 hour (Zif268, Homer 1a) after methylphenidate injection (MP, 5 mg/kg, i.p.; right halfbrain) (Yano and Steiner, 2005a,b). Bottom: Time course of methylphenidate (5 mg/kg)-induced expression of c-Fos, Zif268, and Homer 1a for the dorsal striatal sector on the middle level (Yano and Steiner, 2005a,b). LS, lateral shell of nucleus accumbens

Graybiel and colleagues (1990) first showed that c-Fos induction by acute cocaine treatment, although widespread in the striatum, had a distinctive topography. It was most pronounced in the dorsal central portion of the sensorimotor striatum and was fairly limited (or absent) in parts of the ventral (limbic) striatum, including the nucleus accumbens (Figure 2A). Other early studies confirmed this general pattern for c-Fos and other genes (e.g., Young et al., 1991; Hope et al., 1992; Moratalla et al., 1992; Bhat and Baraban, 1993; Steiner and Gerfen, 1993; Johansson et al., 1994; see Harlan and Garcia, 1998 for a review of the early work).

Many investigators found an overall similar (dorsal-ventral) distribution for amphetamine effects (Graybiel et al., 1990; Moratalla et al., 1992; Wang et al., 1994a; Badiani et al., 1998; Adams et al., 2001), but there are also differences between amphetamine and cocaine effects, for example, in their distribution across the striatal patch/matrix compartments (Harlan and Garcia, 1998). In contrast to the rather uniform gene induction by cocaine in terms of patch/matrix distribution, the IEG response to amphetamine appears reduced in the matrix relative to that in the patches (striosomes) (Graybiel et al., 1990; Graybiel et al., 2000). This finding was confirmed by many (e.g., Moratalla et al., 1992; Nguyen et al., 1992; Wang et al., 1995) but not all (Johansson et al., 1994; Wang et al., 1994b; Jaber et al., 1995) subsequent studies.

To better relate psychostimulant-induced molecular changes to specific corticostriatal circuits/functional domains in the rat, we mapped striatal gene regulation using 23 sampling areas (sectors)—based largely on their predominant cortical inputs—on three rostrocaudal levels (Figure 1A) (Willuhn et al., 2003; Yano and Steiner, 2005a, b; Cotterly et al., 2007; Unal et al., 2009). These studies revealed the following patterns (see Steiner, 2010 for review):

  1. The most robust cocaine-induced changes in gene regulation occur in sensorimotor sectors of the middle and caudal striatum (Figure 2A) (e.g., Steiner and Gerfen, 1993; Willuhn et al., 2003; Unal et al., 2009). A similar regional distribution has been shown for amphetamine (e.g., Badiani et al., 1998).

  2. Within the sensorimotor striatum, maximal changes occur in the dorsal sectors (approximately the dorsal third) (Figure 2A). These sectors are unique in that they receive the densest input from the medial agranular cortex (M2; Figure 1A) (Reep et al., 2003) in addition to convergent inputs from the somatosensory (or visual) and primary motor cortex (cf. Willuhn et al., 2003). Surrounding tissue that is, to a lesser extent, also targeted by medial agranular projections also shows robust changes in gene expression. The rat medial agranular cortex has mixed prefrontal/premotor features (Reep et al., 1987; Passingham et al., 1988; Preuss, 1995; Reep et al., 2003; Uylings et al., 2003) and can therefore be considered a prefrontal/motor interface. Our findings thus indicate that sensorimotor striatal circuits under the influence of medial agranular (prefrontal/premotor) input are particularly prone to psychostimulant-induced neuroplasticity.

  3. Medial and rostral striatal sectors (associative sectors) were affected to a lesser degree (Figure 2A). These sectors receive inputs from prefrontal regions including the cingulate, prelimbic, and orbital cortex (Figure 1A) (e.g., Berendse et al., 1992).

  4. On all three rostrocaudal levels, minimal or no changes in gene regulation were seen in ventral striatal sectors (Figure 2A) that receive inputs mostly from the dorsal agranular insular cortex (Figure 1A) (e.g., Berendse et al., 1992).

  5. Psychostimulant-induced molecular changes in the nucleus accumbens are well appreciated in the addiction literature (e.g., Graybiel et al., 1990; Hope et al., 1992; Hope et al., 1994; for reviews, see Berke and Hyman, 2000; Nestler, 2001) because they are implicated in motivational (reward) processes (Pierce and Kalivas, 1997). However, consistent with the earlier literature (see above), our studies show that gene regulation effects of cocaine in the nucleus accumbens (Figure 2A) are modest compared with those in the sensorimotor striatum (Steiner and Gerfen, 1993; Willuhn et al., 2003; Unal et al., 2009). This effect reflects the finding that cocaine strongly activates only a small proportion of sparsely distributed neurons in the nucleus accumbens (as well as in the most rostral striatum) (Mattson et al., 2008). The nucleus accumbens shell appears more affected than the core, and the most robust effects were seen in the lateral part of the shell (Unal et al., 2009), which also receives medial agranular input (Reep et al., 1987) in addition to inputs from the ventral agranular insular cortex (Berendse et al., 1992) and other limbic areas (e.g., McGeorge and Faull, 1989; Brog et al., 1993; Wright and Groenewegen, 1996). The functional significance of these lateral shell effects is not known, but it is of interest to note that the insular cortex, one of the input regions of that part of the nucleus accumbens, is associated with craving in drug addiction (Naqvi et al., 2007), which often drives relapse.

In summary, amphetamine and cocaine produce changes in gene regulation in limbic striatal regions, but these are relatively modest. These changes are likely involved in altered reward processing in addiction (e.g., Belin and Everitt, 2010). Studies that compared effects in different striatal regions demonstrate that more robust drug-induced changes in gene regulation occur in the sensorimotor striatum. These molecular changes probably mediate the functional changes seen in these regions as the addiction disorder progresses (Porrino et al., 2007). Behaviorally, sensorimotor striatal changes may be responsible for habitual and compulsive aspects of drug taking (Berke and Hyman, 2000; Gerdeman et al., 2003; Everitt and Robbins, 2005; Belin and Everitt, 2010), and are likely also important for relapse to drug seeking after abstinence (Vanderschuren et al., 2005; Fuchs et al., 2006; See et al., 2007).

3.1.3. Striatal cell types

The main cell type of the striatum is the medium-sized spiny projection neuron (“medium spiny neuron”); in the rat, interneurons account for less than 3% of striatal neurons (Oorschot, 2010, 2013). Colocalization studies indicate that psychostimulants affect gene regulation in projection neurons but have minimal or no effect in interneurons (Berretta et al., 1992). For example, cholinergic interneurons showed cocaine-induced IEG expression only in the ventromedial striatum and the medial shell of the nucleus accumbens but not in the core or in the dorsolateral striatum (Berlanga et al., 2003).

Striatal projection neurons are divided into two subtypes that are intermingled and approximately equal in number and that give rise to two different striatal output pathways (Figure 1B). The “direct pathway” (striatonigral neurons) connects the striatum directly to the basal ganglia output nuclei (substantia nigra pars reticulata, entopeduncular nucleus/internal pallidum); the “indirect pathway” begins with the striatopallidal neurons and projects to the output nuclei indirectly via the globus pallidus (external pallidum) and subthalamic nucleus (Gerfen and Bolam, 2010).

Both subtypes of striatal projection neurons use γ-aminobutyric acid as their main neurotransmitter, but they differ in a number of receptors and neuropeptides they express (Steiner and Gerfen, 1998; Heiman et al., 2008). Striatonigral neurons contain predominantly the D1 receptor subtype and the neuropeptides substance P and dynorphin (Figure 1B), whereas striatopallidal neurons mostly express the D2 receptor and the neuropeptide enkephalin. (Because of this differential receptor/neuropeptide distribution, these neurons are sometimes referred to as D1 or D2 neurons, respectively, and these neuropeptides often serve as markers to differentiate effects of drug treatments between these striatal output pathways).

These two striatal output pathways have opposite effects on basal ganglia output and motor control. According to current models of basal ganglia function (Figure 1B), activity in the direct pathway inhibits basal ganglia output, thus disinhibiting thalamocortical (and brainstem) activity (Chevalier and Deniau, 1990) and facilitating behavior, whereas activity in the indirect pathway (i.e., in striatopallidal neurons) results in disinhibition of basal ganglia output, thus arresting behavior (Albin et al., 1989; DeLong, 1990; Redgrave et al., 2010). It is thought that activity in the direct pathway (the “Go pathway”) functions to initiate (or facilitate selection of) motor programs, whereas activity in the indirect pathway (the “Stop pathway”) interrupts motor programs and/or suppresses unwanted (or incompatible) movement. (For an elegant recent demonstration of this oppositional movement control, see Kravitz et al., 2010.) Although these concepts were initially derived from anatomical findings in the dorsal/sensorimotor striatum (Alexander et al., 1986; Albin et al., 1989), recent results confirmed such antagonistic functions of these two pathways for cocaine-induced behavior mediated by the dorsal striatum (Ferguson et al., 2011) and reward processes mediated by the limbic/ventral striatum (Lobo et al., 2010).

Given the differential functional roles of the two subtypes of striatal projection neurons, it was of considerable interest to determine whether psychostimulants alter gene regulation in both subtypes or whether one is preferentially affected. Early clues were obtained from drug effects on the expression of the neuropeptides that are differentially localized in these neurons and thus serve as cell type markers (Steiner and Gerfen, 1998). Many studies showed that amphetamine and cocaine robustly induce expression of substance P and dynorphin (e.g., Hanson et al., 1987; Sivam, 1989; Hurd and Herkenham, 1992; Hurd and Herkenham, 1993; Steiner and Gerfen, 1993; Daunais and McGinty, 1994; Wang and McGinty, 1995a; Drago et al., 1996; Adams et al., 2001; Frankel et al., 2008), which are contained in striatonigral neurons. In contrast, enkephalin expression (in striatopallidal neurons) is only modestly affected by psychostimulants (Steiner and Gerfen, 1993; Jaber et al., 1995; Wang and McGinty, 1996a; Spangler et al., 1997; Mathieu-Kia and Besson, 1998). (It should be noted, however, that enkephalin expression is readily induced by glutamate receptor stimulation or D2 receptor blockade [e.g., Steiner and Gerfen, 1999].)

Colocalization studies using neuropeptide messenger RNAs (mRNAs) or tract tracers as markers confirmed this differential gene regulation for IEGs as well. Amphetamine and cocaine induce IEGs predominantly in striatonigral neurons (Berretta et al., 1992; Cenci et al., 1992; Johansson et al., 1994; Jaber et al., 1995; Kosofsky et al., 1995; Badiani et al., 1999). However, depending on the treatment conditions (i.e., with enough cortical activation/glutamate input; see Steiner, 2010), some IEG induction also occurs in striatopallidal neurons (e.g., Jaber et al., 1995; Badiani et al., 1999; Uslaner et al., 2001; Ferguson and Robinson, 2004).

3.1.4. Dopamine receptor subtypes

The differential effects on striatonigral versus striatopallidal neurons are likely based on the differential distribution of dopamine receptor subtypes between the two projection neuron subtypes (Figure 1B): As mentioned above, D1 receptors are predominantly expressed in striatonigral neurons, and D2 receptors mostly in striatopallidal neurons (Gerfen et al., 1990; Le Moine et al., 1990; Le Moine et al., 1991; Curran and Watson, 1995; Le Moine and Bloch, 1995). Numerous studies show that D1 receptor stimulation and resulting activation of second messenger signaling cascades (Bronson and Konradi, 2010; Caboche et al., 2010) are critical for psychostimulant-induced gene regulation in striatal neurons. Thus, IEG expression induced by amphetamine and cocaine is eliminated either by systemic or intrastriatal administration of D1 receptor antagonists (Graybiel et al., 1990; Young et al., 1991; Moratalla et al., 1992; Cole et al., 1992; Steiner and Gerfen, 1995) or by targeted deletion of the D1 receptor (D1 receptor knockouts) (Drago et al., 1996; Moratalla et al., 1996b; Zhang et al., 2004).

D2 receptors also affect gene regulation in striatal neurons. In contrast to D1 receptors, however, stimulation of D2 receptors inhibits gene expression in striatopallidal neurons (e.g., Gerfen et al., 1990; Le Moine et al., 1997; Pinna et al., 1997), whereas blockade of D2 receptors (e.g., by antipsychotic drugs) increases gene expression in these neurons (e.g., Steiner and Gerfen, 1998). This difference in effect presumably reflects the fact that D2 receptors inhibit second messenger signaling, as opposed to the stimulatory action of D1 receptors (Bronson and Konradi, 2010). However, stimulation of D2 plus D1 receptors potentiates D1 receptor–mediated gene regulation in striatonigral neurons (D1–D2 receptor synergy; e.g., Paul et al., 1992; LaHoste et al., 1993; Gerfen et al., 1995). Consistent with this observation, a full gene response to psychostimulants requires combined stimulation of D1 and D2 receptors (Ruskin and Marshall, 1994). This interaction between D1 and D2 receptors is thought to be mediated by cholinergic interneurons (Wang and McGinty, 1996b; Pisani et al., 2007)—for example, via a D2 receptor–mediated inhibition of inhibitory cholinergic input to striatonigral neurons (Wang and McGinty, 1996b).

In addition, D3 receptors modify such molecular effects. These receptors are predominantly present in ventral striatal regions where they are partly coexpressed with D1 receptors in striatonigral neurons (Le Moine and Bloch, 1996; Schwartz et al., 1998). Because they also exert opposite (inhibitory) effects on second messenger signaling (Zhang et al., 2004), D3 receptors dampen gene induction by D1 receptor stimulation (Carta et al., 2000; Zhang et al., 2004).

In summary, these findings demonstrate that (1) amphetamine- and cocaine-induced changes in gene regulation in the striatum occur preferentially (but not exclusively) in direct pathway (striatonigral) neurons (see also Lobo and Nestler, 2011), and (2) D1 receptors (and their downstream signaling cascades; Caboche et al., 2010) are critical for these molecular changes.

3.2. Relationship between gene regulation in striatum and cortex

Imaging studies in humans and other primates show that exposure to psychostimulants such as cocaine and amphetamine produces functional changes also in various regions of the cortex (e.g., London et al., 1990; Breiter et al., 1997; Beveridge et al., 2006; Porrino et al., 2007). Similarly, systemic administration of cocaine, amphetamine, and other dopamine agonists causes increases in gene expression in the cortex (Figure 2) (e.g., Graybiel et al., 1990; Paul et al., 1992; Dilts et al., 1993; Johansson et al., 1994; Steiner and Gerfen, 1994; Wang and McGinty, 1995a; LaHoste et al., 1996; Badiani et al., 1998). These cortical effects are widespread (Harlan and Garcia, 1998), but a recent detailed mapping study showed that acute and repeated cocaine treatments produce the most robust changes in IEG regulation in sensory and motor regions of the cortex (Unal et al., 2009), thus mirroring the distribution of such molecular changes across striatal functional domains. Other studies have also revealed preferential gene regulation in the sensorimotor cortex for cocaine (e.g., Daunais and McGinty, 1994; Johansson et al., 1994) and amphetamine (e.g., Wang et al., 1994a, 1995; Curran et al., 1996; Badiani et al., 1998; Uslaner et al., 2001).

Some of these cortical effects may be a consequence of drug action directly in the cortex. However, consistent with the models of cortico-basal ganglia-cortical circuits (Figure 1B), many of the cortical changes are caused by drug-induced alterations in basal ganglia output as a consequence of changed activity in the D1 receptor-regulated direct striatal output pathway (for review, see Steiner, 2007). Thus, stimulation of striatal D1 receptors produces widespread increases in gene expression throughout the cortex (Steiner and Kitai, 2000; Gross and Marshall, 2009; see Steiner, 2007).

True to the loop architecture of these circuits, reentrant activity from the cortex (or thalamus; Cotterly et al., 2007) to the striatum is also important for psychostimulant-induced gene regulation in the striatum. Studies demonstrated that blockade of glutamate (N-methyl-D-aspartate) receptors (e.g., Johnson et al., 1991; Torres and Rivier, 1993; Wang et al., 1994a; Hanson et al., 1995) or elimination of corticostriatal afferents (Cenci and Björklund, 1993; Vargo and Marshall, 1995; Ferguson and Robinson, 2004) attenuates psychostimulant-induced gene expression in striatal neurons. Therefore, striatal effects of psychostimulants are a consequence of drug-induced overstimulation of striatal D1 receptors in interaction with cortical (glutamate) input (Hyman et al., 1996; Wang and McGinty, 1996b).

Other findings demonstrate that psychostimulants engage cortical and striatal nodes of corticostriatal circuits in a coordinated manner; gene induction in cortical areas and in their associated functional domains in the striatum is correlated (Figure 1A; Cotterly et al., 2007; Yano and Steiner, 2005a). Given their role in neuroplasticity, these gene regulation effects indicate coordinated neuroplastic changes in cortical and functionally related striatal areas.

3.3. Molecular effects of repeated amphetamine and cocaine exposure

Repeated psychostimulant exposure produces a variety of neuroadaptations and other neuronal changes in the basal ganglia (e.g., Hyman and Nestler, 1996; Kuhar and Pilotte, 1996; Berke and Hyman, 2000; Nestler, 2001; Kelley, 2004). In this section, we provide a few examples of such molecular changes for comparison with similar changes induced by cognitive enhancers, presented later. These examples involve the same IEG and neuropeptide markers as discussed in the previous sections.

As would be expected for molecular adaptations, changes after repeated treatments occur in the same striatal regions and neurons that display the acute drug effects and are directly correlated in magnitude with that of the acute effects (Steiner and Gerfen, 1993; Willuhn et al., 2003; Unal et al., 2009).

3.3.1. Blunted gene inducibility

One of the best-established molecular consequences of repeated psychostimulant treatment is blunting (repression) of gene inducibility in the striatum. Thus, after repeated treatments, genes are still inducible by a drug challenge, but this induction is typically attenuated compared with acute induction. Such blunting was first demonstrated after repeated amphetamine and cocaine treatment for several transcription factor IEGs (e.g., c-Fos and Zif268; Hope et al., 1992, 1994; Persico et al., 1993; Steiner and Gerfen, 1993; Daunais and McGinty, 1994; Moratalla et al., 1996a). Other genes are similarly affected—for example, the effector IEG Homer 1a (Unal et al., 2009) and the neuropeptide substance P (Steiner and Gerfen, 1993; Jaber et al., 1995).

Blunting of gene induction is long-lasting. A recent study showed marked attenuation in Zif268 and Homer 1a inducibility even 3 weeks after a 5-day repeated cocaine treatment (Unal et al., 2009).

Consistent with a compensatory neuroadaptation, the degree of blunting is directly related to the magnitude of the initial (acute) gene induction in a given striatal region—the greater the induction after the first drug administration, the more blunted the induction after chronic treatment (Willuhn et al., 2003; Unal et al., 2009). Mapping studies showed that repeated cocaine treatment produces the most robust blunting in the dorsal/lateral (sensorimotor) striatum at middle to caudal striatal levels (Willuhn et al., 2003; Unal et al., 2009). [It should be noted, however, that gene induction is not universally blunted in all striatal areas after repeated psychostimulant treatments; in parts of the nucleus accumbens, increased rather than reduced gene induction has been demonstrated in several studies (Crombag et al., 2002; Todtenkopf et al., 2002; Brandon and Steiner, 2003; Cotterly et al., 2007; Damez-Werno et al., 2012).]

Various mechanisms may contribute to blunting of gene induction after repeated drug treatment, some shorter-lasting, some long-lasting. Investigators have proposed systems-level neuroadaptations as well as intracellular (epigenetic) adaptations. Examples are:

  1. Given the importance of excitatory inputs for striatal gene regulation, blunted gene induction may partly reflect dampened inputs from the cortex (and/or thalamus), perhaps involving long-term depression-like synapse modifications (see Graybiel et al., 2000; Unal et al., 2009, for discussion).

  2. Neuropeptides such as dynorphin modulate dopamine and glutamate input to striatal neurons and thus indirectly also affect gene regulation (Steiner, 2010). For example, acute IEG induction by cocaine and D1 receptor agonist treatment is inhibited by stimulation of dynorphin (kappa opioid) receptors in the striatum (Steiner and Gerfen, 1995, 1996). Blunting of gene induction may thus at least in part reflect increased dynorphin function and resulting inhibition of dopamine or glutamate action after repeated psychostimulant treatment (as we discuss below).

  3. Epigenetic regulation of gene expression involving chromatin modifications (e.g., histone acetylation and methylation) may best explain the endurance of gene blunting (for reviews see Renthal and Nestler, 2008; Caboche et al., 2010). For example, chromatin modification has been shown to contribute to blunting of c-Fos expression after repeated amphetamine treatment (e.g., Renthal et al., 2008) and to blunting/priming of FosB expression after repeated cocaine treatment (Damez-Werno et al., 2012).

The exact consequences of blunted gene induction for basal ganglia function are unknown. However, the functional integrity of neurons depends on balanced regulation of gene expression because cellular components have limited half-lives and must be replenished. It is assumed that disruption of such homeostatic regulation by psychostimulants results in deficient neuronal function that contributes to behavioral manifestations of psychostimulant addiction (e.g., Hyman and Nestler, 1996; Nestler, 2001).

3.3.2. Alternative splicing: accumulation of deltaFosB

Another often described molecular change caused by psychostimulants is accumulation of the transcription factor deltaFosB in striatonigral neurons (McClung et al., 2004). DeltaFosB is induced by many manipulations that involve excessive neuronal activation (McClung et al., 2004). DeltaFosB is a truncated isoform of FosB (member of the AP-1 family of transcription factors) that is produced by alternative splicing (Nakabeppu and Nathans, 1991). The truncation renders the molecule highly stable. With repeated drug treatments, deltaFosB accumulates in cells and displaces other members of the AP-1 family from the AP-1 transcriptional complex, thus altering the function of this complex (Nakabeppu and Nathans, 1991; McClung et al., 2004).

DeltaFosB accumulation is well established for repeated amphetamine and cocaine treatments (Hope et al., 1994; Nye et al., 1995; Renthal et al., 2008). A recent mapping study showed increased deltaFosB levels in many striatal regions, with maximal increases in the dorsal/lateral striatum, after repeated cocaine treatment (Sato et al., 2011).

Findings indicate that many genes (e.g., those with AP-1 and CRE binding sites in their promoter) are affected by this abnormal transcription factor; some are activated and some are repressed, depending in part also on the length of the drug treatment (McClung and Nestler, 2003; McClung et al., 2004). For example, deltaFosB action appears to upregulate dynorphin expression (Andersson et al., 1999; but see McClung et al., 2004), while playing a role in blunting of c-Fos induction after repeated amphetamine treatment (Renthal et al., 2008).

3.3.3. Increased dynorphin expression

A third widely demonstrated consequence of repeated psychostimulant treatment is increased dynorphin expression in the striatonigral (direct) pathway (Steiner, 2010; Butelman et al., 2012; Yoo et al., 2012). Many laboratories have reported elevated dynorphin mRNA (e.g., Hurd and Herkenham, 1992; Spangler et al., 1993; Steiner and Gerfen, 1993; Daunais and McGinty, 1994; Wang et al., 1994a; Adams et al., 2003; Willuhn et al., 2003) or peptide levels (e.g., Hanson et al., 1987; Li et al., 1988; Sivam, 1989; Smiley et al., 1990) after repeated amphetamine and cocaine treatment. Notably, increased dynorphin expression has also been found in human cocaine addicts (Hurd and Herkenham, 1993; Frankel et al., 2008).

After a single psychostimulant administration, elevated dynorphin mRNA levels persist in the rat for at least 18 to 30 hours (Smith and McGinty, 1994; Wang and McGinty, 1995). Thus, this mRNA accumulates with daily drug treatments. Indeed, after repeated treatment with a dopamine agonist, elevated dynorphin mRNA levels in the striatum lasted several weeks past cessation of the treatment (Andersson et al., 2003). Again, repeated cocaine treatment produces maximally increased dynorphin expression in the dorsal/lateral (sensorimotor) sectors of the middle to caudal striatum (Steiner and Gerfen, 1993; Willuhn et al., 2003).

What is the functional significance of increased dynorphin expression in the striatum? Findings indicate that opioid peptides such as dynorphin (striatonigral neurons) and enkephalin (striatopallidal neurons) act, at least in part, as negative feedback mechanisms (Steiner and Gerfen, 1998) to limit dopamine and glutamate input to these neurons (Steiner, 2010). Repeated excessive activation of these neurons by pharmacological treatments (or other experimental manipulations) is thought to trigger compensatory upregulation of opioid peptide function to counteract the activation (i.e., to act as a “brake”) and maintain systems homeostasis (Hyman and Nestler, 1996).

In the case of upregulated dynorphin function after repeated psychostimulant exposure, it is thus to be expected that during early withdrawal from drug use the “brake” is still on for some time given the relatively long half-life of changes in dynorphin expression. The increased dynorphin signaling would then excessively inhibit inputs to striatal neurons (Hyman and Nestler, 1996; Steiner and Gerfen, 1998; Shippenberg et al., 2007). There is good evidence that increased dynorphin function in this manner contributes to somatic signs of withdrawal such as dysphoria, anxiety, anhedonia, and depression after discontinuation of drug use (Nestler and Carlezon, 2006; Shippenberg et al., 2007; Butelman et al., 2012; Yoo et al., 2012). These effects are thought to contribute to maintenance of drug use or relapse during abstinence.

4. Gene regulation by oral Adderall

The above reviewed effects of amphetamine in animal studies were mostly obtained with intraperitoneal (i.p.) or subcutaneous (s.c.) administration of relatively high doses (~3–10 mg/kg). How relevant are these findings for therapeutic use of Adderall, which involves lower doses and predominantly oral administration?

Psychostimulant effects on gene regulation are dose-dependent. Higher doses produce greater increases in gene expression across a wide range of doses (e.g., Steiner and Gerfen, 1993; Wang and McGinty, 1995b, 1997; Brandon and Steiner, 2003; Chase et al., 2003; Chase et al., 2005a; Yano and Steiner, 2005b). [Occasionally, very high doses have been found to result in attenuated expression (Wang and McGinty, 1995b, 1997), similar to other neuronal effects (e.g., Hanson et al., 2002), possibly due to receptor inactivation (internalization) by the high dose or other mechanisms.]

Gene regulation is also under the control of the drug delivery rate. For example, fast intravenous (i.v.) delivery of a certain cocaine dose (2 mg/kg) produced greater c-Fos induction in the striatum than slower delivery of the same dose (Samaha et al., 2004; Samaha and Robinson, 2005). Similarly, with repeated treatment (self-administration model), faster drug delivery produced more robust blunting of c-Fos inducibility (Wakabayashi et al., 2010). These enhanced neuronal changes were associated with indices of a greater addiction liability (greater escalation of drug intake and propensity to relapse; Wakabayashi et al., 2010).

Conversely, oral (or intragastric) administration of drugs produces slower (and lower) uptake (Swanson and Volkow, 2003; Kuczenski and Segal, 2005; Yano and Steiner, 2007), which would thus be expected to produce less molecular changes. Few studies have investigated the molecular effects of oral amphetamine administration in a therapeutic dose range. Researchers recently used a model with prepubertal rats to assess whether a low dose (1.6 mg/kg) of orally (p.o.) administered Adderall (mix of D- and L-amphetamine), which resulted in amphetamine levels in the blood close to those of children treated with amphetamine, caused changes in c-Fos expression in corticostriatal circuits (Allen et al., 2010). The results showed that despite the low dose and oral route, acute Adderall administration produced significant c-Fos induction in the striatum and cortex (Allen et al., 2010). Moreover, repeated treatment (1.6 mg/kg, p.o., once daily for 14 days) resulted in blunting of c-Fos inducibility in these brain regions (Allen et al., 2010). An earlier study in adult cats reported that 1 mg/kg (p.o.) of amphetamine induced c-Fos in the cortex and striatum (Lin et al., 1996). Consistent with these findings, another study in young rats showed that repeated treatment with a low dose of amphetamine (0.5 mg/kg, s.c., twice daily for 13 days), which resulted in amphetamine plasma concentrations corresponding to the clinical range used in the treatment of ADHD, produced altered dendritic architecture in the prefrontal cortex (Diaz Heijtz et al., 2003.

In summary, the findings obtained with faster administration and higher doses of amphetamine (and cocaine) may be more relevant for abuse of psychostimulants. But the results described above indicate that therapeutically relevant amphetamine doses and routes of administration can produce qualitatively similar molecular changes in neurons of corticostriatal circuits (Carrey and Wilkinson, 2011).

5. Gene regulation by methylphenidate

Methylphenidate, widely used in the treatment of ADHD and other mental disorders, is also popular as a cognitive enhancer (see Introduction). Although methylphenidate has been effective in the clinic for several decades, assessment of its molecular impacts began only about 10 years ago (Yano and Steiner, 2007). Because both clinical and recreational exposure to methylphenidate occurs predominantly in children and adolescents, preclinical studies often focus on the effects in prepubertal/adolescent animals (Yano and Steiner, 2007; Carrey and Wilkinson, 2011; Marco et al., 2011).

Early microarray studies in adolescent rats showed that acute and repeated treatment with 2 mg/kg (i.p.) of methylphenidate altered the expression of more than 2,000 genes in the striatum (Adriani et al., 2006a,b). Similar to other psychostimulants (Section 3), methylphenidate affected genes that encode transcription factors, neurotransmitter receptors, ion channels, postsynaptic density proteins, and other signaling-related molecules as well as many other classes (e.g., molecules involved in cell migration, survival, maturation, and other forms of neuroplasticity) (Adriani et al., 2006a,b; see also Yano and Steiner, 2007; Carrey and Wilkinson, 2011; Marco et al., 2011). Some of the molecular changes persisted well past the termination of the drug treatment, into the adulthood of the animals (Adriani et al., 2006a,b; for similarly long-lasting changes, see Chase et al., 2007; Warren et al., 2011).

Most of these wide-ranging effects will have to be confirmed in follow-up studies; but the effects on the expression of transcription factors/IEGs and neuropeptides in corticostriatal circuits are well established. We summarize these findings here for comparison with the effects of amphetamine and cocaine described above.

5.1. Regulation of immediate-early genes and neuropeptides

The first demonstration of gene regulation by methylphenidate was through oral administration in adult cats (Lin et al., 1996). This study showed that 2.5 mg/kg (p.o.) induced c-Fos expression in many brain areas, including the cortex and striatum. The regional patterns were described as “highly similar” to those produced by 1 mg/kg (p.o.) of amphetamine (these doses were compared because of their similar effects on wakefulness). Importantly, the c-Fos expression patterns of both drugs were different from those induced by modafinil (5 mg/kg, p.o.), also chosen for a similar waking effect (Lin et al., 1996). This comparison indicates that these regional patterns reflect the pharmacological targets of these drugs rather than their waking effect. Acute c-Fos induction by methylphenidate in the striatum (Figure 2B) was confirmed by many subsequent studies in both mice (Penner et al., 2002; Trinh et al., 2003; Hawken et al., 2004) and rats (Brandon and Steiner, 2003; Chase et al., 2003; Chase et al., 2005b; Yano and Steiner, 2005b).

Examples of other IEGs induced in the striatum include the transcription factors Zif268 (Figure 2B) (Brandon and Steiner, 2003; Yano and Steiner, 2005a) and FosB (Chase et al., 2005a) as well as the effector IEGs Arc (Chase et al., 2007; Banerjee et al., 2009) and Homer 1a (Figure 2B) (Yano and Steiner, 2005a; Adriani et al., 2006a; Cotterly et al., 2007).

A few studies show that methylphenidate also affects neuropeptide markers in striatal output neurons. These studies indicate that methylphenidate increases substance P expression (striatonigral neurons) (Figure 3) in a manner similar to other psychostimulants, whereas the opioid peptides dynorphin (striatonigral neurons) and enkephalin (striatopallidal neurons) appear to be less affected (Yano and Steiner, 2007).

An external file that holds a picture, illustration, etc.
Object name is nihms419755f3.jpg

Methylphenidate-induced neuropeptide expression. Top: Film autoradiograms depict substance P expression in the middle striatum at 0 minutes (control, left halfbrain) and 1 hour after injection of methylphenidate (MP, 5 mg/kg, i.p.; right halfbrain). Bottom: Time course of methylphenidate (5 mg/kg)-induced expression of substance P (SP), dynorphin (DYN), and enkephalin (ENK) (in percentage of basal expression) for the dorsal striatal sector on the middle level (Yano and Steiner, 2005b). Note that substance P expression increased in 13 of the 23 striatal sectors, whereas dynorphin and enkephalin expression significantly increased in only two and one sector, respectively (values in area of maximal increase are shown here; see Yano and Steiner, 2005b, for details). ** p < 0.01, * p < 0.05 versus 0 minutes.

We directly compared methylphenidate effects on the expression of these genes by monitoring their mRNA levels between 20 minutes and 24 hours after acute injection of methylphenidate (2–10 mg/kg, i.p., adult rats; Yano and Steiner, 2005b) (Figure 3). Similar to the effects of cocaine/amphetamine (see Section 3.1.3), we found that substance P expression increased in many striatal sectors in a dose-dependent and very robust manner, with elevated mRNA levels present within 20 minutes and lasting for more than 3 hours. Conversely, for dynorphin mRNA, we detected a statistically significant, if modest, increase only in two sectors at 1 hour (Figure 3) (Yano and Steiner, 2005b). This latter finding contrasts with studies on amphetamine and cocaine effects, which showed that significantly increased dynorphin mRNA levels are present within 30 minutes (Willuhn et al., 2003), are prominent at 2 to 3 hours (Hurd and Herkenham, 1992; Smith and McGinty, 1994), and last 18 to 30 hours (Smith and McGinty, 1994; Wang et al., 1995) after acute drug administration.

Enkephalin, which is strongly induced, for example, by D2 receptor antagonists (Steiner and Gerfen, 1998), is only moderately affected by acute cocaine and amphetamine treatments (Hurd and Herkenham, 1992; Steiner and Gerfen, 1993; Wang and McGinty, 1995, 1996a). Acute methylphenidate did not produce consistent effects on enkephalin expression (Yano and Steiner, 2005b).

Neurotensin is another neuropeptide expressed in striatal output pathways; it is contained in both types of projection neurons, and its expression is also regulated by D1, D2 and glutamate receptors (see Hanson et al., 1992; Alburges et al., 2011). Its apparent interactions with the mesolimbic and mesostriatal dopamine systems suggest that neurotensin may influence the addictive properties of psychostimulants (cf. Alburges et al., 2011). Both cocaine and amphetamine treatments produce increased neurotensin expression (e.g., Letter et al., 1987; Hanson et al., 1989; Gygi et al., 1994), and a recent study shows that methylphenidate increases neurotensin expression as well (Alburges et al., 2011).

Overall, the molecular effects of methylphenidate described above typically emerged with doses of ≥2 mg/kg (i.p. or s.c.), and juvenile/adolescent rodents tended to be more sensitive than adults, as is true with other psychostimulants (Yano and Steiner, 2007; Carrey and Wilkinson, 2011).

5.2. Corticostriatal circuits affected

5.2.1. Functional domains of the striatum

To determine which corticostriatal circuits/functional domains are affected by methylphenidate treatment, we first mapped gene regulation throughout the striatum (Yano and Steiner, 2005a,b; Cotterly et al., 2007). We assessed the same 23 striatal sectors (Figure 1A), reflecting specific corticostriatal circuits from the rostral to the caudal striatum, as in our cocaine studies (Willuhn et al., 2003; Unal et al., 2009) to allow direct comparisons. Our findings show that, overall, methylphenidate- and cocaine-induced gene regulation in the striatum display a similar but not identical topography (Figure 2), as follows:

  1. Similar to cocaine, methylphenidate produces the most robust changes in gene expression in sensorimotor sectors of the middle and caudal striatum. Maximal effects are present in the dorsal sectors (Figure 2B) that receive the densest input from the medial agranular cortex. However, unlike cocaine-induced gene regulation, which peaks in the postcommissural caudal striatum (corresponding to the middle-to-caudal putamen) (Willuhn et al., 2003; Unal et al., 2009), methylphenidate-induced gene regulation peaks in somewhat more rostral parts of the sensorimotor striatum (Figure 2B) (Yano and Steiner, 2005a,b; Cotterly et al., 2007).

  2. For both drugs, medial and rostral (associative) sectors are also affected to some extent, although they appear to be more changed by methylphenidate than by cocaine (Figure 2).

  3. Similar to cocaine (Willuhn et al., 2003; Unal et al., 2009), small or no effects are seen in ventral striatal sectors on all rostrocaudal levels (Figure 2B). In the nucleus accumbens, methylphenidate-induced gene regulation appears even less robust than that induced by cocaine; however, the most prominent effects were again found in the lateral part of the shell (Figure 2B) (Brandon and Steiner, 2003; Yano and Steiner, 2005a,b; Cotterly et al., 2007). Such differential gene regulation between sensorimotor striatum and nucleus accumbens, with pronounced effects in the former and minor or no effects in the latter, was also found by others (neurotensin, Alburges et al., 2011; IEGs, e.g., Lin et al., 1996; Trinh et al., 2003; Chase et al., 2005b).

5.2.2. Striatal cell types and dopamine receptors

The striatal output pathways affected by methylphenidate require confirmation by double-labeling studies, but the robust effects on substance P expression (Figure 3) (Brandon and Steiner, 2003; Yano and Steiner, 2005b) strongly indicate that, in line with other psychostimulants, methylphenidate alters gene expression in neurons of the D1 receptor-regulated direct pathway (Figure 1B). This conclusion is supported by dopamine receptor antagonist studies showing that blockade of D1 receptors in the striatum eliminates methylphenidate-induced IEG (Yano et al., 2006) and neuropeptide expression (Alburges et al., 2011). Again similar to other psychostimulants (Ruskin and Marshall, 1994; see Section 3.1.4), D2 receptor stimulation also appears to facilitate such gene regulation (Alburges et al., 2011).

Consistent with the above findings, a recent study in bacterial artificial chromosome-transgenic D1- or D2-EGFP-expressing3 mice found that repeated methylphenidate treatment increased FosB expression in neurons of the direct pathway (D1) but not the indirect pathway (D2) (Kim et al., 2009). However, dendritic spine densities in the nucleus accumbens were increased in both subtypes of projection neurons (Kim et al., 2009).

Together with the dearth of effects on enkephalin expression (indirect pathway) (Brandon and Steiner, 2003; Yano and Steiner, 2005b; Van Waes et al., 2012a), these findings indicate that methylphenidate may more selectively affect direct pathway neurons than do amphetamine and cocaine (for possible mechanisms, see Van Waes et al., 2012a).

5.3. Relationship between gene regulation in striatum and cortex

As mentioned above, and as with other psychostimulants, methylphenidate produces IEG induction also in other brain areas, particularly the cortex (Figure 2B) (Lin et al., 1996; Chase et al., 2005b; Yano and Steiner, 2005a; Banerjee et al., 2009).

We mapped methylphenidate-induced IEG expression throughout the major functional subdivisions of the rat cortex (22 areas on four rostrocaudal levels; Figure 1A) (Yano and Steiner, 2005a; Cotterly et al., 2007). Our results show that acute methylphenidate induces IEG expression most robustly in the medial agranular (M2; premotor) and cingulate cortex (Figure 2B), followed closely by motor and somatosensory areas, with minor effects in the insular cortex (Yano and Steiner, 2005a). Although the overall topography of these methylphenidate-induced cortical changes was thus similar to that of cocaine, the methylphenidate effects tended to spread more into rostral and medial cortical areas (Yano and Steiner, 2005a; Cotterly et al., 2007) than the effects of cocaine (Unal et al., 2009).

Our results indicate that cortical gene regulation by methylphenidate occurs in similar functional domains as IEG regulation in the striatum (Figure 2B). Indeed, our regional analysis determined that these IEG responses were positively correlated between cortical areas and their striatal target sectors, confirming that specific corticostriatal projections are affected (Figure 1A) (Yano and Steiner, 2005a; Cotterly et al., 2007). Cortical IEG regulation was also correlated with striatal substance P and dynorphin induction (striatonigral neurons) but not with enkephalin expression (striatopallidal neurons) (Yano and Steiner, 2005a). These findings thus indicate coordinated methylphenidate-induced neuroplasticity between the cortex and neurons of the direct (but not indirect) striatal output pathway.

5.4. Effects of repeated methylphenidate treatment

5.4.1. Blunted gene inducibility

As discussed in the section on amphetamine and cocaine effects, a well-established neuroadaptation that occurs during repeated psychostimulant treatment is blunting (repression) of gene inducibility. Repeated methylphenidate treatment produces a similar effect. Methylphenidate-induced blunting of gene induction in the striatum has been demonstrated, for example, for c-Fos, Zif268, Arc, and substance P (Brandon and Steiner, 2003; Chase et al., 2003, 2007; Hawken et al., 2004; Cotterly et al., 2007) and can last more than 4 weeks (Chase et al., 2005a). As with repeated cocaine treatment (Unal et al., 2009), the degree of blunting after repeated methylphenidate treatment is directly related to the strength of the acute gene response in a particular striatal region (Cotterly et al., 2007).

Most often gene blunting after repeated methylphenidate treatment has been demonstrated by the (reduced) response to a subsequent methylphenidate challenge. However, given that these drugs share some of their neurochemical effects (see Section 2.1), it is not surprising that repeated methylphenidate pretreatment also results in blunted gene induction by a cocaine challenge (Brandon and Steiner, 2003).

The mechanisms underlying gene blunting by these two drugs, however, may not be identical. For example, repeated cocaine treatment blunted striatal Zif268 and Homer 1a induction to a similar extent (Unal et al., 2009), while repeated methylphenidate treatment (10 mg/kg, i.p., 7 days) produced significant blunting of striatal Zif268 induction but minimal changes in Homer 1a induction (Cotterly et al., 2007).

5.4.2. Alternative splicing: accumulation of deltaFosB

DeltaFosB accumulation in striatal neurons after repeated amphetamine and cocaine treatment is well established (e.g., Hope et al., 1994; Nye et al., 1995; McClung et al., 2004). Repeated methylphenidate treatment also increases levels of FosB immunoreactivity in the striatum and cortex (Chase et al., 2005a,b; Kim et al., 2009). In the striatum, the increased FosB signal was selectively present in striatonigral (D1) neurons (Kim et al., 2009). This immunoreactivity is thought to reflect deltaFosB (Kim et al., 2009), but this remains to be confirmed.

5.4.3. Increased dynorphin expression

As mentioned above, in contrast to cocaine and amphetamine, a single methylphenidate injection caused only a modest increase in dynorphin expression in the striatum (Yano and Steiner, 2005b). Consistent with this finding, recent studies indicate that repeated methylphenidate treatment also produces more modest upregulation of dynorphin expression compared with cocaine and amphetamine. For example, a study using reverse-transcription polymerase chain reaction to measure gene expression failed to find altered striatal dynorphin expression after daily methylphenidate treatment with a low dose (2 mg/kg, i.p.; adolescent rats) for 16 days (Adriani et al., 2006a). Another investigation demonstrated that methylphenidate treatment with a high dose (10 mg/kg, i.p.; adolescent rats) once daily for 7 days, which produced robust blunting of IEG and substance P induction, resulted in a significant but more limited (compared with cocaine and amphetamine effects) increase in dynorphin expression (Brandon and Steiner, 2003). A more aggressive methylphenidate treatment (four injections of 10 mg/kg, s.c., over 6 hours) produced increased dynorphin peptide levels (immunoreactivity) in the striatum and substantia nigra 18 hours later (Alburges et al., 2011). This treatment also enhanced neurotensin expression in striatal output pathways (Alburges et al., 2011).

The findings above, together with the unchanged Homer 1a regulation (Cotterly et al., 2007), indicate that some genes are less affected by methylphenidate than by cocaine or amphetamine. Below (Section 7) we discuss potential mechanisms underlying these differential effects and possible clinical relevance.

5.5. Gene regulation by oral methylphenidate treatment

Since the first demonstration of gene regulation by methylphenidate through oral administration (Lin et al., 1996), few studies have assessed oral effects (Carrey and Wilkinson, 2011). Given that therapeutic use of methylphenidate typically involves oral administration, a recent study investigated whether oral treatment (in freely moving prepubertal rats) with doses that produced clinically relevant methylphenidate blood levels would alter gene regulation in the striatum (Chase et al., 2007). The results showed that acute administration of 7.5 to 10 mg/kg (p.o.), but not 2.5 to 5 mg/kg, induced robust IEG expression (Arc) in the striatum. However, although repeated s.c. injections of 7.5 mg/kg of methylphenidate did cause blunting of Arc induction, 14 days of daily oral treatment with this threshold dose did not attenuate Arc inducibility (Chase et al., 2007). The investigators did not examine higher doses (or other genes).

Future studies will have to determine whether this effect reflected a qualitatively different potential for neuroadaptations by oral treatment compared with injected methylphenidate, or was, more likely, simply a consequence of slower uptake and too low methylphenidate plasma levels after oral administration of this threshold dose.

5.6. Methylphenidate effects: conclusions

The reviewed findings indicate that methylphenidate, even in therapeutically relevant doses, can produce changes in gene regulation in cortical and striatal neurons that are qualitatively similar to those of cocaine and amphetamine, although some of the investigated genes seem to be less affected. Methylphenidate also appears to alter the same corticostriatal circuits/functional domains; these are mostly sensorimotor and to some degree associative domains. Overall, these findings are consistent with an addiction liability for methylphenidate, if reduced compared with cocaine and amphetamine (Svetlov et al., 2007).

6. Gene regulation by modafinil

Modafinil is a relatively novel agent that promotes wakefulness and is thus widely used to treat excessive daytime sleepiness associated with narcolepsy and other sleep disorders, but it is also gaining popularity as a cognitive enhancer (see Introduction). Its effects on gene regulation have been described in only a handful of studies.

To our knowledge, the first study to indicate such molecular effects showed increased expression of glutamine synthetase, an enzyme involved in brain metabolism, after a single injection of modafinil in rats (Touret et al., 1994). More recently, a gene microarray study identified several molecule classes affected by modafinil, including transcription factors such as c-Fos (Hasan et al., 2009), and thus suggested that modafinil exposure may alter various neuronal processes. However, most studies to date have assessed only c-Fos expression (Fos immunoreactivity) as a marker to identify neuronal systems involved in the regulation of sleep and wakefulness.

The 1996 study by Lin and colleagues showed that in adult cats modafinil (5 mg/kg, p.o.) produced minor c-Fos induction in striatum and cortex (i.e., considerably less than induced by methylphenidate [2.5 mg/kg, p.o.] and amphetamine [1 mg/kg, p.o.], despite causing similar wakefulness), but induced pronounced c-Fos expression in the hypothalamus and other brain regions (Lin et al., 1996). A study in rats (300 mg/kg, i.p.; Engber et al., 1998) confirmed these effects.

More recent work found c-Fos induction by modafinil (75–300 mg/kg, i.p.) in various nuclei from the hypothalamus to the brainstem, but also described considerable induction in the cortex and striatum in mice (Willie et al., 2005; Hasan et al., 2009) and rats (Scammell et al., 2000; Fiocchi et al., 2009). Based on findings with other psychostimulants, which typically show correlated regulation of several genes (Steiner and Gerfen, 1993; Willuhn et al., 2003; Yano and Steiner, 2005a,b; Unal et al., 2009), it is likely that modafinil alters the expression of various genes in concert in these brain regions.

Little is known about regional variations in the cortex and striatum or about the cell types and receptors involved. Given that less than 3% of striatal neurons are interneurons (Oorschot, 2010), it is clear that striatal c-Fos induction in these studies also predominantly occurred in projection neurons. In most studies, c-Fos induction in the dorsal striatum was abundant, while the nucleus accumbens showed only modest (Willie et al., 2005) or no induction (Scammell et al., 2000; Fiocchi et al., 2009). A recent mapping study in rats used a modafinil dose (10 mg/kg, i.v.) that produced clinically relevant plasma levels and found the most robust increase in c-Fos expression in the dorsomedial striatum, with a significant c-Fos response also in the nucleus accumbens shell (but not core) and the cingulate cortex (Gozzi et al., 2012).

Based on the mechanisms underlying gene regulation by other dopamine-enhancing drugs, it can be assumed that dopamine receptors are also important for modafinil-induced gene regulation, although this remains to be determined. However, in support of this notion, recent studies showed that D1 (and D2) receptors are important for modafinil-induced increases in motivation and arousal (Qu et al., 2008; Young and Geyer, 2010). Future studies will have to elucidate which corticostriatal circuits are affected and identify the mechanisms underlying these molecular changes.

In summary, knowledge on the molecular effects of modafinil in corticostriatal circuits is currently limited, but early findings suggest that this psychostimulant may have the potential to produce effects that are qualitatively similar to those of cocaine, amphetamine, and methylphenidate.

7. Drug interactions: SSRI antidepressants potentiate methylphenidate-induced gene regulation

As discussed, drug-induced gene regulation in neurons critically depends on the neurochemical effects of the drug. In this section we address an aspect of drug treatments that is often overlooked in the assessment of addiction liability: drug interactions based on the neurochemical effects.

If a combination drug treatment results in altered net neurochemical effects, modified gene regulation and thus presumably addiction liability should be expected. Such drug interactions in gene regulation have recently been shown for methylphenidate and certain prescription medications that modify serotonin transmission. Methylphenidate alone increases dopamine overflow but does not affect serotonin (e.g., Kuczenski and Segal, 1997; Borycz et al., 2008; see Section 2.1) and appears to have a reduced propensity to produce neuroadaptations compared with cocaine and amphetamine. In contrast to methylphenidate, cocaine and amphetamine elevate extracellular serotonin levels as well (Yano and Steiner, 2007). Would a combination treatment of methylphenidate with a drug that enhances serotonin action therefore produce more cocaine-/amphetamine-like gene regulation?

A host of findings support this possibility. For example, studies have shown that serotonin contributes significantly to various behavioral effects of cocaine (for reviews, see Filip et al., 2005; Muller and Huston, 2006; Carey et al., 2008). Similarly, whereas dopamine is critical for cocaine-induced gene regulation in the striatum (see Section 3.1.4), serotonin facilitates such effects (Bhat and Baraban, 1993). Thus, attenuation of the serotonin transmission by transmitter depletion (Bhat and Baraban, 1993), receptor antagonism (Lucas et al., 1997; Castanon et al., 2000), or receptor deletion (Lucas et al., 1997) reduces IEG induction by cocaine in the striatum. Conversely, direct and indirect serotonin receptor agonists increase the expression of IEGs (Li and Rowland, 1993; Torres and Rivier, 1994; Wirtshafter and Cook, 1998; Gardier et al., 2000) and other genes (e.g., Mijnster et al., 1998; Morris et al., 1988; Walker et al., 1996) in the striatum.

We therefore investigated whether enhancing serotonin transmission by an SSRI (selective serotonin reuptake inhibitor) antidepressant in conjunction with methylphenidate treatment would modify methylphenidate-induced gene regulation. Our results show that this is indeed the case: adding an SSRI (fluoxetine or citalopram) to methylphenidate treatment potentiates acute induction of IEGs (Steiner et al., 2010; Van Waes et al., 2010), and substance P and dynorphin (but not enkephalin) (Van Waes et al., 2012a) in the striatum (Figure 4). Moreover, repeated treatment with the methylphenidate+SSRI combination produced potentiated blunting of IEG inducibility and increased dynorphin expression (Van Waes et al., 2012b). This SSRI potentiation of methylphenidate-induced gene regulation was present in most striatal sectors (Figure 4B) but was maximal in the lateral sensorimotor striatum (Van Waes et al., 2010; Van Waes et al., 2012a), mimicking cocaine effects. Behaviorally, these SSRIs potentiated methylphenidate-induced locomotion (Borycz et al., 2008) and stereotypies (Van Waes et al., 2010), and produced other behavioral changes (e.g., enhanced sensitivity to cocaine reward and stress-eliciting situations; Warren et al., 2011; see Section 8.1).

An external file that holds a picture, illustration, etc.
Object name is nihms419755f4.jpg

Fluoxetine potentiates methylphenidate-induced gene regulation. (A) Film autoradiograms depict expression of Zif268 in the middle striatum for rats that received a single injection of vehicle (V), methylphenidate (MP, 5 mg/kg), fluoxetine (FLX, 5 mg/kg), or a combination of methylphenidate plus fluoxetine (Van Waes et al., 2010). (B) Association between the potentiation of Zif268 expression (at 40 minutes; Van Waes et al., 2010) and that of substance P expression (at 90 minutes after drug injection; Van Waes et al., 2012a) in the 23 striatal sectors (open diamonds, sensorimotor; full circles, non-sensorimotor; data expressed as percentage of maximal increase). Potentiation is the difference between MP+FLX and MP groups. The potentiation was most robust in sectors of the sensorimotor striatum (open diamonds). *** p < 0.001

The potential significance of these findings relates to the medical use of methylphenidate and SSRIs. SSRIs such as fluoxetine are among the first-line treatments for several depressive and anxiety disorders (Petersen et al., 2002) and are given to millions of patients in the United States alone every year. As discussed, methylphenidate is used both in the treatment of conditions such as ADHD (Biederman et al., 2007; Swanson and Volkow, 2008) and as a recreational drug and cognitive enhancer (Greely et al., 2008; Kollins, 2008; Wilens et al., 2008). The rate of accidental coexposure due to such overlapping drug use/treatments is unclear, but combination therapies of methylphenidate and an SSRI are indicated for several conditions, including ADHD and anxiety/depression comorbidity (Safer et al., 2003; Bhatara et al., 2004; Kollins, 2008). Methylphenidate is also combined with SSRIs as augmentation therapy in major depressive disorder (e.g., Nelson, 2007; Ishii et al., 2008; Ravindran et al., 2008), as acceleration treatment for SSRIs (e.g., Lavretsky et al., 2003), and as treatment for sexual dysfunction related to SSRIs (e.g., Csoka et al., 2008).

Further studies are necessary to determine how much methylphenidate-SSRI coexposure occurs due to clinical administration or as a result of uncontrolled cognitive enhancer use by patients on SSRIs and whether such coexposure enhances the addiction liability of methylphenidate, as the potentiated gene regulation effects might suggest.

8. Behavioral consequences and clinical considerations

While the potential benefits and ethics of cognitive enhancer use are being debated (e.g., Farah et al., 2004; Greely et al., 2008; Chatterjee, 2009; Harris, 2009; Outram, 2010; Hyman, 2011), developmental neurobiologists and addiction researchers warn that the long-term consequences of protracted use of these psychostimulants, especially during brain development, are hardly understood (e.g., Carlezon and Konradi, 2004; Andersen, 2005; Swanson and Volkow, 2008; Berman et al., 2009) (for reviews of neurobehavioral effects of SSRI exposure during development, see, e.g., Oberlander et al., 2009; Olivier et al., 2011). What are the known behavioral consequences of exposure to psychostimulant cognitive enhancers?

8.1. Findings in animal models

Results from animal studies suggest that repeated psychostimulant exposure during preadolescence and adolescence may predispose the individual to substance use or other mental disorders later in life (Brandon et al., 2001; Bolanos et al., 2003; Carlezon et al., 2003; Wiley et al., 2009). It is clear that methylphenidate, for example, produces behavioral changes in animals that mimic those induced by cocaine and amphetamine (for reviews, see Kollins et al., 2001; Carlezon and Konradi, 2004; Kuczenski and Segal, 2005; Yano and Steiner, 2007). Best established is that, similar to cocaine and amphetamine, repeated methylphenidate pretreatment increases locomotor activity/stereotypy levels induced by a subsequent methylphenidate or cocaine/amphetamine challenge (“sensitization”) (e.g., Kollins et al., 2001; Yano and Steiner, 2007).

Conditioned place preference (CPP) and drug self-administration in animals are two behavioral models that rank among the most relevant for addiction research. The CPP model determines the conditioned rewarding effects of a drug by assessing whether an animal seeks out/prefers (or avoids) a specific environment in which it previously experienced this drug (Tzschentke, 2007). Psychostimulants typically produce conditioned preference, although high doses can be aversive. Pretreatment with cocaine, for example, either facilitates (e.g., Shippenberg and Heidbreder, 1995) or attenuates (or even produces aversion in) subsequent preference conditioning by cocaine (Carlezon et al., 2003), depending on factors such as the conditioning dose and the age of the animal during the pretreatment. Methylphenidate alone also produces conditioned place preference (e.g., Meririnne et al., 2001; Zhu et al., 2011), and methylphenidate pretreatment in adult rats enhances subsequent preference conditioning by methylphenidate (Meririnne et al., 2001). In contrast, studies have shown that methylphenidate pretreatment in preadolescent rats (postnatal day [PND] 20–35) produces place aversion or attenuates preference conditioning by cocaine (Andersen et al., 2002; Carlezon et al., 2003; Wiley et al., 2009), similar to pretreatment with cocaine (Carlezon et al., 2003).

The latter findings are sometimes interpreted as indicating a protective effect of methylphenidate pretreatment during development against psychostimulant abuse later in life. However, according to recent research (Wiley et al., 2009; Warren et al., 2011), such early-life exposure may result in behavioral abnormalities suggestive of impaired mood functions. These include generally decreased responsiveness to rewarding stimuli (similar to anhedonia; Nestler and Carlezon, 2006) and depression-like states (enhanced sensitivity to anxiety- and stress-inducing situations) (Wiley et al., 2009; Warren et al., 2011). Interestingly, combined treatment with SSRIs appears to enhance some and reverse others of these methylphenidate-induced behavioral deficits (Warren et al., 2011).

Most drugs of abuse are also self-administered by animals, and methylphenidate is no exception (Kollins et al., 2001). Pretreatment with cocaine or amphetamine facilitates the animal’s subsequent psychostimulant seeking and self-administration (reviewed in Vezina, 2004). This is also the case for methylphenidate. Thus, repeated methylphenidate pretreatment in preweanling (2 mg/kg, PND 11–20; Crawford et al., 2011), adolescent (2 mg/kg, PND 36–42; Brandon et al., 2001), and adult rats (20 mg/kg; Schenk and Izenwasser, 2002) facilitates subsequent cocaine seeking and self-administration. These findings suggest an enhanced risk for psychostimulant abuse in humans after methylphenidate pretreatment (O’Connor et al., 2011).

Future studies will have to clarify whether the apparently contradictory behavioral findings (aversion in the CPP paradigm vs. enhanced drug self-administration) are related to age differences (developmental stage) during drug exposure, specifics of drug treatments (e.g., dose), or other experimental variables. Alternatively, given access to cocaine, such rats may be more likely to seek and consume the drug despite a diminished rewarding effect—a characteristic of compulsive drug seeking, which is thought to be mediated by the sensorimotor striatum (Vanderschuren and Everitt, 2005).

8.2. Findings in human studies

Do human studies support an increased risk for drug abuse/addiction (substance use disorder, SUD) after exposure to medical psychostimulants? This question has been investigated in young ADHD patients. Despite the remarkably consistent results in animal models, early clinical findings remain equivocal. With the possible exception of an increased risk for smoking, such studies indicated that the risk for SUD was unchanged or even decreased after treatment with psychostimulant medications (e.g., Barkley et al., 2003; Wilens et al., 2003; Kollins, 2008).

However, several issues complicate interpretation of these findings. For one, successful control of symptoms in an ADHD patient likely improves the patient’s educational and societal functioning, and resulting socioeconomic advantages may outweigh (and mask) a treatment-inherent biological risk. There are also technical issues. For example, (unmedicated) ADHD patients already show an enhanced risk for SUD (comorbidity; Kollins, 2008), which statistically would be expected to increase the variance, thus favoring the null hypothesis. Another caveat is the often early assessment of outcomes in the clinical studies (a few years after treatment onset in young patients), whereas there is increasing evidence that neurobiological manifestations of early psychostimulant exposure may appear only later in life (e.g., Bolanos et al., 2003; Tropea et al., 2008; Warren et al., 2011). Thus, conclusions will have to await follow-up studies at an older age (see also Kollins, 2008; Wilens et al., 2008; Berman et al., 2009, for further discussions). Importantly, similar studies in healthy humans who were exposed to psychostimulants, either due to ADHD misdiagnosis or because of cognitive enhancer use, have yet to be conducted.

The reviewed molecular findings indicate that risks may be more serious for abuse of medical psychostimulants. As discussed, the molecular changes induced by psychostimulants are dependent on dose and route of administration (Yano and Steiner, 2007). Proper medical psychostimulant treatment almost always involves oral drug administration, which results in lower drug levels in the brain (Swanson and Volkow, 2003; Kuczenski and Segal, 2005; Carrey and Wilkinson, 2011) and thus a lower risk for neuroadaptations. This contrasts with cognitive enhancer use/abuse. For example, studies show that in recreational settings, intranasal use (snorting of ground-up pills) is not uncommon (e.g., 38.1% prevalence in users among college students; Teter et al., 2006), and intravenous administration also occurs (Parran and Jasinski, 1991; Babcock and Byrne, 2000; Barrett et al., 2005; Teter et al., 2006; White et al., 2006; for a review, see Kollins et al., 2001). These latter routes of administration result in exposure to much faster and higher drug peak levels and are thus expected to have a greater potential for inducing maladaptive neuronal plasticity (Samaha and Robinson, 2005) and enhanced addiction liability. Studies in healthy subjects are needed to evaluate the safety concerns related to cognitive enhancer use and abuse.

9. Conclusions

The findings we have summarized show that psychostimulants such as cocaine and amphetamine produce changes in gene regulation in specific corticostriatal circuits. These effects are most robust in sensorimotor circuits (which are implicated in habit formation and compulsive aspects of drug taking), less pronounced in associative circuits, and more modest in limbic circuits (where they are thought to contribute to altered reward processing in addiction). At the cellular level, psychostimulants alter predominantly neurons of the direct striatal output pathway (“Go pathway”), while the indirect pathway is less affected, and those changes appear to depend on the treatment context (i.e., arousal and associated changes in excitatory striatal input). Overall, these findings support the notion that action selection and initiation are compromised in addiction.

Comparative studies on cognitive enhancers (Adderall, methylphenidate, modafinil) show that they can induce largely similar molecular changes in corticostriatal circuits. Moreover, the same functional domains, cell types, neurotransmitters, and receptors appear to be affected. These effects were mostly explored with high drug doses, but qualitatively similar molecular changes were evident with drug treatments that mimicked medical treatments (oral administration and low drug plasma levels), although it is not clear whether these molecular changes are robust enough to alter behavior. Drug levels associated with cognitive enhancer abuse, however, are likely higher and, with protracted use, likely contribute to molecular changes that increase the addiction liability of these drugs. Further research is necessary to resolve these questions.

Highlights

  • Medical psychostimulants (methylphenidate, amphetamine, modafinil) are used as cognitive enhancers

  • These drugs can alter gene regulation in corticostriatal circuits similar to cocaine

  • These molecular changes may contribute to the addiction liability of cognitive enhancers

Acknowledgments

Our research summarized in this review was supported by grant DA011261 from the National Institute on Drug Abuse (H.S.). We thank the reviewers for their thoughtful suggestions that have helped improve our paper.

Abbreviations

ADHDattention-deficit hyperactivity disorder
IEGimmediate-early gene
M2medial agranular cortex
mRNAmessenger RNA
SUDsubstance use disorder

Footnotes

1Data available at the DEA website, www.deadiversion.usdoj.gov/fed_regs/quotas/2009/fr10212.htm, accessed August 2, 2012.

2We do not specifically address the molecular effects of methamphetamine and related compounds, which are also approved by the US Food and Drug Administration for the treatment of ADHD and other conditions (Berman et al., 2009). Many methamphetamine effects on gene regulation are similar to those of amphetamine; we refer the interested reader to a recent review by Keefe and Horner (2010[0]) on this topic.

3EGFP, enhanced green fluorescent protein

Conflict of interest statement

There is no conflict of interest.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • Adams DH, Hanson GR, Keefe KA. Differential effects of cocaine and methamphetamine on neurotensin/neuromedin N and preprotachykinin messenger RNA expression in unique regions of the striatum. Neuroscience. 2001;102:843–851. [PubMed] [Google Scholar]
  • Adams DH, Hanson GR, Keefe KA. Distinct effects of methamphetamine and cocaine on preprodynorphin messenger RNA in rat striatal patch and matrix. J Neurochem. 2003;84:87–93. [PubMed] [Google Scholar]
  • Adriani W, Leo D, Greco D, Rea M, di Porzio U, Laviola G, Perrone-Capano C. Methylphenidate administration to adolescent rats determines plastic changes in reward-related behavior and striatal gene expression. Neuropsychopharmacology. 2006a;31:1946–1956. [PubMed] [Google Scholar]
  • Adriani W, Leo D, Guarino M, Natoli A, Di Consiglio E, De Angelis G, Traina E, Testai E, Perrone-Capano C, Laviola G. Short-term effects of adolescent methylphenidate exposure on brain striatal gene expression and sexual/endocrine parameters in male rats. Ann N Y Acad Sci. 2006b;1074:52–73. [PubMed] [Google Scholar]
  • Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989;12:366–375. [PubMed] [Google Scholar]
  • Alburges ME, Hoonakker AJ, Horner KA, Fleckenstein AE, Hanson GR. Methylphenidate alters basal ganglia neurotensin systems through dopaminergic mechanisms: A comparison with cocaine treatment. J Neurochem. 2011;117:470–478. [PubMed] [Google Scholar]
  • Alexander GE, Crutcher MD, DeLong MR. Basal ganglia-thalamocortical circuits: parallel substrates for motor, oculomotor, “prefrontal” and “limbic” functions. Prog Brain Res. 1990;85:119–146. [PubMed] [Google Scholar]
  • Alexander GE, DeLong MR, Strick PL. Parallel organization of functionally segregated circuits linking basal ganglia and cortex. Annu Rev Neurosci. 1986;9:357–381. [PubMed] [Google Scholar]
  • Allen JK, Wilkinson M, Soo EC, Hui JP, Chase TD, Carrey N. Chronic low dose Adderall XR down-regulates cfos expression in infantile and prepubertal rat striatum and cortex. Neuroscience. 2010;169:1901–1912. [PubMed] [Google Scholar]
  • Andersen SL. Stimulants and the developing brain. Trends Pharmacol Sci. 2005;26:237–243. [PubMed] [Google Scholar]
  • Andersen SL, Arvanitogiannis A, Pliakas AM, LeBlanc C, Carlezon WAJ. Altered responsiveness to cocaine in rats exposed to methylphenidate during development. Nat Neurosci. 2002;5:13–14. [PubMed] [Google Scholar]
  • Andersson M, Hilbertson A, Cenci MA. Striatal fosB expression is causally linked with l-DOPA-induced abnormal involuntary movements and the associated upregulation of striatal prodynorphin mRNA in a rat model of Parkinson’s disease. Neurobiol Dis. 1999;6:461–474. [PubMed] [Google Scholar]
  • Andersson M, Westin JE, Cenci MA. Time course of striatal DeltaFosB-like immunoreactivity and prodynorphin mRNA levels after discontinuation of chronic dopaminomimetic treatment. Eur J Neurosci. 2003;17:661–666. [PubMed] [Google Scholar]
  • Babcock Q, Byrne T. Student perceptions of methylphenidate abuse at a public liberal arts college. J Am Coll Health. 2000;49:143–145. [PubMed] [Google Scholar]
  • Badiani A, Oates MM, Day HE, Watson SJ, Akil H, Robinson TE. Amphetamine-induced behavior, dopamine release, and c-fos mRNA expression: modulation by environmental novelty. J Neurosci. 1998;18:10579–10593. [PMC free article] [PubMed] [Google Scholar]
  • Badiani A, Oates MM, Day HE, Watson SJ, Akil H, Robinson TE. Environmental modulation of amphetamine-induced c-fos expression in D1 versus D2 striatal neurons. Behav Brain Res. 1999;103:203–209. [PubMed] [Google Scholar]
  • Banerjee PS, Aston J, Khundakar AA, Zetterström TS. Differential regulation of psychostimulant-induced gene expression of brain derived neurotrophic factor and the immediate-early gene Arc in the juvenile and adult brain. Eur J Neurosci. 2009;29:465–476. [PubMed] [Google Scholar]
  • Barkley RA, Fischer M, Smallish L, Fletcher K. Does the treatment of attention-deficit/hyperactivity disorder with stimulants contribute to drug use/abuse? A 13-year prospective study. Pediatrics. 2003;111:97–109. [PubMed] [Google Scholar]
  • Barrett SP, Darredeau C, Bordy LE, Pihl RO. Characteristics of methylphenidate misuse in a university student sample. Can J Psychiatry. 2005;51:126–127. [PubMed] [Google Scholar]
  • Belin D, Everitt BJ. Drug addiction: The neural and psychological basis of a compulsive incentive habit. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Elsevier; London: 2010. pp. 571–592. [Google Scholar]
  • Berendse HW, Galis-de Graaf Y, Groenewegen HJ. Topographical organization and relationship with ventral striatal compartments of prefrontal corticostriatal projections in the rat. J Comp Neurol. 1992;316:314–347. [PubMed] [Google Scholar]
  • Berke JD, Hyman SE. Addiction, dopamine, and the molecular mechanisms of memory. Neuron. 2000;25:515–532. [PubMed] [Google Scholar]
  • Berke JD, Paletzki RF, Aronson GJ, Hyman SE, Gerfen CR. A complex program of striatal gene expression induced by dopaminergic stimulation. J Neurosci. 1998;18:5301–5310. [PMC free article] [PubMed] [Google Scholar]
  • Berlanga ML, Olsen CM, Chen V, Ikegami A, Herring BE, Duvauchelle CL, Alcantara AA. Cholinergic interneurons of the nucleus accumbens and dorsal striatum are activated by the self-administration of cocaine. Neuroscience. 2003;120:1149–1156. [PubMed] [Google Scholar]
  • Berman SM, Kuczenski R, McCracken JT, London ED. Potential adverse effects of amphetamine treatment on brain and behavior: a review. Mol Psychiatry. 2009;14:123–142. [PMC free article] [PubMed] [Google Scholar]
  • Berretta S, Robertson HA, Graybiel AM. Dopamine and glutamate agonists stimulate neuron-specific expression of Fos-like protein in the striatum. J Neurophysiol. 1992;68:767–777. [PubMed] [Google Scholar]
  • Berridge CW, Devilbiss DM. Psychostimulants as cognitive enhancers: The prefrontal cortex, catecholamines, and attention-deficit/hyperactivity disorder. Biol Psychiatry. 2011;69:e101–111. [PMC free article] [PubMed] [Google Scholar]
  • Berridge CW, Devilbiss DM, Andrzejewski ME, Arnsten AF, Kelley AE, Schmeichel B, Hamilton C, Spencer RC. Methylphenidate preferentially increases catecholamine neurotransmission within the prefrontal cortex at low doses that enhance cognitive function. Biol Psychiatry. 2006;60:1111–1120. [PubMed] [Google Scholar]
  • Beveridge TJ, Smith HR, Daunais JB, Nader MA, Porrino LJ. Chronic cocaine self-administration is associated with altered functional activity in the temporal lobes of non human primates. Eur J Neurosci. 2006;23:3109–3118. [PubMed] [Google Scholar]
  • Bhat RV, Baraban JM. Activation of transcription factor genes in striatum by cocaine: role of both serotonin and dopamine systems. J Pharmacol Exp Ther. 1993;267:496–505. [PubMed] [Google Scholar]
  • Bhatara V, Feil M, Hoagwood K, Vitiello B, Zima B. National trends in concomitant psychotropic medication with stimulants in pediatric visits: practice versus knowledge. J Atten Disord. 2004;7:217–226. [PubMed] [Google Scholar]
  • Biederman J, Wilens TE, Spencer TJ, Adler LA. Diagnosis and treatment of adults with attention-deficit/hyperactivity disorder. CNS Spectr. 2007;12:1–15. [Google Scholar]
  • Black YD, Maclaren FR, Naydenov AV, Carlezon WAJ, Baxter MG, Konradi C. Altered attention and prefrontal cortex gene expression in rats after binge-like exposure to cocaine during adolescence. J Neurosci. 2006;26:9656–9665. [PMC free article] [PubMed] [Google Scholar]
  • Bolanos CA, Barrot M, Berton O, Wallace-Black D, Nestler EJ. Methylphenidate treatment during pre- and periadolescence alters behavioral responses to emotional stimuli at adulthood. Biol Psychiatry. 2003;54:1317–1329. [PubMed] [Google Scholar]
  • Borycz J, Zapata A, Quiroz C, Volkow ND, Ferré S. 5-HT(1B) receptor-mediated serotoninergic modulation of methylphenidate-induced locomotor activation in rats. Neuropsychopharmacology. 2008;33:619–626. [PubMed] [Google Scholar]
  • Brady KT, Gray KM, Tolliver BK. Cognitive enhancers in the treatment of substance use disorders: clinical evidence. Pharmacol Biochem Behav. 2011;99:285–294. [PMC free article] [PubMed] [Google Scholar]
  • Brandon CL, Marinelli M, Baker LK, White FJ. Enhanced reactivity and vulnerability to cocaine following methylphenidate treatment in adolescent rats. Neuropsychopharmacology. 2001;25:651–661. [PubMed] [Google Scholar]
  • Brandon CL, Steiner H. Repeated methylphenidate treatment in adolescent rats alters gene regulation in the striatum. Eur J Neurosci. 2003;18:1584–1592. [PubMed] [Google Scholar]
  • Breiter HC, Gollub RL, Weisskoff RM, Kennedy DN, Makris N, Berke JD, Goodman JM, Kantor HL, Gastfriend DR, Riorden JP, Mathew RT, Rosen BR, Hyman SE. Acute effects of cocaine on human brain activity and emotion. Neuron. 1997;19:591–611. [PubMed] [Google Scholar]
  • Brog JS, Salyapongse A, Deutch AY, Zahm DS. The patterns of afferent innervation of the core and shell in the “accumbens” part of the rat ventral striatum: immunohistochemical detection of retrogradely transported fluoro-gold. J Comp Neurol. 1993;338:255–278. [PubMed] [Google Scholar]
  • Bronson SE, Konradi C. Second-messenger cascades. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 447–460. [Google Scholar]
  • Butelman ER, Yuferov V, Kreek MJ. κ-opioid receptor/dynorphin system: genetic and pharmacotherapeutic implications for addiction. Trends Neurosci. 2012 Epub. [PMC free article] [PubMed] [Google Scholar]
  • Bymaster FP, Katner JS, Nelson DL, Hemrick-Luecke SK, Threlkeld PG, Heiligenstein JH, Morin SM, Gehlert DR, Perry KW. Atomoxetine increases extracellular levels of norepinephrine and dopamine in prefrontal cortex of rat: a potential mechanism for efficacy in attention deficit/hyperactivity disorder. Neuropsychopharmacology. 2002;27:699–711. [PubMed] [Google Scholar]
  • Caboche J, Roze E, Brami-Cherrier K, Betuing S. Chromatin remodeling: role in neuropathologies of the basal ganglia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 527–545. [Google Scholar]
  • Carey RJ, Huston JP, Müller CP. Pharmacological inhibition of dopamine and serotonin activity blocks spontaneous and cocaine-activated behaviour. Prog Brain Res. 2008;172:347–360. [PubMed] [Google Scholar]
  • Carlezon WAJ, Konradi C. Understanding the neurobiological consequences of early exposure to psychotropic drugs: linking behavior with molecules. Neuropharmacology. 2004;47:47–60. [PMC free article] [PubMed] [Google Scholar]
  • Carlezon WAJ, Mague SD, Andersen SL. Enduring behavioral effects of early exposure to methylphenidate in rats. Biol Psychiatry. 2003;54:1330–1337. [PubMed] [Google Scholar]
  • Carrey N, Wilkinson M. A review of psychostimulant-induced neuroadaptation in developing animals. Neurosci Bull. 2011;27:197–214. [PMC free article] [PubMed] [Google Scholar]
  • Carta AR, Gerfen CR, Steiner H. Cocaine effects on gene regulation in the striatum and behavior: increased sensitivity in D3 dopamine receptor-deficient mice. Neuroreport. 2000;11:2395–2399. [PubMed] [Google Scholar]
  • Castanon N, Scearce-Levie K, Lucas JJ, Rocha B, Hen R. Modulation of the effects of cocaine by 5-HT1B receptors: a comparison of knockouts and antagonists. Pharmacol Biochem Behav. 2000;67:559–566. [PubMed] [Google Scholar]
  • Cenci MA, Björklund A. Transection of corticostriatal afferents reduces amphetamine-and apomorphine-induced striatal Fos expression and turning behaviour in unilaterally 6-hydroxydopamine-lesioned rats. Eur J Neurosci. 1993;5:1062–1070. [PubMed] [Google Scholar]
  • Cenci MA, Campbell K, Wictorin K, Björklund A. Striatal c-fos induction by cocaine or apomorphine occurs preferentially in output neurons projecting to the substantia nigra in the rat. Eur J Neurosci. 1992;4:376–380. [PubMed] [Google Scholar]
  • Chase T, Carrey N, Soo E, Wilkinson M. Methylphenidate regulates activity regulated cytoskeletal associated but not brain-derived neurotrophic factor gene expression in the developing rat striatum. Neuroscience. 2007;144:969–984. [PubMed] [Google Scholar]
  • Chase TD, Brown RE, Carrey N, Wilkinson M. Daily methylphenidate administration attenuates c-fos expression in the striatum of prepubertal rats. Neuroreport. 2003;14:769–772. [PubMed] [Google Scholar]
  • Chase TD, Carrey N, Brown RE, Wilkinson M. Methylphenidate differentially regulates c-fos and fosB expression in the developing rat striatum. Dev Brain Res. 2005a;157:181–191. [PubMed] [Google Scholar]
  • Chase TD, Carrey N, Brown RE, Wilkinson M. Methylphenidate regulates c-fos and fosB expression in multiple regions of the immature rat brain. Dev Brain Res. 2005b;156:1–12. [PubMed] [Google Scholar]
  • Chatterjee A. Is it acceptable for people to take methylphenidate to enhance performance? No BMJ. 2009;338:b1956. [PubMed] [Google Scholar]
  • Chaudhuri A. Neural activity mapping with inducible transcription factors. Neuroreport. 1997;8:v–ix. [PubMed] [Google Scholar]
  • Chevalier G, Deniau JM. Disinhibition as a basic process in the expression of striatal functions. Trends Neurosci. 1990;13:277–280. [PubMed] [Google Scholar]
  • Cole AJ, Bhat RV, Patt C, Worley PF, Baraban JM. D1 dopamine receptor activation of multiple transcription factor genes in rat striatum. J Neurochem. 1992;58:1420–1426. [PubMed] [Google Scholar]
  • Cotterly L, Beverley JA, Yano M, Steiner H. Dysregulation of gene induction in corticostriatal circuits after repeated methylphenidate treatment in adolescent rats: Differential effects on zif 268 and homer 1a. Eur J Neurosci. 2007;25:3617–3628. [PubMed] [Google Scholar]
  • Crawford CA, Baella SA, Farley CM, Herbert MS, Horn LR, Campbell RH, Zavala AR. Early methylphenidate exposure enhances cocaine self-administration but not cocaine-induced conditioned place preference in young adult rats. Psychopharmacology. 2011;213:43–52. [PMC free article] [PubMed] [Google Scholar]
  • Crombag HS, Jedynak JP, Redmond K, Robinson TE, Hope BT. Locomotor sensitization to cocaine is associated with increased Fos expression in the accumbens, but not in the caudate. Behav Brain Res. 2002;136:455–462. [PubMed] [Google Scholar]
  • Csoka A, Bahrick A, Mehtonen OP. Persistent sexual dysfunction after discontinuation of selective serotonin reuptake inhibitors. J Sex Med. 2008;5:227–233. [PubMed] [Google Scholar]
  • Curran EJ, Akil H, Watson SJ. Psychomotor stimulant- and opiate-induced c-fos mRNA expression patterns in the rat forebrain: comparisons between acute drug treatment and a drug challenge in sensitized animals. Neurochem Res. 1996;21:1425–1435. [PubMed] [Google Scholar]
  • Curran EJ, Watson SJ. Dopamine receptor mRNA expression patterns by opioid peptide cells in the nucleus accumbens of the rat: a double in situ hybridization study. J Comp Neurol. 1995;361:57–76. [PubMed] [Google Scholar]
  • Damez-Werno D, Laplant Q, Sun H, Scobie KN, Dietz DM, Walker IM, Koo JW, Vialou VF, Mouzon E, Russo SJ, Nestler EJ. Drug experience epigenetically primes fosb gene inducibility in rat nucleus accumbens. J Neurosci. 2012;32:10267–10272. [PMC free article] [PubMed] [Google Scholar]
  • Daunais JB, McGinty JF. Acute and chronic cocaine administration differentially alters striatal opioid and nuclear transcription factor mRNAs. Synapse. 1994;18:35–45. [PubMed] [Google Scholar]
  • DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990;13:281–285. [PubMed] [Google Scholar]
  • Di Chiara G, Imperato A. Drugs abused by humans preferentially increase synaptic dopamine concentrations in the mesolimbic system of freely moving rats. Proc Natl Acad Sci USA. 1988;85:5274–5278. [PMC free article] [PubMed] [Google Scholar]
  • Diaz Heijtz R, Kolb B, Forssberg H. Can a therapeutic dose of amphetamine during pre-adolescence modify the pattern of synaptic organization in the brain? Eur J Neurosci. 2003;18:3394–3399. [PubMed] [Google Scholar]
  • Dilts RPJ, Helton TE, McGinty JF. Selective induction of Fos and FRA immunoreactivity within the mesolimbic and mesostriatal dopamine terminal fields. Synapse. 1993;13:251–263. [PubMed] [Google Scholar]
  • Drago J, Gerfen CR, Westphal H, Steiner H. D1 dopamine receptor-deficient mouse: Cocaine-induced regulation of immediate-early gene and substance P expression in the striatum. Neuroscience. 1996;74:813–823. [PubMed] [Google Scholar]
  • Engber TM, Koury EJ, Dennis SA, Miller MS, Contreras PC, Bhat RV. Differential patterns of regional c-Fos induction in the rat brain by amphetamine and the novel wakefulness-promoting agent modafinil. Neurosci Lett. 1998;241:95–98. [PubMed] [Google Scholar]
  • Everitt BJ, Robbins TW. Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nat Neurosci. 2005;8:1481–1489. [PubMed] [Google Scholar]
  • Farah MJ, Illes J, Cook-Deegan R, Gardner H, Kandel E, King P, Parens E, Sahakian B, Wolpe PR. Neurocognitive enhancement: what can we do and what should we do? Nat Rev Neurosci. 2004;5:421–425. [PubMed] [Google Scholar]
  • Ferguson SM, Eskenazi D, Ishikawa M, Wanat MJ, Phillips PE, Dong Y, Roth BL, Neumaier JF. Transient neuronal inhibition reveals opposing roles of indirect and direct pathways in sensitization. Nat Neurosci. 2011;14:22–24. [PMC free article] [PubMed] [Google Scholar]
  • Ferguson SM, Robinson TE. Amphetamine-evoked gene expression in striatopallidal neurons: regulation by corticostriatal afferents and the ERK/MAPK signaling cascade. J Neurochem. 2004;91:337–348. [PubMed] [Google Scholar]
  • Ferrario CR, Gorny G, Crombag HS, Li Y, Kolb B, Robinson TE. Neural and behavioral plasticity associated with the transition from controlled to escalated cocaine use. Biol Psychiatry. 2005;58:751–759. [PubMed] [Google Scholar]
  • Filip M, Frankowska M, Zaniewska M, Gołda A, Przegaliński E. The serotonergic system and its role in cocaine addiction. Pharmacol Rep. 2005;57:685–700. [PubMed] [Google Scholar]
  • Fiocchi EM, Lin YG, Aimone L, Gruner JA, Flood DG. Armodafinil promotes wakefulness and activates Fos in rat brain. Pharmacol Biochem Behav. 2009;92:549–557. [PubMed] [Google Scholar]
  • Fleckenstein AE, Volz TJ, Hanson GR. Psychostimulant-induced alterations in vesicular monoamine transporter-2 function: neurotoxic and therapeutic implications. Neuropharmacology. 2009;56(Suppl 1):133–138. [PMC free article] [PubMed] [Google Scholar]
  • Frankel PS, Alburges ME, Bush L, Hanson GR, Kish SJ. Striatal and ventral pallidum dynorphin concentrations are markedly increased in human chronic cocaine users. Neuropharmacology. 2008;55:41–46. [PMC free article] [PubMed] [Google Scholar]
  • Fuchs RA, Branham RK, See RE. Different neural substrates mediate cocaine seeking after abstinence versus extinction training: A critical role for the dorsolateral caudate–putamen. J Neuroscience. 2006;26:3584–3588. [PMC free article] [PubMed] [Google Scholar]
  • Fukui R, Svenningsson P, Matsuishi T, Higashi H, Nairn AC, Greengard P, Nishi A. Effect of methylphenidate on dopamine/DARPP signalling in adult, but not young, mice. J Neurochem. 2003;87:1391–1401. [PubMed] [Google Scholar]
  • Gardier AM, Moratalla R, Cuellar B, Sacerdote M, Guibert B, Lebrec H, Graybiel AM. Interaction between the serotoninergic and dopaminergic systems in d-fenfluramine-induced activation of c-fos and jun B genes in rat striatal neurons. J Neurochem. 2000;74:1363–1373. [PubMed] [Google Scholar]
  • Gatley SJ, Pan D, Chen R, Chaturvedi G, Ding YS. Affinities of methylphenidate derivatives for dopamine, norepinephrine and serotonin transporters. Life Sci. 1996;58:231–239. [PubMed] [Google Scholar]
  • Gerasimov MR, Franceschi M, Volkow ND, Gifford A, Gatley SJ, Marsteller D, Molina PE, Dewey SL. Comparison between intraperitoneal and oral methylphenidate administration: A microdialysis and locomotor activity study. J Pharmacol Exp Ther. 2000;295:51–57. [PubMed] [Google Scholar]
  • Gerdeman GL, Partridge JG, Lupica CR, Lovinger DM. It could be habit forming: drugs of abuse and striatal synaptic plasticity. Trends Neurosci. 2003;26:184–192. [PubMed] [Google Scholar]
  • Gerfen CR, Bolam JP. The neuroanatomical organization of the basal ganglia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 3–28. [Google Scholar]
  • Gerfen CR, Engber TM, Mahan LC, Susel Z, Chase TN, Monsma FJ, Jr, Sibley DR. D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science. 1990;250:1429–1432. [PubMed] [Google Scholar]
  • Gerfen CR, Keefe KA, Gauda EB. D1 and D2 dopamine receptor function in the striatum: coactivation of D1- and D2-dopamine receptors on separate populations of neurons results in potentiated immediate-early gene response in D1-containing neurons. J Neurosci. 1995;15:8167–8176. [PMC free article] [PubMed] [Google Scholar]
  • Goldstein RZ, Woicik PA, Maloney T, Tomasi D, Alia-Klein N, Shan J, Honorio J, Samaras D, Wang R, Telang F, Wang GJ, Volkow ND. Oral methylphenidate normalizes cingulate activity in cocaine addiction during a salient cognitive task. Proc Natl Acad Sci USA. 2010;107:16667–16672. [PMC free article] [PubMed] [Google Scholar]
  • Gozzi A, Colavito V, Seke Etet PF, Montanari D, Fiorini S, Tambalo S, Bifone A, Zucconi GG, Bentivoglio M. Modulation of fronto-cortical activity by modafinil: a functional imaging and fos study in the rat. Neuropsychopharmacology. 2012;37:822–837. [PMC free article] [PubMed] [Google Scholar]
  • Graybiel AM, Canales JJ, Capper-Loup C. Levodopa-induced dyskinesias and dopamine-dependent stereotypies: a new hypothesis. Trends Neurosci. 2000;23:S71–S77. [PubMed] [Google Scholar]
  • Graybiel AM, Moratalla R, Robertson HA. Amphetamine and cocaine induce drug-specific activation of the c-fos gene in striosome-matrix compartments and limbic subdivisions of the striatum. Proc Natl Acad Sci USA. 1990;87:6912–6916. [PMC free article] [PubMed] [Google Scholar]
  • Greely H, Sahakian B, Harris J, Kessler RC, Gazzaniga M, Campbell P, Farah MJ. Towards responsible use of cognitive-enhancing drugs by the healthy. Nature. 2008;456:702–705. [PubMed] [Google Scholar]
  • Groenewegen HJ, Berendse HW, Wolters JG, Lohman AH. The anatomical relationship of the prefrontal cortex with the striatopallidal system, the thalamus and the amygdala: evidence for a parallel organization. Prog Brain Res. 1990;85:95–116. [PubMed] [Google Scholar]
  • Gross NB, Marshall JF. Striatal dopamine and glutamate receptors modulate methamphetamine-induced cortical Fos expression. Neuroscience. 2009;161:1114–1125. [PMC free article] [PubMed] [Google Scholar]
  • Gygi SP, Gibb JW, Hanson GR. Differential effects of antipsychotic and psychotomimetic drugs on neurotensin systems of discrete extrapyramidal and limbic regions. J Pharmacol Exp Ther. 1994;270:192–197. [PubMed] [Google Scholar]
  • Haber SN. The primate basal ganglia: parallel and integrative networks. J Chem Neuroanat. 2003;26:317–330. [PubMed] [Google Scholar]
  • Hanson GR, Bush L, Keefe KA, Alburges ME. Distinct responses of basal ganglia substance P systems to low and high doses of methamphetamine. J Neurochem. 2002;82:1171–1178. [PubMed] [Google Scholar]
  • Hanson GR, Merchant KM, Letter AA, Bush L, Gibb JW. Methamphetamine-induced changes in the striatal-nigral dynorphin system: role of D-1 and D-2 receptors. European J Pharmacol. 1987;144:245–246. [PubMed] [Google Scholar]
  • Hanson GR, Midgley LP, Bush LG, Johnson M, Gibb JW. Comparison of responses by neuropeptide systems in rat to the psychotropic drugs, methamphetamine, cocaine and PCP. NIDA Res Monogr. 1989;95:348. [PubMed] [Google Scholar]
  • Hanson GR, Singh N, Merchant K, Johnson M, Bush L, Gibb JW. Responses of limbic and extrapyramidal neurotensin systems to stimulants of abuse. Involvement of dopaminergic mechanisms. Ann N Y Acad Sci. 1992;668:165–172. [PubMed] [Google Scholar]
  • Hanson GR, Singh N, Merchant K, Johnson M, Gibb JW. The role of NMDA receptor systems in neuropeptide responses to stimulants of abuse. Drug Alcohol Depend. 1995;37:107–110. [PubMed] [Google Scholar]
  • Harlan RE, Garcia MM. Drugs of abuse and immediate-early genes in the forebrain. Mol Neurobiol. 1998;16:221–267. [PubMed] [Google Scholar]
  • Harris J. Is it acceptable for people to take methylphenidate to enhance performance? Yes BMJ. 2009;338:b1955. [PubMed] [Google Scholar]
  • Hasan S, Pradervand S, Ahnaou A, Drinkenburg W, Tafti M, Franken P. How to keep the brain awake? The complex molecular pharmacogenetics of wake promotion. Neuropsychopharmacology. 2009;34:1625–1640. [PubMed] [Google Scholar]
  • Hawken CM, Brown RE, Carrey N, Wilkinson M. Long-term methylphenidate treatment down-regulates c-fos in the striatum of male CD-1 mice. Neuroreport. 2004;15:1045–1048. [PubMed] [Google Scholar]
  • Heiman M, Schaefer A, Gong S, Peterson JD, Day M, Ramsey KE, Suárez-Fariñas M, Schwarz C, Stephan DA, Surmeier DJ, Greengard P, Heintz N. A translational profiling approach for the molecular characterization of CNS cell types. Cell. 2008;135:738–748. [PMC free article] [PubMed] [Google Scholar]
  • Hope B, Kosofsky B, Hyman SE, Nestler EJ. Regulation of immediate early gene expression and AP-1 binding in the rat nucleus accumbens by chronic cocaine. Proc Natl Acad Sci USA. 1992;89:5764–5768. [PMC free article] [PubMed] [Google Scholar]
  • Hope BT, Nye HE, Kelz MB, Self DW, Iadarola MJ, Nakabeppu Y, Duman RS, Nestler EJ. Induction of a long-lasting AP-1 complex composed of altered Fos-like proteins in brain by chronic cocaine and other chronic treatments. Neuron. 1994;13:1235–1244. [PubMed] [Google Scholar]
  • Hurd YL, Herkenham M. Influence of a single injection of cocaine, amphetamine or GBR 12909 on mRNA expression of striatal neuropeptides. Mol Brain Res. 1992;16:97–104. [PubMed] [Google Scholar]
  • Hurd YL, Herkenham M. Molecular alterations in the neostriatum of human cocaine addicts. Synapse. 1993;13:357–369. [PubMed] [Google Scholar]
  • Hurd YL, Ungerstedt U. In vivo neurochemical profile of dopamine uptake inhibitors and releasers in rat caudate-putamen. Eur J Pharmacol. 1989;166:251–260. [PubMed] [Google Scholar]
  • Husain M, Mehta MA. Cognitive enhancement by drugs in health and disease. Trends Cogn Sci. 2011;15:28–36. [PMC free article] [PubMed] [Google Scholar]
  • Hyman SE. Addiction: a disease of learning and memory. Am J Psychiatry. 2005;162:1414–1422. [PubMed] [Google Scholar]
  • Hyman SE. Cognitive enhancement: promises and perils. Neuron. 2011;69:595–598. [PubMed] [Google Scholar]
  • Hyman SE, Cole RL, Schwarzschild M, Cole D, Hope B, Konradi C. Molecular mechanisms of striatal gene regulation: a critical role for glutamate in dopamine-mediated gene induction. In: Merchant KM, editor. Pharmacological Regulation of Gene Expression in the CNS. CRC; Boca Raton: 1996. pp. 115–131. [Google Scholar]
  • Hyman SE, Malenka RC. Addiction and the brain: the neurobiology of compulsion and its persistence. Nat Rev Neurosci. 2001;2:695–703. [PubMed] [Google Scholar]
  • Hyman SE, Nestler EJ. Initiation and adaptation: a paradigm for understanding psychotropic drug action. Am J Psychiatry. 1996;153:151–162. [PubMed] [Google Scholar]
  • Ishii M, Tatsuzawa Y, Yoshino A, Nomura S. Serotonin syndrome induced by augmentation of SSRI with methylphenidate. Psychiatry Clin Neurosci. 2008;62:246. [PubMed] [Google Scholar]
  • Jaber M, Cador M, Dumartin B, Normand E, Stinus L, Bloch B. Acute and chronic amphetamine treatments differently regulate neuropeptide messenger RNA levels and Fos immunoreactivity in rat striatal neurons. Neuroscience. 1995;65:1041–1050. [PubMed] [Google Scholar]
  • Jedynak JP, Uslaner JM, Esteban JA, Robinson TE. Methamphetamine-induced structural plasticity in the dorsal striatum. Eur J Neurosci. 2007;25:847–853. [PubMed] [Google Scholar]
  • Joel D, Weiner I. The organization of the basal ganglia-thalamocortical circuits: open interconnected rather than closed segregated. Neuroscience. 1994;63:363–379. [PubMed] [Google Scholar]
  • Johansson B, Lindström K, Fredholm BB. Differences in the regional and cellular localization of c-fos messenger RNA induced by amphetamine, cocaine and caffeine in the rat. Neuroscience. 1994;59:837–849. [PubMed] [Google Scholar]
  • Johnson M, Bush LG, Gibb JW, Hanson GR. Role of N-methyl-D-aspartate (NMDA) receptors in the response of extrapyramidal neurotensin and dynorphin A systems to cocaine and GBR 12909. Biochem Pharmacol. 1991;41:649–652. [PubMed] [Google Scholar]
  • Kankaanpaa A, Meririnne E, Seppala T. 5-HT3 receptor antagonist MDL 72222 attenuates cocaine- and mazindol-, but not methylphenidate-induced neurochemical and behavioral effects in the rat. Psychopharmacology. 2002;159:341–350. [PubMed] [Google Scholar]
  • Keefe KA, Horner KA. Neurotransmitter regulation of basal ganglia gene expression. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 461–490. [Google Scholar]
  • Kelley AE. Memory and addiction: shared neural circuitry and molecular mechanisms. Neuron. 2004;44:161–179. [PubMed] [Google Scholar]
  • Kim Y, Teylan MA, Baron M, Sands A, Nairn AC, Greengard P. Methylphenidate-induced dendritic spine formation and DeltaFosB expression in nucleus accumbens. Proc Natl Acad Sci U S A. 2009;106:2915–2920. [PMC free article] [PubMed] [Google Scholar]
  • Knapska E, Kaczmarek L. A gene for neuronal plasticity in the mammalian brain: Zif268/Egr-1/NGFI-A/Krox-24/TIS8/ZENK? Prog Neurobiol. 2004;74:183–211. [PubMed] [Google Scholar]
  • Kollins SH. ADHD, substance use disorders, and psychostimulant treatment: current literature and treatment guidelines. J Atten Disord. 2008;12:115–125. [PubMed] [Google Scholar]
  • Kollins SH, MacDonald EK, Rush CR. Assessing the abuse potential of methylphenidate in nonhuman and human subjects: a review. Pharmacol Biochem Behav. 2001;68:611–627. [PubMed] [Google Scholar]
  • Konradi C, Westin JE, Carta M, Eaton ME, Kuter K, Dekundy A, Lundblad M, Cenci MA. Transcriptome analysis in a rat model of L-DOPA-induced dyskinesia. Neurobiol Dis. 2004;17:219–236. [PMC free article] [PubMed] [Google Scholar]
  • Kosofsky BE, Genova LM, Hyman SE. Substance P phenotype defines specificity of c-fos induction by cocaine in developing rat striatum. J Comp Neurol. 1995;351:41–50. [PubMed] [Google Scholar]
  • Kravitz AV, Freeze BS, Parker PR, Kay K, Thwin MT, Deisseroth K, Kreitzer AC. Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature. 2010;466:622–626. [PMC free article] [PubMed] [Google Scholar]
  • Kuczenski R, Segal DS. Effects of methylphenidate on extracellular dopamine, serotonin, and norepinephrine: comparison with amphetamine. J Neurochem. 1997;68:2032–2037. [PubMed] [Google Scholar]
  • Kuczenski R, Segal DS. Locomotor effects of acute and repeated threshold doses of amphetamine and methylphenidate: relative roles of dopamine and norepinephrine. J Pharmacol Exp Ther. 2001;296:876–883. [PubMed] [Google Scholar]
  • Kuczenski R, Segal DS. Stimulant actions in rodents: implications for attention-deficit/hyperactivity disorder treatment and potential substance abuse. Biol Psychiatry. 2005;57:1391–1396. [PubMed] [Google Scholar]
  • Kuhar MJ, Pilotte NS. Neurochemical changes in cocaine withdrawal. Trends Pharmacol Sci. 1996;17:260–264. [PubMed] [Google Scholar]
  • LaHoste GJ, Ruskin DN, Marshall JF. Cerebrocortical Fos expression following dopaminergic stimulation: D1/D2 synergism and its breakdown. Brain Res. 1996;728:97–104. [PubMed] [Google Scholar]
  • LaHoste GJ, Yu J, Marshall JF. Striatal Fos expression is indicative of dopamine D1/D2 synergism and receptor supersensitivity. Proc Natl Acad Sci USA. 1993;90:7451–7455. [PMC free article] [PubMed] [Google Scholar]
  • Lanni C, Lenzken SC, Pascale A, Del Vecchio I, Racchi M, Pistoia F, Govoni S. Cognition enhancers between treating and doping the mind. Pharmacol Res. 2008;57:196–213. [PubMed] [Google Scholar]
  • Lavretsky H, Kim MD, Kumar A, Reynolds CF. Combined treatment with methylphenidate and citalopram for accelerated response in the elderly: an open trial. J Clin Psychiatry. 2003;64:1410–1414. [PubMed] [Google Scholar]
  • Le Moine C, Bloch B. D1 and D2 dopamine receptor gene expression in the rat striatum: sensitive cRNA probes demonstrate prominent segregation of D1 and D2 mRNAs in distinct neuronal populations of the dorsal and ventral striatum. J Comp Neurol. 1995;355:418–426. [PubMed] [Google Scholar]
  • Le Moine C, Bloch B. Expression of the D3 dopamine receptor in peptidergic neurons of the nucleus accumbens: comparison with the D1 and D2 dopamine receptors. Neuroscience. 1996;73:131–143. [PubMed] [Google Scholar]
  • Le Moine C, Normand E, Bloch B. Phenotypical characterization of the rat striatal neurons expressing the D1 dopamine receptor gene. Proc Natl Acad Sci USA. 1991;88:4205–4209. [PMC free article] [PubMed] [Google Scholar]
  • Le Moine C, Normand E, Guitteny AF, Fouque B, Teoule R, Bloch B. Dopamine receptor gene expression by enkephalin neurons in rat forebrain. Proc Natl Acad Sci USA. 1990;87:230–234. [PMC free article] [PubMed] [Google Scholar]
  • Le Moine C, Svenningsson P, Fredholm BB, Bloch B. Dopamine-adenosine interactions in the striatum and the globus pallidus: inhibition of striatopallidal neurons through either D2 or A2A receptors enhances D1 receptor-mediated effects on c-fos expression. J Neurosci. 1997;17:8038–8048. [PMC free article] [PubMed] [Google Scholar]
  • Leonard BE, McCartan D, White J, King DJ. Methylphenidate: a review of its neuropharmacological, neuropsychological and adverse clinical effects. Hum Psychopharmacol. 2004;19:151–180. [PubMed] [Google Scholar]
  • Letter AA, Merchant K, Gibb JW, Hanson GR. Effect of methamphetamine on neurotensin concentrations in rat brain regions. J Pharmacol Exp Ther. 1987;241:443–447. [PubMed] [Google Scholar]
  • Li BH, Rowland NE. Dexfenfluramine induces Fos-like immunoreactivity in discrete brain regions in rats. Brain Res Bull. 1993;31:43–48. [PubMed] [Google Scholar]
  • Li SJ, Sivam SP, McGinty JF, Jiang HK, Douglass J, Calavetta L, Hong JS. Regulation of the metabolism of striatal dynorphin by the dopaminergic system. J Pharmacol Exp Ther. 1988;246:403–408. [PubMed] [Google Scholar]
  • Lin JS, Hou Y, Jouvet M. Potential brain neuronal targets for amphetamine-, methylphenidate-, and modafinil-induced wakefulness, evidenced by c-fos immunocytochemistry in the cat. Proc Natl Acad Sci USA. 1996;93:14128–14133. [PMC free article] [PubMed] [Google Scholar]
  • Lobo MK, Covington HE, Chaudhury D, Friedman AK, Sun H, Damez-Werno D, Dietz DM, Zaman S, Koo JW, Kennedy PJ, Mouzon E, Mogri M, Neve RL, Deisseroth K, Han MH, Nestler EJ. Cell type-specific loss of BDNF signaling mimics optogenetic control of cocaine reward. Science. 2010;330:385–390. [PMC free article] [PubMed] [Google Scholar]
  • Lobo MK, Nestler EJ. The striatal balancing act in drug addiction: distinct roles of direct and indirect pathway medium spiny neurons. Front Neuroanat. 2011;5:41. [PMC free article] [PubMed] [Google Scholar]
  • Loland CJ, Mereu M, Okunola OM, Cao J, Prisinzano TE, Mazier S, Kopajtic T, Shi L, Katz JL, Tanda G, Newman AH. R-Modafinil (Armodafinil): A Unique Dopamine Uptake Inhibitor and Potential Medication for Psychostimulant Abuse. Biol Psychiatry. 2012 in press. [PMC free article] [PubMed] [Google Scholar]
  • London ED, Cascella NG, Wong DF, Phillips RL, Dannals RF, Links JM, Herning R, Grayson R, Jaffe JH, Wagner HNJ. Cocaine-induced reduction of glucose utilization in human brain. A study using positron emission tomography and [fluorine 18]-fluorodeoxyglucose. Arch Gen Psychiatry. 1990;47:567–574. [PubMed] [Google Scholar]
  • Lucas JJ, Segu L, Hen R. 5-Hydroxytryptamine1B receptors modulate the effect of cocaine on c-fos expression: converging evidence using 5-hydroxytryptamine1B knockout mice and the 5-hydroxytryptamine1B/1D antagonist GR127935. Mol Pharmacol. 1997;51:755–763. [PubMed] [Google Scholar]
  • Mache S, Eickenhorst P, Vitzthum K, Klapp BF, Groneberg DA. Cognitive-enhancing substance use at German universities: frequency, reasons and gender differences. Wien Med Wochenschr. 2012;162:262–271. [PubMed] [Google Scholar]
  • Madras BK, Xie Z, Lin Z, Jassen A, Panas H, Lynch L, Johnson R, Livni E, Spencer TJ, Bonab AA, Miller GM, Fischman AJ. Modafinil occupies dopamine and norepinephrine transporters in vivo and modulates the transporters and trace amine activity in vitro. J Pharmacol Exp Ther. 2006;319:561–569. [PubMed] [Google Scholar]
  • Maher B. Poll results: look who’s doping. Nature. 2008;452:674–675. [PubMed] [Google Scholar]
  • Marco EM, Adriani W, Ruocco LA, Canese R, Sadile AG, Laviola G. Neurobehavioral adaptations to methylphenidate: the issue of early adolescent exposure. Neurosci Biobehav Rev. 2011;35:1722–1739. [PubMed] [Google Scholar]
  • Mathieu-Kia AM, Besson MJ. Repeated administration of cocaine, nicotine and ethanol: effects on preprodynorphin, preprotachykinin A and preproenkephalin mRNA expression in the dorsal and the ventral striatum of the rat. Mol Brain Res. 1998;54:141–151. [PubMed] [Google Scholar]
  • Mattson BJ, Koya E, Simmons DE, Mitchell TB, Berkow A, Crombag HS, Hope BT. Context-specific sensitization of cocaine-induced locomotor activity and associated neuronal ensembles in rat nucleus accumbens. Eur J Neurosci. 2008;27:202–212. [PubMed] [Google Scholar]
  • McClung CA, Nestler EJ. Regulation of gene expression and cocaine reward by CREB and DeltaFosB. Nat Neurosci. 2003;6:1208–1215. [PubMed] [Google Scholar]
  • McClung CA, Ulery PG, Perrotti LI, Zachariou V, Berton O, Nestler EJ. DeltaFosB: a molecular switch for long-term adaptation in the brain. Mol Brain Res. 2004;132:146–154. [PubMed] [Google Scholar]
  • McGeorge AJ, Faull RLM. The organization of the projection from the cerebral cortex to the striatum in the rat. Neuroscience. 1989;29:503–537. [PubMed] [Google Scholar]
  • McGinty JF, Shi XD, Schwendt M, Saylor A, Toda S. Regulation of psychostimulant-induced signaling and gene expression in the striatum. J Neurochem. 2008;104:1440–1449. [PMC free article] [PubMed] [Google Scholar]
  • Meririnne E, Kankaanpaa A, Seppala T. Rewarding properties of methylphenidate: sensitization by prior exposure to the drug and effects of dopamine D1- and D2-receptor antagonists. J Pharmacol Exp Ther. 2001;298:539–550. [PubMed] [Google Scholar]
  • Mijnster MJ, Galis-de Graaf Y, Voorn P. Serotonergic regulation of neuropeptide and glutamic acid decarboxylase mRNA levels in the rat striatum and globus pallidus: studies with fluoxetine and DOI. Mol Brain Res. 1998;54:64–73. [PubMed] [Google Scholar]
  • Minzenberg MJ, Carter CS. Modafinil: a review of neurochemical actions and effects on cognition. Neuropsychopharmacology. 2008;33:1477–1502. [PubMed] [Google Scholar]
  • Moratalla R, Elibol B, Vallejo M, Graybiel AM. Network-level changes in expression of inducible Fos-Jun proteins in the striatum during chronic cocaine treatment and withdrawal. Neuron. 1996a;17:147–156. [PubMed] [Google Scholar]
  • Moratalla R, Robertson HA, Graybiel AM. Dynamic regulation of NGFI-A (zif268, egr1) gene expression in the striatum. J Neurosci. 1992;12:2609–2622. [PMC free article] [PubMed] [Google Scholar]
  • Moratalla R, Xu M, Tonegawa S, Graybiel AM. Cellular responses to psychomotor stimulant and neuroleptic drugs are abnormal in mice lacking the D1 dopamine receptor. Proc Natl Acad Sci USA. 1996b;93:14928–14933. [PMC free article] [PubMed] [Google Scholar]
  • Morris BJ, Reimer S, Hollt V, Herz A. Regulation of striatal prodynorphin mRNA levels by the raphe-striatal pathway. Brain Res. 1988;464:15–22. [PubMed] [Google Scholar]
  • Muller CP, Huston JP. Determining the region-specific contributions of 5-HT receptors to the psychostimulant effects of cocaine. Trends Pharmacol Sci. 2006;27:105–112. [PubMed] [Google Scholar]
  • Nakabeppu Y, Nathans D. A naturally occurring truncated form of FosB that inhibits Fos/Jun transcriptional activity. Cell. 1991;64:751–759. [PubMed] [Google Scholar]
  • Naqvi NH, Rudrauf D, Damasio H, Bechara A. Damage to the insula disrupts addiction to cigarette smoking. Science. 2007;315:531–534. [PMC free article] [PubMed] [Google Scholar]
  • Natarajan R, Yamamoto BK. The basal ganglia as a substrate for the multiple actions of amphetamines. Basal Ganglia. 2011;1:49–57. [PMC free article] [PubMed] [Google Scholar]
  • Nelson JC. Augmentation strategies in the treatment of major depressive disorder. Recent findings and current status of augmentation strategies. CNS Spectr. 2007;12(Suppl 22):6–9. [PubMed] [Google Scholar]
  • Nestler EJ. Molecular basis of long-term plasticity underlying addiction. Nat Rev Neurosci. 2001;2:119–128. [PubMed] [Google Scholar]
  • Nestler EJ, Carlezon WAJ. The mesolimbic dopamine reward circuit in depression. Biol Psychiatry. 2006;59:1151–1159. [PubMed] [Google Scholar]
  • Nguyen TV, Kosofsky BE, Birnbaum R, Cohen BM, Hyman SE. Differential expression of c-Fos and Zif268 in rat striatum after haloperidol, clozapine, and amphetamine. Proc Natl Acad Sci USA. 1992;89:4270–4274. [PMC free article] [PubMed] [Google Scholar]
  • Nye HE, Hope BT, Kelz MB, Iadarola M, Nestler EJ. Pharmacological studies of 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]
  • 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. [PubMed] [Google Scholar]
  • Oberlander TF, Gingrich JA, Ansorge MS. Sustained neurobehavioral effects of exposure to SSRI antidepressants during development: molecular to clinical evidence. Clin Pharmacol Ther. 2009;86:672–677. [PMC free article] [PubMed] [Google Scholar]
  • Olivier JD, Blom T, Arentsen T, Homberg JR. The age-dependent effects of selective serotonin reuptake inhibitors in humans and rodents: A review. Prog Neuropsychopharmacol Biol Psychiatry. 2011;35:1400–1408. [PubMed] [Google Scholar]
  • Oorschot DE. Cell types in the different nuclei of the basal ganglia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 63–74. [Google Scholar]
  • Oorschot DE. The percentage of interneurons in the dorsal striatum of the rat, cat, monkey and human: A critique of the evidence. Basal Ganglia. 2013;3 in press. [Google Scholar]
  • Outram SM. The use of methylphenidate among students: the future of enhancement? J Med Ethics. 2010;36:198–202. [PubMed] [Google Scholar]
  • Pan D, Gatley SJ, Dewey SL, Chen R, Alexoff DA, Ding YS, Fowler JS. Binding of bromine-substituted analogs of methylphenidate to monoamine transporters. Eur J Neurosci. 1994;264:177–182. [PubMed] [Google Scholar]
  • Parran TVJ, Jasinski DR. Intravenous methylphenidate abuse. Prototype for prescription drug abuse. Arch Intern Med. 1991;151:781–783. [PubMed] [Google Scholar]
  • Pascoli V, Valjent E, Corbille AG, Corvol JC, Tassin JP, Girault JA, Herve D. cAMP and extracellular signal-regulated kinase signaling in response to d-amphetamine and methylphenidate in the prefrontal cortex in vivo: role of beta 1-adrenoceptors. Mol Pharmacol. 2005;68:421–429. [PubMed] [Google Scholar]
  • Passingham RE, Myers C, Rawlins N, Lightfoot V, Fearn S. Premotor cortex in the rat. Behav Neurosci. 1988;102:101–109. [PubMed] [Google Scholar]
  • Paul ML, Graybiel AM, David J-C, Robertson HA. D1-like and D2-like dopamine receptors synergistically activate rotation and c-fos expression in the dopamine-depleted striatum in a rat model of Parkinson’s disease. J Neurosci. 1992;12:3729–3742. [PMC free article] [PubMed] [Google Scholar]
  • Penner MR, McFadyen MP, Pinaud R, Carrey N, Robertson HA, Brown RE. Age-related distribution of c-fos expression in the striatum of CD-1 mice after acute methylphenidate administration. Dev Brain Res. 2002;135:71–77. [PubMed] [Google Scholar]
  • Persico AM, Schindler CW, Brannock MT, Gonzalez AM, Surratt CK, Uhl GR. Dopaminergic gene expression during amphetamine withdrawal. Neuroreport. 1993;4:41–44. [PubMed] [Google Scholar]
  • Petersen T, Dording C, Neault NB, Kornbluh R, Alpert JE, Nierenberg AA, Rosenbaum JF, Fava M. A survey of prescribing practices in the treatment of depression. Prog Neuropsychopharmacol Biol Psychiatry. 2002;26:177–187. [PubMed] [Google Scholar]
  • Pierce RC, Kalivas PW. A circuitry model of the expression of behavioral sensitization to amphetamine-like psychostimulants. Brain Res Rev. 1997;25:192–216. [PubMed] [Google Scholar]
  • Pinna A, Wardas J, Cristalli G, Morelli M. Adenosine A2A receptor agonists increase Fos-like immunoreactivity in mesolimbic areas. Brain Res. 1997;759:41–49. [PubMed] [Google Scholar]
  • Pisani A, Bernardi G, Ding J, Surmeier DJ. Re-emergence of striatal cholinergic interneurons in movement disorders. Trends Neurosci. 2007;30:545–553. [PubMed] [Google Scholar]
  • Porrino LJ, Smith HR, Nader MA, Beveridge TJ. The effects of cocaine: a shifting target over the course of addiction. Prog Neuropsychopharmacol Biol Psychiatry. 2007;31:1593–1600. [PMC free article] [PubMed] [Google Scholar]
  • Preuss TM. Do rats have prefrontal cortex? The Rose-Woolsey-Akert program reconsidered. J Cogn Neurosci. 1995;7:1–24. [PubMed] [Google Scholar]
  • Qu WM, Huang ZL, Xu XH, Matsumoto N, Urade Y. Dopaminergic D1 and D2 receptors are essential for the arousal effect of modafinil. J Neurosci. 2008;28:8462–8469. [PMC free article] [PubMed] [Google Scholar]
  • Ravindran AV, Kennedy SH, O’Donovan MC, Fallu A, Camacho F, Binder CE. Osmotic-release oral system methylphenidate augmentation of antidepressant monotherapy in major depressive disorder: results of a double-blind, randomized, placebo-controlled trial. J Clin Psychiatry. 2008;69:87–94. [PubMed] [Google Scholar]
  • Redgrave P, Rodriguez M, Smith Y, Rodriguez-Oroz MC, Lehericy S, Bergman H, Agid Y, DeLong MR, Obeso JA. Goal-directed and habitual control in the basal ganglia: implications for Parkinson’s disease. Nat Rev Neurosci. 2010;11:760–772. [PMC free article] [PubMed] [Google Scholar]
  • Reep RL, Cheatwood JL, Corwin JV. The associative striatum: organization of cortical projections to the dorsocentral striatum in rats. J Comp Neurol. 2003;467:271–292. [PubMed] [Google Scholar]
  • Reep RL, Corwin JV, Hashimoto A, Watson RT. Efferent connections of the rostral portion of medial agranular cortex in rats. Brain Res Bull. 1987;19:203–221. [PubMed] [Google Scholar]
  • Reichel CM, See RE. Chronic modafinil effects on drug-seeking following methamphetamine self-administration in rats. Int J Neuropsychopharmacol. 2012;15:919–929. [PMC free article] [PubMed] [Google Scholar]
  • Renthal W, Carle TL, Maze I, Covington HEr, Truong HT, Alibhai I, Kumar A, Montgomery RL, Olson EN, Nestler EJ. Delta FosB mediates epigenetic desensitization of the c-fos gene after chronic amphetamine exposure. J Neurosci. 2008;28:7344–7349. [PMC free article] [PubMed] [Google Scholar]
  • Renthal W, Nestler EJ. Epigenetic mechanisms in drug addiction. Trends Mol Med. 2008;14:341–350. [PMC free article] [PubMed] [Google Scholar]
  • Ritz MC, Cone EJ, Kuhar MJ. Cocaine inhibition of ligand binding at dopamine, norepinephrine and serotonin transporters: a structure-activity study. Life Sci. 1990;46:635–645. [PubMed] [Google Scholar]
  • Robbins TW, Everitt BJ. Drug addiction: bad habits add up. Nature. 1999;398:567–570. [PubMed] [Google Scholar]
  • Robbins TW, Granon S, Muir JL, Durantou F, Harrison A, Everitt BJ. Neural systems underlying arousal and attention. Implications for drug abuse. Ann N Y Acad Sci. 1998;846:222–237. [PubMed] [Google Scholar]
  • Robinson TE, Kolb B. Persistent structural modifications in nucleus accumbens and prefrontal cortex neurons produced by previous experience with amphetamine. J Neurosci. 1997;17:8491–8497. [PMC free article] [PubMed] [Google Scholar]
  • Ruskin DN, Marshall JF. Amphetamine- and cocaine-induced fos in the rat striatum depends on D2 dopamine receptor activation. Synapse. 1994;18:233–240. [PubMed] [Google Scholar]
  • Safer DJ, Zito JM, DosReis S. Concomitant psychotropic medication for youths. Am J Psychiatry. 2003;160:438–449. [PubMed] [Google Scholar]
  • Samaha AN, Mallet N, Ferguson SM, Gonon F, Robinson TE. The rate of cocaine administration alters gene regulation and behavioral plasticity: implications for addiction. J Neurosci. 2004;24:6362–6370. [PMC free article] [PubMed] [Google Scholar]
  • Samaha AN, Robinson TE. Why does the rapid delivery of drugs to the brain promote addiction? Trends Pharmacol Sci. 2005;26:82–87. [PubMed] [Google Scholar]
  • Sandoval V, Riddle EL, Hanson GR, Fleckenstein AE. Methylphenidate redistributes vesicular monoamine transporter-2: role of dopamine receptors. J Neurosci. 2002;22:8705–8710. [PMC free article] [PubMed] [Google Scholar]
  • Sandoval V, Riddle EL, Hanson GR, Fleckenstein AE. Methylphenidate alters vesicular monoamine transport and prevents methamphetamine-induced dopaminergic deficits. J Pharmacol Exp Ther. 2003;304:1181–1187. [PubMed] [Google Scholar]
  • Sato SM, Wissman AM, McCollum AF, Woolley CS. Quantitative mapping of cocaine-induced ΔFosB expression in the striatum of male and female rats. PLoS One. 2011;6:e21783. [PMC free article] [PubMed] [Google Scholar]
  • Scammell TE, Estabrooke IV, McCarthy MT, Chemelli RM, Yanagisawa M, Miller MS, Saper CB. Hypothalamic arousal regions are activated during modafinil-induced wakefulness. J Neurosci. 2000;20:8620–8628. [PMC free article] [PubMed] [Google Scholar]
  • Schenk S, Izenwasser S. Pretreatment with methylphenidate sensitizes rats to the reinforcing effects of cocaine. Pharmacol Biochem Behav. 2002;72:651–657. [PubMed] [Google Scholar]
  • Schmitt KC, Reith ME. The atypical stimulant and nootropic modafinil interacts with the dopamine transporter in a different manner than classical cocaine-like inhibitors. PLoS One. 2011;6:e25790. [PMC free article] [PubMed] [Google Scholar]
  • Schwartz J-C, Diaz J, Bordet R, Griffon N, Perachon S, Pilon C, Ridray S, Sokoloff P. Functional implications of multiple dopamine receptor subtypes: the D1/D3 receptor coexistence. Brain Res Rev. 1998;26:236–242. [PubMed] [Google Scholar]
  • Schweri MM, Skolnick P, Rafferty MF, Rice KC, Janowsky AJ, Paul SM. [3H]Threo-(+/−)-methylphenidate binding to 3,4-dihydroxyphenylethylamine uptake sites in corpus striatum: correlation with the stimulant properties of ritalinic acid esters. J Neurochem. 1985;45:1062–1070. [PubMed] [Google Scholar]
  • See RE, Elliott JC, Feltenstein MW. The role of dorsal vs ventral striatal pathways in cocaine-seeking behavior after prolonged abstinence in rats. Psychopharmacology. 2007;194:321–331. [PubMed] [Google Scholar]
  • Segal DS, Kuczenski R. Escalating dose-binge treatment with methylphenidate: role of serotonin in the emergent behavioral profile. J Pharmacol Exp Ther. 1999;291:19–30. [PubMed] [Google Scholar]
  • Sharp FR, Sagar SM, Swanson RA. Metabolic mapping with cellular resolution: c-fos vs. 2-deoxyglucose. Crit Rev Neurobiol. 1993;7:205–228. [PubMed] [Google Scholar]
  • Shippenberg TS, Heidbreder C. Sensitization to the conditioned rewarding effects of cocaine: pharmacological and temporal characteristics. J Pharmacol Exp Ther. 1995;273:808–815. [PubMed] [Google Scholar]
  • Shippenberg TS, Zapata A, Chefer VI. Dynorphin and the pathophysiology of drug addiction. Pharmacol Ther. 2007;116:306–321. [PMC free article] [PubMed] [Google Scholar]
  • Sivam SP. Cocaine selectively increases striatonigral dynorphin levels by a dopaminergic mechanism. J Pharmacol Exp Ther. 1989;250:818–824. [PubMed] [Google Scholar]
  • Smiley PL, Johnson M, Bush L, Gibb JW, Hanson GR. Effects of cocaine on extrapyramidal and limbic dynorphin systems. J Pharmacol Exp Ther. 1990;253:938–943. [PubMed] [Google Scholar]
  • Smith AJW, McGinty JF. Acute amphetamine or methamphetamine alters opioid peptide mRNA expression in rat striatum. Mol Brain Res. 1994;21:359–362. [PubMed] [Google Scholar]
  • Sofuoglu M, Devito EE, Waters AJ, Carroll KM. Cognitive enhancement as a treatment for drug addictions. Neuropharmacology. 2012 in press. [PMC free article] [PubMed] [Google Scholar]
  • Solanto MV. Dopamine dysfunction in AD/HD: integrating clinical and basic neuroscience research. Behav Brain Res. 2002;130:65–71. [PubMed] [Google Scholar]
  • Spangler R, Unterwald EM, Kreek MJ. ‘Binge’ cocaine administration induces a sustained increase of prodynorphin mRNA in rat caudate-putamen. Mol Brain Res. 1993;19:323–327. [PubMed] [Google Scholar]
  • Spangler R, Zhou Y, Maggos CE, Schlussman SD, Ho A, Kreek MJ. Prodynorphin, proenkephalin and kappa opioid receptor mRNA responses to acute “binge” cocaine. Mol Brain Res. 1997;44:139–142. [PubMed] [Google Scholar]
  • Steiner H. Basal ganglia – cortex interactions: Regulation of cortical function by D1 dopamine receptors in the striatum. In: Tseng KY, Atzori M, editors. Monoaminergic Modulation of Cortical Excitability. Springer; Berlin: 2007. pp. 265–285. [Google Scholar]
  • Steiner H. Psychostimulant-induced gene regulation in corticostriatal circuits. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Academic Press/Elsevier; London: 2010. pp. 501–525. [Google Scholar]
  • Steiner H, Gerfen CR. Cocaine-induced c-fos messenger RNA is inversely related to dynorphin expression in striatum. J Neurosci. 1993;13:5066–5081. [PMC free article] [PubMed] [Google Scholar]
  • Steiner H, Gerfen CR. Tactile sensory input regulates basal and apomorphine-induced immediate-early gene expression in rat barrel cortex. J Comp Neurol. 1994;344:297–304. [PubMed] [Google Scholar]
  • Steiner H, Gerfen CR. Dynorphin opioid inhibition of cocaine-induced, D1 dopamine receptor-mediated immediate-early gene expression in the striatum. J Comp Neurol. 1995;353:200–212. [PubMed] [Google Scholar]
  • Steiner H, Gerfen CR. Dynorphin regulates D1 dopamine receptor-mediated responses in the striatum: relative contributions of pre- and postsynaptic mechanisms in dorsal and ventral striatum demonstrated by altered immediate-early gene induction. J Comp Neurol. 1996;376:530–541. [PubMed] [Google Scholar]
  • Steiner H, Gerfen CR. Role of dynorphin and enkephalin in the regulation of striatal output pathways and behavior. Exp Brain Res. 1998;123:60–76. [PubMed] [Google Scholar]
  • Steiner H, Gerfen CR. Enkephalin regulates acute D2 dopamine receptor antagonist-induced immediate-early gene expression in striatal neurons. Neuroscience. 1999;88:795–810. [PubMed] [Google Scholar]
  • Steiner H, Kitai ST. Regulation of rat cortex function by D1 dopamine receptors in the striatum. J Neurosci. 2000;20:5449–5460. [PMC free article] [PubMed] [Google Scholar]
  • Steiner H, Van Waes V, Marinelli M. Fluoxetine potentiates methylphenidate-induced gene regulation in addiction-related brain regions: Concerns for use of cognitive enhancers? Biol Psychiatry. 2010;67:592–594. [PMC free article] [PubMed] [Google Scholar]
  • Sulzer D. How addictive drugs disrupt presynaptic dopamine neurotransmission. Neuron. 2011;69:628–649. [PMC free article] [PubMed] [Google Scholar]
  • Svetlov SI, Kobeissy FH, Gold MS. Performance enhancing, non-prescription use of Ritalin: a comparison with amphetamines and cocaine. J Addict Dis. 2007;26:1–6. [PubMed] [Google Scholar]
  • Swanson JM, Volkow ND. Serum and brain concentrations of methylphenidate: implications for use and abuse. Neurosci Biobehav Rev. 2003;27:615–621. [PubMed] [Google Scholar]
  • Swanson JM, Volkow ND. Increasing use of stimulants warns of potential abuse. Nature. 2008;453:586. [PMC free article] [PubMed] [Google Scholar]
  • Teter CJ, McCabe SE, LaGrange K, Cranford JA, Boyd CJ. Illicit use of specific prescription stimulants among college students: prevalence, motives, and routes of administration. Pharmacotherapy. 2006;26:1501–1510. [PMC free article] [PubMed] [Google Scholar]
  • Thomas U. Modulation of synaptic signalling complexes by Homer proteins. J Neurochem. 2002;81:407–413. [PubMed] [Google Scholar]
  • Todtenkopf MS, Mihalakopoulos A, Stellar JR. Withdrawal duration differentially affects c-fos expression in the medial prefrontal cortex and discrete subregions of the nucleus accumbens in cocaine-sensitized rats. Neuroscience. 2002;114:1061–1069. [PubMed] [Google Scholar]
  • Torres G, Horowitz JM. Drugs of abuse and brain gene expression. Psychosom Med. 1999;61:630–650. [PubMed] [Google Scholar]
  • Torres G, Rivier C. Cocaine-induced expression of striatal c-fos in the rat is inhibited by NMDA receptor antagonists. Brain Res Bull. 1993;30:173–176. [PubMed] [Google Scholar]
  • Torres G, Rivier C. Induction of c-fos in rat brain by acute cocaine and fenfluramine exposure: a comparison study. Brain Res. 1994;647:1–9. [PubMed] [Google Scholar]
  • Touret M, Sallanon-Moulin M, Fages C, Roudier V, Didier-Bazes M, Roussel B, Tardy M, Jouvet M. Effects of modafinil-induced wakefulness on glutamine synthetase regulation in the rat brain. Mol Brain Res. 1994;26:123–128. [PubMed] [Google Scholar]
  • Trinh JV, Nehrenberg DL, Jacobsen JP, Caron MG, Wetsel WC. Differential psychostimulant-induced activation of neural circuits in dopamine transporter knockout and wild type mice. Neuroscience. 2003;118:297–310. [PubMed] [Google Scholar]
  • Tropea TF, Guerriero RM, Willuhn I, Unterwald EM, Ehrlich ME, Steiner H, Kosofsky BE. Augmented D1 dopamine receptor signaling and immediate-early gene induction in adult striatum after prenatal cocaine. Biol Psychiatry. 2008;63:1066–1074. [PMC free article] [PubMed] [Google Scholar]
  • Tzschentke TM. Measuring reward with the conditioned place preference (CPP) paradigm: update of the last decade. Addict Biol. 2007;12:227–462. [PubMed] [Google Scholar]
  • Unal CT, Beverley JA, Willuhn I, Steiner H. Long-lasting dysregulation of gene expression in corticostriatal circuits after repeated cocaine treatment in adult rats: Effects on zif 268 and homer 1a. Eur J Neurosci. 2009;29:1615–1626. [PMC free article] [PubMed] [Google Scholar]
  • Uslaner J, Badiani A, Norton CS, Day HE, Watson SJ, Akil H, Robinson TE. Amphetamine and cocaine induce different patterns of c-fos mRNA expression in the striatum and subthalamic nucleus depending on environmental context. Eur J Neurosci. 2001;13:1977–1983. [PubMed] [Google Scholar]
  • Uylings HB, Groenewegen HJ, Kolb B. Do rats have a prefrontal cortex? Behav Brain Res. 2003;146:3–17. [PubMed] [Google Scholar]
  • Van Waes V, Beverley J, Marinelli M, Steiner H. Selective serotonin reuptake inhibitor antidepressants potentiate methylphenidate (Ritalin)-induced gene regulation in the adolescent striatum. Eur J Neurosci. 2010;32:435–447. [PMC free article] [PubMed] [Google Scholar]
  • Van Waes V, Carr B, Beverley JA, Steiner H. Fluoxetine potentiation of methylphenidate-induced neuropeptide expression in the striatum occurs selectively in direct pathway (striatonigral) neurons. J Neurochem. 2012a in press. [PMC free article] [PubMed] [Google Scholar]
  • Van Waes V, Vandrevala M, Beverley J, Steiner H. SSRIs potentiate methylphenidate-induced blunting of gene expression in the adolescent striatum. Soc Neurosci Abstr. 2012b;42:778.15. [Google Scholar]
  • Vanderschuren LJ, Di Ciano P, Everitt BJ. Involvement of the dorsal striatum in cue-controlled cocaine seeking. J Neurosci. 2005;25:8665–8670. [PMC free article] [PubMed] [Google Scholar]
  • Vanderschuren LJ, Everitt BJ. Behavioral and neural mechanisms of compulsive drug seeking. Eur J Pharmacol. 2005;526:77–88. [PubMed] [Google Scholar]
  • Vargo JM, Marshall JF. Time-dependent changes in dopamine agonist-induced striatal Fos immunoreactivity are related to sensory neglect and its recovery after unilateral prefrontal cortex injury. Synapse. 1995;20:305–315. [PubMed] [Google Scholar]
  • Vezina P. Sensitization of midbrain dopamine neuron reactivity and the self-administration of psychomotor stimulant drugs. Neurosci Biobehav Rev. 2004;27:827–839. [PubMed] [Google Scholar]
  • Volkow ND, Fowler JS, Logan J, Alexoff D, Zhu W, Telang F, Wang GJ, Jayne M, Hooker JM, Wong C, Hubbard B, Carter P, Warner D, King P, Shea C, Xu Y, Muench L, Apelskog-Torres K. Effects of modafinil on dopamine and dopamine transporters in the male human brain: clinical implications. JAMA. 2009;301:1148–1154. [PMC free article] [PubMed] [Google Scholar]
  • Volkow ND, Wang G, Fowler JS, Logan J, Gerasimov M, Maynard L, Ding Y, Gatley SJ, Gifford A, Franceschi D. Therapeutic doses of oral methylphenidate significantly increase extracellular dopamine in the human brain. J Neurosci. 2001;21(RC121):1–5. [PMC free article] [PubMed] [Google Scholar]
  • Volkow ND, Wang GJ, Fowler JS, Gatley SJ, Logan J, Ding YS, Hitzemann R, Pappas N. Dopamine transporter occupancies in the human brain induced by therapeutic doses of oral methylphenidate. Am J Psychiatry. 1998;155:1325–1331. [PubMed] [Google Scholar]
  • Volkow ND, Wang GJ, Tomasi D, Telang F, Fowler JS, Pradhan K, Jayne M, Logan J, Goldstein RZ, Alia-Klein N, Wong C. Methylphenidate attenuates limbic brain inhibition after cocaine-cues exposure in cocaine abusers. PLoS One. 2010;5:e11509. [PMC free article] [PubMed] [Google Scholar]
  • Wakabayashi KT, Weiss MJ, Pickup KN, Robinson TE. Rats markedly escalate their intake and show a persistent susceptibility to reinstatement only when cocaine is injected rapidly. J Neurosci. 2010;30:11346–11355. [PMC free article] [PubMed] [Google Scholar]
  • Walker PD, Capodilupo JG, Wolf WA, Carlock LR. Preprotachykinin and preproenkephalin mRNA expression within striatal subregions in response to altered serotonin transmission. Brain Res. 1996;732:25–35. [PubMed] [Google Scholar]
  • Wall SC, Gu H, Rudnick G. Biogenic amine flux mediated by cloned transporters stably expressed in cultured cell lines: amphetamine specificity for inhibition and efflux. Mol Pharmacol. 1995;47:544–550. [PubMed] [Google Scholar]
  • Wang JQ, Daunais JB, McGinty JF. NMDA receptors mediate amphetamine-induced upregulation of zif/268 and preprodynorphin mRNA expression in rat striatum. Synapse. 1994a;18:343–353. [PubMed] [Google Scholar]
  • Wang JQ, Daunais JB, McGinty JF. Role of kainate/AMPA receptors in induction of striatal zif/268 and preprodynorphin mRNA by a single injection of amphetamine. Mol Brain Res. 1994b;27:118–126. [PubMed] [Google Scholar]
  • Wang JQ, McGinty JF. Alterations in striatal zif/268, preprodynorphin and preproenkephalin mRNA expression induced by repeated amphetamine administration in rats. Brain Res. 1995a;673:262–274. [PubMed] [Google Scholar]
  • Wang JQ, McGinty JF. Dose-dependent alteration in zif/268 and preprodynorphin mRNA expression induced by amphetamine or methamphetamine in rat forebrain. J Pharmacol Exp Ther. 1995b;273:909–917. [PubMed] [Google Scholar]
  • Wang JQ, McGinty JF. D1 and D2 receptor regulation of preproenkephalin and preprodynorphin mRNA in rat striatum following acute injection of amphetamine or methamphetamine. Synapse. 1996a;22:114–122. [PubMed] [Google Scholar]
  • Wang JQ, McGinty JF. Glutamatergic and cholinergic regulation of immediate-early gene and neuropeptide gene expression in the striatum. In: Merchant KM, editor. Pharmacological Regulation of Gene Expression in the CNS. CRC; Boca Raton: 1996b. pp. 81–113. [Google Scholar]
  • Wang JQ, McGinty JF. The full D1 dopamine receptor agonist SKF-82958 induces neuropeptide mRNA in the normosensitive striatum of rats: regulation of D1/D2 interactions by muscarinic receptors. J Pharmacol Exp Ther. 1997;281:972–982. [PubMed] [Google Scholar]
  • Wang JQ, Smith AJW, McGinty JF. A single injection of amphetamine or methamphetamine induces dynamic alterations in c-fos, zif/268 and preprodynorphin messenger RNA expression in rat forebrain. Neuroscience. 1995;68:83–95. [PubMed] [Google Scholar]
  • Warren BL, Iñiguez SD, Alcantara LF, Wright KN, Parise EM, Weakley SK, Bolaños-Guzmán CA. Juvenile administration of concomitant methylphenidate and fluoxetine alters behavioral reactivity to reward- and mood-related stimuli and disrupts ventral tegmental area gene expression in adulthood. J Neurosci. 2011;31:10347–10358. [PMC free article] [PubMed] [Google Scholar]
  • White BP, Becker-Blease KA, Grace-Bishop K. Stimulant medication use, misuse, and abuse in an undergraduate and graduate student sample. J Am Coll Health. 2006;54:261–268. [PubMed] [Google Scholar]
  • Wilens TE, Adler LA, Adams J, Sgambati S, Rotrosen J, Sawtelle R, Utzinger L, Fusillo S. Misuse and diversion of stimulants prescribed for ADHD: a systematic review of the literature. J Am Acad Child Adolesc Psychiatry. 2008;47:21–31. [PubMed] [Google Scholar]
  • Wilens TE, Faraone SV, Biederman J, Gunawardene S. Does stimulant therapy of attention-deficit/hyperactivity disorder beget later substance abuse? A meta-analytic review of the literature. Pediatrics. 2003;111:179–185. [PubMed] [Google Scholar]
  • Wiley MD, Poveromo LB, Antapasis J, Herrera CM, Bolaños Guzmán CA. Kappa-opioid system regulates the long-lasting behavioral adaptations induced by early-life exposure to methylphenidate. Neuropsychopharmacology. 2009;34:1339–1350. [PMC free article] [PubMed] [Google Scholar]
  • Willie JT, Renthal W, Chemelli RM, Miller MS, Scammell TE, Yanagisawa M, Sinton CM. Modafinil more effectively induces wakefulness in orexin-null mice than in wild-type littermates. Neuroscience. 2005;130:983–995. [PubMed] [Google Scholar]
  • Willuhn I, Sun W, Steiner H. Topography of cocaine-induced gene regulation in the rat striatum: Relationship to cortical inputs and role of behavioural context. Eur J Neurosci. 2003;17:1053–1066. [PubMed] [Google Scholar]
  • Wirtshafter D, Cook DF. Serotonin-1B agonists induce compartmentally organized striatal Fos expression in rats. Neuroreport. 1998;9:1217–1221. [PubMed] [Google Scholar]
  • Wright CI, Groenewegen HJ. Patterns of overlap and segregation between insular cortical, intermediodorsal thalamic and basal amygdaloid afferents in the nucleus accumbens of the rat. Neuroscience. 1996;73:359–373. [PubMed] [Google Scholar]
  • Xiao B, Tu JC, Worley PF. Homer: a link between neural activity and glutamate receptor function. Curr Opin Neurobiol. 2000;10:370–374. [PubMed] [Google Scholar]
  • Yano M, Beverley JA, Steiner H. Inhibition of methylphenidate-induced gene expression in the striatum by local blockade of D1 dopamine receptors: Interhemispheric effects. Neuroscience. 2006;140:699–709. [PubMed] [Google Scholar]
  • Yano M, Steiner H. Methylphenidate (Ritalin) induces Homer 1a and zif 268 expression in specific corticostriatal circuits. Neuroscience. 2005a;132:855–865. [PubMed] [Google Scholar]
  • Yano M, Steiner H. Topography of methylphenidate (Ritalin)-induced gene regulation in the striatum: differential effects on c-fos, substance P and opioid peptides. Neuropsychopharmacology. 2005b;30:901–915. [PubMed] [Google Scholar]
  • Yano M, Steiner H. Methylphenidate and cocaine: the same effects on gene regulation? Trends Pharmacol Sci. 2007;28:588–596. [PubMed] [Google Scholar]
  • Yoo JH, Kitchen I, Bailey A. The endogenous opioid system in cocaine addiction: what lessons have opioid peptide and receptor knockout mice taught us? Br J Pharmacol. 2012;166:1993–2014. [PMC free article] [PubMed] [Google Scholar]
  • Young JW, Geyer MA. Action of modafinil--increased motivation via the dopamine transporter inhibition and D1 receptors? Biol Psychiatry. 2010;67:784–787. [PMC free article] [PubMed] [Google Scholar]
  • Young ST, Porrino LJ, Iadarola MJ. Cocaine induces striatal c-Fos-immunoreactive proteins via dopaminergic D1 receptors. Proc Natl Acad Sci USA. 1991;88:1291–1295. [PMC free article] [PubMed] [Google Scholar]
  • Yuferov V, Nielsen D, Butelman E, Kreek MJ. Microarray studies of psychostimulant-induced changes in gene expression. Addict Biol. 2005;10:101–118. [PubMed] [Google Scholar]
  • Zhang L, Lou D, Jiao H, Zhang D, Wang X, Xia Y, Zhang J, Xu M. Cocaine-induced intracellular signaling and gene expression are oppositely regulated by the dopamine D1 and D3 receptors. J Neurosci. 2004;24:3344–3354. [PMC free article] [PubMed] [Google Scholar]
  • Zhu J, Spencer TJ, Liu-Chen LY, Biederman J, Bhide PG. Methylphenidate and μ opioid receptor interactions: a pharmacological target for prevention of stimulant abuse. Neuropharmacology. 2011;61:283–292. [PMC free article] [PubMed] [Google Scholar]
Feedback