Neuroscience and Biobehavioral Reviews: Stephen V. Faraone

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


journal homepage: www.elsevier.com/locate/neubiorev

The pharmacology of amphetamine and methylphenidate: Relevance to the T


neurobiology of attention-deficit/hyperactivity disorder and other
psychiatric comorbidities

Stephen V. Faraonea,b,
a
Departments of Psychiatry and of Neuroscience and Physiology, State University of New York (SUNY) Upstate Medical University, Syracuse, NY, United States
b
K.G. Jebsen Centre for Research on Neuropsychiatric Disorders, University of Bergen, Bergen, Norway

A R T I C L E I N F O A B S T R A C T

Keywords: Psychostimulants, including amphetamines and methylphenidate, are first-line pharmacotherapies for in-
Amphetamine dividuals with attention-deficit/hyperactivity disorder (ADHD). This review aims to educate physicians re-
Attention-deficit/hyperactivity disorder garding differences in pharmacology and mechanisms of action between amphetamine and methylphenidate,
Methylphenidate thus enhancing physician understanding of psychostimulants and their use in managing individuals with ADHD
Pharmacology
who may have comorbid psychiatric conditions. A systematic literature review of PubMed was conducted in
April 2017, focusing on cellular- and brain system–level effects of amphetamine and methylphenidate. The
primary pharmacologic effect of both amphetamine and methylphenidate is to increase central dopamine and
norepinephrine activity, which impacts executive and attentional function. Amphetamine actions include do-
pamine and norepinephrine transporter inhibition, vesicular monoamine transporter 2 (VMAT-2) inhibition, and
monoamine oxidase activity inhibition. Methylphenidate actions include dopamine and norepinephrine trans-
porter inhibition, agonist activity at the serotonin type 1A receptor, and redistribution of the VMAT-2. There is
also evidence for interactions with glutamate and opioid systems. Clinical implications of these actions in in-
dividuals with ADHD with comorbid depression, anxiety, substance use disorder, and sleep disturbances are
discussed.

1. Introduction (GLU) pathways (Cortese, 2012; Maltezos et al., 2014; Blum et al.,
2008; Potter et al., 2014; Elia et al., 2011). The alterations in these
Attention-deficit/hyperactivity disorder (ADHD) was initially iden- neurotransmitter systems affect the function of brain structures that
tified in children (Lahey et al., 1994) but is now understood to persist moderate executive function, working memory, emotional regulation,
into adulthood in about two thirds of cases (Barkley et al., 2002; Weiss and reward processing (Fig. 1) (Faraone et al., 2015).
et al., 1985; Faraone et al., 2006). In a 2007 meta-analysis that included Individuals with ADHD are often diagnosed with additional psy-
more than 100 studies, the estimated worldwide prevalence of ADHD in chiatric comorbidities, including anxiety, mood, substance use, sleep
individuals < 18 years old was 5.29% (Polanczyk et al., 2007). The disturbances, and antisocial personality disorders (Mao and Findling,
estimated prevalence in adults was 4.4% in a national survey in the 2014; Kooij et al., 2012; Konofal et al., 2010). Importantly, the neu-
United States (Kessler et al., 2006), 3.4% in a 10-nation survey (Fayyad robiological substrates that mediate behaviors associated with ADHD
et al., 2007), and 2.5% in a meta-regression analysis of 6 studies (Simon share commonalities to some extent with those involved in these co-
et al., 2009). morbid disorders (Farb and Ratner, 2014; Sternat and Katzman, 2016;
A large body of evidence suggests that multiple neurotransmitters Schwartz and Kilduff, 2015). Genetic studies have also identified shared
and brain structures play a role in ADHD (Purper-Ouakil et al., 2011; genetic risk factors between ADHD and associated comorbid disorders
Cortese, 2012; Faraone et al., 2015). Although a substantial amount of (Sharp et al., 2014; Carey et al., 2016). As such, comorbidities need to
research has focused on dopamine (DA) and norepinephrine (NE), be taken into account when considering pharmacotherapy in an in-
ADHD has also been linked to dysfunction in serotonin (5hydro- dividual with ADHD.
xytryptamine [5-HT]), acetylcholine (ACH), opioid, and glutamate Although stimulants (including amphetamine [AMP]–based and


Correspondence to: Departments of Psychiatry and of Neuroscience and Physiology, State University of New York (SUNY) Upstate Medical University, Syracuse, NY 13210, United
States.
E-mail address: sfaraone@childpsychresearch.org.

https://doi.org/10.1016/j.neubiorev.2018.02.001
Received 2 November 2017; Received in revised form 25 January 2018; Accepted 5 February 2018
Available online 08 February 2018
0149-7634/ © 2018 The Author. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/BY-NC-ND/4.0/).
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Fig. 1. Brain Mechanisms in ADHD*.


(a) The cortical regions (lateral view) of the brain have a role in attention-deficit/hyperactivity disorder (ADHD). The dorsolateral prefrontal cortex is linked to working memory, the
ventromedial prefrontal cortex to complex decision making and strategic planning, and the parietal cortex to orientation of attention. (b) ADHD involves the subcortical structures
(medial view) of the brain. The ventral anterior cingulate cortex and the dorsal anterior cingulate cortex subserve affective and cognitive components of executive control. Together with
the basal ganglia (comprising the nucleus accumbens, caudate nucleus, and putamen), they form the frontostriatal circuit. Neuroimaging studies show structural and functional ab-
normalities in all of these structures in patients with ADHD, extending into the amygdala and cerebellum. (c) Neurotransmitter circuits in the brain are involved in ADHD. The dopamine
system plays an important part in planning and initiation of motor responses, activation, switching, reaction to novelty, and processing of reward. The noradrenergic system influences
arousal modulation, signal-to-noise ratios in cortical areas, state-dependent cognitive processes, and cognitive preparation of urgent stimuli. (d) Executive control networks are affected in
patients with ADHD. The executive control and cortico-cerebellar networks coordinate executive functioning (i.e., planning, goal-directed behavior, inhibition, working memory, and the
flexible adaptation to context). These networks are underactivated and have lower internal functional connectivity in individuals with ADHD compared with individuals without the
disorder. (e) ADHD involves the reward network. The ventromedial prefrontal cortex, orbitofrontal cortex, and ventral striatum are at the center of the brain network that responds to
anticipation and receipt of reward. Other structures involved are the thalamus, the amygdala, and the cell bodies of dopaminergic neurons in the substantia nigra, which, as indicated by
the arrows, interact in a complex manner. Behavioral and neural responses to reward are abnormal in ADHD. (f) The alerting network is impaired in ADHD. The frontal and parietal
cortical areas and the thalamus intensively interact in the alerting network (indicated by the arrows), which supports attentional functioning and is weaker in individuals with ADHD than
in controls. (g) ADHD involves the default-mode network (DMN). The DMN consists of the medial prefrontal cortex and the posterior cingulate cortex (medial view) as well as the lateral
parietal cortex and the medial temporal lobe (lateral view). DMN fluctuations are 180° out of phase with fluctuations in networks that become activated during externally oriented tasks,
presumably reflecting competition between opposing processes for processing resources. Negative correlations between the DMN and the frontoparietal control network are weaker in
patients with ADHD than in people who do not have the disorder.
*
Reprinted with permission from Macmillan Publishers Ltd: [NAT REV DIS PRIMERS] (Faraone et al., 2015), copyright 2015.

methylphenidate [MPH]–based agents) and nonstimulants (e.g., ato- of exposure to both AMP and MPH (dela Pena et al., 2015; Zhu and
moxetine, clonidine, and guanfacine) are approved for the treatment of Reith, 2008), differences in the specific cellular mechanisms of action of
ADHD (Thomas et al., 2013), stimulants are considered first-line AMP and MPH may influence their effects on the neurobiological sub-
therapy in children, adolescents, and adults with ADHD because of their strates of ADHD and response to treatment in individuals with ADHD as
greater efficacy (Atkinson and Hollis, 2010; Subcommittee on well as their effects on common comorbidities, such as depression and
Attention-Deficit/Hyperactivity Disorder et al., 2011; Rostain, 2008; anxiety.
Bolea-Alamanac et al., 2014; Kooij et al., 2010; Canadian Attention The objective of this review is to educate physicians about the
Deficit Hyperactivity Disorder Resource Alliance (CADDRA), 2011). mechanisms of action of AMP and MPH and the implications of these
AMP and MPH have been shown to exhibit comparable efficacy in 2 actions on the management of ADHD and its comorbidities. To achieve
meta-analyses (Faraone and Glatt, 2010; Catala-Lopez et al., 2017), this goal, a systematic review of the published literature was conducted
with other analyses reporting that AMP has moderately greater effects to obtain articles describing the cellular- and brain system–level effects
than MPH (Faraone and Buitelaar, 2010; Joseph et al., 2017; Stuhec of AMP and MPH. The results of relevant studies are described and
et al., 2015). The tolerability and safety profiles of AMP and MPH in interpreted in the context of the treatment of ADHD and in light of the
terms of adverse events, treatment discontinuation, and cardiovascular comorbidities associated with ADHD.
effects are also generally comparable (Duong et al., 2012; Vaughan and
Kratochvil, 2012; Martinez-Raga et al., 2017), although weight loss and
2. Methods
insomnia have been reported to be more common with AMP than with
MPH (Catala-Lopez et al., 2017). Additional work is planned that will
A systematic literature review of PubMed was conducted on April
further compare the efficacy, tolerability, and safety profiles of different
24, 2017; no limits were included for publication year or the language
pharmacologic interventions in children, adolescents, and adults with
of publication. The search consisted of titles and abstracts and used the
ADHD (Cortese et al., 2017).
following search string: (amphetamine [MESH term] OR methylpheni-
Although increased synaptic availability of DA and NE is a key result
date [MESH term]) AND (cellular OR receptor binding OR

256
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Fig. 2. PRISMA Flow Diagram of the Literature Search.


*Publications were excluded if they did not focus on the mechanism of action or pharmacologic effects of amphetamine or methylphenidate or if the publications were imaging studies in
individuals who were not healthy (i.e., studies in those with psychiatric disorders were not included).

neuroimaging OR FMRI OR SPECT OR PET OR positron emission to- expression of the DAT and decreases in DA uptake (Chen et al., 2009).
mography OR magnetic resonance OR tomography OR spectroscopy) In a dose-dependent and region-specific manner, AMP also increases
AND (dopamine OR serotonin OR norepinephrine OR acetylcholine OR vesicular DA release via inhibition of the vesicular monoamine trans-
glutamate OR opioid OR opiate) NOT (methamphetamine OR MDMA porter 2 (VMAT-2), which releases DA from vesicular storage, and the
OR ecstasy OR addiction OR abuse). The literature search included both concomitant release of cytosolic DA via reverse transport by the DAT
animal and human studies. (Easton et al., 2007bb; Riddle et al., 2007; Sulzer et al., 1995). Fur-
Additional articles of interest were obtained via an assessment of the thermore, AMP inhibits monoamine oxidase (MAO) activity (Miller
relevant articles obtained through the literature review and based on et al., 1980; Robinson, 1985), which decreases cytosolic monoamine
author knowledge. A publication was excluded if it did not specifically breakdown. A wide array of studies using positron emission tomo-
focus on the mechanism of action or pharmacologic effects of AMP or graphy (PET) or single-photon emission computed tomography (SPECT)
MPH or if it was an imaging study in individuals who were other than have demonstrated that AMP produced reductions in the binding po-
healthy (i.e., studies in those with psychiatric disorders were not in- tential of ligands for DA receptors (al-Tikriti et al., 1994; Carson et al.,
cluded). 1997; Castner et al., 2000; Chou et al., 2000; Dewey et al., 1993;
Drevets et al., 1999; Gallezot et al., 2014; Ginovart et al., 1999; Howlett
3. Results and Nahorski, 1979; Laruelle et al., 1997; Le Masurier et al., 2004; Lind
et al., 2005; Mukherjee et al., 1997; Pedersen et al., 2007; Saelens et al.,
The literature search yielded 673 articles (Fig. 2). Of these, 137 1980; Schiffer et al., 2006; Seneca et al., 2006; Sun et al., 2003; Tomic
were considered relevant to the topic of this review. Additional articles et al., 1997; Tomic and Joksimovic, 2000; van Berckel et al., 2006) and
(n = 13) were identified based on author knowledge and assessment of NE receptors (Finnema et al., 2015; Landau et al., 2012), which is an
the reference sections of relevant citations. The articles included in this indirect indicator of increased competition for binding sites resulting
review are summarized in Table 1. from increased extracellular DA or NE. The striatum, which contains
most of the DATs in the brain (Volkow et al., 1996bb; Fischman et al.,
1997), appears to be a principal site of action of AMP (Avelar et al.,
3.1. Preclinical studies 2013; Kilbourn and Domino, 2011), but direct effects in the cortex and
the ventral tegmental area have also been reported (Pum et al., 2007;
3.1.1. Amphetamine Ren et al., 2009; Schwarz et al., 2007bb). The effects of AMP extend to
The main mechanism of action of AMP is to increase synaptic ex- and are modulated by other neurotransmitter systems (Choe et al.,
tracellular DA and NE levels (Avelar et al., 2013; Covey et al., 2013; 2002; Duttaroy et al., 1992; Inderbitzin et al., 1997; Konradi et al.,
Finnema et al., 2015; Floor and Meng, 1996; Jedema et al., 2014; Joyce 1996; Liu et al., 2003; Pum et al., 2007; Quelch et al., 2014; Ritz and
et al., 2007; Kuczenski and Segal, 1997; May et al., 1988; Mukherjee Kuhar, 1989; Shaffer et al., 2010; Smith et al., 2005; Yin et al., 2010; Yu
et al., 1997; Pum et al., 2007; Ren et al., 2009; Schiffer et al., 2006; Xiao et al., 2003), including ACH, 5-HT, opioid, and GLU, either directly
and Becker, 1998; Young et al., 2011; Wall et al., 1995). This effect is through enhanced release from presynaptic terminals or via down-
mediated by inhibition of DA transporters (DAT) and NE transporters stream effects.
(NET) (Avelar et al., 2013; Covey et al., 2013; Easton et al., 2007bb), In other studies, AMP has been shown to produce changes at a more
which reduces the reuptake of these molecules from the synapse. In global level. In studies of cerebral blood flow (CBF), AMP increased
wild-type mice, AMP initially increases surface trafficking of DAT and whole brain CBF in rats and baboons (Chen et al., 2008; Kashiwagi
DA uptake, but continued AMP exposure results in decreased surface

257
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Table 1
Studies of the Actions of AMP and MPH.

Preclinical studies

Amphetamine
al-Tikriti et al. (1994) • AMP 123
increased the washout rate of [ I]IBF (a D receptor antagonist) from the striatum of baboons, as measured using SPECT
• AMP-stimulated
2
Annamalai et al. (2010) downregulation of the NET was linked to the PKC-resistant T258/S259 structural motif and mediated by reduced plasma
membrane insertion and enhanced endocytosis
Avelar et al. (2013) • AMP increased electrically evoked DA levels, inhibited DA uptake, and upregulated DA vesicular release in rat striatum
Bjorklund et al. (2008) • Adenosine A receptor knockout mice exhibited reduced locomotor responses to AMP, suggesting potential alterations in monoaminergic
3
(DA, 5-HT, or NE) systems
Carson et al. (1997) • AMP 11
reduced [ C]raclopride (a D receptor antagonist) binding in the striatum of rhesus monkeys, as measured by PET
• AMP
2
123
Castner et al. (2000) reduced [ I]IBZM (a D receptor antagonist) binding in the striatum of rhesus monkeys, as measured by SPECT
• AMP-induced
2
Chen et al. (2008) increases in whole-brain relative cerebral blood volume were attenuated by electroacupuncture (electrical paw stimulation) in
rats
Chen et al. (2009) • Rapid AMP-induced increases in striatal surface DAT levels and DA release in mice were not observed in PKC-β knockout mice
Choe et al. (2002) • AMP increased phosphorylation of CREB in rats; intrastriatal administration of PHCCC (an mGluR type 1 antagonist) and systemic
administration of MPEP (an mGluR type 5 antagonist) attenuated this effect
Chou et al. (2000) • PET 11
AMP reduced [ C]FLB 457 (a D receptor antagonist) binding in the thalamus, cortex, and striatum of cynomolgus monkeys, as measured by
2

Covey et al. (2013) • AMP increased DA release in rat striatum when administered with a short-duration electrical stimulus and decreased DA release when
administered with a long-duration electrical stimulus
Dewey et al. (1993) • AMP 11
decreased [ C]raclopride (a D receptor antagonist) binding in the striatum of baboons, as measured by PET
• AMP
2
Dixon et al. (2005) caused widespread increases in fMRI BOLD signal intensity in subcortical structures containing rich DA innervation, with decreases in
BOLD signal observed in the superficial layers of the cortex in rats
Drevets et al. (1999) • the 11
AMP-induced reductions in [ C]raclopride (a D receptor antagonist) binding in baboons were greater in the anteroventral striatum than in
dorsal striatum, as measured by PET
2

Duttaroy et al. (1992) • upregulation


Chronic AMP exposure blocked naloxone-induced supersensitivity of μ- and δ-opioid receptors without altering naloxone-induced
of these receptors in mice
Easton et al. (2007a) • AMP isomers increased DA release (basal and electrically stimulated) in the nucleus accumbens, medial entorhinal cortex, colliculi,
hippocampal area CA , and thalamic nuclei of rats
• AMP
1
Easton et al. (2007b) inhibited accumulation of DA or NE in rat brain synaptosomes and vesicles, which was mitigated by inhibitors of the DAT and NET
Finnema et al. (2015) • PET,
AMP reduced binding of [ C]ORM-13070 (an adrenergic α receptor antagonist) in the striatum of cynomolgus monkeys, as measured by
11

and increased rat striatal NE and DA concentrations


2c

Floor and Meng (1996) • Low AMP concentrations increased DA release from rat brain synaptic vesicles independent of vesicular alkalinization; high-AMP-
concentration DA release was reduced as a function of the proton gradient
Gallezot et al. (2014) • and 11 11
AMP reduced [ C]raclopride (a D receptor antagonist) and [ C]PHNO (a D /D receptor agonist) binding in the striatum, globus pallidus,
2
substantia nigra of rhesus monkeys, as measured by PET
2 3

Ginovart et al. (1999) • day 11


Chronic AMP administration (14 days) was associated with an increased [ C]raclopride (a D receptor antagonist) binding affinity after 1
of drug withdrawal and with binding after 7 and 14 days of withdrawal in cynomolgus monkeys, as measured by PET
2

Hartvig et al. (1997) • AMP increased DA biosynthesis in the brain in nonhuman primates, as measured by PET
Helmeste and Seeman (1982) • mice
Mice with high D receptor density in the striatum and olfactory tubercle exhibited hypolocomotion in response to low-dose AMP, whereas
2
with low D receptor density were nonresponsive to AMP
• binding
2
3
Howlett and Nahorski (1979) AMP exposure increased [ H]spiperone (a D /D receptor antagonist) binding affinity and density after 4 days of exposure and decreased
2 3
density after 20 days of exposure in rat striatum
Inderbitzin et al. (1997) • AMP increased preprodynorphin mRNA expression and κ-opioid receptor binding in the basal ganglia of rats
Jedema et al. (2014) • AMP increased DA release in the caudate and prefrontal cortex of rhesus macaques, as measured using microdialysis, with DA levels
remaining elevated for a longer duration in the prefrontal cortex than in the caudate
Joyce et al. (2007) • Administration of a combination of mixed amphetamine salts directly into the striatum of rats stimulated DA release
Kashiwagi et al. (2015) • AMP increased relative cerebral blood volume in awake and anesthetized rats, with changes correlated to changes in striatal DA
concentration
Konradi et al. (1996) • AMP-induced activation of immediate early genes in dissociated rat striatal cultures was blocked by MK-801 (an NMDA receptor antagonist)
Kuczenski and Segal (1997) • AMP increased DA and 5-HT efflux in the caudate and NE efflux in the hippocampus of rats following systemic administration
Lahti and Tamminga (1999) • AMP increased D receptor occupancy in the cortex and caudate of rats, as measured by receptor autoradiography
• byAMPPETdecreased binding of [ C]yohimbine (an adrenergic ɑ receptor antagonist) in the thalamus, cortex, and caudate of pigs, as measured
2
11
Landau et al. (2012) 2

Laruelle et al. (1997) • changes 123 123


AMP reduced [ I]IBZM and [ I]IBF (D receptor antagonists) in vervets (nonhuman primates), as measured by SPECT, with binding
2
correlating with peak DA release as measured by microdialysis
Le Masurier et al. (2004) • with 11
AMP-reduced [ C]raclopride (a D receptor antagonist) binding in the striatum of rats, as measured by PET, was attenuated by pretreatment
a tyrosine-free amino acid mixture
2

Lind et al. (2005) • AMP 11


decreased [ C]raclopride (a D receptor antagonist) binding in the caudate, putamen, and ventral striatum of pigs, as measured by PET
• AMP
2
Liu et al. (2003) increased cytosolic calcium concentrations and NE release in a dose-dependent and extracellular calcium-dependent manner in bovine
adrenal chromaffin cells; nicotinic receptor antagonists suppressed these effects
• AMP acted as a nicotinic receptor agonist to induce calcium increases and NE release in bovine adrenal chromaffin cells
May et al. (1988) • AMP stimulated DA release in the caudate nucleus of rats, as measured by in vivo fast-scan cyclic voltammetry
Miller et al. (1980) • AMP inhibited MAO type A activity
Mukherjee et al. (1997) • AMP 18
reduced [ F]fallypride (a D receptor antagonist) binding in the striatum of rats and monkeys, as measured by PET
• Stimulation
2
Patrick et al. (1981) of DA synthesis in rat brain striatal synaptosomes produced by the depolarizing agent veratridine markedly reduced AMP,
suggesting AMP altered interactions between tyrosine hydroxylase and the synaptosomes regulating catecholamine formation
Pedersen et al. (2007) • pretreatment 11
AMP-induced decreases in [ C]raclopride (a D receptor antagonist) binding in the dorsal and ventral striatum of rats were not altered by
with pargyline, as measured using PET
2

Preece et al. (2007) • AMP reduced BOLD signal intensity in the nucleus accumbens and prefrontal cortex and increased signal intensity in the motor cortex of rats
as measured by fMRI; the signal intensity changes in the nucleus accumbens and caudate (but not in the motor cortex) were attenuated by
pretreatment with a tyrosine-free amino acid mixture
Price et al. (2002) • After administration of AMP, a general increase in CBF that gradually declined toward baseline values was observed using a bolus injection
PET method in baboons
Pum et al. (2007) • AMP dose-dependently increased DA and 5-HT levels in the perirhinal, entorhinal, and prefrontal cortices (continued on next page)

258
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Table 1 (continued)

Preclinical studies

Quelch et al. (2014) • AMP reduced [ H]carfentanil (a μ-opioid receptor antagonist) binding in rat total brain homogenate, as measured by subcellular
3

fractionation, and in the superior colliculi, hypothalamus, and amygdala of rats, as measured by in vivo receptor binding
Ren et al. (2009) • AMP increased DA release in the caudate/putamen and cAMP activity in the caudate/putamen, nucleus accumbens, and medial prefrontal
cortex in rats
Riddle et al. (2007) • Administration of low-dose AMP altered VMAT-2 distribution within nerve terminals selectively in monoaminergic neurons
Ritz and Kuhar (1989) • rats
AMP exhibited high affinity for DA, NE, and 5-HT reuptake sites and for α adrenergic receptor sites, as measured using in vivo binding in
2

Robinson (1985) • AMP enantiomers (d and l) inhibited MAO type A and MAO type B activity in rat liver mitochondria
Saelens et al. (1980) • cortex, 3
AMP increased [ H]spiroperidol (a D receptor antagonist) binding in regions containing DA terminals (e.g., the septum, nucleus accumbens,
2
caudate-putamen) or dendrites (e.g., ventral tegmental area, substantia nigra), as measured by in vivo receptor binding
Schiffer et al. (2006) • binding 11
AMP elevated extracellular DA in the striatum but to a greater degree than MPH; reductions in [ C]raclopride (a D receptor antagonist)
were similar with AMP and MPH
2

Schwarz et al. (2007a) • AMP produced widespread increases in regional CBF in multiple brain region clusters that are involved in primary DA pathways in rats
Schwendt et al. (2006) • AMP decreased RGS4 mRNA in the caudate-putamen and cortex and RGS4 protein levels in the caudate-putamen of rats but did not modify
the effects of SCH 23390 or eticlopride (D receptor antagonists) on RGS4 mRNA or protein levels in these same brain regions
• PET
2
11 11
Seneca et al. (2006) AMP reduced binding of [ C]MNPA and [ C]raclopride (D receptor antagonists) in the striatum of cynomolgus monkeys, as measured by
2

Shaffer et al. (2010) • AMP reduced striatal and increased medial prefrontal cortical protein expression of mGluR type 5 in rats
Skinbjerg et al. (2010) • PET 18
AMP-induced decreases in [ F]fallypride (a D receptor antagonist) binding were associated with receptor internalization, as assessed using
2
in knockout mice lacking the ability to internalize D receptors
• Melanin-concentrating
2
Smith et al. (2005) hormone type 1 receptor knockout mice exhibited an increased locomotor response to AMP but did not exhibit
altered DA, NE, or 5-HT release in the striatum in response to AMP
Sulzer et al. (1995) • AMP produced a rapid release of DA from vesicles, resulting in redistribution to the cytosol
Sulzer and Rayport (1990) • AMP reduced the pH gradient in rat chromaffin cells, resulting in reduced DA uptake
Sun et al. (2003) • AMP 3 3
decreased binding of [ H]raclopride but not of [ H]spiperone (D receptor antagonists) in rat striatum
• inAMPthedecreased
2
3 3
Tomic et al. (1997) [ H]SCH 23390 (a D receptor antagonist) binding in the caudate and nucleus accumbens, increased [ H]SCH 23390 binding
1
substantia nigra, and decreased [3H]spiperone (a D receptor antagonist) binding in the striatum and nucleus accumbens in rats, as
2
measured using autoradiography
Tomic and Joksimovic (2000) • striatum 3 3
Acute treatment with AMP reduced [ H]SCH 23390 (a D receptor antagonist) and [ H]spiperone (a D receptor antagonist) binding in the
and nucleus accumbens in rats, as measured by autoradiography
1 2

van Berckel et al. (2006) • mGluR 3


AMP-induced decrease in [ H]raclopride (a D receptor antagonist) binding in the striatum of baboons was enhanced by LY354740 (an
type 2/3 receptor agonist), as measured by PET
2

Xiao and Becker (1998) • AMP-induced striatal DA release in rats was increased by catechol estrogens, as measured using in vivo microdialysis
Yin et al. (2010) • AMP produced time-dependent changes in levels of GABA, 5-HT, and NE in the cerebellar vermis of mice
Young et al. (2011) • AMP increased DA concentrations in the striatum, but not in the cortex or ventral tegmental area, in prairie voles
Yu et al. (2003) • AMP enhanced the cellular response of cortical and hippocampal neurons to CHPG (an mGluR type 5 agonist) in rats
Methylphenidate
Andrews and Lavin (2006) • Intracortical MPH increased cortical cell excitability in rats, an effect that was lost following catecholamine depletion and was mediated via
stimulation of α adrenergic receptors but not D receptors
• MPH
2 1
Bartl et al. (2010) exerted diverse cellular effects including increased neurotransmitter levels, downregulated synaptic gene expression, and enhanced cell
proliferation in rat pheochromocytoma cells
Ding et al. (1994) • MPH 11
reduced binding of [ C]dl-threo-MPH in the striatum of baboons, as measured by PET
Ding et al. (1997) • MPH 11 11
reduced binding of [ C]d-threo-MPH (but not [ C]l-threo-MPH) in the striatum of baboons, as measured by PET
Dresel et al. (1999) • baboons, 99
MPH reduced [ Tc]TRODAT-1 (a DAT/SERT ligand) binding to DAT in the striatum, but not to the SERT in the midbrain/hypothalamus in
as measured by PET
Easton et al. (2007b) • MPH inhibited accumulation of DA or NE in rat brain synaptosomes and vesicles, with greater potency observed for synaptosomal inhibition
Federici et al. (2005) • MPH reduced spontaneous firing of DA neurons, as measured by electrophysiologic recordings in rat brain slices, via blockade of the DAT
Gamo et al. (2010) • idazoxan
MPH increased performance on a spatial working memory task in rhesus monkeys, an effect that was blocked by the α adrenergic antagonist 2

Gatley et al. (1995) • [PET;C]d-threo-MPH


11
exhibited high affinity for the DAT that was insensitive to competition with endogenous DA in baboons, as measured by
11
the highest concentrations of [ C]d-threo-MPH binding in mouse brain were found in striatum
Gatley et al. (1999) • MPH 3
reduced [ H]cocaine (a DAT agonist) binding in the olfactory tubercle and striatum in mice, as measured by PET
Gatley et al. (1996) • MPH exhibited higher affinities for the DAT and NET than for the serotonin transporter in rat brain
Kuczenski and Segal (1997) • MPH increased DA efflux in the caudate and NE efflux in the hippocampus of rats following systemic administration
Markowitz et al. (2009) • MPH acted as an agonist at the 5-HT receptor in guinea pig ileum
• MPH
1A
Markowitz et al. (2006) exhibited affinity at the NET and DAT and at 5-HT and 5-HT receptors
• MPH
1A 2B
Michaelides et al. (2010) differentially altered metabolism in the prefrontal cortex and cerebellar vermis as a function of DA D receptor functionality in mice
• MPH
4
123
Nikolaus et al. (2005) reduced [ I]FP-CIT (a DAT ligand) binding in rat striatum, as measured by SPECT
Nikolaus et al. (2007) • MPH 123
dose-dependently reduced [ I]FP-CIT (a DAT ligand) binding in rat striatum, as measured by SPECT
Nikolaus et al. (2011) • MPH 123
decreased [ I]IBZM (a D receptor antagonist) binding in the striatum of rats, as measured by SPECT
• Administration
2
Riddle et al. (2007) of low-dose MPH altered VMAT-2 distribution within nerve terminals selectively in monoaminergic neurons
Sandoval et al. (2002) • MPH increased vesicular DA uptake and binding of VMAT-2 and altered VMAT-2 cellular distribution
Schiffer et al. (2006) • binding 11
MPH elevated extracellular DA in the striatum but to a lesser degree than AMP; reductions in [ C]raclopride (a D receptor antagonist)
were similar with MPH and AMP
2

Somkuwar et al. (2013) • MPH administered during the adolescent period in rats produced strain-dependent alterations in cortical DAT activity at adulthood
Volkow et al. (1999a) • MPH 11
reduced [ C]raclopride (a D receptor antagonist) binding potential but not binding affinity in baboons, as measured by PET
• MPH caused little or no change in DA, NE, or 5-HT efflux, although it blocked uptake mediated by both NET and DAT in transfected cell lines
2
Wall et al. (1995)

Human neuroimaging studies


Amphetamine
Aalto et al. (2009) • AMP 11
did not alter [ C]FLB 457 (a D /D receptor ligand) binding in the cortex of healthy adults, as measured by PET
• conditioned
2 3
11
Boileau et al. (2007) AMP decreased [ C]raclopride (a D receptor antagonist) binding in ventral striatum and putamen of healthy adults in response to
2
stimuli, as measured by PET
Buckholtz et al. (2010) • increased 18
AMP-induced reductions in [ F]fallypride (a D receptor antagonist) binding in the striatum of healthy adults were associated with
trait impulsivity
2

(continued on next page)

259
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Table 1 (continued)

Preclinical studies

Cardenas et al. (2004) • measured 11


AMP produced decreases in [ C]raclopride (a D receptor antagonist) binding that were sustained for up to 6 h in healthy adults, as
using PET studies
2

Colasanti et al. (2012) • anterior


AMP reduced [ C]carfentanil (a μ-opioid receptor antagonist) binding in the caudate, putamen, frontal cortex, thalamus, insula, and
11

cingulate cortex of healthy adult males, as measured by PET


Cropley et al. (2008) • healthy 18
AMP reduced [ F]fallypride (a D receptor antagonist) binding in the substantia nigra, medial prefrontal and orbital cortices, and caudate of
adults, as measured by PET
2

Devous et al. (2001) • AMP increased regional CBF to the prefrontal cortex, inferior orbital frontal cortex, ventral tegmentum, amygdala, and anterior thalamus,
and decreased regional CBF to the motor and visual cortices, fusiform gyrus, and regions of the temporal lobe of healthy adults, as measured
by SPECT
Drevets et al. (2001) • byAMP-induced 11
reductions in [ C]raclopride (a D receptor antagonist) binding in the anteroventral striatum of healthy adults, as measured
PET, were negatively correlated with feelings of euphoria
2

Garrett et al. (2015) • AMP increased fMRI-based BOLD signal variability in healthy adults, with the effects being more pronounced in older adults (60–70 years
old) than younger adults (20–30 years old)
Guterstam et al. (2013) • adult
AMP had no effect on [ C]carfentanil (a μ-opioid receptor antagonist) binding in the striatum, cortex, amygdala, or hippocampus of healthy
11

males, as measured by PET


Hariri et al. (2002) • AMP potentiated responses of the right amygdala to angry and fearful facial expressions of healthy adults, as measured by fMRI, without
producing changes in performance of an emotional recognition task or a sensorimotor control task
Kegeles et al. (1999) • AMP 123
decreased [ I]IBZM (a D receptor antagonist) binding in the striatum of healthy adults, as measured by SPECT
• AMP
2
Knutson et al. (2004) exerted an equalizing influence on activity in the ventral striatum by enhancing tonic activity over phasic activity during anticipation of
positive and negative incentives in healthy adults, as measured by fMRI
Laruelle et al. (1995) • AMP 123
decreased [ I]IBZM (a D receptor antagonist) binding in the striatum of healthy adults, as measured by SPECT
• AMP
2
123
Laruelle and Innis (1996) decreased [ I]IBZM (a D receptor antagonist) binding in the striatum of healthy adults, as measured by SPECT
• correlated
2
11
Leyton et al. (2002) AMP-decreased [ C]raclopride (a D receptor antagonist) binding in the ventral striatum of healthy adults, as measured by PET, was
2
with increases in drug wanting and novelty seeking
Leyton et al. (2004) • were 11
AMP-induced decreases in [ C]raclopride (a D receptor antagonist) binding in the ventral striatum of healthy adults, as measured by PET,
attenuated by acute depletion of phenylalanine and tyrosine
2

Martinez et al. (2003) • the 11


AMP decreased [ C]raclopride (a D receptor antagonist) binding in healthy adults, as measured by PET, with larger reductions observed in
2
limbic (ventral) and the sensorimotor (postcommissural) striatal regions than associative (caudate and precommissural) striatal regions
Mick et al. (2014) • insula,
AMP reduced [ C]carfentanil (a μ-opioid receptor antagonist) binding in the caudate, putamen, frontal lobe, thalamus, nucleus accumbens,
11

amygdala, and anterior cingulate cortex of healthy adults, as measured by PET


Narendran et al. (2009) • anterior 11 11
AMP decreased [ C]FLB 457 (a D /D receptor ligand) and [ C]fallypride (a D receptor antagonist) binding in the medial temporal lobe,
2 3 2
cingulate cortex, dorsolateral and medial prefrontal cortices, and parietal cortex of healthy adults, as measured by PET
Narendran et al. (2010) • asAMPmeasured 11
2 3
11
11
decreased [ C]NPA (a D /D receptor agonist) and [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults,
by PET, with reductions of [ C]NPA binding being greater than those of [ C]raclopride
2
11

Oswald et al. (2005) • correlated 11


AMP-induced reductions in [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults, as measured by PET, were
2
with increased cortisol release and positive subjective ratings of AMP
Oswald et al. (2015) • caudate 11
AMP decreased [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults, with lower binding in the right dorsal
2
being associated with lower winnings in the Iowa Gambling Task
Riccardi et al. (2006a) • regions 18
AMP reduced [ F]fallypride (a D receptor antagonist) binding in healthy adults, as measured by PET, with the reductions across brain
differing as a function of gender
2

Riccardi et al. (2006b) • ofAMPhealthy 18


reduced [ F]fallypride (a D receptor antagonist) binding in the striatum, substantia nigra, amygdala, temporal cortex, and thalamus
adults, as measured by PET
2

Riccardi et al. (2011) • regions 18


AMP reduced [ F]fallypride (a D receptor antagonist) binding in healthy adults, as measured by PET, with the reductions in selected baring
2
being correlated with Stroop test scores, Stroop test interference, and affective state
Rose et al. (2006) • AMP
by MRI
increased CBF in the cerebellum, brainstem, temporal lobe, striatum, and prefrontal and parietal cortices of healthy adults, as measured

Schouw et al. (2013) • AMP-induced increases in regional CBF in the striatum, anterior cingulate cortex, thalamus, and cerebellum of healthy adults, as measured
123
by MRI, were not correlated with reductions in [ I]IBZM receptor binding in the striatum, as measured by SPECT
Shotbolt et al. (2012) • putamen 11
AMP induced reductions in [ C]−(+)-PHNO (a D /D receptor agonist) binding in the ventral pallidum, ventral striatum, thalamus, and
2 3
11
of healthy adults, as measured by PET, which were greater than the reductions in [ C]raclopride (a D receptor antagonist) binding
• AMP
2
Silverstone et al. (2002) increased myo-inositol and phosphomonoester concentrations in the temporal lobe of healthy adults, as measured by MRS
Slifstein et al. (2010) • healthy 18
AMP reduced [ F]fallypride (a D receptor antagonist) binding in the striatum, globus pallidus, midbrain, hippocampus, and amygdala of
adults, as measured by PET
2

Vollenweider et al. (1998) • AMP increased regional cerebral glucose metabolism in the anterior and posterior cingulate cortex, caudate nucleus, putamen, and thalamus
of healthy adults, as measured by PET with [ F]-FDG 18

Wand et al. (2007) • associated 11


AMP-induced reductions in [ C]raclopride (a D receptor antagonist) binding in striatum of healthy adults, as measured by PET, were
with stress-induced cortisol levels
2

Willeit et al. (2008) • measured 11


AMP reduced [ C]–(+)-PHNO (a D /D receptor agonist) binding in the striatum but not in the globus pallidus of healthy adults, as
by PET
2 3

Wolkin et al. (1987) • AMP


PET
decreased regional cerebral glucose metabolism in the frontal cortex, temporal cortex, and striatum of healthy adults, as measured by

Woodward et al. (2011) • positively 18


AMP-induced reductions in [ F]fallypride (a D2 receptor antagonist) binding in the striatum of healthy adults, as measured by PET, were
correlated with overall schizotypal traits

Methylphenidate
Booij et al. (1997) • MPH 123
reduced [ I]IBZM (a D receptor antagonist) binding in the striatum of healthy adults, as measured by SPECT
• adults,
2
11
Clatworthy et al. (2009) MPH reduced [ C]raclopride (a D receptor antagonist) binding in the putamen, ventral striatum, and postcommissural caudate of healthy
2
as measured by PET, with reductions in the postcommissural caudate being negatively correlated with reversal learning performance
Costa et al. (2013) • MPH increased neuronal activation in the putamen of healthy adults as measured by fMRI, during a go/no-go task when a response inhibition
error occurred but not when a response was successfully inhibited
Ding et al. (1997) • MPH 11 11
reduced binding of [ C]d-threo-MPH to a greater degree than [ C]l-threo-MPH in the striatum of healthy adults, as measured by PET
Hannestad et al. (2010) • MPH dose-dependently reduced [11C]MRB (a NET ligand) binding across the brain (e.g., locus coeruleus, raphe nucleus, hypothalamus,
thalamus) of healthy adults, as measured using PET
Kasparbauer et al. (2015) • MPH increased BOLD signal during successful go/no-go trials in healthy adult carriers of the SLC6A3 9R allele of the DAT gene but a decrease
in 10/10 allele homozygotes, as measured by fMRI
(continued on next page)

260
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Table 1 (continued)

Preclinical studies

Moeller et al. (2014) • MPH improved Stroop color-word task performance and concurrently reduced dorsal anterior cingulate cortex responses of healthy adults, as
measured by fMRI
Montgomery et al. (2007) • ofMPHhealthy 11
reduced [ C]FLB 457 (a D /D receptor ligand) binding in the frontal cortex, anterior cingulate cortex, temporal cortex, and thalamus
2
adults, as measured by PET
3

Mueller et al. (2014) • MPH increased connectivity strength between the dorsal attention network and thalamus, with the left and right frontoparietal networks and
the executive control networks also showing increased connectivity to sensory-motor and visual cortex regions, and decreased connectivity
to cortical and subcortical components of cortico-striato-thalamo-cortical circuits in healthy adults, as measured by fMRI
Ramaekers et al. (2013) • MPH reduced functional connectivity between the nucleus accumbens and the basal ganglia, medial prefrontal cortex, and temporal cortex
without changing functional connectivity between the medial dorsal nucleus and the limbic circuit, as measured by fMRI
Ramasubbu et al. (2012) • MPH decreased oxy-hemoglobin levels, an index of decreased neural activation, in the right lateral prefrontal cortex of healthy adults, as
measured by fNIRS
Schabram et al. (2014) • MPH 18
decreased DA turnover in the caudate and putamen of healthy adult females, as measured by PET using [ F]FDOPA
Schlosser et al. (2009) • Processing of uncertain information was associated with higher activation of the parietal association cortex and posterior cingulate cortex
after placebo relative to MPH, but with higher left and right parahippocampal and cerebellar activation after MPH relative to placebo, as
measured by fMRI
Schweitzer et al. (2004) • MPH improved Paced Auditory Serial Addition Test performance, reduced regional CBF in the prefrontal cortex, and increased regional CBF
in the right thalamus and precentral gyrus of healthy adults, as measured by PET
Spencer et al. (2006) • measured 11
Both short-acting and long-acting MPH formulations decreased [ C]altropane (a DAT ligand) binding in the striatum of healthy adults, as
by PET
Spencer et al. (2010) • PET 11
Plasma MPH concentrations were correlated with [ C]altropane (a DAT ligand) binding in the striatum of healthy adults, as measured by

Tomasi et al. (2011) • MPH increased the activation of the parietal and prefrontal cortices and increased the deactivation of the insula and posterior cingulate
cortex in healthy adults, as measured by fMRI during visual attention and working memory tasks
Udo de Haes et al. (2005) • striatum 11
MPH reduced [ C]raclopride (a D receptor antagonist) binding in healthy adults, as measured by PET, with binding changes in the dorsal
2
correlating with MPH-induced increases in euphoria
Volkow et al. (1994) • MPH 11
decreased [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults, as measured by PET
• MPH
2
11
Volkow et al. (1996a) decreased [ C]d-threo-MPH binding in the striatum of healthy adults, as measured by PET
Volkow et al. (1997) • increases 18
MPH produced variable changes in brain metabolism in healthy adults, as measured by PET using [ F]FDG, but produced consistent
in cerebellar metabolism and decreases in the basal ganglia
Volkow et al. (1998) • observed 11
MPH dose-dependently reduced [ C]cocaine (a DAT ligand) binding in the striatum of healthy adults, as measured by PET, with effects
at therapeutic doses used for ADHD
Volkow et al. (1999a,b) • ofMPHMPH-induced 11
dose-dependently reduced [ C]cocaine (a DAT ligand) binding in the striatum of healthy adults, as measured by PET, with self-reports
“high” and “rush” being significantly correlated with [ C]cocaine binding
11

Volkow et al. (2001) • MPH 11


reduced [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults, as measured by PET
• MPH
2
11
Volkow et al. (2002a) reduced [ C]cocaine (a DAT ligand) binding in the striatum of healthy adults, as measured by PET
Volkow et al. (2002b) • signal-to-noise
A review of PET studies suggested that mechanism of action of MPH was related to binding of MPH to the DAT levels sufficient to increase
ratios and to increase the salience of stimuli, with variability in MPH response being related to differences in DAT level and
DA release across individuals
Volkow et al. (2004) • measured 11
MPH reduced [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults in a context-dependent manner, as
2
by PET, with reductions observed when a remunerated mathematical task was performed but not when scenery cards were
passively viewed
Wang et al. (1999) • reproducible 11
MPH reduced [ C]raclopride (a D receptor antagonist) binding in the striatum of healthy adults, as measured by PET, with changes being
2
in individuals at 1- to 2-week intervals

5-HT = 5-hydroxytryptamine (serotonin); ADHD = attention-deficit/hyperactivity disorder; AMP = amphetamine; BOLD = blood-oxygen-level dependent; cAMP = cyclic adenosine
monophosphate; CBF = cerebral blood flow; CHPG = 2-chloro-5-hydroxyphenylglycine; CREB = cAMP-responsive element binding protein; DA = dopamine; DAT = dopamine trans-
porter; FDG = 2-deoxy-2-fluoro-D-glucose; FDOPA = fluorodopamine; fMRI = functional magnetic resonance imaging; fNIRS = functional near infrared spectroscopy; FP-CIT = N-ω-
fluoropropyl-2β-carbomethoxy-3β-(4-iodophenyl)-nortropane; GABA = gamma-aminobutyric acid; GLU = glutamate; IBF = iodobenzofuran; IBZM = iodobenzamide;
MAO = monoamine oxidase; mGluR = metabotropic glutamate receptor; MNPA = (R)-2-CH3O-N-n-propylnorapomorphine; MPEP = 2-methyl-6-(phenylethynyl)pyridine hydro-
chloride; MPH = methylphenidate; MRB = methylreboxetine; MRI = magnetic resonance imaging; mRNA = messenger ribonucleic acid; MRS = magnetic resonance spectroscopy;
NE = norepinephrine; NET = norepinephrine transporter; NMDA = N-methyl-D-aspartate; NPA = 2-methoxy-N-propyl-norapomorphine; PET = positron emission tomography;
PHCCC = N-phenyl-7-(hydroxyimino)cyclopropa[b]chromen-1a-carboxamide; PHNO = (+)-4-propyl-3,4,4a,5,6,10b-hexahydro-2H-naphtho[1,2-b][1,4]oxazin-9-ol; PKC = protein ki-
nase C; RGS4 = regulator of G-protein signaling 4; SERT = serotonin transporter; SPECT = single-photon emission computed tomography; TRODAT = ([2-[2-[3-(4-chlorophenyl)-8-
methyl-8-azabicyclo[3,2,1]oct-2-yl]methyl](2-mercaptoethyl)amino]ethyl-amino-ethan-etio-lato-(3-)-oxo-[1R-(exo-exo)]; VMAT-2 = vesicular monoamine transporter 2.

et al., 2015; Price et al., 2002; Schwarz et al., 2007aa), with changes in et al., 2006; Nikolaus et al., 2007; Wall et al., 1995), an affinity for and
CBF being correlated with striatal DA concentration (Kashiwagi et al., agonist activity at the 5-HT1A receptor (Markowitz et al., 2009;
2015). In a functional magnetic resonance imaging (fMRI) study in rats, Markowitz et al., 2006), and redistribution of VMAT-2 (Riddle et al.,
AMP caused widespread increases in blood-oxygen-level–dependent 2007; Sandoval et al., 2002). As a consequence of these interactions,
(BOLD) signal intensity in subcortical structures with rich DA in- MPH elevates extracellular DA and NE levels (Easton et al., 2007bb;
nervation and with decreases in BOLD signal in the superficial layers of Kuczenski and Segal, 1997; Schiffer et al., 2006; Young et al., 2011).
the cortex (Dixon et al., 2005). In another fMRI study in rats, AMP The enhanced efflux of DA and NE associated with MPH exposure re-
reduced BOLD signal intensity in the nucleus accumbens and prefrontal sults in increased availability of DA and NE to bind to their respective
cortex and increased signal intensity in the motor cortex of rats, with transporters (i.e., the DAT or NET) or to DA or NE receptors, as evi-
signal intensity changes in the nucleus accumbens and caudate (but not denced by reductions in ligand binding in PET and SPECT studies (Ding
in the motor cortex) being attenuated by pretreatment with a tyrosine- et al., 1997; Dresel et al., 1999; Gatley et al., 1999; Nikolaus et al.,
free amino acid mixture (Preece et al., 2007). 2005, Nikolaus et al., 2007, 2011; Volkow et al., 1999a).
Although increases in extracellular levels of striatal DA in rats
measured using microdialysis are less pronounced with MPH than with
3.1.2. Methylphenidate AMP, both compounds have been shown to exhibit similar magnitude of
The direct effects of MPH include inhibition of the DAT and NET effects with regard to reductions in DA binding potential as measured
(Dresel et al., 1999; Federici et al., 2005; Gatley et al., 1996; Markowitz

261
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

by PET in rodents and nonhuman primates (Schiffer et al., 2006). Assessments of functional activity using fMRI have provided evi-
Multiple studies have demonstrated that MPH also directly interacts dence for the widespread functional effects of MPH (Costa et al., 2013;
with adrenergic receptors (Andrews and Lavin, 2006; Gamo et al., Moeller et al., 2014; Mueller et al., 2014; Ramaekers et al., 2013;
2010; Markowitz et al., 2009, 2006). Through activation of α2 adre- Schlosser et al., 2009; Tomasi et al., 2011). Using fMRI, it has been
nergic receptors, MPH has been demonstrated to stimulate cortical shown that MPH increases activation of the parietal and prefrontal
excitability (Andrews and Lavin, 2006). Further evidence for the in- cortices and increases deactivation of the insula and posterior cingulate
teraction of MPH with α2 adrenergic receptors comes from data in- cortex during visual attention and working memory tasks (Tomasi et al.,
dicating that the procognitive effects of MPH in a working memory task 2011). Another fMRI study reported MPH-induced activation in the
are blocked by the α2 adrenergic antagonist idazoxan (Gamo et al., putamen during a go/no-go task when a response inhibition error oc-
2010). The effects of MPH on α2 adrenergic receptors are notable given curred but not when a response was successfully inhibited (Costa et al.,
that two α2 adrenergic receptor agonist drugs (extended-release forms 2013), suggesting that the effects of MPH are context dependent. Fur-
of guanfacine and clonidine) are indicated for the treatment of ADHD thermore, MPH exposure altered connectivity strength across various
(Jain and Katic, 2016). cortical and subcortical networks (Mueller et al., 2014) and shifted
brain activation under conditions of uncertainty to higher levels of
3.2. Human neuroimaging studies activation in left and right parahippocampal regions and cerebellar
regions (Schlosser et al., 2009). Lastly, MPH-associated decreases in
3.2.1. Amphetamine task-related errors on the Stroop color-word task were associated with
Reductions in ligand binding in PET (Boileau et al., 2007; Buckholtz concurrent decreases in anterior cingulate cortex activity (Moeller
et al., 2010; Cardenas et al., 2004; Cropley et al., 2008; Drevets et al., et al., 2014). MPH has also been shown to reduce regional CBF in the
2001; Leyton et al., 2002, 2004; Martinez et al., 2003; Narendran et al., prefrontal cortex and increase regional CBF in the thalamus and pre-
2009, 2010; Oswald et al., 2005, 2015; Riccardi et al., 2006aa,b, 2011; central gyrus (Schweitzer et al., 2004). In another study that used
Shotbolt et al., 2012; Slifstein et al., 2010; Wand et al., 2007; Willeit functional near-infrared spectroscopy, MPH-associated improvements
et al., 2008; Woodward et al., 2011) and SPECT (Kegeles et al., 1999; in the performance of a working memory task corresponded with de-
Laruelle et al., 1995; Laruelle and Innis, 1996; Schouw et al., 2013) creased oxy-hemoglobin levels in the right lateral prefrontal cortex,
studies in healthy humans indicate that AMP increases DA release which is a surrogate for decreased neural activation (Ramasubbu et al.,
across multiple brain regions, including the dorsal and ventral striatum, 2012).
substantia nigra, and regions of the cortex. AMP also has been shown to
alter regional CBF to areas of the brain with DA innervation, including 4. Discussion
the striatum, anterior cingulate cortex, prefrontal and parietal cortex,
inferior orbital cortex, thalamus, cerebellum, and amygdala (Devous Although their mechanisms of action differ, the primary central
et al., 2001; Rose et al., 2006; Schouw et al., 2013; Vollenweider et al., nervous system effects of AMP and MPH within the brain include in-
1998; Wolkin et al., 1987). The effects of AMP on regional CBF appear creased catecholamine availability in striatal and cortical regions, as
to be dependent on the dose, with lower doses decreasing rates of blood evidenced in preclinical (Avelar et al., 2013; Covey et al., 2013; Easton
flow in the frontal and temporal cortices and in the striatum (Wolkin et al., 2007bb; Finnema et al., 2015; Floor and Meng, 1996; Jedema
et al., 1987) and higher doses increasing blood flow in the anterior et al., 2014; Joyce et al., 2007; Kuczenski and Segal, 1997; May et al.,
cingulate cortex, caudate nucleus, putamen, and thalamus 1988; Mukherjee et al., 1997; Pum et al., 2007; Ren et al., 2009;
(Vollenweider et al., 1998). In fMRI studies in healthy adults, AMP Schiffer et al., 2006; Xiao and Becker, 1998; Young et al., 2011; Wall
increased BOLD signal variability (Garrett et al., 2015) and exerted an et al., 1995) and human (Boileau et al., 2007; Buckholtz et al., 2010;
“equalizing” effect on ventral striatum activity during incentive pro- Cardenas et al., 2004; Cropley et al., 2008; Drevets et al., 2001; Kegeles
cessing (Knutson et al., 2004). In addition, AMP was shown to et al., 1999; Laruelle et al., 1995; Laruelle and Innis, 1996; Leyton et al.,
strengthen amygdalar responses during the processing of angry and 2002, 2004; Martinez et al., 2003; Narendran et al., 2009, 2010;
fearful facial expressions (Hariri et al., 2002). Oswald et al., 2005, 2015; Riccardi et al., 2006aa,b, 2011; Schouw
Changes in neuronal activity have been shown to correlate with et al., 2013; Shotbolt et al., 2012; Slifstein et al., 2010; Wand et al.,
various behavioral traits (Buckholtz et al., 2010; Drevets et al., 2001; 2007; Willeit et al., 2008; Woodward et al., 2011; Booij et al., 1997;
Leyton et al., 2002; Woodward et al., 2011). In PET studies, changes in Clatworthy et al., 2009; Montgomery et al., 2007; Spencer et al., 2006,
the binding potential of [11C]raclopride (a D2 receptor antagonist) in 2010; Udo de Haes et al., 2005; Volkow et al., 1994, 2001, 2004; Wang
regions of the ventral striatum of healthy adults associated with AMP et al., 1999) studies. These increases in DA and NE availability affect
binding have been reported to be negatively correlated with changes in corticostriatal systems that subserve behaviors related to cognition and
AMP-associated euphoria (Drevets et al., 2001) and with increases in executive function (Moeller et al., 2014; Schlosser et al., 2009; Tomasi
drug wanting and novelty seeking (Leyton et al., 2002). et al., 2011), risky decision making (Oswald et al., 2015), emotional
responsivity (Hariri et al., 2002), and the regulation of reward pro-
3.2.2. Methylphenidate cesses (Haber, 2016). Importantly, ADHD has been associated with
Methylphenidate has also been shown to increase striatal DA structural and functional alterations in regions of the brain where AMP
availability, as measured by reductions in ligand binding potential in and MPH have been shown to alter DA and NE activity.
PET studies (Booij et al., 1997; Clatworthy et al., 2009; Montgomery
et al., 2007; Spencer et al., 2006, 2010; Udo de Haes et al., 2005; 4.1. Structural alterations in ADHD
Volkow et al., 1994, 2001, 2004; Wang et al., 1999), with evidence to
indicate that this effect is related to binding to the DAT (Volkow et al., A meta-analysis of imaging data from individuals with ADHD across
1998, Volkow et al., 1999a,b, 2002a). MPH-induced reductions in all age groups revealed altered white matter integrity in diverse brain
striatal [11C]raclopride binding were associated with MPH-induced areas, including the striatum and the frontal, temporal, and parietal
changes in euphoria and anxiety and were correlated to age (Udo de lobes (van Ewijk et al., 2012). In a meta-analysis of imaging data from
Haes et al., 2005; Volkow et al., 1994). In addition, NE systems have children and adults (Nakao et al., 2011), global gray matter volume was
been implicated as key targets for MPH, with MPH dose-dependently significantly smaller in those with ADHD, especially in basal ganglia
blocking the NET in the thalamus and other NET-rich regions; the es- structures integral to executive function. In adults with ADHD, reduced
timated occupancy of the NET at therapeutic doses of 0.35–0.55 mg/kg gray matter volume in the caudate and parts of the dorsolateral pre-
MPH is 70%–80% (Hannestad et al., 2010). frontal cortex, inferior parietal lobe, anterior cingulate cortex,

262
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

putamen, and cerebellum were observed; increased volume was noted adults with ADHD (Volkow et al., 2012). In another study, in adults
in other parts of the dorsolateral prefrontal cortex and inferior parietal with ADHD, long-term MPH treatment increased striatal DAT avail-
lobe (Seidman et al., 2011). A meta-analysis of 1713 persons with ability (Wang et al., 2013). In another PET study in adults with ADHD,
ADHD and 1529 controls found volumetric reductions in the ac- decreased DAT and D2/D3 receptor availability in the nucleus ac-
cumbens, amygdala, caudate, hippocampus, and putamen (Hoogman cumbens and midbrain compared with individuals without ADHD was
et al., 2017). reported, and reduced availability of DAT and D2/D3 receptor avail-
In a meta-analysis of 9 PET and SPECT studies, which included 169 ability was significantly correlated with lower indices of motivation
patients with ADHD and 173 healthy controls, it was reported that only in those with ADHD (Volkow et al., 2011). Also, a PET study in
striatal DAT density in patients with ADHD was 14% higher than in young male adults with ADHD has also reported dysfunctional DA
healthy controls (Fusar-Poli et al., 2012). Meta-regression analysis metabolism in the putamen, amygdala, and dorsal midbrain relative to
further revealed that previous exposure to ADHD medication influenced healthy controls regardless of treatment status (naive vs previously
striatal DAT density, with lower DAT density being associated with a treated with MPH) and that a history of MPH treatment resulted in a
lack of medication exposure (Fusar-Poli et al., 2012). As the correlation down-regulation of DA turnover (Ludolph et al., 2008). Despite the
between medication exposure and striatal DAT density accounted for neuroimaging evidence that supports a role for the DAT in adult ADHD,
48% of the variance across studies (Fusar-Poli et al., 2012), it was a systematic literature review examining the pharmacogenetics of adult
suggested that the higher striatal DAT density in individuals with ADHD ADHD found that only 1 of 5 identified studies reported finding a
was a neuroadaptive response to stimulant exposure. polymorphism at the DAT gene associated with ADHD (Contini et al.,
2013).
4.2. Functional alterations in ADHD Studies of NET availability have not consistently reported altered
NET availability in individuals with ADHD (Sigurdardottir et al., 2016;
A meta-analysis of imaging data focusing specifically on timing Vanicek et al., 2014), but one study reported genotype-dependent in-
function, which is important for impulsiveness in ADHD, showed con- creases in NET binding in the thalamus and cerebellum of adults with
sistent deficits in the left inferior prefrontal, parietal, and cerebellar ADHD compared with controls; this effect was largely due to the effect
regions of individuals with ADHD (Hart et al., 2012). A meta-analysis of of NET gene polymorphisms on NET binding potential (Sigurdardottir
24 task-related fMRI studies coupled with functional decoding based on et al., 2016). Another study reported no NET differences across multiple
the BrainMap database reported hypoactivation in the left putamen, brain regions, including the hippocampus, thalamus, and midbrain, in
inferior frontal gyrus, temporal pole, and right caudate of individuals individuals with ADHD compared with controls (Vanicek et al., 2014).
with ADHD (Cortese et al., 2016). When examining these deficits in Beyond the changes observed in DA and NE systems in individuals
regard to the BrainMap database, it was suggested that individuals with with ADHD, there is evidence that GLU, 5-HT, ACH, and opioid systems
ADHD may exhibit deficits in the cognitive aspects of music, perception play a role in ADHD. Studies examining neurometabolism using proton
and audition, speech and language, and executive function (Cortese MRIs have reported glutamatergic deficits in the frontal cortical and
et al., 2016). striatal regions in individuals with ADHD that may be related to cog-
In high-functioning, drug-naive young adults with ADHD, resting- nitive control and symptom severity (Maltezos et al., 2014; Dramsdahl
state fMRIs revealed altered connectivity in the orbitofrontal-temporal- et al., 2011; Arcos-Burgos et al., 2012). Regarding the 5-HT systems, it
occipital and frontal-amygdala-occipital networks (relating to in- has been reported that increased methylation of the 5-HT transporter is
attentive and hyperactive/impulsive symptoms, respectively) compared associated with worse clinical presentation and reduced cortical
with matched controls; these abnormalities were not related to devel- thickness in children with ADHD (Park et al., 2015). In adults, sig-
opmental delays, impaired cognition, or use of pharmacotherapy nificant differences in 5-HT transporter interregional correlations be-
(Cocchi et al., 2012). Imaging studies have also reported hypoactivity in tween the precuneus and hippocampus have been reported in adults
the prefrontal cortex and weak connections to other brain regions in with ADHD compared with controls without ADHD (Vanicek et al.,
individuals with ADHD (Arnsten, 2009). A review of progress in neu- 2017). Functionally, it has been reported that decreased levels of acti-
roimaging indicated that the initial focus on frontostriatal dysfunction vation are observed in the precuneus of adolescents with ADHD com-
has given way to a broader understanding of the complex interactions pared with individuals without ADHD; however, there is a significantly
of various regions of the brain in which alterations may contribute to greater increase in the activation of the precuneus following the ad-
ADHD symptoms across the life span (Cortese and Castellanos, 2012). ministration of fluoxetine in adolescents with ADHD compared with
A substantial body of literature has examined the role of DA systems controls (Chantiluke et al., 2015). At present, there are no published
in the neurobiology of ADHD and ADHD-related symptoms and beha- neuroimaging studies of endogenous opioid or ACH systems in in-
viors. Among treatment-naive adolescents with ADHD, DAT density dividuals with ADHD. However, altered function in both systems has
shows a significant inverse relationship with blood flow in the cingulate been implicated in ADHD (Blum et al., 2008; Potter et al., 2014).
cortex, the frontal and temporal lobes, and the cerebellum, brain re-
gions which are involved in modulating attention (da Silva et al., 2011). 4.3. Co-occurring psychiatric conditions and ADHD
In a PET study of treatment-naive men with ADHD and men without
ADHD, more pronounced reductions in AMP-induced reductions in Attention-deficit/hyperactivity disorder is often associated with
striatal [11C]raclopride binding were associated with worse response comorbid psychiatric disorders (Mao and Findling, 2014; Kooij et al.,
inhibition, and those with ADHD had the highest-magnitude reductions 2012; Konofal et al., 2010), such as anxiety and mood disorders. Im-
and AMP-induced reductions in striatal [11C]raclopride binding portantly, the same brain regions and neurotransmitter systems that
(Cherkasova et al., 2014). Two SPECT studies and 1 PET study showed underlie ADHD are also implicated in the psychiatric disorders that are
that adults with ADHD had higher DAT concentrations than adults frequently comorbid with ADHD (Farb and Ratner, 2014; Sternat and
without ADHD (Dresel et al., 2000; Krause et al., 2000; Spencer et al., Katzman, 2016; Schwartz and Kilduff, 2015). Thus, it is important to
2007), with the SPECT studies further reporting that MPH reduced DAT understand how ADHD therapies might influence psychiatric co-
availability in adults with ADHD (Dresel et al., 2000; Krause et al., morbidities.
2000). Furthermore, studies have shown significant correlations be-
tween global clinical improvement in ADHD symptoms following MPH 4.3.1. Anxiety disorders
treatment and striatal DAT availability (Krause et al., 2005; la Fougere In healthy human volunteers, AMP has been reported to potentiate
et al., 2006), and the MPH-induced increases in DA availability in the amygdalar activity in response to the processing of angry and fearful
ventral striatum are associated with improved ADHD symptomology in facial expressions (Hariri et al., 2002). These data provide a potential

263
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

neurobiologic basis for the anxiogenic effects of AMP (Hariri et al., 4.3.3. Substance use disorders
2002). However, it has been theorized that stimulant-associated aug- The neurobiology of reward and addiction and the key role of me-
mentation of serotonergic drive could ameliorate the comorbid anxiety solimbic DA systems have been described in great detail (Volkow and
associated with ADHD (Heal et al., 2013). In practice, the effects of Morales, 2015; Koob, 2006). Although associations have been made
stimulants on anxiety can be complex, with acute administration of between ADHD and substance abuse, their relationship is complex.
MPH reducing anxiety in adults and chronic treatment during early life Several reviews have emphasized that substance use disorders can be
increasing anxiety during adulthood (Sanchez-Perez et al., 2012). comorbid with ADHD (Mao and Findling, 2014; Kooij et al., 2012). For
example, in a study of 208 adults diagnosed with ADHD and treated
with psychostimulants as youths, the relative risk of having a diagnosis
4.3.2. Depressive disorders of substance use disorder or alcohol abuse, respectively, compared with
Psychostimulants have been used in the treatment of major de- the general population was 7.7 or 5.2 (Dalsgaard et al., 2014). Fur-
pressive disorder (MDD) since the early 1950s, when the use of MPH thermore, a 2012 review noted that there was evidence for increased
was first examined for MDD (Robin and Wiseberg, 1958). The rationale rates of substance abuse in individuals with ADHD treated with psy-
for examining the potential utility of psychostimulants in depressive chostimulants (Nelson and Galon, 2012). Given the known abuse lia-
disorders is based on preclinical and clinical evidence implicating DA in bility of psychostimulants and data indicating that psychostimulant
depressive symptomatology (Treadway and Zald, 2011). However, the medications are associated with misuse and diversion (Rabiner et al.,
rapid onset of action of psychostimulants suggests the mechanisms by 2009; Garnier et al., 2010), it is not surprising that psychostimulant
which they may influence depressive symptoms is likely to differ from medications approved for use in ADHD are schedule II medications with
that of antidepressants (Malhi et al., 2016). black box warnings for potential drug dependence (Panagiotou et al.,
Based on the published literature, the effects of psychostimulants on 2011). Some treatment guidelines suggest that nonstimulant alter-
depressive symptoms appear equivocal. Although one meta-analysis natives be considered as therapies for ADHD when issues related to
published in 2008 based on 3 short-term trials reported statistically abuse and dependence are a concern (Atkinson and Hollis, 2010; Bolea-
significant improvements favoring monotherapy with psychostimulants Alamanac et al., 2014; Pliszka and AACAP Work Group on Quality
compared with placebo for depressive symptoms (Candy et al., 2008), a Issues, 2007).
systematic review of augmentation therapy for MDD published in 2009 However, multiple studies have provided evidence that psychosti-
based on MPH (2 studies) and modafinil (2 studies) reported that nei- mulant treatment in individuals with substance use disorders and
ther augmentation strategy was clinically superior to antidepressant ADHD is not associated with a significant worsening of substance abuse
monotherapy in reducing depressive symptoms (Fleurence et al., 2009). (Konstenius et al., 2010; Levin et al., 2015). In a study that examined
In addition, 2 large phase 3 studies and a phase 2 study of lisdex- the efficacy of extended-release mixed AMP salts in the treatment of
amfetamine dimesylate augmentation in adults with MDD and in- ADHD symptoms and cocaine abuse in 126 cocaine-dependent adults
adequate response to antidepressant therapy failed to meet their pri- with ADHD (Levin et al., 2015), significantly greater reductions in
mary efficacy endpoint versus placebo (Richards et al., 2016; Richards ADHD symptoms and higher abstinence from cocaine use were ob-
et al., 2017). The lack of treatment effect in these recently published served with extended-release mixed AMP salts than placebo. In another
studies, taken together with inconsistent findings based on meta-ana- study, osmotic-controlled release oral delivery system (OROS) MPH did
lyses and systematic reviews, suggests that psychostimulants are in- not produce significantly greater reductions in ADHD symptoms than
effective in treating the undifferentiated symptoms of depression. Psy- placebo in 24 AMP-dependent adults with newly diagnosed ADHD.
chostimulants have also been examined in other mood disorders, However, OROS MPH treatment also was not associated with evidence
including treatment-resistant depression, bipolar depression, and de- of increased AMP abuse, as measured by self-reported days of AMP use
pression associated with specific medical conditions (Dell’Osso and or craving for AMP, time to relapse, or cumulative abstinence duration
Ketter, 2013). For bipolar depression, there is some evidence sup- (Konstenius et al., 2010).
porting the efficacy of psychostimulant augmentation, but the quality
and quantity of studies does not allow for a strong evidence-based re- 4.3.4. Sleep disturbances
commendation for their use to be put forth (Dell’Osso and Ketter, The neurobiologic substrates of sleep are diverse and distributed
2013). throughout the brain, with monoaminergic systems playing an im-
Despite inconclusive evidence regarding the efficacy of psychosti- portant role in wakefulness (Schwartz and Kilduff, 2015). Substantial
mulants in treating depressive symptoms, the continued publication of literature exists regarding the sleep disturbances associated with ADHD,
review articles on this topic (Malhi et al., 2016; McIntyre et al., 2017; which include insomnia, disordered sleep, difficulty falling asleep, sleep
Hegerl and Hensch, 2017) demonstrates continued interest in this area apnea, daytime somnolence, and increased nocturnal motor activity
of research. These review articles hypothesize that the lack of consistent (see (Konofal et al., 2010; Cohen-Zion and Ancoli-Israel, 2004;
clinical efficacy of psychostimulant augmentation could be due to Lecendreux and Cortese, 2007; Snitselaar et al., 2017) for reviews).
poorly defined psychopathology and that psychostimulant effects may Evidence suggests that impaired and/or disordered sleep is present in
be more pronounced in selected symptom domains (McIntyre et al., individuals not being treated with psychostimulants (Konofal et al.,
2017; Hegerl and Hensch, 2017) or that the effects of psychostimulants 2010). For example, in a study of the effects of MPH on sleep in children
are short-lasting (Malhi et al., 2016). It has also been speculated that with ADHD, parents reported that approximately 10% of study parti-
the use of clinician-rated scales (rather than patient-rated scales) in cipants had sleep problems before starting their medication (Becker
some studies (Richards et al., 2016, 2017) may not have adequately et al., 2016). However, in regard to the reported effects of psychosti-
captured the effects of psychostimulant treatment. mulants on sleep in individuals with ADHD, there are some dis-
It should also be noted that the use of stimulants as augmentation crepancies. In a meta-analysis of 9 articles, the use of psychostimulant
agents in combination with tricyclic antidepressants and MAO in- medication was associated with longer sleep latency, worse sleep effi-
hibitors is controversial, with issues concerning the possible develop- ciency, and shorter sleep duration (Kidwell et al., 2015). A review of the
ment of an adrenergic crisis, the emergence of serotonin syndrome, or a safety and tolerability of ADHD medications noted that insomnia was
hypertensive crisis being raised (Stotz et al., 1999). In fact, both MPH- one of the most commonly reported adverse events associated with
based agents and AMP-based agents are contraindicated in individuals psychostimulant treatment (Duong et al., 2012). In contrast, some
taking an MAO inhibitor (currently or within the preceding 2 weeks) studies have shown that psychostimulants have no significant negative
because a hypertensive crisis may result (Thomas et al., 2015). impact on sleep (Becker et al., 2016; Owens et al., 2016; Surman and
Roth, 2011). A post hoc analysis of the effects of lisdexamfetamine or

264
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

SHP465 mixed amphetamine salts in adults with ADHD demonstrated 5. Conclusions


that the proportions of participants exhibiting a worsening of sleep
during treatment, as measured by the Pittsburgh Sleep Quality Index, Based on the published literature, the primary pharmacologic ef-
did not differ from that of placebo (Surman and Roth, 2011). Dis- fects of both AMP and MPH are related to increased central DA and NE
crepancies in the effects of stimulants on sleep in individuals with activity in brain regions that include the cortex and striatum. These
ADHD might be attributable to various factors, including sleep quality regions are involved in the regulation of executive and attentional
prior to treatment, the stimulant formulation, the length of treatment, function (Faraone et al., 2015). In ADHD, dysfunction in the DA and NE
and the method of sleep assessment (Cohen-Zion and Ancoli-Israel, systems, which are critical to proper cortical and striatal function, likely
2004; Becker et al., 2016; Kidwell et al., 2015). For example, in a study account for some of the pathophysiology of ADHD (Cortese, 2012;
of MPH in children with ADHD, 23% of participants without preexisting Faraone et al., 2015; Arnsten, 2009). Although it is a limitation of the
sleep problems developed sleep problems while taking MPH, whereas review that the only database searched was PubMed, it is unlikely that
68.5% of those with preexisting sleep problems no longer experienced important studies were not captured.
sleep problems after taking MPH (Becker et al., 2016). It has been speculated that the moderately greater efficacy of AMP-
based agents compared with MPH-based agents in ADHD may be re-
lated to differences in their molecular actions (Faraone and Buitelaar,
4.4. Other safety concerns 2010), but to date there is no conclusive clinical evidence to support
this speculation. Furthermore, there is no conclusive clinical evidence
In addition to considering the potential effect of AMP and MPH in supporting a prospective choice for an AMP-based agent over an MPH-
individuals with other comorbid psychiatric disorders, the peripheral based agent (or vice versa) based on the mechanisms of action of these
effects of AMP and MPH related to their pharmacology need to be drug classes. As such, the current understanding of differences in the
considered, particularly in regard to their effects on cardiovascular mechanisms of action of AMP and MPH has not led to clinical guide-
function. Increased heart rate and blood pressure are among the most lines regarding their use in specific patient populations. Furthermore, it
frequent treatment-emergent adverse events reported with psychosti- is possible that differences in the pharmacologic profile between AMP
mulant treatment (Duong et al., 2012; Vaughan and Kratochvil, 2012); and MPH, in combination with the complexities associated with the
increases in pulse and blood pressure are also frequently reported etiology of ADHD (Faraone et al., 2015), contribute to individual dif-
(Duong et al., 2012; Vaughan and Kratochvil, 2012; Santosh et al., ferences in treatment response to AMP-based agents or MPH-based
2011). There is also a safety concern related to the potential for adverse agents in individuals with ADHD. Interactions among these factors
cardiovascular outcomes, including ischemic attacks, myocardial in- might explain why some patients have a differential response to these
farction, and stroke (Westover and Halm, 2012). In a self-controlled drugs.
case series analysis, treatment with MPH was associated with an in- When contemplating pharmacotherapy for ADHD, in addition to
creased overall risk of arrhythmia and with an increased risk of myo- taking into account the potential for adverse cardiovascular outcomes
cardial infarction from 1 week to 2 months after treatment initiation in (Westover and Halm, 2012), the presence of comorbid psychiatric dis-
children and adolescents with ADHD (Shin et al., 2016). Another study orders should be considered. Multiple psychiatric comorbidities, in-
reported that the risk for an emergency department visit for cardiac- cluding depression and anxiety (Mao and Findling, 2014; Kooij et al.,
related reasons among youths did not differ between those being 2012), are thought to be mediated in part by shared neurobiological
treated with AMP versus MPH (Winterstein et al., 2009). However, pathways that are also implicated in the pathophysiology of ADHD
these findings should be considered in light of data from 2 large studies (Farb and Ratner, 2014; Sternat and Katzman, 2016). As such, the effect
that reported no significant increase in risk for cardiovascular events in of psychostimulant treatment on the symptoms of these disorders and
current users of ADHD medications compared with nonusers (Cooper the potential interactions with medications used to treat these disorders
et al., 2011; Habel et al., 2011). need to be considered.
The neurotransmitter systems responsible for stimulant-associated
adverse events and safety concerns are in large part related to stimu- Author contribution
lation of peripheral NE activity (Duong et al., 2012; Westover and
Halm, 2012). Based on these concerns, package inserts for stimulants Dr. Faraone designed the systematic literature review, is responsible
include black box warnings regarding the potential for serious adverse for the content of the manuscript, and approved the final draft.
cardiovascular events (Panagiotou et al., 2011). Assessments of the
risks and benefits of stimulant therapy for ADHD should be made on an Disclosures
individual basis, and individuals on psychostimulant treatment should
be monitored. In the past year, Dr. Faraone received income, potential income,
Another issue of potential concern is the possible neuroin- travel expenses, continuing education support, and/or research support
flammatory effects of psychostimulants. In rats, MPH administration from Otsuka, Lundbeck, Kenpharma, Rhodes, Arbor, Ironshore, Shire,
has been reported to produce neuroinflammation and oxidative stress in Akili Interactive Labs, CogCubed, Alcobra, VAYA, Sunovion, Genomind,
the hippocampus and cerebral cortex, as measured by the inflammatory and NeuroLifeSciences. With his institution, he holds US patent
markers tumor necrosis factor α and interleukin 1β (Motaghinejad US20130217707 A1 for the use of sodium-hydrogen exchange in-
et al., 2017, 2016). In a systematic review of 14 studies (Anand et al., hibitors in the treatment of ADHD.
2017), no study assessed the relationship between psychostimulant
treatment and neuroinflammation so the relevance of preclinical Acknowledgments
models of neuroinflammation to psychostimulant treatment in in-
dividuals with ADHD is unknown. The same review did find evidence Shire Development LLC (Lexington, MA) provided funding to
suggesting a role for inflammation in the pathogenesis of ADHD (Anand Complete Healthcare Communications, LLC (CHC; West Chester, PA),
et al., 2017), which is consistent with a meta-analysis finding elevated an ICON plc company, for support in writing and editing this manu-
levels of oxidative stress in patients diagnosed with ADHD (Joseph script. Under the direction of the author, writing assistance was pro-
et al., 2015). Oxidative stress has also been implicated in the lower vided by Madhura Mehta, PhD, and Craig Slawecki, PhD, employees of
brain volumes seen in ADHD patients (Hess et al., 2017). CHC. Editorial assistance in the form of proofreading, copyediting, and
fact checking was also provided by CHC. The author exercised full
control over the content throughout the development and had final

265
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

approval of the manuscript for submission. amphetamine-induced dopamine release in rhesus monkeys exposed to subchronic
Dr. Faraone is supported by the K. G. Jebsen Centre for Research on amphetamine. Neuropsychopharmacology 22 (1), 4–13.
Catala-Lopez, F., Hutton, B., Nunez-Beltran, A., Page, M.J., Ridao, M., Macias Saint-
Neuropsychiatric Disorders, University of Bergen, Bergen, Norway, the Gerons, D., et al., 2017. The pharmacological and non-pharmacological treatment of
European Union’s Seventh Framework Programme for research, tech- attention deficit hyperactivity disorder in children and adolescents: a systematic re-
nological development, and demonstration under grant agreement no view with network meta-analyses of randomised trials. PLoS One 12 (7), e0180355.
Chantiluke, K., Barrett, N., Giampietro, V., Brammer, M., Simmons, A., Murphy, D.G.,
602805, the European Union’s Horizon 2020 research and innovation et al., 2015. Inverse effect of fluoxetine on medial prefrontal cortex activation during
programme under grant agreement No 667302, and NIMH grants reward reversal in ADHD and autism. Cereb. Cortex 25 (7), 1757–1770.
5R01MH101519 and U01 MH109536-01. Chen, Y.I., Wang, F.N., Nelson, A.J., Xu, H., Kim, Y., Rosen, B.R., et al., 2008. Electrical
stimulation modulates the amphetamine-induced hemodynamic changes: an fMRI
study to compare the effect of stimulating locations and frequencies on rats. Neurosci.
References Lett. 444 (2), 117–121.
Chen, R., Furman, C.A., Zhang, M., Kim, M.N., Gereau, R.W., Leitges, M., et al., 2009.
Protein kinase Cbeta is a critical regulator of dopamine transporter trafficking and
Aalto, S., Hirvonen, J., Kaasinen, V., Hagelberg, N., Kajander, J., Nagren, K., et al., 2009.
regulates the behavioral response to amphetamine in mice. J. Pharmacol. Exp. Ther.
The effects of d-amphetamine on extrastriatal dopamine D2/D3 receptors: a rando-
328 (3), 912–920.
mized, double-blind, placebo-controlled PET study with [11C]FLB 457 in healthy
Cherkasova, M.V., Faridi, N., Casey, K.F., O’Driscoll, G.A., Hechtman, L., Joober, R., et al.,
subjects. Eur. J. Nucl. Med. Mol. Imaging 36 (3), 475–483.
2014. Amphetamine-induced dopamine release and neurocognitive function in
al-Tikriti, M.S., Baldwin, R.M., Zea-Ponce, Y., Sybirska, E., Zoghbi, S.S., Laruelle, M.,
treatment-naive adults with ADHD. Neuropsychopharmacology 39 (6), 1498–1507.
et al., 1994. Comparison of three high affinity SPECT radiotracers for the dopamine
Choe, E.S., Chung, K.T., Mao, L., Wang, J.Q., 2002. Amphetamine increases phosphor-
D2 receptor. Nucl. Med. Biol. 21 (2), 179–188.
ylation of extracellular signal-regulated kinase and transcription factors in the rat
Anand, D., Colpo, G.D., Zeni, G., Zeni, C.P., Teixeira, A.L., 2017. Attention-deficit/hy-
striatum via group I metabotropic glutamate receptors. Neuropsychopharmacology
peractivity disorder and inflammation: what does current knowledge tell us? A sys-
27 (4), 565–575.
tematic review. Front. Psychiatry 8, 228.
Chou, Y.H., Halldin, C., Farde, L., 2000. Effect of amphetamine on extrastriatal D2 do-
Andrews, G.D., Lavin, A., 2006. Methylphenidate increases cortical excitability via acti-
pamine receptor binding in the primate brain: a PET study. Synapse 38 (2), 138–143.
vation of alpha-2 noradrenergic receptors. Neuropsychopharmacology 31 (3),
Clatworthy, P.L., Lewis, S.J., Brichard, L., Hong, Y.T., Izquierdo, D., Clark, L., et al., 2009.
594–601.
Dopamine release in dissociable striatal subregions predicts the different effects of
Annamalai, B., Mannangatti, P., Arapulisamy, O., Ramamoorthy, S., Jayanthi, L.D., 2010.
oral methylphenidate on reversal learning and spatial working memory. J. Neurosci.
Involvement of threonine 258 and serine 259 motif in amphetamine-induced nor-
29 (15), 4690–4696.
epinephrine transporter endocytosis. J. Neurochem. 115 (1), 23–35.
Cocchi, L., Bramati, I.E., Zalesky, A., Furukawa, E., Fontenelle, L.F., Moll, J., et al., 2012.
Arcos-Burgos, M., Londono, A.C., Pineda, D.A., Lopera, F., Palacio, J.D., Arbelaez, A.,
Altered functional brain connectivity in a non-clinical sample of young adults with
et al., 2012. Analysis of brain metabolism by proton magnetic resonance spectroscopy
attention-deficit/hyperactivity disorder. J. Neurosci. 32 (49), 17753–17761.
(1H-MRS) in attention-deficit/hyperactivity disorder suggests a generalized differ-
Cohen-Zion, M., Ancoli-Israel, S., 2004. Sleep in children with attention-deficit hyper-
ential ontogenic pattern from controls. Atten. Defic. Hyperact. Disord. 4 (4),
activity disorder (ADHD): a review of naturalistic and stimulant intervention studies.
205–212.
Sleep Med. Rev. 8 (5), 379–402.
Arnsten, A.F., 2009. Toward a new understanding of attention-deficit hyperactivity dis-
Colasanti, A., Searle, G.E., Long, C.J., Hill, S.P., Reiley, R.R., Quelch, D., et al., 2012.
order pathophysiology: an important role for prefrontal cortex dysfunction. CNS
Endogenous opioid release in the human brain reward system induced by acute
Drugs 23 (Suppl. 1), 33–41.
amphetamine administration. Biol. Psychiatry 72 (5), 371–377.
Atkinson, M., Hollis, C., 2010. NICE guideline: attention deficit hyperactivity disorder.
Contini, V., Rovaris, D.L., Victor, M.M., Grevet, E.H., Rohde, L.A., Bau, C.H., 2013.
Arch. Dis. Child Educ. Pract. Ed. 95 (1), 24–27.
Pharmacogenetics of response to methylphenidate in adult patients with attention-
Avelar, A.J., Juliano, S.A., Garris, P.A., 2013. Amphetamine augments vesicular dopa-
deficit/hyperactivity disorder (ADHD): a systematic review. Eur.
mine release in the dorsal and ventral striatum through different mechanisms. J.
Neuropsychopharmacol. 23 (6), 555–560.
Neurochem. 125 (3), 373–385.
Cooper, W.O., Habel, L.A., Sox, C.M., Chan, K.A., Arbogast, P.G., Cheetham, T.C., et al.,
Barkley, R.A., Fischer, M., Smallish, L., Fletcher, K., 2002. The persistence of attention-
2011. ADHD drugs and serious cardiovascular events in children and young adults. N.
deficit/hyperactivity disorder into young adulthood as a function of reporting source
Engl. J. Med. 365 (20), 1896–1904.
and definition of disorder. J. Abnorm. Psychol. 111 (2), 279–289.
Cortese, S., 2012. The neurobiology and genetics of attention-deficit/hyperactivity dis-
Bartl, J., Link, P., Schlosser, C., Gerlach, M., Schmitt, A., Walitza, S., et al., 2010. Effects
order (ADHD): what every clinician should know. Eur. J. Paediatr. Neurol. 16 (5),
of methylphenidate: the cellular point of view. Atten. Defic. Hyperact. Disord. 2 (4),
422–433.
225–232.
Cortese, S., Castellanos, F.X., 2012. Neuroimaging of attention-deficit/hyperactivity dis-
Becker, S.P., Froehlich, T.E., Epstein, J.N., 2016. Effects of methylphenidate on sleep
order: current neuroscience-informed perspectives for clinicians. Curr. Psychiatry
functioning in children with attention-deficit/hyperactivity disorder. J. Dev. Behav.
Rep. 14 (5), 568–578.
Pediatr. 37 (5), 395–404.
Cortese, S., Adamo, N., Mohr-Jensen, C., Hayes, A.J., Bhatti, S., Carucci, S., et al., 2017.
Bjorklund, O., Halldner-Henriksson, L., Yang, J., Eriksson, T.M., Jacobson, M.A., Dare, E.,
Comparative efficacy and tolerability of pharmacological interventions for attention-
et al., 2008. Decreased behavioral activation following caffeine, amphetamine and
deficit/hyperactivity disorder in children, adolescents and adults: protocol for a
darkness in A3 adenosine receptor knock-out mice. Physiol. Behav. 95 (5), 668–676.
systematic review and network meta-analysis. BMJ Open 7 (1), e013967.
Blum, K., Chen, A.L., Braverman, E.R., Comings, D.E., Chen, T.J., Arcuri, V., et al., 2008.
Cortese, S., Castellanos, F.X., Eickhoff, C.R., D’Acunto, G., Masi, G., Fox, P.T., et al., 2016.
Attention-deficit-hyperactivity disorder and reward deficiency syndrome.
Functional decoding and meta-analytic connectivity modeling in adult attention-
Neuropsychiatr. Dis. Treat. 4 (5), 893–918.
deficit/hyperactivity disorder. Biol. Psychiatry 80 (12), 896–904.
Boileau, I., Dagher, A., Leyton, M., Welfeld, K., Booij, L., Diksic, M., et al., 2007.
Costa, A., Riedel, M., Pogarell, O., Menzel-Zelnitschek, F., Schwarz, M., Reiser, M., et al.,
Conditioned dopamine release in humans: a positron emission tomography [11C]
2013. Methylphenidate effects on neural activity during response inhibition in
raclopride study with amphetamine. J. Neurosci. 27 (15), 3998–4003.
healthy humans. Cereb. Cortex 23 (5), 1179–1189.
Bolea-Alamanac, B., Nutt, D.J., Adamou, M., Asherson, P., Bazire, S., Coghill, D., et al.,
Covey, D.P., Juliano, S.A., Garris, P.A., 2013. Amphetamine elicits opposing actions on
2014. Evidence-based guidelines for the pharmacological management of attention
readily releasable and reserve pools for dopamine. PLoS One 8 (5), e60763.
deficit hyperactivity disorder: update on recommendations from the British
Cropley, V.L., Innis, R.B., Nathan, P.J., Brown, A.K., Sangare, J.L., Lerner, A., et al., 2008.
Association for Psychopharmacology. J. Psychopharmacol. 28 (3), 179–203.
Small effect of dopamine release and no effect of dopamine depletion on [18F]fall-
Booij, J., Korn, P., Linszen, D.H., van Royen, E.A., 1997. Assessment of endogenous do-
ypride binding in healthy humans. Synapse 62 (6), 399–408.
pamine release by methylphenidate challenge using iodine-123 iodobenzamide
da Silva Jr., N., Szobot, C.M., Anselmi, C.E., Jackowski, A.P., Chi, S.M., Hoexter, M.Q.,
single-photon emission tomography. Eur. J. Nucl. Med. 24 (6), 674–677.
et al., 2011. Attention deficit/hyperactivity disorder: is there a correlation between
Buckholtz, J.W., Treadway, M.T., Cowan, R.L., Woodward, N.D., Li, R., Ansari, M.S.,
dopamine transporter density and cerebral blood flow? Clin. Nucl. Med. 36 (8),
et al., 2010. Dopaminergic network differences in human impulsivity. Science 329
656–660.
(5991), 532.
Dalsgaard, S., Mortensen, P.B., Frydenberg, M., Thomsen, P.H., 2014. ADHD, stimulant
Canadian Attention Deficit Hyperactivity Disorder Resource Alliance (CADDRA), 2011.
treatment in childhood and subsequent substance abuse in adulthood - a naturalistic
Canadian ADHD Practice Guidelines, third ed. CADDRA, Toronto, ON, Canada.
long-term follow-up study. Addict. Behav. 39 (1), 325–328.
Candy, M., Jones, L., Williams, R., Tookman, A., King, M., 2008. Psychostimulants for
dela Pena, I., Gevorkiana, R., Shi, W.X., 2015. Psychostimulants affect dopamine trans-
depression. Cochrane Database Syst. Rev. (2), CD006722.
mission through both dopamine transporter-dependent and independent mechan-
Cardenas, L., Houle, S., Kapur, S., Busto, U.E., 2004. Oral D-amphetamine causes pro-
isms. Eur. J. Pharmacol. 764, 562–570.
longed displacement of [11C]raclopride as measured by PET. Synapse 51 (1), 27–31.
Dell’Osso, B., Ketter, T.A., 2013. Use of adjunctive stimulants in adult bipolar depression.
Carey, C.E., Agrawal, A., Bucholz, K.K., Hartz, S.M., Lynskey, M.T., Nelson, E.C., et al.,
Int. J. Neuropsychopharmacol. 16 (1), 55–68.
2016. Associations between polygenic risk for psychiatric disorders and substance
Devous Sr., M.D., Trivedi, M.H., Rush, A.J., 2001. Regional cerebral blood flow response
involvement. Front. Genet. 7, 149.
to oral amphetamine challenge in healthy volunteers. J. Nucl. Med. 42 (4), 535–542.
Carson, R.E., Breier, A., de Bartolomeis, A., Saunders, R.C., Su, T.P., Schmall, B., et al.,
Dewey, S.L., Smith, G.S., Logan, J., Brodie, J.D., Fowler, J.S., Wolf, A.P., 1993. Striatal
1997. Quantification of amphetamine-induced changes in [11C]raclopride binding
binding of the PET ligand 11C-raclopride is altered by drugs that modify synaptic
with continuous infusion. J. Cereb. Blood Flow. Metab. 17 (4), 437–447.
dopamine levels. Synapse 13 (4), 350–356.
Castner, S.A., al-Tikriti, M.S., Baldwin, R.M., Seibyl, J.P., Innis, R.B., Goldman-Rakic, P.S.,
Ding, Y.S., Fowler, J.S., Volkow, N.D., Gatley, S.J., Logan, J., Dewey, S.L., et al., 1994.
2000. Behavioral changes and [123I]IBZM equilibrium SPECT measurement of
Pharmacokinetics and in vivo specificity of [11C]dl-threo-methylphenidate for the

266
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

presynaptic dopaminergic neuron. Synapse 18 (2), 152–160. 2015. Amphetamine modulates brain signal variability and working memory in
Ding, Y.S., Fowler, J.S., Volkow, N.D., Dewey, S.L., Wang, G.J., Logan, J., et al., 1997. younger and older adults. Proc. Natl. Acad. Sci. U. S. A. 112 (24), 7593–7598.
Chiral drugs: comparison of the pharmacokinetics of [11C]d-threo and L-threo-me- Gatley, S.J., Ding, Y.S., Volkow, N.D., Chen, R., Sugano, Y., Fowler, J.S., 1995. Binding of
thylphenidate in the human and baboon brain. Psychopharmacology (Berl.) 131 (1), d-threo-[11C]methylphenidate to the dopamine transporter in vivo: insensitivity to
71–78. synaptic dopamine. Eur. J. Pharmacol. 281 (2), 141–149.
Dixon, A.L., Prior, M., Morris, P.M., Shah, Y.B., Joseph, M.H., Young, A.M., 2005. Gatley, S.J., Volkow, N.D., Gifford, A.N., Fowler, J.S., Dewey, S.L., Ding, Y.S., et al., 1999.
Dopamine antagonist modulation of amphetamine response as detected using phar- Dopamine-transporter occupancy after intravenous doses of cocaine and methyl-
macological MRI. Neuropharmacology 48 (2), 236–245. phenidate in mice and humans. Psychopharmacology (Berl.) 146 (1), 93–100.
Dramsdahl, M., Ersland, L., Plessen, K.J., Haavik, J., Hugdahl, K., Specht, K., 2011. Adults Gatley, S.J., Pan, D., Chen, R., Chaturvedi, G., Ding, Y.S., 1996. Affinities of methyl-
with attention-deficit/hyperactivity disorder - a brain magnetic resonance spectro- phenidate derivatives for dopamine, norepinephrine and serotonin transporters. Life
scopy study. Front. Psychiatry 2, 65. Sci. 58 (12), 231–239.
Dresel, S.H., Kung, M.P., Huang, X., Plossl, K., Hou, C., Shiue, C.Y., et al., 1999. In vivo Ginovart, N., Farde, L., Halldin, C., Swahn, C.G., 1999. Changes in striatal D2-receptor
imaging of serotonin transporters with [99mTc]TRODAT-1 in nonhuman primates. density following chronic treatment with amphetamine as assessed with PET in
Eur. J. Nucl. Med. 26 (4), 342–347. nonhuman primates. Synapse 31 (2), 154–162.
Dresel, S., Krause, J., Krause, K.H., LaFougere, C., Brinkbaumer, K., Kung, H.F., et al., Guterstam, J., Jayaram-Lindstrom, N., Cervenka, S., Frost, J.J., Farde, L., Halldin, C.,
2000. Attention deficit hyperactivity disorder: binding of [99mTc]TRODAT-1 to the et al., 2013. Effects of amphetamine on the human brain opioid system-a positron
dopamine transporter before and after methylphenidate treatment. Eur J. Nucl. Med. emission tomography study. Int. J. Neuropsychopharmacol. 16 (4), 763–769.
27 (10), 1518–1524. Habel, L.A., Cooper, W.O., Sox, C.M., Chan, K.A., Fireman, B.H., Arbogast, P.G., et al.,
Drevets, W.C., Price, J.C., Kupfer, D.J., Kinahan, P.E., Lopresti, B., Holt, D., et al., 1999. 2011. ADHD medications and risk of serious cardiovascular events in young and
PET measures of amphetamine-induced dopamine release in ventral versus dorsal middle-aged adults. JAMA 306 (24), 2673–2683.
striatum. Neuropsychopharmacology 21 (6), 694–709. Haber, S.N., 2016. Corticostriatal circuitry. Dialogues Clin. Neurosci. 18 (1), 7–21.
Drevets, W.C., Gautier, C., Price, J.C., Kupfer, D.J., Kinahan, P.E., Grace, A.A., et al., Hannestad, J., Gallezot, J.D., Planeta-Wilson, B., Lin, S.F., Williams, W.A., van Dyck, C.H.,
2001. Amphetamine-induced dopamine release in human ventral striatum correlates et al., 2010. Clinically relevant doses of methylphenidate significantly occupy nor-
with euphoria. Biol. Psychiatry 49 (2), 81–96. epinephrine transporters in humans in vivo. Biol. Psychiatry 68 (9), 854–860.
Duong, S., Chung, K., Wigal, S.B., 2012. Metabolic, toxicological, and safety considera- Hariri, A.R., Mattay, V.S., Tessitore, A., Fera, F., Smith, W.G., Weinberger, D.R., 2002.
tions for drugs used to treat ADHD. Expert Opin. Drug Metab. Toxicol. 8 (5), Dextroamphetamine modulates the response of the human amygdala.
543–552. Neuropsychopharmacology 27 (6), 1036–1040.
Duttaroy, A., Billings, B., Candido, J., Yoburn, B.C., 1992. Chronic d-amphetamine in- Hart, H., Radua, J., Mataix-Cols, D., Rubia, K., 2012. Meta-analysis of fMRI studies of
hibits opioid receptor antagonist-induced supersensitivity. Eur. J. Pharmacol. 221 (2- timing in attention-deficit hyperactivity disorder (ADHD). Neurosci. Biobehav. Rev.
3), 211–215. 36 (10), 2248–2256.
Easton, N., Marshall, F., Fone, K.C., Marsden, C.A., 2007a. Differential effects of the D- Hartvig, P., Torstenson, R., Tedroff, J., Watanabe, Y., Fasth, K.J., Bjurling, P., et al., 1997.
and L-isomers of amphetamine on pharmacological MRI BOLD contrast in the rat. Amphetamine effects on dopamine release and synthesis rate studied in the rhesus
Psychopharmacology (Berl.) 193 (1), 11–30. monkey brain by positron emission tomography. J. Neural Transm. (Vienna) 104
Easton, N., Steward, C., Marshall, F., Fone, K., Marsden, C., 2007b. Effects of ampheta- (4–5), 329–339.
mine isomers, methylphenidate and atomoxetine on synaptosomal and synaptic ve- Heal, D.J., Smith, S.L., Gosden, J., Nutt, D.J., 2013. Amphetamine, past and present–a
sicle accumulation and release of dopamine and noradrenaline in vitro in the rat pharmacological and clinical perspective. J. Psychopharmacol. 27 (6), 479–496.
brain. Neuropharmacology 52 (2), 405–414. Hegerl, U., Hensch, T., 2017. Why do stimulants not work in typical depression? Aust. N.
Elia, J., Glessner, J.T., Wang, K., Takahashi, N., Shtir, C.J., Hadley, D., et al., 2011. Z. J. Psychiatry 51 (1), 20–22.
Genome-wide copy number variation study associates metabotropic glutamate re- Helmeste, D.M., Seeman, P., 1982. Amphetamine-induced hypolocomotion in mice with
ceptor gene networks with attention deficit hyperactivity disorder. Nat. Genet. 44 (1), more brain D2 dopamine receptors. Psychiatry Res. 7 (3), 351–359.
78–84. Hess, J.L., Akutagava-Martins, G.C., Patak, J.D., Glatt, S.J., Faraone, S.V., 2017. Why is
Faraone, S.V., Buitelaar, J., 2010. Comparing the efficacy of stimulants for ADHD in there selective subcortical vulnerability in ADHD? Clues from postmortem brain gene
children and adolescents using meta-analysis. Eur. Child Adolesc. Psychiatry 19 (4), expression data. Mol. Psychiatry. http://dx.doi.org/10.1038/mp.2017.242. (Epub
353–364. ahead of print).
Faraone, S.V., Glatt, S.J., 2010. A comparison of the efficacy of medications for adult Hoogman, M., Bralten, J., Hibar, D.P., Mennes, M., Zwiers, M.P., Schweren, L.S., et al.,
attention-deficit/hyperactivity disorder using meta-analysis of effect sizes. J. Clin. 2017. Subcortical brain volume differences in participants with attention deficit
Psychiatry 71, 754–763. hyperactivity disorder in children and adults: a cross-sectional mega-analysis. Lancet
Faraone, S.V., Biederman, J., Mick, E., 2006. The age-dependent decline of attention Psychiatry 4 (4), 310–319.
deficit hyperactivity disorder: a meta-analysis of follow-up studies. Psychol. Med. 36 Howlett, D.R., Nahorski, S.R., 1979. Acute and chronic amphetamine treatments mod-
(2), 159–165. ulate striatal dopamine receptor binding sites. Brain Res. 161 (1), 173–178.
Faraone, S.V., Asherson, P., Banaschewski, T., Biederman, J., Buitelaar, J.K., Ramos- Inderbitzin, S., Lauber, M.E., Schlumpf, M., Lichtensteiger, W., 1997. Amphetamine-in-
Quiroga, J.A., et al., 2015. Attention-deficit/hyperactivity disorder. Nat. Rev. Dis. duced preprodynorphin mRNA expression and kappa-opioid receptor binding in basal
Primers 1, 15020. ganglia of adult rats after prenatal exposure to diazepam. Brain Res. Dev. Brain Res.
Farb, D.H., Ratner, M.H., 2014. Targeting the modulation of neural circuitry for the 98 (1), 114–124.
treatment of anxiety disorders. Pharmacol. Rev. 66 (4), 1002–1032. Jain, R., Katic, A., 2016. Current and investigational medication delivery systems for
Fayyad, J., De Graaf, R., Kessler, R., Alonso, J., Angermeyer, M., Demyttenaere, K., et al., treating attention-deficit/hyperactivity disorder. Prim. Care Companion CNS Disord.
2007. Cross-national prevalence and correlates of adult attention-deficit hyper- 18 (4). http://dx.doi.org/10.4088/PCC.16r01979.
activity disorder. Br. J. Psychiatry 190 (5), 402–409. Jedema, H.P., Narendran, R., Bradberry, C.W., 2014. Amphetamine-induced release of
Federici, M., Geracitano, R., Bernardi, G., Mercuri, N.B., 2005. Actions of methylpheni- dopamine in primate prefrontal cortex and striatum: striking differences in magni-
date on dopaminergic neurons of the ventral midbrain. Biol. Psychiatry 57 (4), tude and timecourse. J. Neurochem. 130 (4), 490–497.
361–365. Joseph, A., Ayyagari, R., Xie, M., Cai, S., Xie, J., Huss, M., et al., 2017. Comparative
Finnema, S.J., Hughes, Z.A., Haaparanta-Solin, M., Stepanov, V., Nakao, R., Varnas, K., efficacy and safety of attention-deficit/hyperactivity disorder pharmacotherapies,
et al., 2015. Amphetamine decreases alpha2C-adrenoceptor binding of [11C]ORM- including guanfacine extended release: a mixed treatment comparison. Eur. Child
13070: a PET study in the primate brain. Int. J. Neuropsychopharmacol. 18 (3), 1–10. Adolesc. Psychiatry 26 (8), 875–897.
Fischman, A.J., Babich, J.W., Elmaleh, D.R., Barrow, S.A., Meltzer, P., Hanson, R.N., Joseph, N., Zhang-James, Y., Perl, A., Faraone, S.V., 2015. Oxidative stress and ADHD: a
et al., 1997. SPECT imaging of dopamine transporter sites in normal and MPTP- meta-analysis. J. Atten. Disord. 19 (11), 915–924.
treated rhesus monkeys. J. Nucl. Med. 38 (1), 144–150. Joyce, B.M., Glaser, P.E., Gerhardt, G.A., 2007. Adderall produces increased striatal do-
Fleurence, R., Williamson, R., Jing, Y., Kim, E., Tran, Q.V., Pikalov, A.S., et al., 2009. A pamine release and a prolonged time course compared to amphetamine isomers.
systematic review of augmentation strategies for patients with major depressive Psychopharmacology (Berl.) 191 (3), 669–677.
disorder. Psychopharmacol. Bull. 42 (3), 57–90. Kashiwagi, Y., Rokugawa, T., Yamada, T., Obata, A., Watabe, H., Yoshioka, Y., et al.,
Floor, E., Meng, L., 1996. Amphetamine releases dopamine from synaptic vesicles by dual 2015. Pharmacological MRI response to a selective dopamine transporter inhibitor,
mechanisms. Neurosci. Lett. 215 (1), 53–56. GBR12909, in awake and anesthetized rats. Synapse 69 (4), 203–212.
Fusar-Poli, P., Rubia, K., Rossi, G., Sartori, G., Balottin, U., 2012. Striatal dopamine Kasparbauer, A.M., Rujescu, D., Riedel, M., Pogarell, O., Costa, A., Meindl, T., et al., 2015.
transporter alterations in ADHD: pathophysiology or adaptation to psychostimulants? Methylphenidate effects on brain activity as a function of SLC6A3 genotype and
A meta-analysis. Am. J. Psychiatry 169 (3), 264–272. striatal dopamine transporter availability. Neuropsychopharmacology 40 (3),
Gallezot, J.D., Kloczynski, T., Weinzimmer, D., Labaree, D., Zheng, M.Q., Lim, K., et al., 736–745.
2014. Imaging nicotine- and amphetamine-induced dopamine release in rhesus Kegeles, L.S., Zea-Ponce, Y., Abi-Dargham, A., Rodenhiser, J., Wang, T., Weiss, R., et al.,
monkeys with [(11)C]PHNO vs [(11)C]raclopride PET. Neuropsychopharmacology 1999. Stability of [123I]IBZM SPECT measurement of amphetamine-induced striatal
39 (4), 866–874. dopamine release in humans. Synapse 31 (4), 302–308.
Gamo, N.J., Wang, M., Arnsten, A.F., 2010. Methylphenidate and atomoxetine enhance Kessler, R.C., Adler, L., Barkley, R., Biederman, J., Conners, C.K., Demler, O., et al., 2006.
prefrontal function through alpha2-adrenergic and dopamine D1 receptors. J. Am. The prevalence and correlates of adult ADHD in the United States: results from the
Acad. Child Adolesc. Psychiatry 49 (10), 1011–1023. National Comorbidity survey replication. Am. J. Psychiatry 163 (4), 716–723.
Garnier, L.M., Arria, A.M., Caldeira, K.M., Vincent, K.B., O’Grady, K.E., Wish, E.D., 2010. Kidwell, K.M., Van Dyk, T.R., Lundahl, A., Nelson, T.D., 2015. Stimulant medications and
Sharing and selling of prescription medications in a college student sample. J. Clin. sleep for youth with ADHD: a meta-analysis. Pediatrics 136 (6), 1144–1153.
Psychiatry 71 (3), 262–269. Kilbourn, M.R., Domino, E.F., 2011. Increased in vivo [11C]raclopride binding to brain
Garrett, D.D., Nagel, I.E., Preuschhof, C., Burzynska, A.Z., Marchner, J., Wiegert, S., et al., dopamine receptors in amphetamine-treated rats. Eur. J. Pharmacol. 654 (3),

267
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

254–257. 42–51.
Knutson, B., Bjork, J.M., Fong, G.W., Hommer, D., Mattay, V.S., Weinberger, D.R., 2004. Markowitz, J.S., DeVane, C.L., Ramamoorthy, S., Zhu, H.J., 2009. The psychostimulant d-
Amphetamine modulates human incentive processing. Neuron 43 (2), 261–269. threo-(R,R)-methylphenidate binds as an agonist to the 5HT(1A) receptor. Pharmazie
Konofal, E., Lecendreux, M., Cortese, S., 2010. Sleep and ADHD. Sleep Med. 11 (7), 64 (2), 123–125.
652–658. Markowitz, J.S., DeVane, C.L., Pestreich, L.K., Patrick, K.S., Muniz, R., 2006. A compre-
Konradi, C., Leveque, J.C., Hyman, S.E., 1996. Amphetamine and dopamine-induced hensive in vitro screening of d-, l-, and dl-threo-methylphenidate: an exploratory
immediate early gene expression in striatal neurons depends on postsynaptic NMDA study. J. Child Adolesc. Psychopharmacol. 16 (6), 687–698.
receptors and calcium. J. Neurosci. 16 (13), 4231–4239. Martinez, D., Slifstein, M., Broft, A., Mawlawi, O., Hwang, D.R., Huang, Y., et al., 2003.
Konstenius, M., Jayaram-Lindstrom, N., Beck, O., Franck, J., 2010. Sustained release Imaging human mesolimbic dopamine transmission with positron emission tomo-
methylphenidate for the treatment of ADHD in amphetamine abusers: a pilot study. graphy. Part II: amphetamine-induced dopamine release in the functional subdivi-
Drug Alcohol. Depend. 108 (1-2), 130–133. sions of the striatum. J. Cereb. Blood Flow. Metab. 23 (3), 285–300.
Koob, G.F., 2006. The neurobiology of addiction: a neuroadaptational view relevant for Martinez-Raga, J., Ferreros, A., Knecht, C., de Alvaro, R., Carabal, E., 2017. Attention-
diagnosis. Addiction 101 (Suppl. 1), 23–30. deficit hyperactivity disorder medication use: factors involved in prescribing, safety
Kooij, J.J., Huss, M., Asherson, P., Akehurst, R., Beusterien, K., French, A., et al., 2012. aspects and outcomes. Ther. Adv. Drug Saf. 8 (3), 87–99.
Distinguishing comorbidity and successful management of adult ADHD. J. Atten. May, L.J., Kuhr, W.G., Wightman, R.M., 1988. Differentiation of dopamine overflow and
Disord. 16 (5 Suppl), 3S–19S. uptake processes in the extracellular fluid of the rat caudate nucleus with fast-scan in
Kooij, S.J., Bejerot, S., Blackwell, A., Caci, H., Casas-Brugue, M., Carpentier, P.J., et al., vivo voltammetry. J. Neurochem. 51 (4), 1060–1069.
2010. European consensus statement on diagnosis and treatment of adult ADHD: the McIntyre, R.S., Lee, Y., Zhou, A.J., Rosenblat, J.D., Peters, E.M., Lam, R.W., et al., 2017.
European Network Adult ADHD. BMC Psychiatry 10, 67. The efficacy of psychostimulants in major depressive episodes: a systematic review
Krause, K.H., Dresel, S.H., Krause, J., Kung, H.F., Tatsch, K., 2000. Increased striatal and meta-analysis. J. Clin. Psychopharmacol. 37 (4), 412–418.
dopamine transporter in adult patients with attention deficit hyperactivity disorder: Michaelides, M., Pascau, J., Gispert, J.D., Delis, F., Grandy, D.K., Wang, G.J., et al., 2010.
effects of methylphenidate as measured by single photon emission computed tomo- Dopamine D4 receptors modulate brain metabolic activity in the prefrontal cortex
graphy. Neurosci. Lett. 285 (2), 107–110. and cerebellum at rest and in response to methylphenidate. Eur. J. Neurosci. 32 (4),
Krause, J., la Fougere, C., Krause, K.H., Ackenheil, M., Dresel, S.H., 2005. Influence of 668–676.
striatal dopamine transporter availability on the response to methylphenidate in Mick, I., Myers, J., Stokes, P.R., Erritzoe, D., Colasanti, A., Bowden-Jones, H., et al., 2014.
adult patients with ADHD. Eur. Arch. Psychiatry Clin. Neurosci. 255 (6), 428–431. Amphetamine induced endogenous opioid release in the human brain detected with
Kuczenski, R., Segal, D.S., 1997. Effects of methylphenidate on extracellular dopamine, [11C]carfentanil PET: replication in an independent cohort. Int. J.
serotonin, and norepinephrine: comparison with amphetamine. J. Neurochem. 68 Neuropsychopharmacol. 17 (12), 2069–2074.
(5), 2032–2037. Miller, H.H., Shore, P.A., Clarke, D.E., 1980. In vivo monoamine oxidase inhibition by d-
la Fougere, C., Krause, J., Krause, K.H., Josef Gildehaus, F., Hacker, M., Koch, W., et al., amphetamine. Biochem. Pharmacol. 29 (10), 1347–1354.
2006. Value of 99mTc-TRODAT-1 SPECT to predict clinical response to methylphe- Moeller, S.J., Honorio, J., Tomasi, D., Parvaz, M.A., Woicik, P.A., Volkow, N.D., et al.,
nidate treatment in adults with attention deficit hyperactivity disorder. Nucl. Med. 2014. Methylphenidate enhances executive function and optimizes prefrontal func-
Commun. 27 (9), 733–737. tion in both health and cocaine addiction. Cereb. Cortex 24 (3), 643–653.
Lahey, B.B., Applegate, B., McBurnett, K., Biederman, J., Greenhill, L., Hynd, G.W., et al., Montgomery, A.J., Asselin, M.C., Farde, L., Grasby, P.M., 2007. Measurement of me-
1994. DSM-IV field trials for attention deficit hyperactivity disorder in children and thylphenidate-induced change in extrastriatal dopamine concentration using [11C]
adolescents. Am. J. Psychiatry 151 (11), 1673–1685. FLB 457 PET. J. Cereb. Blood Flow. Metab. 27 (2), 369–377.
Lahti, R.A., Tamminga, C.A., 1999. Effect of amphetamine, alpha-methyl-p-tyrosine Motaghinejad, M., Motevalian, M., Shabab, B., Fatima, S., 2017. Effects of acute doses of
(alpha-MPT) and antipsychotic agents on dopamine D2-type receptor occupancy in methylphenidate on inflammation and oxidative stress in isolated hippocampus and
rats. Prog. Neuropsychopharmacol. Biol Psychiatry 23 (7), 1277–1283. cerebral cortex of adult rats. J. Neural Transm. (Vienna) 124 (1), 121–131.
Landau, A.M., Doudet, D.J., Jakobsen, S., 2012. Amphetamine challenge decreases yo- Motaghinejad, M., Motevalian, M., Shabab, B., 2016. Effects of chronic treatment with
himbine binding to alpha2 adrenoceptors in Landrace pig brain. methylphenidate on oxidative stress and inflammation in hippocampus of adult rats.
Psychopharmacology (Berl.) 222 (1), 155–163. Neurosci. Lett. 619, 106–113.
Laruelle, M., Innis, R.B., 1996. Images in neuroscience. SPECT imaging of synaptic do- Mueller, S., Costa, A., Keeser, D., Pogarell, O., Berman, A., Coates, U., et al., 2014. The
pamine. Am. J. Psychiatry 153 (10), 1249. effects of methylphenidate on whole brain intrinsic functional connectivity. Hum.
Laruelle, M., Iyer, R.N., al-Tikriti, M.S., Zea-Ponce, Y., Malison, R., Zoghbi, S.S., et al., Brain Mapp. 35 (11), 5379–5388.
1997. Microdialysis and SPECT measurements of amphetamine-induced dopamine Mukherjee, J., Yang, Z.Y., Lew, R., Brown, T., Kronmal, S., Cooper, M.D., et al., 1997.
release in nonhuman primates. Synapse 25 (1), 1–14. Evaluation of d-amphetamine effects on the binding of dopamine D-2 receptor
Laruelle, M., Abi-Dargham, A., van Dyck, C.H., Rosenblatt, W., Zea-Ponce, Y., Zoghbi, radioligand, 18F-fallypride in nonhuman primates using positron emission tomo-
S.S., et al., 1995. SPECT imaging of striatal dopamine release after amphetamine graphy. Synapse 27 (1), 1–13.
challenge. J. Nucl. Med. 36 (7), 1182–1190. Nakao, T., Radua, J., Rubia, K., Mataix-Cols, D., 2011. Gray matter volume abnormalities
Le Masurier, M., Houston, G., Cowen, P., Grasby, P., Sharp, T., Hume, S., 2004. Tyrosine- in ADHD: voxel-based meta-analysis exploring the effects of age and stimulant
free amino acid mixture attenuates amphetamine-induced displacement of [11C]ra- medication. Am. J. Psychiatry 168 (11), 1154–1163.
clopride in striatum in vivo: a rat PET study. Synapse 51 (2), 151–157. Narendran, R., Frankle, W.G., Mason, N.S., Rabiner, E.A., Gunn, R.N., Searle, G.E., et al.,
Lecendreux, M., Cortese, S., 2007. Sleep problems associated with ADHD: a review of 2009. Positron emission tomography imaging of amphetamine-induced dopamine
current therapeutic options and recommendations for the future. Expert Rev. release in the human cortex: a comparative evaluation of the high affinity dopamine
Neurother. 7 (12), 1799–1806. D2/3 radiotracers [11C]FLB 457 and [11C]fallypride. Synapse 63 (6), 447–461.
Levin, F.R., Mariani, J.J., Specker, S., Mooney, M., Mahony, A., Brooks, D.J., et al., 2015. Narendran, R., Mason, N.S., Laymon, C.M., Lopresti, B.J., Velasquez, N.D., May, M.A.,
Extended-release mixed amphetamine salts vs placebo for comorbid adult attention- et al., 2010. A comparative evaluation of the dopamine D(2/3) agonist radiotracer
deficit/hyperactivity disorder and cocaine use disorder: a randomized clinical trial. [11C](-)-N-propyl-norapomorphine and antagonist [11C]raclopride to measure am-
JAMA Psychiatry 72 (6), 593–602. phetamine-induced dopamine release in the human striatum. J. Pharmacol. Exp.
Leyton, M., Boileau, I., Benkelfat, C., Diksic, M., Baker, G., Dagher, A., 2002. Ther. 333 (2), 533–539.
Amphetamine-induced increases in extracellular dopamine, drug wanting, and no- Nelson, A., Galon, P., 2012. Exploring the relationship among ADHD, stimulants, and
velty seeking: a PET/[11C]raclopride study in healthy men. substance abuse. J. Child Adolesc. Psychiatr. Nurs. 25 (3), 113–118.
Neuropsychopharmacology 27 (6), 1027–1035. Nikolaus, S., Wirrwar, A., Antke, C., Arkian, S., Schramm, N., Muller, H.W., et al., 2005.
Leyton, M., Dagher, A., Boileau, I., Casey, K., Baker, G.B., Diksic, M., et al., 2004. Quantitation of dopamine transporter blockade by methylphenidate: first in vivo
Decreasing amphetamine-induced dopamine release by acute phenylalanine/tyrosine investigation using [123I]FP-CIT and a dedicated small animal SPECT. Eur. J. Nucl.
depletion: a PET/[11C]raclopride study in healthy men. Neuropsychopharmacology Med. Mol. Imaging 32 (3), 308–313.
29 (2), 427–432. Nikolaus, S., Antke, C., Beu, M., Kley, K., Larisch, R., Wirrwar, A., et al., 2007. In-vivo
Lind, N.M., Olsen, A.K., Moustgaard, A., Jensen, S.B., Jakobsen, S., Hansen, A.K., et al., quantification of dose-dependent dopamine transporter blockade in the rat striatum
2005. Mapping the amphetamine-evoked dopamine release in the brain of the with small animal SPECT. Nucl. Med. Commun. 28 (3), 207–213.
Gottingen minipig. Brain Res. Bull. 65 (1), 1–9. Nikolaus, S., Antke, C., Beu, M., Kley, K., Wirrwar, A., Huston, J.P., et al., 2011. Binding
Liu, P.S., Liaw, C.T., Lin, M.K., Shin, S.H., Kao, L.S., Lin, L.F., 2003. Amphetamine en- of [123I]iodobenzamide to the rat D2 receptor after challenge with various doses of
hances Ca2+ entry and catecholamine release via nicotinic receptor activation in methylphenidate: an in vivo imaging study with dedicated small animal SPECT. Eur.
bovine adrenal chromaffin cells. Eur. J. Pharmacol. 460 (1), 9–17. J. Nucl. Med. Mol. Imaging 38 (4), 694–701.
Ludolph, A.G., Kassubek, J., Schmeck, K., et al., 2008. Dopaminergic dysfunction in at- Oswald, L.M., Wong, D.F., McCaul, M., Zhou, Y., Kuwabara, H., Choi, L., et al., 2005.
tention deficit hyperactivity disorder (ADHD), differences between pharmacologi- Relationships among ventral striatal dopamine release, cortisol secretion, and sub-
cally treated and never treated young adults: a 3,4-dihdroxy-6-[18F]fluorophenyl-l- jective responses to amphetamine. Neuropsychopharmacology 30 (4), 821–832.
alanine PET study. Neuroimage 41, 718–727. Oswald, L.M., Wand, G.S., Wong, D.F., Brown, C.H., Kuwabara, H., Brasic, J.R., 2015.
Malhi, G.S., Byrow, Y., Bassett, D., Boyce, P., Hopwood, M., Lyndon, W., et al., 2016. Risky decision-making and ventral striatal dopamine responses to amphetamine: a
Stimulants for depression: on the up and up? Aust. N. Z. J. Psychiatry 50 (3), positron emission tomography [(11)C]raclopride study in healthy adults.
203–207. Neuroimage 113, 26–36.
Maltezos, S., Horder, J., Coghlan, S., Skirrow, C., O’Gorman, R., Lavender, T.J., et al., Owens, J., Weiss, M., Nordbrock, E., Mattingly, G., Wigal, S., Greenhill, L.L., et al., 2016.
2014. Glutamate/glutamine and neuronal integrity in adults with ADHD: a proton Effect of Aptensio XR (methylphenidate HCl extended-release) capsules on sleep in
MRS study. Transl. Psychiatry 4, e373. children with attention-deficit/hyperactivity disorder. J. Child Adolesc.
Mao, A.R., Findling, R.L., 2014. Comorbidities in adult attention-deficit/hyperactivity Psychopharmacol. 26 (10), 873–881.
disorder: a practical guide to diagnosis in primary care. Postgrad. Med. 126 (5), Panagiotou, O.A., Contopoulos-Ioannidis, D.G., Papanikolaou, P.N., Ntzani, E.E.,

268
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

Ioannidis, J.P., 2011. Different black box warning labeling for same-class drugs. J. Saelens, J.K., Simke, J.P., Neale, S.E., Weeks, B.J., Selwyn, M., 1980. Effects of halo-
Gen. Intern. Med. 26 (6), 603–610. peridol and d-amphetamine on in vivo 3H-spiroperiodol binding in the rat forebrain.
Park, S., Lee, J.M., Kim, J.W., Cho, D.Y., Yun, H.J., Han, D.H., et al., 2015. Associations Arch. Int. Pharmacodyn. Ther. 246 (1), 98–107.
between serotonin transporter gene (SLC6A4) methylation and clinical characteristics Sanchez-Perez, A.M., Garcia-Aviles, A., Albert Gasco, H., Sanjuan, J., Olucha-Bordonau,
and cortical thickness in children with ADHD. Psychol. Med. 45 (14), 3009–3017. F.E., 2012. [Effects of methylphenidate on anxiety]. Rev. Neurol. 55 (8), 499–506.
Patrick, R.L., Berkowitz, A.L., Regenstein, A.C., 1981. Effects of in vivo amphetamine Sandoval, V., Riddle, E.L., Hanson, G.R., Fleckenstein, A.E., 2002. Methylphenidate re-
administration on dopamine synthesis regulation in rat brain striatal synaptosomes. distributes vesicular monoamine transporter-2: role of dopamine receptors. J.
J. Pharmacol. Exp. Ther. 217 (3), 686–691. Neurosci. 22 (19), 8705–8710.
Pedersen, K., Simonsen, M., Ostergaard, S.D., Munk, O.L., Rosa-Neto, P., Olsen, A.K., Santosh, P.J., Sattar, S., Canagaratnam, M., 2011. Efficacy and tolerability of pharma-
et al., 2007. Mapping the amphetamine-evoked changes in [11C]raclopride binding cotherapies for attention-deficit hyperactivity disorder in adults. CNS Drugs 25 (9),
in living rat using small animal PET: modulation by MAO-inhibition. Neuroimage 35 737–763.
(1), 38–46. Schabram, I., Henkel, K., Mohammadkhani Shali, S., Dietrich, C., Schmaljohann, J., Winz,
Pliszka, S., AACAP Work Group on Quality Issues, 2007. Practice parameter for the as- O., et al., 2014. Acute and sustained effects of methylphenidate on cognition and
sessment and treatment of children and adolescents with attention-deficit/hyper- presynaptic dopamine metabolism: an [18F]FDOPA PET study. J. Neurosci. 34 (44),
activity disorder. J. Am. Acad. Child Adolesc. Psychiatry 46 (7), 894–921. 14769–14776.
Polanczyk, G., de Lima, M.S., Horta, B.L., Biederman, J., Rohde, L.A., 2007. The world- Schiffer, W.K., Volkow, N.D., Fowler, J.S., Alexoff, D.L., Logan, J., Dewey, S.L., 2006.
wide prevalence of ADHD: a systematic review and metaregression analysis. Am. J. Therapeutic doses of amphetamine or methylphenidate differentially increase sy-
Psychiatry 164 (6), 942–948. naptic and extracellular dopamine. Synapse 59 (4), 243–251.
Potter, A.S., Schaubhut, G., Shipman, M., 2014. Targeting the nicotinic cholinergic system Schlosser, R.G., Nenadic, I., Wagner, G., Zysset, S., Koch, K., Sauer, H., 2009.
to treat attention-deficit/hyperactivity disorder: rationale and progress to date. CNS Dopaminergic modulation of brain systems subserving decision making under un-
Drugs 28 (12), 1103–1113. certainty: a study with fMRI and methylphenidate challenge. Synapse 63 (5),
Preece, M.A., Sibson, N.R., Raley, J.M., Blamire, A., Styles, P., Sharp, T., 2007. Region- 429–442.
specific effects of a tyrosine-free amino acid mixture on amphetamine-induced Schouw, M.L., Kaag, A.M., Caan, M.W., Heijtel, D.F., Majoie, C.B., Nederveen, A.J., et al.,
changes in BOLD fMRI signal in the rat brain. Synapse 61 (11), 925–932. 2013. Mapping the hemodynamic response in human subjects to a dopaminergic
Price, J.C., Drevets, W.C., Ruszkiewicz, J., Greer, P.J., Villemagne, V.L., Xu, L., et al., challenge with dextroamphetamine using ASL-based pharmacological MRI.
2002. Sequential H(2)(15)O PET studies in baboons: before and after amphetamine. Neuroimage 72, 1–9.
J. Nucl. Med. 43 (8), 1090–1100. Schwartz, M.D., Kilduff, T.S., 2015. The neurobiology of sleep and wakefulness. Psychiatr.
Pum, M., Carey, R.J., Huston, J.P., Muller, C.P., 2007. Dissociating effects of cocaine and Clin. N. Am. 38 (4), 615–644.
d-amphetamine on dopamine and serotonin in the perirhinal, entorhinal, and pre- Schwarz, A.J., Gozzi, A., Reese, T., Bifone, A., 2007a. Functional connectivity in the
frontal cortex of freely moving rats. Psychopharmacology (Berl.) 193 (3), 375–390. pharmacologically activated brain: resolving networks of correlated responses to D-
Purper-Ouakil, D., Ramoz, N., Lepagnol-Bestel, A.M., Gorwood, P., Simonneau, M., 2011. amphetamine. Magn. Reson. Med. 57 (4), 704–713.
Neurobiology of attention deficit/hyperactivity disorder. Pediatr. Res. 69 (5 Pt 2), Schwarz, A.J., Gozzi, A., Reese, T., Heidbreder, C.A., Bifone, A., 2007b. Pharmacological
69R–76R. modulation of functional connectivity: the correlation structure underlying the
Quelch, D.R., Katsouri, L., Nutt, D.J., Parker, C.A., Tyacke, R.J., 2014. Imaging en- phMRI response to D-amphetamine modified by selective dopamine D3 receptor an-
dogenous opioid peptide release with [11C]carfentanil and [3H]diprenorphine: in- tagonist SB277011A. Magn. Reson. Imaging 25 (6), 811–820.
fluence of agonist-induced internalization. J. Cereb. Blood Flow. Metab. 34 (10), Schweitzer, J.B., Lee, D.O., Hanford, R.B., Zink, C.F., Ely, T.D., Tagamets, M.A., et al.,
1604–1612. 2004. Effect of methylphenidate on executive functioning in adults with attention-
Rabiner, D.L., Anastopoulos, A.D., Costello, E.J., Hoyle, R.H., McCabe, S.E., deficit/hyperactivity disorder: normalization of behavior but not related brain ac-
Swartzwelder, H.S., 2009. The misuse and diversion of prescribed ADHD medications tivity. Biol. Psychiatry 56 (8), 597–606.
by college students. J. Atten. Disord. 13 (2), 144–153. Schwendt, M., Gold, S.J., McGinty, J.F., 2006. Acute amphetamine down-regulates RGS4
Ramaekers, J.G., Evers, E.A., Theunissen, E.L., Kuypers, K.P., Goulas, A., Stiers, P., 2013. mRNA and protein expression in rat forebrain: distinct roles of D1 and D2 dopamine
Methylphenidate reduces functional connectivity of nucleus accumbens in brain re- receptors. J. Neurochem. 96 (6), 1606–1615.
ward circuit. Psychopharmacology (Berl.) 229 (2), 219–226. Seidman, L.J., Biederman, J., Liang, L., Valera, E.M., Monuteaux, M.C., Brown, A., et al.,
Ramasubbu, R., Singh, H., Zhu, H., Dunn, J.F., 2012. Methylphenidate-mediated reduc- 2011. Gray matter alterations in adults with attention-deficit/hyperactivity disorder
tion in prefrontal hemodynamic responses to working memory task: a functional identified by voxel based morphometry. Biol. Psychiatry 69 (9), 857–866.
near-infrared spectroscopy study. Hum. Psychopharmacol. 27 (6), 615–621. Seneca, N., Finnema, S.J., Farde, L., Gulyas, B., Wikstrom, H.V., Halldin, C., et al., 2006.
Ren, J., Xu, H., Choi, J.K., Jenkins, B.G., Chen, Y.I., 2009. Dopaminergic response to Effect of amphetamine on dopamine D2 receptor binding in nonhuman primate brain:
graded dopamine concentration elicited by four amphetamine doses. Synapse 63 (9), a comparison of the agonist radioligand [11C]MNPA and antagonist [11C]raclopride.
764–772. Synapse 59 (5), 260–269.
Riccardi, P., Park, S., Anderson, S., Doop, M., Ansari, M.S., Schmidt, D., et al., 2011. Sex Shaffer, C., Guo, M.L., Fibuch, E.E., Mao, L.M., Wang, J.Q., 2010. Regulation of group I
differences in the relationship of regional dopamine release to affect and cognitive metabotropic glutamate receptor expression in the rat striatum and prefrontal cortex
function in striatal and extrastriatal regions using positron emission tomography and in response to amphetamine in vivo. Brain Res. 1326, 184–192.
[18F]fallypride. Synapse 65 (2), 99–102. Sharp, S.I., McQuillin, A., Marks, M., Hunt, S.P., Stanford, S.C., Lydall, G.J., et al., 2014.
Riccardi, P., Zald, D., Li, R., Park, S., Ansari, M.S., Dawant, B., et al., 2006a. Sex differ- Genetic association of the tachykinin receptor 1 TACR1 gene in bipolar disorder,
ences in amphetamine-induced displacement of [18F]fallypride in striatal and ex- attention deficit hyperactivity disorder, and the alcohol dependence syndrome. Am.
trastriatal regions: a PET study. Am. J. Psychiatry 163 (9), 1639–1641. J. Med. Genet. B Neuropsychiatr. Genet. 165B (4), 373–380.
Riccardi, P., Li, R., Ansari, M.S., Zald, D., Park, S., Dawant, B., et al., 2006b. Shin, J.Y., Roughead, E.E., Park, B.J., Pratt, N.L., 2016. Cardiovascular safety of me-
Amphetamine-induced displacement of [18F] fallypride in striatum and extrastriatal thylphenidate among children and young people with attention-deficit/hyperactivity
regions in humans. Neuropsychopharmacology 31 (5), 1016–1026. disorder (ADHD): nationwide self controlled case series study. BMJ 353, i2550.
Richards, C., McIntyre, R.S., Weisler, R., Sambunaris, A., Brawman-Mintzer, O., Gao, J., Shotbolt, P., Tziortzi, A.C., Searle, G.E., Colasanti, A., van der Aart, J., Abanades, S., et al.,
et al., 2016. Lisdexamfetamine dimesylate augmentation for adults with major de- 2012. Within-subject comparison of [11C]-(+)-PHNO and [11C]raclopride sensi-
pressive disorder and inadequate response to antidepressant monotherapy: results tivity to acute amphetamine challenge in healthy humans. J. Cereb. Blood Flow.
from 2 phase 3, multicenter, randomized, double-blind, placebo-controlled studies. J. Metab. 32 (1), 127–136.
Affect. Disord. 206, 151–160. Sigurdardottir, H.L., Kranz, G.S., Rami-Mark, C., James, G.M., Vanicek, T., Gryglewski,
Richards, C., Iosifescu, D.V., Mago, R., Sarkis, E., Reynolds, J., Geibel, B., et al., 2017. A G., et al., 2016. Effects of norepinephrine transporter gene variants on NET binding in
randomized, double-blind, placebo-controlled, dose-ranging study of lisdex- ADHD and healthy controls investigated by PET. Hum. Brain Mapp. 37 (3), 884–895.
amfetamine dimesylate augmentation for major depressive disorder in adults with Silverstone, P.H., O’Donnell, T., Ulrich, M., Asghar, S., Hanstock, C.C., 2002. Dextro-
inadequate response to antidepressant therapy. J. Clin. Psychopharmacol. 31 (9), amphetamine increases phosphoinositol cycle activity in volunteers: an MRS study.
1190–1203. Hum. Psychopharmacol. 17 (8), 425–429.
Riddle, E.L., Hanson, G.R., Fleckenstein, A.E., 2007. Therapeutic doses of amphetamine Simon, V., Czobor, P., Balint, S., Meszaros, A., Bitter, I., 2009. Prevalence and correlates
and methylphenidate selectively redistribute the vesicular monoamine transporter-2. of adult attention-deficit hyperactivity disorder: meta-analysis. Br. J. Psychiatry 194
Eur. J. Pharmacol. 571 (1), 25–28. (3), 204–211.
Ritz, M.C., Kuhar, M.J., 1989. Relationship between self-administration of amphetamine Skinbjerg, M., Liow, J.S., Seneca, N., Hong, J., Lu, S., Thorsell, A., et al., 2010. D2 do-
and monoamine receptors in brain: comparison with cocaine. J. Pharmacol. Exp. pamine receptor internalization prolongs the decrease of radioligand binding after
Ther. 248 (3), 1010–1017. amphetamine: a PET study in a receptor internalization-deficient mouse model.
Robin, A.A., Wiseberg, S., 1958. A controlled trial of methyl phenidate (ritalin) in the Neuroimage 50 (4), 1402–1407.
treatment of depressive states. J. Neurol. Neurosurg. Psychiatry 21 (1), 55–57. Slifstein, M., Kegeles, L.S., Xu, X., Thompson, J.L., Urban, N., Castrillon, J., et al., 2010.
Robinson, J.B., 1985. Stereoselectivity and isoenzyme selectivity of monoamine oxidase Striatal and extrastriatal dopamine release measured with PET and [18F] fallypride.
inhibitors. Enantiomers of amphetamine, N-methylamphetamine and deprenyl. Synapse 64 (5), 350–362.
Biochem. Pharmacol. 34 (23), 4105–4108. Smith, D.G., Tzavara, E.T., Shaw, J., Luecke, S., Wade, M., Davis, R., et al., 2005.
Rose, S.E., Janke, A.L., Strudwick, M.W., McMahon, K.L., Chalk, J.B., Snyder, P., et al., Mesolimbic dopamine super-sensitivity in melanin-concentrating hormone-1 re-
2006. Assessment of dynamic susceptibility contrast cerebral blood flow response to ceptor-deficient mice. J. Neurosci. 25 (4), 914–922.
amphetamine challenge: a human pharmacological magnetic resonance imaging Snitselaar, M.A., Smits, M.G., van der Heijden, K.B., Spijker, J., 2017. Sleep and circadian
study at 1.5 and 4 T. Magn. Reson. Med. 55 (1), 9–15. rhythmicity in adult ADHD and the effect of stimulants. J. Atten. Disord. 21 (1),
Rostain, A.L., 2008. Attention-deficit/hyperactivity disorder in adults: evidence-based 14–26.
recommendations for management. Postgrad. Med. 120 (3), 27–38. Somkuwar, S.S., Darna, M., Kantak, K.M., Dwoskin, L.P., 2013. Adolescence

269
S.V. Faraone Neuroscience and Biobehavioral Reviews 87 (2018) 255–270

methylphenidate treatment in a rodent model of attention deficit/hyperactivity dis- Volkow, N.D., Wang, G.J., Fowler, J.S., Gatley, S.J., Logan, J., Ding, Y.S., et al., 1998.
order: dopamine transporter function and cellular distribution in adulthood. Dopamine transporter occupancies in the human brain induced by therapeutic doses
Biochem. Pharmacol. 86 (2), 309–316. of oral methylphenidate. Am. J. Psychiatry 155 (10), 1325–1331.
Spencer, T.J., Biederman, J., Ciccone, P.E., Madras, B.K., Dougherty, D.D., Bonab, A.A., Volkow, N.D., Wang, G., Fowler, J.S., Logan, J., Gerasimov, M., Maynard, L., et al., 2001.
et al., 2006. PET study examining pharmacokinetics, detection and likeability, and Therapeutic doses of oral methylphenidate significantly increase extracellular do-
dopamine transporter receptor occupancy of short- and long-acting oral methylphe- pamine in the human brain. J. Neurosci. 21 (2), Rc121.
nidate. Am. J. Psychiatry 163 (3), 387–395. Volkow, N.D., Wang, G.J., Fowler, J.S., Telang, F., Maynard, L., Logan, J., et al., 2004.
Spencer, T.J., Bonab, A.A., Dougherty, D.D., Martin, J., McDonnell, T., Fischman, A.J., Evidence that methylphenidate enhances the saliency of a mathematical task by in-
2010. A PET study examining pharmacokinetics and dopamine transporter occu- creasing dopamine in the human brain. Am. J. Psychiatry 161 (7), 1173–1180.
pancy of two long-acting formulations of methylphenidate in adults. Int. J. Mol. Med. Volkow, N.D., Wang, G.J., Tomasi, D., Kollins, S.H., Wigal, T.L., Newcorn, J.H., et al.,
25 (2), 261–265. 2012. Methylphenidate-elicited dopamine increases in ventral striatum are associated
Spencer, T.J., Biederman, J., Madras, B.K., Dougherty, D.D., Bonab, A.A., Livni, E., et al., with long-term symptom improvement in adults with attention deficit hyperactivity
2007. Further evidence of dopamine transporter dysregulation in ADHD: a controlled disorder. J. Neurosci. 32 (3), 841–849.
PET imaging study using altropane. Biol. Psychiatry 62 (9), 1059–1061. Volkow, N.D., Wang, G.J., Newcorn, J.H., Kollins, S.H., Wigal, T.L., Telang, F., et al.,
Sternat, T., Katzman, M.A., 2016. Neurobiology of hedonic tone: the relationship between 2011. Motivation deficit in ADHD is associated with dysfunction of the dopamine
treatment-resistant depression, attention-deficit hyperactivity disorder, and sub- reward pathway. Mol. Psychiatry 16 (11), 1147–1154.
stance abuse. Neuropsychiatr. Dis. Treat. 12, 2149–2164. Volkow, N.D., Wang, G.J., Fowler, J.S., Gatley, S.J., Ding, Y.S., Logan, J., et al., 1996a.
Stotz, G., Woggon, B., Angst, J., 1999. Psychostimulants in the therapy of treatment- Relationship between psychostimulant-induced "high" and dopamine transporter
resistant depression: review of the literature and findings from a retrospective study occupancy. Proc. Natl. Acad. Sci. U. S. A. 93 (19), 10388–10392.
in 65 depressed patients. Dialogues Clin. Neurosci. 1 (3), 165–174. Volkow, N.D., Fowler, J.S., Gatley, S.J., Logan, J., Wang, G.J., Ding, Y.S., et al., 1996b.
Stuhec, M., Munda, B., Svab, V., Locatelli, I., 2015. Comparative efficacy and accept- PET evaluation of the dopamine system of the human brain. J. Nucl. Med. 37 (7),
ability of atomoxetine, lisdexamfetamine, bupropion and methylphenidate in treat- 1242–1256.
ment of attention deficit hyperactivity disorder in children and adolescents: a meta- Volkow, N.D., Fowler, J.S., Gatley, S.J., Dewey, S.L., Wang, G.J., Logan, J., et al., 1999a.
analysis with focus on bupropion. J. Affect. Disord. 178, 149–159. Comparable changes in synaptic dopamine induced by methylphenidate and by co-
Subcommittee on Attention-Deficit/Hyperactivity Disorder, Steering Committee on caine in the baboon brain. Synapse 31 (1), 59–66.
Quality Improvement and Management, Wolraich, M., Brown, L., Brown, R.T., Volkow, N.D., Wang, G.J., Fowler, J.S., Gatley, S.J., Logan, J., Ding, Y.S., et al., 1999b.
DuPaul, G., et al., 2011. ADHD: clinical practice guideline for the diagnosis, eva- Blockade of striatal dopamine transporters by intravenous methylphenidate is not
luation, and treatment of attention-deficit/hyperactivity disorder in children and sufficient to induce self-reports of “high”. J. Pharmacol. Exp. Ther. 288 (1), 14–20.
adolescents. Pediatrics 128 (5), 1007–1022. Volkow, N.D., Wang, G.J., Fowler, J.S., Logan, J., Franceschi, D., Maynard, L., et al.,
Sulzer, D., Rayport, S., 1990. Amphetamine and other psychostimulants reduce pH gra- 2002a. Relationship between blockade of dopamine transporters by oral methyl-
dients in midbrain dopaminergic neurons and chromaffin granules: a mechanism of phenidate and the increases in extracellular dopamine: therapeutic implications.
action. Neuron 5 (6), 797–808. Synapse 43 (3), 181–187.
Sulzer, D., Chen, T.K., Lau, Y.Y., Kristensen, H., Rayport, S., Ewing, A., 1995. Volkow, N.D., Fowler, J.S., Wang, G., Ding, Y., Gatley, S.J., 2002b. Mechanism of action
Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and pro- of methylphenidate: insights from PET imaging studies. J. Atten. Disord. 6 (Suppl. 1),
motes reverse transport. J. Neurosci. 15 (5 Pt 2), 4102–4108. S31–S43.
Sun, W., Ginovart, N., Ko, F., Seeman, P., Kapur, S., 2003. In vivo evidence for dopamine- Vollenweider, F.X., Maguire, R.P., Leenders, K.L., Mathys, K., Angst, J., 1998. Effects of
mediated internalization of D2-receptors after amphetamine: differential findings high amphetamine dose on mood and cerebral glucose metabolism in normal vo-
with [3H]raclopride versus [3H]spiperone. Mol. Pharmacol. 63 (2), 456–462. lunteers using positron emission tomography (PET). Psychiatry Res. 83 (3), 149–162.
Surman, C.B., Roth, T., 2011. Impact of stimulant pharmacotherapy on sleep quality: post Wall, S.C., Gu, H., Rudnick, G., 1995. Biogenic amine flux mediated by cloned trans-
hoc analyses of 2 large, double-blind, randomized, placebo-controlled trials. J. Clin. porters stably expressed in cultured cell lines: amphetamine specificity for inhibition
Psychiatry 72 (7), 903–908. and efflux. Mol. Pharmacol. 47 (3), 544–550.
Thomas, M., Rostain, A., Prevatt, F., 2013. ADHD diagnosis and treatment in college Wand, G.S., Oswald, L.M., McCaul, M.E., Wong, D.F., Johnson, E., Zhou, Y., et al., 2007.
students and young adults. Adolesc. Med. State Art Rev. 24 (3), 659–679. Association of amphetamine-induced striatal dopamine release and cortisol responses
Thomas, S.J., Shin, M., McInnis, M.G., Bostwick, J.R., 2015. Combination therapy with to psychological stress. Neuropsychopharmacology 32 (11), 2310–2320.
monoamine oxidase inhibitors and other antidepressants or stimulants: strategies for Wang, G.J., Volkow, N.D., Fowler, J.S., Logan, J., Pappas, N.R., Wong, C.T., et al., 1999.
the management of treatment-resistant depression. Pharmacotherapy 35 (4), Reproducibility of repeated measures of endogenous dopamine competition with
433–449. [11C]raclopride in the human brain in response to methylphenidate. J. Nucl. Med. 40
Tomasi, D., Volkow, N.D., Wang, G.J., Wang, R., Telang, F., Caparelli, E.C., et al., 2011. (8), 1285–1291.
Methylphenidate enhances brain activation and deactivation responses to visual at- Wang, G.J., Volkow, N.D., Wigal, T., Kollins, S.H., Newcorn, J.H., Telang, F., et al., 2013.
tention and working memory tasks in healthy controls. Neuroimage 54 (4), Long-term stimulant treatment affects brain dopamine transporter level in patients
3101–3110. with attention deficit hyperactive disorder. PLoS One 8 (5), e63023.
Tomic, M., Joksimovic, J., 2000. Psychotomimetics moderately affect dopamine receptor Weiss, G., Hechtman, L., Milroy, T., Perlman, T., 1985. Psychiatric status of hyperactives
binding in the rat brain. Neurochem. Int. 36 (2), 137–142. as adults: a controlled prospective 15-year follow-up of 63 hyperactive children. J.
Tomic, M., Vukosavic, S., Joksimovic, J., 1997. Acute amphetamine and/or phencyclidine Am. Acad. Child Psychiatry 24 (2), 211–220.
effects on the dopamine receptor specific binding in the rat brain. Eur. Westover, A.N., Halm, E.A., 2012. Do prescription stimulants increase the risk of adverse
Neuropsychopharmacol. 7 (4), 295–301. cardiovascular events? A systematic review. BMC Cardiovasc. Disord. 12, 41.
Treadway, M.T., Zald, D.H., 2011. Reconsidering anhedonia in depression: lessons from Willeit, M., Ginovart, N., Graff, A., Rusjan, P., Vitcu, I., Houle, S., et al., 2008. First human
translational neuroscience. Neurosci. Biobehav. Rev. 35 (3), 537–555. evidence of d-amphetamine induced displacement of a D2/3 agonist radioligand: a
Udo de Haes, J.I., Kortekaas, R., Van Waarde, A., Maguire, R.P., Pruim, J., den Boer, J.A., [11C]-(+)-PHNO positron emission tomography study. Neuropsychopharmacology
2005. Assessment of methylphenidate-induced changes in binding of continuously 33 (2), 279–289.
infused [11C]-raclopride in healthy human subjects: correlation with subjective ef- Winterstein, A.G., Gerhard, T., Shuster, J., Saidi, A., 2009. Cardiac safety of methylphe-
fects. Psychopharmacology (Berl.) 183 (3), 322–330. nidate versus amphetamine salts in the treatment of ADHD. Pediatrics 124 (1),
van Berckel, B.N., Kegeles, L.S., Waterhouse, R., Guo, N., Hwang, D.R., Huang, Y., et al., e75–80.
2006. Modulation of amphetamine-induced dopamine release by group II metabo- Wolkin, A., Angrist, B., Wolf, A., Brodie, J., Wolkin, B., Jaeger, J., et al., 1987. Effects of
tropic glutamate receptor agonist LY354740 in non-human primates studied with amphetamine on local cerebral metabolism in normal and schizophrenic subjects as
positron emission tomography. Neuropsychopharmacology 31 (5), 967–977. determined by positron emission tomography. Psychopharmacology (Berl.) 92 (2),
van Ewijk, H., Heslenfeld, D.J., Zwiers, M.P., Buitelaar, J.K., Oosterlaan, J., 2012. 241–246.
Diffusion tensor imaging in attention deficit/hyperactivity disorder: a systematic Woodward, N.D., Cowan, R.L., Park, S., Ansari, M.S., Baldwin, R.M., Li, R., et al., 2011.
review and meta-analysis. Neurosci. Biobehav. Rev. 36 (4), 1093–1106. Correlation of individual differences in schizotypal personality traits with ampheta-
Vanicek, T., Spies, M., Rami-Mark, C., Savli, M., Hoflich, A., Kranz, G.S., et al., 2014. The mine-induced dopamine release in striatal and extrastriatal brain regions. Am. J.
norepinephrine transporter in attention-deficit/hyperactivity disorder investigated Psychiatry 168 (4), 418–426.
with positron emission tomography. JAMA Psychiatry 71 (12), 1340–1349. Xiao, L., Becker, J.B., 1998. Effects of estrogen agonists on amphetamine-stimulated
Vanicek, T., Kutzelnigg, A., Philippe, C., Sigurdardottir, H.L., James, G.M., Hahn, A., striatal dopamine release. Synapse 29 (4), 379–391.
et al., 2017. Altered interregional molecular associations of the serotonin transporter Yin, H.S., Lai, C.C., Tien, T.W., Han, S.K., Pu, X.L., 2010. Differential changes in cerebellar
in attention deficit/hyperactivity disorder assessed with PET. Hum. Brain Mapp. 38 transmitter content and expression of calcium binding proteins and transcription
(2), 792–802. factors in mouse administered with amphetamine. Neurochem. Int. 57 (3), 288–296.
Vaughan, B., Kratochvil, C.J., 2012. Pharmacotherapy of pediatric attention-deficit/hy- Young, K.A., Liu, Y., Gobrogge, K.L., Dietz, D.M., Wang, H., Kabbaj, M., et al., 2011.
peractivity disorder. Child Adolesc. Psychiatr. Clin. N. Am. 21 (4), 941–955. Amphetamine alters behavior and mesocorticolimbic dopamine receptor expression
Volkow, N.D., Morales, M., 2015. The brain on drugs: from reward to addiction. Cell. 162 in the monogamous female prairie vole. Brain Res. 1367, 213–222.
(4), 712–725. Yu, M.F., Lin, W.W., Li, L.T., Yin, H.S., 2003. Activation of metabotropic glutamate re-
Volkow, N.D., Wang, G.J., Fowler, J.S., Logan, J., Schlyer, D., Hitzemann, R., et al., 1994. ceptor 5 is associated with effect of amphetamine on brain neurons. Synapse 50 (4),
Imaging endogenous dopamine competition with [11C]raclopride in the human 334–344.
brain. Synapse 16 (4), 255–262. Zhu, J., Reith, M.E., 2008. Role of the dopamine transporter in the action of psychosti-
Volkow, N.D., Wang, G.J., Fowler, J.S., Logan, J., Angrist, B., Hitzemann, R., et al., 1997. mulants, nicotine, and other drugs of abuse. CNS Neurol. Disord. Drug Targets 7 (5),
Effects of methylphenidate on regional brain glucose metabolism in humans: re- 393–409.
lationship to dopamine D2 receptors. Am. J. Psychiatry 154 (1), 50–55.

270

You might also like