Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Apr 9.
Published in final edited form as: Prog Mol Biol Transl Sci. 2011;98:1–46. doi: 10.1016/B978-0-12-385506-0.00001-6

Monoamine Transporters: Vulnerable and Vital Doorkeepers

Zhicheng Lin 1,*, Juan J Canales 2, Thröstur Björgvinsson 3, Morgane M Thomsen 1, Hong Qu 4, Qing-Rong Liu 5, Gonzalo E Torres 6, S Barak Caine 1
PMCID: PMC3321928  NIHMSID: NIHMS361090  PMID: 21199769

Abstract

Transporters of dopamine, serotonin and norepinephrine have been empirically used as medication targets for several mental illnesses for the last decades. These protein-targeted medications are effective only for subpopulations of patients with transporter-related brain disorders. Since the cDNA clonings in early 1990’s, molecular studies of these transporters have revealed a wealth of information about the transporters’ structure activity relationship (SAR), neuropharmacology, cell biology, biochemistry, pharmacogenetics and the diseases related to the human genes encoding these transporters among related regulators. Such new information creates a unique opportunity to develop transporter-specific medications based on SAR, mRNA, DNA and perhaps transporter trafficking regulation, for a list of highly relevant diseases including substance abuse, depression, schizophrenia and Parkinson’s disease.

I. Introduction

Monoamine transporters are transmembrane proteins located in plasma membranes of monoaminergic neurons, including the dopamine transporter (DAT), serotonin transporter (SERT, also expressed in platelets), and norepinephrine transporter (NET) (1, 2). These proteins use ion (Na+, Cl) gradients as energy sources to move monoamines into or out of neurons. The major function of these transporters is to terminate monoamine transmission by inward transport of substrates away from the synaptic cleft. In the membrane of intracellular synaptic vesicles is the vesicular monoamine transporters 1 and 2 (VMAT1 and VMAT2), which use a proton gradient as the energy source to sequester cytosolic monoamines into the vesicles and then release the monoamines into synaptic cleft by exocytosis. Therefore, the overall function of these four transporters is to regulate tempo-spatial components of monoamine transmission. Loss of a transporter could cause severe disease or lethality. For instance, two loss-of-function DAT mutants, L368Q and P395L, cause infantile parkinsonism-dystonia in humans (3). Complete deletion of the VMAT2 gene causes developmental defect and embryonic lethality in mice (4-6).

Because of the exclusive expression of each transporter in the corresponding neurons, these transporters are often used as markers of specific neurons. DAT is expressed in dopaminergic neurons that project mainly from VTA and substantia nigra to pre-frontal cortex, nucleus accumbens and striatum; SERT plays its role in the pons and upper brain stem; NET is localized in the locus coeruleus and the lateral tegmental group that project to many other brain regions. VMAT1 is expressed transiently during brain development and VMAT2 is the main vesicular transporter in these monoaminergic neurons (7). Importantly, these monoaminergic neurons intervene with each other and with many other types of neurons and innervate various brain regions including cortex, hippocampus, amygdala and hypothalamus.

The extensive distribution of these transporters determines their central roles in neurotransmission and ideal medication targets for a spectrum of monoamine-related neuropsychiatric disorders, including attention deficit hyperactivity disorder (ADHD), depression, anxiety, addiction, narcolepsy, fatigue, obesity, eating disorder, other mood disorders, schizophrenia (SCZ), bipolar disorder and Parkinson’s disease. On the other hand, the central roles have also presented these plasma membrane proteins as functional targets for drugs of abuse such as alcohol, cocaine, methamphetamine and MDMA (3,4-methylenedioxymethamphetamine or Ecstasy). In this chapter, we summarize the recent progress in our understanding of the contribution of these monoamine transporters to brain function and diseases.

II. Clinical benefits: demonstration of the medical roles of monoamine transporters

Due to amino acid sequence and proposed structural similarity among the three plasma membrane transporters (DAT, SERT and NET), many monoamine transporter inhibitors have affinity for all three transporters. Unlike the other sections below that discuss individual transporters, this section categorizes the main diseases and their treatments with monoamine transporter inhibitors.

Depression

Depression is the most common disease that is treated by directly targeting the norepinephrine (8), serotonin (9), dopamine (e.g. 10-11) transporters, and/or some combination of the three (e.g., 12).

First developed in the 1950s in an attempt to improve the effectiveness of chlorpromazine and tricyclic antidepressants (TCA) function by inhibiting the reuptake of serotonin, norepinephrine, and dopamine through blocking each respective neurotransmitter transporter (SERT, NET, DAT) (13). Each class of drugs acts on all three of these monoamine systems, with most TCAs primarily inhibiting NET and SERT (14). These medications were then superseded by the selective serotonin reuptake inhibitors (SSRI) as antidepressants. As the most commonly prescribed antidepressant medication, SSRIs are posited to work more effectively within the complex central nervous-neural circuit-gene system in the epidemiology of depression (15), and thus have far less adverse side effects in comparison to TCAs and MAOIs (9). Hundreds of placebo controlled trials have demonstrated benefits in moderate to severe depression, particularly in those with symptoms of more acute major depressive episodes and dysthymia (9, 16, 17) and melancholic depression (18, 19). SSRIs also possess strong therapeutic activity for various DSM-IV-TR disorders (e.g. panic) as described below.

The most recently utilized class of antidepressants falls under selective norepinephrine/dopamine reuptake inhibitors (SNDRIs), with bupropion (Welbutrin) as the most commonly prescribed one. Bupropion is an effective and generally well-tolerated option in the treatment of moderate to severe Major Depressive Disorder (e.g., 20). In addition, bupropion has been shown to be as effective as many common psychopharmacological medications in managing symproms of depression (21). Trials have demonstrated that SSRIs appear to be more effective in the treatment of moderate/acute depression while SNDRIs may be advantageous in the treatment of chronic depression (14).

Obsessive Compulsive Disorder (OCD) and other Anxiety Disorders

Clomipramine, an inhibitor of SERT > NET > DAT and some receptors, was discovered by Spanish psychiatrist Lopez-Ibor in 1960’s to be effective for treating OCD symptoms and the efficacy was subsequently confirmed by many other trials (see 22-24). SSRIs have been used for OCD but the clomipramine effect size appeared to be larger than the SSRIs (e.g., 25). Regardless, clomipramine and SSRIs remain an integral part of “best practice” management of OCD.

In fact, drugs acting on the monoamine transporters, specifically SERT, have efficacy in other anxiety disorders too. Panic Disorder w/agoraphobia was originally named in DSM III in 1980 following research in the USA described as “pharmacodissection” using the NET inhibitor imipramine (26). Recent studies show benefits for norepinephrine and serotonin reuptake inhibitors in treatment of panic agoraphobia (27, 28), Social Anxiety Disorder (29), Generalized Anxiety Disorder (30), and Post Traumatic Stress Disorder (31).

Chronic pain syndrome

In addition to new interventions (e.g. the use sodium oxybate in the treatment of fibromyalgia pain and insomnia), TCAs, SNDRIs and SSRIs are promising medications for fibromyalgia pain (32, 33). TCAs have been shown to improve the symptoms of pain, poor sleep and fatigue associated with fibromyalgia but still possess greater side effects (e.g., 34). The results of SNRIs are more promising (35), although current effect sizes are smaller than trials using TCAs. The efficacy of pain treatments seemed to be better achieved by balanced inhibition of multiple monoamine transporters (33).

Attention Deficit Hyperactivity Disorder (ADHD)

Atomoxetine, a NET inhibitor, has been shown in randomized, clinical trials to significantly reduce ADHD symptoms in both comorbid and noncomorbid children (36), adolescents, and adults with ADHD (37). While stimulants including the DAT inhibitors methylphenidate and amphetamine remain the most frequently prescribed medication in treating ADHD, SNDRIs are currently the leading second line alternatives (38, 39).

Cigarette Smoking

Reviews on the effect of cigarette smoking postulate that chronic exposure to nicotine elicits depressogenic changes in serotonin formation and release in the hippocampus (40). These changes may contribute to the symptoms of depression experienced by many smokers when they first quit. The research examining this relationship has resulted in clinically significant findings. For instance, medications such as bupropion and nortriptyline have been shown to be efficacious in the treatment of cigarette smoking (41). Moreover, it has been shown that a treatment modality which includes nicotine replacement/cessation therapy is recommended for individuals who are highly nicotine dependent and who have a current or past history of major depressive disorder (42).

In summary, DAT serves as a medication target for ADHD, depression, OCD, smoking cessation, as well as narocolepsy and Parkinson’s disease that are not mentioned here; SERT for depression, anxiety, OCD, pain; and NET for depression, OCD, anxiety and ADHD, diseases that affect approximately 20% of the populations.

III.Preclinical indications – behavioral pharmacology

During the last 40 years of preclinical and human psychopharmacology research, the monoamine transporters have been recognized to play a central role in modulating a wide variety of physiological and behavioral functions, including locomotion, autonomic function, and hormone regulation, and to make a fundamental contribution to emotional and cognitive function, neurotoxicity and mental disease. Such a central role is paralleled by the breadth of monoaminergic projections to the neocortex, basal ganglia and limbic forebrain. These pharmacologic and anatomical findings have also been instrumental for the identification and development of medications that are currently used as pharmacotherapies to treat a variety of behavioral disorders, such as major depression, OCD, anxiety, ADHD, and addiction.

DAT

The pioneering investigations of Arvid Carlsson and Kjell Fuxe established DA as a neurotransmitter in the late 1950s (43) and the topography of the dopaminergic innervation of the central nervous system was soon delineated. The mesolimbic DA system originating in the ventral tegmental area innervates the nucleus accumbens of the ventral striatum, where DA is postulated to participate in the control of exploratory activity, reward-related processes, and reinforcement, both natural and drug-induced (44, 45). In turn, the nigrostriatal contingent of DA fibers projecting to the dorsal striatum (i.e., caudate and putamen nuclei in humans) is linked to the control of movement, as revealed by the clinical phenomenology associated with basal ganglia disorders (46), and for “chunking” action repertoires and habits (47). Further, the innervation of the prefrontal cortex by the mesocortical DA pathway has been proposed to modulate various aspects of executive function, including working memory function, planning, and attention (48). Historically, the emerging view of DA as an important transmitter has been reinforced by two additional findings. First, the clinical efficacy of antipsychotic medications was observed to correlate with the binding affinity for DA D2 receptors. Second, observations showed that most abused drugs, especially psychomotor stimulants such as cocaine and amphetamine (AMPH) exert psychoactive effects through interactions with the DA system, some of which involved mainly the DAT.

Preclinical experiments and human data have demonstrated that the DAT is involved in the behavioral reinforcing and euphorogenic effects of stimulant drugs. The ability of cocaine-like drugs to maintain self-administration in rodents is correlated with their potency in inhibiting the DAT (49). The idea is that cocaine binding to DAT may increase extracellular DA concentration by blocking the reuptake activity and inducing the release of reserve pool of DA (50), activating DA receptors. Moreover, the self-reported “high” induced by stimulants in humans appears to be a function of both the rate of DAT occupancy by the stimulant and the speed of stimulant delivery into the brain (51). This evidence suggested that both the binding affinity for the DAT and the pharmacokinetic/pharmacodynamic properties are important characteristics that predict the psychopharmacological effects of drugs acting at the DAT. In addition, preclinical behavioral assays including tests of locomotor activity, conditioned place preference, drug discrimination and self-administration indicated that various DAT inhibitors differ from prototypical stimulants such as cocaine and AMPH. This evidence fueled speculation that it might be possible to design molecules that bind to the DAT, and prevent the actions of AMPH and cocaine at the DAT but lack psychomotor stimulant-like effects. However, AMPH- or cocaine-antagonism without blocking DA reuptake activity has not been successful due to overlapping DAT sites for stimulant binding and DA recognition.

An alternative to antagonism approach is a substitute approach: a slow onset, long acting competing agonist could be used to treat stimulant addiction. As indicated, the specific pharmacokinetic/dynamic features and rate of DAT occupancy are important factors that influence the cocaine-like properties of DAT inhibitors (51). Different DA uptake inhibitors display specific modes of interaction with the DAT, leading to specific conformational changes of the protein (52, 53), and the different DAT conformations are related to the ability of the inhibitors to induce locomotor activity and substitute for cocaine in discrimination assays (54). It is currently believed that the rational design and development of high-affinity, long acting DAT ligands with specific pharmacokinetic/dynamic properties might lead to the discovery of optimal medications for stimulant addiction (55, 56). A variety of molecules based on the 3-aryltropanes (WIN compounds) (57), the 1,4-dialkylpiperazines (e.g., GBR 12909 and its analogues) (55), and analogues of benztropine (BZT) (56) have all been synthesized, tested in vitro for binding at different transporters and receptors, and evaluated in preclinical models. Of these, BZT analogs exhibit ideal characteristics in preclinical assays, blocking the behavioral effects of cocaine and AMPH and exhibiting weak abuse liability in place preference and self-administration assays (58-60).

SERT

5-HT-containing neurons of the raphe regions located in the pons and upper brain stem extensively innervate the diencephalon and telencephalon, providing input to the hypothalamus, habenula, thalamus, amygdala, striatum, and cortical mantle (61). The 5-HT system is believed to regulate mood, emotion, learning, memory, sleep and appetite. Of the chemical neurotransmitters, 5-HT is perhaps the most widely implicated in the treatment of mental illnesses (62). The 5-HT transporter (SERT) is responsible for taking up 5-HT from the synapse and is indeed the target of an ample range of molecules which inhibit the uptake process. Two aspects have been decisive to confirm the key functions of 5-HT and the importance of inhibiting 5-HT transport as therapeutic principle. First, the discovery that 5-HT is structurally related to many psychotropic agents, including lysergic acid diethylamide (LSD) and psilocin (63) . Second, the finding that many drugs with psychoactive properties, including cocaine, AMPH, ecstasy, TCAs and SSRIs, effectively interact with SERT to block 5-HT uptake (64). These findings, particularly the latter, confirmed a critical role for 5-HT in the regulation of mood and affect, and marked a milestone in neuropsychopharmacology and psychiatry research.

Over the last decades, tremendous strides have been made in the treatment of major depressive illness, anxiety and eating disorders. The introduction of the first generation of monoamine transporter inhibitors, the TCAs, which target both NET and SERT with variable affinity, revolutionized the management of affective disorders. However, the widespread activity of TCAs at multiple biological sites, including not only monoamine transporters but also noradrenergic, histaminergic and muscarinic receptors, associates with unwanted side-effects (65). The development of SSRIs, such as fluoxetine, citalopram and fluvoxamine, led to fewer side-effects, while retaining full clinical efficacy. In animal models of depression, such as the learned helplessness and forced swimming tests, SSRIs have systematically displayed strong activity suggestive of clinical efficacy (66, 67). These models provide opportunities to assay new compounds and investigate the mechanisms underlying preclinical efficacy. Indeed, in spite of adequate treatment with current antidepressant medication, a large proportion of patients do not receive full symptom remission when treated with SSRIs, and new approaches are presently pursued. A promising lead for improved therapeutic effects is the development of triple uptake inhibitors targeting SERT, NET and DAT (SNDRIs). Several of these compounds are in advance stages of development, though none has yet been approved for human use. In animal models, the SNDRI, DOV 21,947, inhibits reuptake of 5-HT, NE and DA (EC50 of 12, 23, and 96 nM at human recombinant transporters, respectively) and exhibits antidepressant effects in the forced swim and tail suspension test (68). DOV 102,677, another SNDRI (EC50 of 129, 103 and 133 nM at human recombinant transporters, respectively) was as effective as methylphenidate in reducing the amplitude of the startle response in juvenile mice, in addition to showing an antidepressant profile (69). Preclinical indications suggest that “triple inhibitors” may produce a more rapid onset of action and greater efficacy than traditional antidepressants (70). In addition, an emerging literature indicates that non-classical transporters such as the plasma membrane monoamine transporter (PMAT) and organic cation transporters (OCTs), which subserve promiscuous uptake of biogenic amines, may constitute new targets for improved antidepressant action, alone or in combination with SSRIs (71).

It is noticed that NET and SERT are also binding sites for psychomotor stimulants including cocaine, AMPH, methamphetamine or ecstasy (3,4-methylenedioxymethamphetamine, MDMA). Although DA seems to underlie the reinforcing effects more closely, several studies have suggested that these binding activities contribute to at least some of the reinforcing properties (72-76), warranting medication targets for drug addiction.

NET

There are two major noradrenergic neuronal clusterings in the brain: the locus coeruleus and the lateral tegmental group, which provide extensive innervation to the striatum, amygdala, hypothalamus, thalamus, cerebellum, and neocortex (77). Such widespread ascending projections have been implicated in the modulation of arousal, sleep, and cognitive processes (78, 79). An important additional function of the NE system is to control the endocrine and the autonomic nervous system, which play a fundamental role in anxiety and the stress response (80).

The NE system is an important target for a wide range of drugs used for the treatment of mood, anxiety and behavioral disorders, including major depression, generalized anxiety disorder and ADHD. The neurobiological links between stress, anxiety and depression have long been postulated, but the identification of such relationship has only begun to emerge. The locus coeruleus is uniquely placed to integrate both exteroceptive cues and internal visceral/endocrine information, and to influence stress- and fear-related anatomical structures, including amygdala, periaqueductal gray and neocortex (81). A great deal of evidence suggests that the NE system and the corticotrophin-releasing factor (CRF) pathways co-regulate their activation in response to fear and stress (82). Pharmacological inhibition of NE transport with NET blockers that display little or no activity at other monoamine transporters exerts activity in several animal models of stress and depression. Atomoxetine and reboxetine, which exhibit 50-to-100-fold preference at human NET versus other monoamine transporters (83), show strong activity in animal models of depression that is predictive of therapeutic effects in humans. For example, the therapeutic-like effects of reboxetine have been assayed in the olfactory bulbectomized (OB) rat model of depression, reducing immobility time in the forced swim test and hyperactivity in the open field (84). Reboxetine also attenuates the physiological responses associated to swim stress, including 5-HT elevations in amygdala and prefrontal cortex and activation of hypothalamic pituitary adrenal axis (85). The anti-depressant-like effects of reboxetine in the forced swim test are blocked by 6-hydroxydopamine lesions of the ventral noradrenergic bundle, suggesting that enhanced noradrenergic activity mediates the effects of reboxetine (86). Chronic social stress can produce depressive-like symptoms in mice and rats including decreased locomotor and exploratory behavior, reduced sucrose preference and increased immobility in the forced swim test. Such disturbances elicited by chronic social stress are ameliorated by reboxetine (87).

Altered NET activity likely contributes to ADHD as well, consistent with the fact that selective NET inhibitors are primarily prescribed for treatment of ADHD (88). Reboxetine and atomoxetine enhance attentional performance, and reverse or attenuate some cognitive deficits in animal models of ADHD. Rats trained in a five-choice serial reaction time task increased the percentage of correct responses and decreased the number of premature responses under the influence of reboxetine (89) or atomoxetine (90). In the spontaneously hypertensive rat, a genetic model for ADHD, atomoxetine ameliorated the learning deficits that typically exhibit these rats (91). Further, atomoxetine is an effective treatment for ADHD, especially in patients with comorbid disorders and those who do not tolerate well stimulant drugs (92).

VMAT2

In virtually all tissues, a large proportion of monoamines present are located within specialized subcellular particles referred to as synaptic vesicles. The VMAT2 is responsible for the translocation of DA, 5-HT and NE among others from the cytoplasm into these synaptic vesicles. Such vital function points to these transporters as important players in monoaminergic transmission and potential target for the development of treatments for neuropsychiatric illnesses.

Some of the neuropharmacological and neurotoxic effects of various stimulant compounds are likely to result from interference with VMATs. In purified striatal tissue, cocaine treatment rapidly and reversibly increased both the V(max) of DA uptake and the B(max) of VMAT2 ligand (dihydrotetrabenazine, DTBZ) binding, with other DAT inhibitors, such as GBR 12935, similarly increasing vesicular DA uptake (93). As assessed in ex vivo fractions from striatal tissue, cocaine shifted VMAT2 protein from a synaptosomal membrane fraction to a vesicle-enriched fraction (94). The cocaine-induced VMAT2 redistribution, reduction in DA release, and decrease in total DA transport are mediated by DA D2 receptors, which are activated following accumulation of extracellular DA (95). However, the effects of AMPH and its analogs, including methamphetamine (METH) and MDMA, are radically different to those of cocaine. AMPH-like molecules enter into the cytoplasm through both uptake by monoamine transporters and diffusion across the cell membranes, impede vesicular DA sequestration and promote reverse DA transport and release (96). Contrary to cocaine, METH exposure rapidly decreased vesicular uptake and DTBZ binding (93) and caused a redistribution of VMAT2 from a vesicular-enriched fraction to a location outside the synaptosomal preparation (94). MDMA also induced an abrupt decrease in vesicular DA transport in striatal vesicles prepared from treated rats (97). The importance of VMAT2 in mediating drug-induced physiological and behavioral effects has been confirmed in mice deficient in VMAT2. VMAT2 heterozygous mice are supersensitive to cocaine and AMPH, but do not develop sensitization to cocaine (4). Collectively, these observations suggest that cocaine and AMPH-related compounds differentially affect vesicular DA transport and subcellular localization of VMAT2. These findings may have important implications for understanding the behavioral effects and neurotoxic profiles of psychomotor stimulants.

VMAT2 may also be a pivotal player regulating the balance between toxicity and neuroprotection. The maintenance of low levels of cytosolic monoamines is important to prevent oxidative damage associated with deamination (98). Such critical patrolling function requires effective reuptake of monoamines by VMAT2. Therefore, inhibitors of VMAT2 could produce toxic effects in monoaminergic neurons, a prediction that is consistent with the distinct neurotoxic profile of METH and MDMA. Moreover, VMAT2 is involved in N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced neurotoxicity. The active metabolite of MPTP, N-methyl-4-phenylpyridinium (MPP+), is a substrate for VMAT2. These data suggest that pharmacological agents that increase VMAT-2 activity may afford neuroprotection and constitute new treatments for neurodegenerative disorders. The quantitative assessment of VMAT2 density by PET scanning has been clinically useful for early diagnosis and monitoring of neurological and neuropsychiatric conditions, including Parkinson’s and Alzheimer’s diseases and drug addiction (99).

In summary, monoamine transporters represent brain pathways underlying many behaviors associated with addiction, depression, anxiety, social management and Parkinson’s disease among others.

IV. Molecular study: cDNA cloning and structure activity relationship (SAR)

Molecular characterization of the transporters was not possible until their cDNAs were cloned in early 1990’s. By using various strategies, several laboratories cloned the cDNAs and revealed the amino acid sequences, enabling the immediate study of structural and activity relationship. With the help of site-directed mutagenesis analyses and elucidation of crystal structure of the bacterial homolog Leucine transporter (LeuTAa) (100), functional residues in monoamine transporters are being uncovered for ion transport, substrate specificities and drug binding sites.

cDNA cloning and characterization

Structural study of monoamine transporters was not possible until early 1990’s when the transporter cDNAs were cloned, which was initiated by successful purification and cloning of GABA transporter GAT 1 (101, 102). By direct expression of human SK-N-SH cell cDNA pools in COS-1 cells and screening for radio-labeled norepinephrine analog, m-iodobenzylguanidine ([125I] mIBG) positive transfected cells, Amara’s laboratory cloned the hNET cDNA (103). Based on conservative amino acid sequences between GAT1 and NET, degenerative oligonucleotides were used to amplify cDNA fragments from brain mRNA pools to screen cDNA libraries and clone the DAT and SERT (104-106). Due to structural differences from the plasma membrane transporters, the vesicular VMAT1 and VMAT2 transporters were cloned based on their abilities to take up neurotoxin (expressing cells survived in the presence of the toxin) or radio-labeled substrate in the expressing cells vs non-expressing cells(107, 108).

With the cDNAs cloned, all transporters were subsequently characterized and their basic features are summarized in Table 1. The plasma membrane and vesicular monoamine transporters belong to different families and have different biophysical properties, such as ion dependence, action as symporter (NET, DAT, and SERT) or action as an antiporters (VMAT1 and VMAT2), rank order of substrates, membrane localization, drug binding sites and affinities. They also share some common biophysical properties, such as uptake activities for monoamines, 12 transmembrane domains (TMs), co-expression in monoamine neurons. The active transport activities of the plasma membrane and the vesicular transporters are unidirectional to the cytoplasm and the vesicular lumen, respectively, under normal neurotransmission because of the extracellular sodium gradient created by Na+/K+ ATPase and the intravesicular proton gradient created by H+ ATPase.

Table 1.

Biochemical and molecular properties of monoamine transporters*

Protein alias DAT SERT NET VMAT1 VMAT2
Substrates dopamine
norepinephrene
amphetamine,
MPP+
Serotonin
MDMA
norepinephrene
dopamine
amphetamine
serotonin
dopamine
norepinephrene
serotonin
dopamine
norepinephrene
Rank DA>NE>5HT 5HT>MDMA DA>NE>5HT 5HT>DA>NE 5HT>DA>NE
K M 885 nM 320 nM 457 nM 1 μM 1 μM
Selected
inhibitors
mazindol
amphetamine
cocaine
paroxetine
citalopram
chlorimipramine
mazindol
amphetamine
cocaine
reserpine
tetrabemazine
reserpine
tetrabemazine
Ion Na+, Cl Na+, Cl, K+ Na+, Cl H+ H+
Amino acids 619 653 617 521 515
Tissue
distribution
mid brain brain stem
mid brain
peripheral
brain stem Adrenal gland
PC-12 cells
brain stem
mid brain
stomach
References (104) (106) (103) (108) (108)
*

All transporters have 12 TMs.

SAR of the plasma membrane transporters

No crystal structures are solved for any of the monoamine transporters. Immediately after the cDNA cloning, a large number of site-directed mutagenesis studies were carried out to identify amino acid residues important for substrate and inhibitor recognition and interaction with ions. Subsequently, a crystal structure of the bacterial transporter, Aquifex aeolicus leucine transporter (leuTAa) (100) that shares amino acid sequence identity of 20-25% with the plasma membrane monoamine transporters (109, 110), was solved and has been used as a template to model the structures of SERT (110-114), DAT (110, 113, 115, 116) and NET (110, 113, 117). These SAR and molecular modeling studies have revealed TMs 1, 3, 6 and 8 as important binding domains. Specifically, residues Asp98, Tyr267 and Tyr289 of SERT are conserved among the three plasma membrane transporters and postulated to comprise the substrate binding cavity. Six SERT-specific residues including Asp98, Tyr95, Gly338, Ile172, Ser336 and Gly442 appear to interact with the substrate (118-121). The acidic side chain of Asp98 is predicted to interact withsubstrate by formation a salt bridge with the positively charged amine of 5-HT (111). Tyr95 and Gly338 are predicted to be involved in interaction with the aliphatic side chain of 5-HT. The residue Tyr95 probably stabilizes the substrate by stacking interaction between the aromatic ring of Tyr95 and the 5-HT (111). Other functional residues interact with ions. Ala96, Asp98 (D79 in DAT and D75 in NET), Asn101, Ser336 and Asn368 of SERT are postulated to coordinate the first Na+ ion binding site (111). Other conserved residues including Gly94, Val97, Leu434, Asp437 and Ser438 of SERT are predicted to coordinate the second Na+ ion (111). The Cl-binding site in SERT consists of Tyr121, Ser336, Asn368 and Ser372 that are conserved among the three plasma membrane transporters (122, 123). Replacements of Asn368 and Ser372 with negatively charged residues Glu or Asp lead to the transportation of substrate 5-HT in the absence of Cl− ion (122, 123), indicating that the negatively charge amino acids can function as the Cl− ion. In particular, Cl-has been shown to be involved in facilitating a conformational change in SERT (124) and DAT (125) for alternating substrate binding.

In summary, the functional cloning and biochemical analysis of both plasma membrane and vesicular monoamine transporters are building up a solid foundation for the future study of those important drug targets and clinical and pathological relevant molecules.

V. Protein regulations

Monoamine transporters share other common features including oligomerization and cytoplasmically localized N- and C-terminals, in addition to 12 TMs (126-131). Oligomerization appears necessary for the transporter complex to efficiently exit the endoplasmic reticulum and cytoplasmic tails mediate subcellular trafficking of the transporter proteins (129, 132). It has also been suggested that the amphetamine-induced substrate efflux depends on the oligomeric nature of monoamine transporters (133).

Sucellular localization

Different transporters are localized at different subcellular sites. NET is primarily expressed within the cytoplasm in the majority of prefrontal cortex nerve terminals (134). Interestingly, the cytoplasmic localization of NET correlates with the absence of detectable Tyrosine Hydroxylase immunoreactivity, suggesting that most noradrenergic neurons in the prefrontal cortex operate in a low-activity state because they can neither synthesize, nor recapture noradrenaline. Manipulations that increase in vivo stress responses result in an increase of both the plasma membrane expression of NET and Tyrosine Hydroxylase immunoreactivity, illustrating the importance of trafficking mechanisms in the regulation of neuronal activity. By contrast, DAT and SERT are largely expressed at the plasma membrane of nerve terminals and specifically to perisynaptic areas (135, 136), supporting volume transmission as the primary transmission mode for dopamine and serotonin. Recent observations suggest that dopamine and serotonin can also be released from dendrites (137-139); however, the role of DAT and SERT in the regulation of postsynaptic monoamine release has not been well established.

Drug regulation

The subcellular localization can be regulated by psychostimulants including cocaine and amphetamine (AMPH) (140). It is reported that DAT cell surface levels are increased in response to cocaine without known mechanisms (141, 142). The transport of AMPH-like drugs results in a complex cascade of events leading to the redistribution of vesicular monoamine into the cytosol, reversal of transport, and consequent transporter-mediated monoamine release (143, 144). Phosphorylation of the amino terminus of DAT by Protein Kinase C (PKC) and Calmodulin-Dependent Kinase II (CAMK II) has recently been shown to be essential for the AMPH-induced DA efflux (145-147). AMPH regulated DAT trafficking bidirectionally between cell surface and intracellular vesicle by mechanisms to be dissected (148, 149).

DAT regulators

At the cell membrane, the function of monoamine transporters can be regulated also by multiple second messenger systems including Protein Kinase A, Protein Kinase C, Protein Kinase G, Tyrosine Kinases, Phosphatases, Calcium and Calmodulin-dependent Kinases, as well as Araquidonic Acid (for review see 150-154). Of these; the best characterized effect is the down-regulation of transporter activity by Protein Kinase C (PKC) activators through internalizing the proteins for degradation (155, 156). Interestingly, the down-regulation of DAT function by PKC activation is not a direct result of transporter phosphorylation as mutant transporter molecules that lack PKC phosphorylation sites still internalize after PKC activation (157). Endocytic signals have been identified in the intracellular carboxy terminus of DAT (158-160). In looking at the mechanisms associated with the internalization of DAT, Sorkin and colleagues (161) found that PKC activation significantly increases transporter ubiquitination through a mechanism involving the E3 ubiquitin ligase Nedd4-2 and the clathrin-coated pits adaptor proteins Epsin, Eps15, and Eps15R. Three lysine residues located in the amino terminus of DAT are required for both the PKC-induced ubiquitination and down-regulation of transporter function (162). Another study showed that the protein-ubiquity E3 ligase Parkin enhances DAT uptake and cell surface expression of the transporter via a mechanism involving an enhanced ubiquitination and degradation of misfolded DAT (163). Thus, a model is emerging in which ubiquitination is an important signal that links the plasma membrane transporter to the endocytic and degradation machinery (164). In addition to second messengers, other proteins also regulate DAT for transporter synthesis and assembly, subcellular localization, or modulating intrinsic activities such as uptake, efflux, and conductances associated with transporters.

The first DAT interacting protein identified was the synaptic Protein Interacting with C Kinase (PICK1) (165). PICK1 is a PDZ domain-containing protein which plays an important role in the targeting and clustering of several receptors and ion channels at synapses. Specifically, PICK1’s PDZ motif binds to a class I PDZ binding site located at the extreme carboxyl terminal of DAT, which may facilitate the targeting of DAT to nerve terminals. Although PICK1 also binds the carboxyl terminus of NET and SERT in vitro, the physiological significance of these interactions is unknown (166). Other regulators include α-Synuclein (167-169), Hic-5 (166), PP2A (170), NEDD4-2 (161, 171), GPR37 (172), Parkin (163, 173), PKC-β (146), and dopamine D2 receptors (174). In all these cases, manipulations that alter the expression levels of the interacting protein resulted in changes in the DAT cell surface levels. Interestingly, the ability of DAT to co-immunoprecipate with the D2 receptor appears to be decreased in post-mortem brain tissue from schizophrenic patients (175). The increasing number of proteins that influence trafficking mechanisms leaves questions as to whether each of these proteins interacts with DAT at different steps in the endocytic, recycling, degradation pathways and/or individual pathway steps require multiple interactions.

More recently identified DAT regulators include PKC-β (146), CAMKII-α (147), Syntaxin 1A (176, 177), and Synaptogyrin-3 (178). PKC-β, CAMKII-α, and Syntaxin 1A all promote DA efflux in response to AMPH. Inhibition of PKC-β blocks AMPH induced DA release in cultured cells (146). Consistently, PKC-β knock-out mice have less AMPH stimulated DA efflux and reduced locomotor activity in response to AMPH (179). On the other hand, studies have shown that CAMKII-α binds directly to the carboxyl terminus of DAT (147). This association between CaMKII-α and the carboxyl terminus of DAT and promotes phosphorylation of the amino terminus of DAT. Interestingly, CAMKII-α activation is necessary to enable Syntaxin 1A to bind directly to the amino-terminus of DAT and promotes DA efflux (177). Synaptogyrin-3 is also a synaptic vesicle protein and it modulates the transporter function via N-terminus-N-terminus binding (178). Functional studies showed that Synaptogyrin-3 overexpression in catecholaminergic PC12 and MN9D cells resulted in increases in DAT activity without changes in cell surface density. These results were not recapitulated in cells devoid of a vesicular dopamine system, suggesting that the DAT/Synaptogyrin-3 interaction play a role in docking synaptic vesicles at the plasma membrane near DAT to facilitate a more efficient loading of the vesicles with extracellular DA after release. Indeed, studies also showed an interaction between DAT and VMAT2 (178), perhaps for a rapid and efficient coupling between the two transporters.

SERT and NET regulators

SERT and NET are also regulated by several proteins in trafficking as well as intrinsic activity. For example, PP2A and PKC interact with SERT, in addition to DAT and NET (170). The PKC-induced internalization of SERT was shown to be prevented by serotonin (180) suggesting a feedback mechanism where the transmitter can control the levels of plasma membrane transporter “on demand.” Syntaxin 1A has also been identified as a SERT/NET interacting protein and might have a dual role regulating both transporter trafficking and intrinsic channel activity (181, 182). Additional regulators include Hic-5 (166, 183), α-Synuclein (168, 169), the Secretory Carrier Membrane Protein 2 (184), Neuronal Nitric Oxide Synthase (185), and the cGMP-dependent Protein Kinase alpha (186). The latter is an important link between the regulation of transporter function and PKG activation. Activation of adenosine A3 receptors results in two PKG-dependent signaling pathways regulating SERT function. One of these mechanisms leads to an increase in catalytic activity; whereas, the second one is p38MAPK-dependent and involves an increase in the cell surface levels of the transporter. In addition to PP2A, other known regulators of NET include PICK1 (165), Hic5 (166), α-Synuclein (169), Syntaxin 1A (187) and 14-3-3 proteins (188).

VMAT2 regulators

Much less is know about regulation of VMAT2, except its physical interaction with dopamine synthesizing enzymes such as tyrosine hydroxylase (TH) and aromatic amino acid decarboxylase (AADC) (189).

In summary, the activities of plasma membrane transporters are regulated by both small molecules and proteins at two main levels: cell surface expression and intrinsic activity. These regulators compromise an expanding community associated with various mechanisms yet to be fully understood. Thus, our current understanding of how monoamine transporters and homeostasis are regulated will surely be dynamic in the years to come as we continue to identify and investigate additional interactions with monoamine transporters.

VI. Animal genetics

The cDNA cloning has enabled animal genetic studies since middle 1990’s. Findings from these genetic and modeling studies are consistent with the results from the pre-clinical research and confirm the roles of monoamine transporters in brain function and behaviors.

DAT

The essential role of DAT in the brain is demonstrated by gene knockout studies. Complete deletion (z−/−) causes dramatic affects including hyperdopaminergia with delayed clearance of extracellular dopamine, abnormal brain development, spontaneous hyperactivity and high response to and/or lack of habituation to novelty, motor and learning deficits, sleep disturbances and loss of cocaine’s abuse-related effects (190-200).

DATkd (knock-down) mice expressing 10% of wild-type DAT levels displayed similar effects including hyperactivity with delayed habituation and increased stereotypy, and deficits in response inhibition, but not prepulse inhibition of startle (201-204). DATkd mice also consumed more food and showed behaviors suggesting increased “motivation” towards food reward (205-207). The DAT-reduced mice thus may model aspects of manic or obsessive-compulsive behavior, Tourette’s syndrome, ADHD, and perhaps of over-eating. Consistently, mice overexpressing DAT by 20-30% showed phenotypes opposite to hyperdomaninergic DAT-deficient strains, including lower spontaneous activity with prompter habituation in a novel environment, increased neurotoxicity of MPTP (208), and increased amphetamine-induced dopamine release, hyperactivity and conditioned place preference (CPP) (208, 209). These findings support the results from the DAT-deficient mice.

Recent DAT transgenic studies have demonstrated that DAT is the functional target of cocaine in the brain. Mice carrying a cocaine-insensitive mutant DAT (DAT-CI) failed to self-administer cocaine under various conditions whereas food, D-amphetamine, and a direct dopamine agonist maintained operant behavior at levels comparable to wild-type mice (210). Cocaine failed to increase extracellular accumbens dopamine or induce locomotor activity stimulation, stereotypies, or conditioned place preferences in the DAT-CI mice (211-213).

SERT

The contribution of SERT to various rodent behaviors is extensively documented by both gene knockout in mice and inactivation (N-ethyl-N-nitrosourea or ENU mutagenesis) in rats. Complete loss of functional SERT caused spontaneous serotonin syndrome-like behaviors (tremor, Straub tail and backward movement) and hypersensitivity to its induction by serotonergic or opioid drugs, reduced motor coordination, strength, and locomotor activity, an anxiety-like and depressive-like phenotype and decreased aggression and other social interactions (214-227). In agreement with the large literature implicating 5-HT in the pathophysiology and treatment of mood and anxiety disorders, SERT-deficient (−/− and +/-) mice showed an anxiogenic phenotype, which was ameliorated by a 5-HT1A antagonist (215, 220, 228, 229). Cognitive assays showed SERT−/− rodents may perform better in emotionally guided/motivated learning, but worse in “neutral” tasks - perhaps related to the anxiety-like phenotype (230, 231). Consistently, a SERT-overexpressing mouse line showed a phenotype opposite to the SERT−/− in many respects: reduced extracellular and total 5-HT levels, and decreased anxiety-like behaviors (232, 233). It is noteworthy that genetic deletion thus had effects opposite to SSRI treatment in many respects, suggesting compensatory mechanisms in the mutant animals, or perhaps complex adaptations to chronic SSRI treatment.

NET

NET−/− mice generally behaved like anti-depressant-treated wild-types in several assays, and showed no or reduced further effect of most antidepressants, corroborating the role of norepinephrine systems in the action of some antidepressant drugs (234-236). Furthermore, NET−/− mice were resistant to depressive-like effects of stressors, behaviorally and neurochemically (236). Also, consistent with the mild antinociceptive effects of α2 agonists and tricyclic antidepressants, NET−/− mice showed α2-mediated enhancement of antinocicepive effects of morphine and endogenous opioids (237). Like DAT−/− mice, NET−/− mice were less vulnerable to MPTP-induced neurotoxicity (238). Studies in NET−/− mice have also implicated noradrenergic systems in epilepsy and/or actions of anticonvulsants (239, 240), as well as confirming roles in cardiovascular and respiratory functions (241). Thus NET−/− mice offer a depression-resistant phenotype, and may also be valuable to the fields of analgesia, epilepsy, and aspects of resistance to neurotoxicity/Parkinson’s.

VMAT2

Complete deletion of VMAT2 is lethal within a few days of birth, and in vivo studies have concentrated on the viable VMAT2+/- mice with reduced VMAT2 expression (4-6). VMAT2+/- mice were behaviorally hypersensitive to the acute stimulant effects of cocaine, amphetamine, methamphetamine and ethanol (4, 6, 242). Paradoxically, dopamine-releasing, rewarding and sensitizing effects of psychostimulant drugs and ethanol were reduced in the VMAT2+/- mice (4, 6, 243, 244). As expected, VMAT2+/- mice were found more vulnerable than wild-types to neurotoxicity/neurodegeneration induced by MPTP, L-DOPA, or methamphetamine (6, 243, 245-247). VMAT2+/- mice also showed depressive-like behaviors (244, 248, 249).

In addition to the knockout approach, a transgenic mouse strain with very low levels of VMAT2 expression (5% of wild-type) was generated (250). Unlike the VMAT2−/− mice, these “VMAT2 LO” (low) mice are viable, but represent a more pronounced VMAT2 deficit than the previously available VMAT2+/- mice. Findings from VMAT2 LO mice corroborated and extended those from VMAT2+/- mice regarding their Parkinson’s-like phenotype, including increased susceptibilit y t o M P T P- or methamphetamine-induced neurotoxicity/ neurodegeneration (250, 251). Further, VMAT2 LO mice showed an age-dependent phenotype reminiscent of pre-parkinsonian symptoms and Parkinson’s disease, including olfactory deficits, anxiety- and depressive-like behaviors, progressive loss of striatal dopamine, nigral dopaminergic cell loss, α-synuclein accumulation and L-DOPA-responsive motor deficits (250, 252-254). Thus VMAT2 mutant mice may be useful to model Parkinson’s and related depressive states, oxidative stress, and amphetamines addiction.

In summary, despite concerns regarding both compensatory mechanisms and conflicting findings, an impressive database now suggests a striking amount of agreement between genetic manipulations in animals and traditional pharmacological approaches. For example, knockout mice lacking monoamine transporters frequently exhibit a phenotype akin to normal animals treated with monoamine reuptake inhibitors. Mice lacking the DAT, but not the SERT or NET, are insensitive to cocaine in assays of hyperactivity and self-administration, and this is even clearer for mice with a “knockin” of a cocaine-insensitive but functional DAT. Such findings bolster the utility of mutant mice generally for behavioral and in vivo pharmacology studies, and suggest that DAT mutants may be used to study aspects of hyperdopaminergic states that may have relevance for ADHD, SCZ, bipolar disease, addiction and Parkinson’s. Correspondingly, SERT mutants may be especially valuable for studying hyper-serotonergic states relevant to anxiolysis and antidepressant treatments, as well as abuse and toxicity of hallucinogens and stimulants such as LSD and MDMA and methamphetamine. Likewise NET mutants offer tools relevant to cardiovascular and respiratory functions, antidepressants, analgesics, anticonvulsants, and neurotoxicity. VMAT mutants have been used mostly to clarify the functions of that protein, and the potential role in Parkinson’s-related neurodegeneration and mood disorders.

VII. The transporter genes as risk factors

Human genetic studies of the transporters did not start until the cDNAs were cloned and the Human Genome Project began in early 1990’s. During the last one and a half decades, a lot of progress has been made to elucidate DNA sequence polymorphisms in the transporter genes and their correlations with gene activities and ultimately with genetic etiologies of diseases. We attempt to summarize the most recent progresses in genetic associations with disease (see 255 for a review on non-synonymous mutations).

Gene structures

The human monoamine transporters are each encoded by single copy genes located on different chromosomes. The human DAT gene or hDAT/SLC6A3 (53.2 kb on Chr 5) is the largest; hSERT/SLC6A4 has 41.4 kb (Chr 17); hNET/SLC6A2 is 48.6 kb (Chr 16), hVMAT1/SLC18A1 (38.4 kb on Chr8) and hVMAT2/SLC18A2 (38.3 kb on Chr 10) the smallest in size, without considering their external regulatory regions such as promoters. Each of first four transporter genes has 15 exons and 14 introns (hVMAT1 has 16 exons and 15 introns).

A remarkable feature of these gene structures is that at a DNA sequence level, these transporter genes carry a lot of polymorphisms in our populations, approximately 7-14 polymorphisms per kb on average based on the current NCBI database information. The average density of polymorphisms is 14.4 in hDAT, 9.0 in hSERT, 7.8 in hNET, 10.5 in hVMAT1 and 8.4 in hVMAT2, expecting that these numbers will be increased by the ongoing 1000 Genomes Project. These high densities of DNA sequence polymorphisms suggest that genetic variation in the transporters contributes to monoamine-related inter-person variation in brain function and diseases in extreme cases.

hDAT

DNA sequence variation in hDAT is correlated with its expressional activity and other brain activity. The well-studied variable number tandem repeats located in the 3′ untranslated region (3′UTR) in Exon 15 (3′ VNTR of 40 bp) has been examined as to which allele of this marker is associated with low expression levels (a risk factor). In vitro and human imaging analyses have consistently shown that the 10-repeat allele is associated with lower expression levels than the 9-repeat allele (256-258). In addition, an Intron 8 VNTR and 5′ haplotypes are also correlated with expressional variations (259, 260). These functional analyses of hDAT genotypes are consistent with genotype-dependent brain activities including those in ADHD patients (261-264), suggesting that genetic variation in the transporter gene may lead to altered brain function and even disorders.

Consistently, association studies have shown that various genetic markers throughout the gene are associated with at least eleven different disorders such as ADHD, substance abuse, depression, SCZ and bipolar disorder in different populations (Table 2). The genetic variation also modulates treatment efficacy for ADHD, depression, SCZ and smoking (312-317). A recent study has suggested that the 3′VNTR and Intron 8 VNTR are differentially involved in ADHD in childhood versus persistent manner (318). Genotype 10/10 and haplotype 6-10 are associated more with childhood ADHD but 9/9 and 6-9 are with a persistent form of this diseases in Europeans. A multiple SNP-formed haplotype is associated with depression in Mexican-Americans (282). Intron 4 rs464049 is associated with SCZ in Croatians (319).

Table 2.

Associations with human diseases

Disease hDAT hSERT hNET hVMAT2 Representitive
reference
1. ADHD 265-267
2. Aggression 268
3. Alcoholism 269-273
4. Alzheimer’s 274
5. Angry 275
6. Anorexia nervosa 276, 277
7. Anxiety 278
8. Autism 279
9. Bipolar disorder * 280
10. Delirium 281
11. Depression 278, 282-286
12. Drugs 287-289
13. IBS in women 290
14. Hypertension 291, 292
15. Insomnia 293
16. Neuroticism 294
17. OCD 295
18. Orthostatic intolerance 296
19. Pain 297
20. Parkinson’s disease 298-300
21. PTSD 301
22. Schizophrenia * 302, 303
23. Smoking (s) 304
24. Suicide behavior 305
25. TCI/TPQ HA 306
26. Panic disorder 307
27. Unipolar disorder 308

Total 11 20 6 4

ADHD, attention deficit hyperactivity disorder; IBS, irritable bowel syndrome; OCD, obsessive compulsive disorder; PTSD, post-traumatic stress disorder; TCI/TPQ HA, TCI/TPQ harm avoidance.

*

hVMAT1 is also reported to be associated (309-311).

Gene-gene interaction is implicated in related disorders. For example, hDAT x DRD2 is found to contribute to smoking in Poles (320) and hDAT x DRD3 to SCZ in Spanish (321). It is also shown to interact with its transcription factor NR4A2 in conferring risk for smoking (322). In addition, epigenetic and environmental involvement further complicates statistical evaluation of genetic associations with ADHD and Parkinson’s disease (298, 323, 324).

hSERT

The differences in hSERT expression levels among different individuals could be as large as 6-fold in selected brain regions (325, 326). A number of lines of evidence have linked hSERT polymorphisms to such differences: the promoter marker 5-HTTLPR has two alleles, s and l and in vitro, postmortem and in vivo studies have suggested that the s allele is associated with lower expression levels of hSERT (326-329).

Consistently with the genotype-dependent expression, genetic markers in hSERT are shown to be associated with 20 different diseases such as ADHD, aggression, alcoholism, Alzheimer’s, autism, substance abuse, anxiety, OCD, depression, suicide, PTSD, SCZ, panic and unipolar disorders (Table 2). In addition, the genetic variations also modulate the treatment efficacy for ADHD and alcoholism and environmental effect (unemployment etc) on PTSD (282, 330-335). Furthermore, the genetic variations may confer risk for depression in patients with different diseases (336, 337).

hNET

While little information is available about genotypic correlation with hNET expression levels in relevant brain regions, approximately 50 association studies have been done and revealed several positive signals. Findings from these studies suggest that this gene is associated with six diseases including ADHD, depression, drug abuse, hypertension, orthostatic intolerance and anorexia nervosa (Table 2). Overall association with ADHD and modulation of the treatment efficacy for this disease has been confirmed by genome-wide association study (GWAS) (338, 339). Only markers located in the 3′ side of the gene did not support hNET association with ADHD (267), suggesting the promoter plays a major role in conferring the risk for ADHD. Future study using the promoter markers are expected to reveal more positive association signals.

hVMATs

Among the transporter genes of main interest here, hVMAT1 and hVMAT2 are the least studied so far. There are fewer than 20 published studies on these genes. Postmortem and imaging analyses suggested that hVMAT2 expression levels vary by more than 10-fold among individuals and were reduced in substance abusers (340, 341).

Limited number of association studies have shown that hVMAT2 is positively associated with four diseases including alcoholism (only disease whose association was replicated), depression, Parkinson’s disease and SCZ (Table 2). Consistently with the fact that hDAT and hVMAT2 are coupled to each other in regulating DA transmission, hDAT and hVMAT2 interact with each other in conferring risk for SCZ (302).

More recently, hVMAT1 is also implicated in brain activity and positively associated with SCZ and bipolar disorder but has not been analyzed for depression yet (309-311, 342).

In summary, the monoamine transporter genes are highly polymorphic in their chromosomal DNA sequences. Such polymorphisms contribute to the variable expressions and risk for at least 27 brain disorders. Remarkably consistent with the preclinical findings, all four studied transporter genes are associated with depression (Table 2). Three of these genes are associated with ADHD, alcoholism, Parkinson’s disease and SCZ.

VIII. Perspectives for medication development

Fifty years of medications demonstrate that the monoamine transporters are effective medication targets for a spectrum of brain disorders including ADHD, depression, OCD, anxiety, smoking and Parkinson’s disease. As illustrated in Figure 1, recent pharmacologic and molecular studies suggest that these transporters can be more widely and more effectively ultilized for medication developments.

Figure 1. Schematic of clinical relevance of monoamine transporters (MATs in blue) or vesicular monoamine transporters (VMATs in brown): regulation (black arrow) of activity and expression by small molecules, proteins and DNA sequence polymorphisms ( see “x”).

Figure 1

Clinical roles of MAT inhibitors are indicated on the left hand side and diseases contributed by genetic variations lisyed on the right side. Green dot, monoamine. Horizontal arrow, MAT gene.

Medication issues

There are several caveats associated with current transporter protein-based medications such as those for depression, OCD and ADHD. First, although proven to effectively treat depression, the use of TCAs and MAOIs has been shown to elicit adverse side effects (343-346) and require high doses to achieve therapeutic effects (14). While comparably effective, more contemporary treatments seem to mollify these issues (e.g., 16). Second, approximately 30% of patients with depression fail to respond to antidepressant drug therapy (e.g., 344). Prospective, longitudinal research has shown that more than 75% of those who experience a first episode will relapse in their lifetimes (e.g., 347). Third, the conflicting body of literature on antidepressant tolerance must also be taken into account. Tolerance can emerge with long-term antidepressant treatment or on retreatment after discontinuation (348). Patients who are treated with antidepressants may be less likely to have a positive response to new antidepressant treatments (18). However, conflicting research has hinted that prior antidepressant treatment may have no effect on new interventions (349, 350) and that discontinuation of maintenance SSRI treatments can result in a far poorer prognosis than continued maintenance (351). Fourth, treatments of OCD with these medications do not display long-term benefits after the medications are terminated (352). Finally, the abuse potential of both AMPH and methylphenidate warrants investigation of novel DAT blockers with low abuse profile to treat ADHD. Similarly, the DAT is likely to become the target for future pharmacotherapies used for the management of several other disorders which feature dopaminergic dysregulation, including Parkinson’s disease, Lesch-Nyhan syndrome, Tourette’s syndrome and obesity (353).

In addition to the limited efficacy, the current medications cover only approximately a third of related diseases. The large body of preclinical neuropharmacologic, animal genetics and human association data suggests that monoamine transporters regulate brain pathways that underlie not only these diseases under transporter protein-based treatments but also twenty others such as addiction, mood disorder, PTSD, stress, hypertension, SCZ, bipolar disorder and anorexia nervosa, etc. (Table 2).

The caveats and limited coverage mirror a fact that the current medications have not benefited much from the molecular studies conducted during the last 15 years, warranting new searches for complementary medications. The new and expanding information from molecular studies suggests that it is very likely for monoamine transporters to serve as more effective medication targets and for more related diseases.

New targets for medications

Substance abuse and depression represent the main focus for future transporter-based medication development. Substance abuse is a worldwide epdemic but the treatment has benefited little from the elucidation that these monoamine transporters are the functional targets for sunstances of abuse. Depression is common and predicted by the World Health Organization (WHO) to be the second most debilitating disease next to heart disease by the year 2020. It is also one of the best studied diseases by different independent disciplines including neuropharmacology, animal genetics and human genetics. Animal genetic data suggested that reduced or eliminated transporter expression causes depressive behaviors, suggesting that medications that up-regulate these transporters’ activity may prove effective in treating this disease. However, the current medications are all based on inhibition of the transporters and such an approach works to certain degree perhaps by inhibiting multiple transporters to balance the neurotransmission of different types at the same time. It remains difficult to boost the transporter activity in the plasma membranes. Understanding the molecular mechanisms could assist in evidence-based medication development.

First, with the help of structural models and SAR information, medicinal chemists are in an exciting position to explore the transport-sparing antagonism and develop more transporter-selective inhibitors. Such new inhibitors could develop into medications for drug abuse among others and would allow testing transporter-specific hypotheses in terms of roles the transporters play in brain function and diseases. Second, we ought to take advantage of the intracellular networks that regulate transporter expression in the plasma membrane. There are more than ten proteins that interact and regulate transporter trafficking and therefore could serve as new targets for medication development. Third, DNA sequence polymorphisms that underlie transporters’ genetic contribution to the diseases provide novel medication targets. These novel medication targets may require the dissection of related signaling cascades in the monoaminergic neurons or can be directly used in cell-based high throughput screening (HTS) for small molecule regulators. Finally, there is paucity of information on translation and translocation efficiency and regulation of stability of the transporter mRNA molecules. mRNA represents an unexplored subject and another medication target. A major advantage of using nucleic acids as medication targets is the increased transporter specificity and target diversity. Therefore, it will not be surprising to see monoamine transporter-based new medication targets to be discovered, tested and utilized in the next decade.

Acknowledgement

This research was supported by Plan Nacional Sobre Drogas (grant PNSD2008-057, Spanish Ministry of Health), and Red de Trastornos Adictivos (grant RD06/0001/0032, RETICS, Instituto de Salud Carlos III) (JJC), the China National High-Tech 863 Programs (2006AA02A312) (HQ), the Intramural Research Program of the NIH, NIDA (QRL), and NIH grants DA021409 (ZL), DA027825 (MMT), and DA016710 (GET).

References

  • 1.Amara SG, Kuhar MJ. Neurotransmitter transporters: recent progress. Annu Rev Neurosci. 1993;16:73–93. doi: 10.1146/annurev.ne.16.030193.000445. [DOI] [PubMed] [Google Scholar]
  • 2.Langer SZ, Galzin AM. Studies on the serotonin transporter in platelets. Experientia. 1988;44:127–30. doi: 10.1007/BF01952194. [DOI] [PubMed] [Google Scholar]
  • 3.Kurian MA, Zhen J, Cheng SY, Li Y, Mordekar SR, Jardine P, et al. Homozygous loss-of-function mutations in the gene encoding the dopamine transporter are associated with infantile parkinsonism-dystonia. J Clin Invest. 2009;119:1595–603. doi: 10.1172/JCI39060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Wang YM, Gainetdinov RR, Fumagalli F, Xu F, Jones SR, Bock CB, et al. Knockout of the vesicular monoamine transporter 2 gene results in neonatal death and supersensitivity to cocaine and amphetamine. Neuron. 1997;19:1285–96. doi: 10.1016/s0896-6273(00)80419-5. [DOI] [PubMed] [Google Scholar]
  • 5.Fon EA, Pothos EN, Sun BC, Killeen N, Sulzer D, Edwards RH. Vesicular transport regulates monoamine storage and release but is not essential for amphetamine action. Neuron. 1997;19:1271–83. doi: 10.1016/s0896-6273(00)80418-3. [DOI] [PubMed] [Google Scholar]
  • 6.Takahashi N, Miner LL, Sora I, Ujike H, Revay RS, Kostic V, et al. VMAT2 knockout mice: heterozygotes display reduced amphetamine-conditioned reward, enhanced amphetamine locomotion, and enhanced MPTP toxicity. Proc Natl Acad Sci USA. 1997;94:9938–43. doi: 10.1073/pnas.94.18.9938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hansson SR, Hoffman BJ, Mezey E. Ontogeny of vesicular monoamine transporter mRNAs VMAT1 and VMAT2. I. The developing rat central nervous system. Brain Res Dev Brain Res. 1998;110:135–58. doi: 10.1016/s0165-3806(98)00104-7. [DOI] [PubMed] [Google Scholar]
  • 8.Gutman DA, Owens MJ. Serotonin and Norepinephrine Transporter Binding Profile of SSRIs. Essential Psychopharmacology. 2006;7:35–41. [PubMed] [Google Scholar]
  • 9.Anderson IM. Selective serotonin reuptake inhibitors versus tricyclic antidepressants: A meta-analysis of efficacy and tolerability. Journal of Affect Disorders. 2000;58:19–36. doi: 10.1016/s0165-0327(99)00092-0. [DOI] [PubMed] [Google Scholar]
  • 10.Dunlop BW, Nemeroff CB. The role of dopamine in the pathophysiology of depression. Archives of General Psychiatry. 2007;64:327–37. doi: 10.1001/archpsyc.64.3.327. [DOI] [PubMed] [Google Scholar]
  • 11.Willner P. Dopaminergic mechanisms in depression and mania. In: Watson, editor. Psychopharmacology: The Fourth Generation of Progress. On-Line edition Lippincott Williams & Wilkins; New York: 2002. [Google Scholar]
  • 12.Shelton RC. The Dual-Action Hypothesis: Does Pharmacology Matter? Journal of Clinical Psychiatry. 2004;65:5–10. [PubMed] [Google Scholar]
  • 13.Zhou Z, Zhen J, Karpowich NK, Goetz RM, Law CJ, Reith MEA, et al. LeuT-desipramine structure reveals how antidepressants block neurotransmitter reuptake. Science. 2007;317:1390–3. doi: 10.1126/science.1147614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Thase ME, Denko T. Pharmacotherapy of mood disorders. Annual Review of Clinical Psychology. 2008;4:53–91. doi: 10.1146/annurev.clinpsy.2.022305.095301. [DOI] [PubMed] [Google Scholar]
  • 15.Delgado PL. How antidepressant help depression: Mechanisms of action and clinical response. The Journal of Clinical Psychiatry. 2004;65:25–30. [PubMed] [Google Scholar]
  • 16.Bauer M, Bschor T, Pfennig A, Whybrow PC, Angst J, Versiani M, et al. World Federation of Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar depressive disorders in primary care. World Journal of Biological Psychiatry. 2007;8:67–104. doi: 10.1080/15622970701227829. [DOI] [PubMed] [Google Scholar]
  • 17.Bauer M, Whybrow PC, Angst J, Versiani M, Moller HJ. World Federation of Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar depressive disorders, Part 1: acute and continuation treatment of major depressive disorder. World Journal of Biological Psychiatry. 2002;3:5–43. doi: 10.3109/15622970209150599. [DOI] [PubMed] [Google Scholar]
  • 18.Amsterdam JD. Selective serotonin reuptake inhibitor in severe and melancholic depression. Special issue: Supplement: Selective serotonin reuptake inhibitors (SSRIs) in depression: A decade of progress. Journal of Psychopharmacology. 1998. pp. S99–S111. [DOI] [PubMed]
  • 19.Whale R, Clifford EM, Bhagwagar Z, Cowen PJ. Decreased sensitivity of 5-HT[sub]ID[/sub] receptors in melancholic depression. British Journal of Psychiatry. 2001;178:454–7. doi: 10.1192/bjp.178.5.454. [DOI] [PubMed] [Google Scholar]
  • 20.Lineberry CG, Johnston JA, Raymond RN, Samara B, et al. A fixed-dose (300 mg) efficacy study of bupropion and placebo in depressed outpatients. Journal of Clinical Psychiatry. 1990;51:194–9. [PubMed] [Google Scholar]
  • 21.Dhillon S, Yang LPH, Curran MP. Spotlight on Bupropion in Major Depressive Disorder. CNS Drugs. 2008;22:613–17. doi: 10.2165/00023210-200822070-00006. [DOI] [PubMed] [Google Scholar]
  • 22.Ananth J. Clomiparmine: An antiobsessive drug. The Canadian Journal of Psychiatry / La Revue canadienne de psychiatrie. 1986;31:253–8. doi: 10.1177/070674378603100314. [DOI] [PubMed] [Google Scholar]
  • 23.Ackerman D, Greenland S. Multivariate meta-analysis of controlled drug studies for obsessive-compulsive disorder. Journal of Clinical Psychopharmacology. 2002;22:309–17. doi: 10.1097/00004714-200206000-00012. [DOI] [PubMed] [Google Scholar]
  • 24.Fineberg NA, Pampaloni I, Pallanti S, Ipser J, Stein DJ. Sustained response versus relapse: The pharmacotherapeutic goal for obessisive-compulsive disorder. International Clinical Psychopharmacology. 2007;22:313–22. doi: 10.1097/YIC.0b013e32825ea312. [DOI] [PubMed] [Google Scholar]
  • 25.Gellar DA, Biederman J, Stewart SE, Mullin B, Martin A, Spencer T, et al. Which SSRI? A Meta-Analysis of Pharmacotherapy Trials in Pediatric Obsessive-Compulsive Disorder. The American Journal of Psychiatry. 2003;160:1919–1928. doi: 10.1176/appi.ajp.160.11.1919. [DOI] [PubMed] [Google Scholar]
  • 26.Klein DF. Delineation of two drug-responsive anxiety syndromes. Psychopharmacologia. 1964;5:397–408. doi: 10.1007/BF02193476. [DOI] [PubMed] [Google Scholar]
  • 27.Benítez CIP, Smith K, Vasile RG, Rende R, Edelen MO, Keller MB. Use of benzodiazepines and selective serotonin reuptake inhibitors in middle-aged and older adults with anxiety disorders: A longitudinal and prospective study. Special issue: Clinical studies for late-life disorders. The American Journal of Geriatric Psychiatry. 2008. pp. 5–13. [DOI] [PubMed]
  • 28.Bruce SE, Vasile RG, Goisman RM, Salzman C, Spencer M, Machan JT, et al. Are benzodiazepines still the medication for patients with panic disorder with or without agoraphobia? The American Journal of Psychiatry. 2003;160:1432–8. doi: 10.1176/appi.ajp.160.8.1432. [DOI] [PubMed] [Google Scholar]
  • 29.van der Linden GJH, Stein DJ, van Balkom AJLM. The efficacy of the selective serotonin reuptake inhibitors for social anxiety disorders (social phobia): A meta-analysis of randomized controlled trials. Special issue: Selective serotonin reuptake inhibitors in the anxiety disorders. International Clinical Psychopharmacology. 2002. pp. S15–S23. [DOI] [PubMed]
  • 30.Anderson IM, Palm ME. Pharmacological Treatments for Worry: Focus on Generalized Anxiety Disorder. In: Davey, Graham CL, Wells Adrian, editors. Worry and its psychological disorders: Theory, assessment and treatment. US; Hoboken, NJ: 2006. pp. 305–334. [Google Scholar]
  • 31.Davis LL, Frazier EC, Williford RB, Newell JM. Long-Term Pharmacotherapy for Post-Traumatic Stress Disorder. CNS Drugs. 2006;20:465–76. doi: 10.2165/00023210-200620060-00003. [DOI] [PubMed] [Google Scholar]
  • 32.Goldenberg DL. Update on the treatment of fibromyalgia. Bulletin of Rheumatoid Disorders. 2004;53:1–7. [Google Scholar]
  • 33.Staud R. Pharmacological treatment of fibromyalgia syndrome: new developments. Drugs. 2010;70:1–14. doi: 10.2165/11530950-000000000-00000. [DOI] [PubMed] [Google Scholar]
  • 34.Anderberg UM, Marteinsdottir I, Von-Knorring L. Citalopram in patients with fibromyalgia: a randomized, double-blind, placebo controlled study. European Journal of Pain. 2000;4:27–35. doi: 10.1053/eujp.1999.0148. [DOI] [PubMed] [Google Scholar]
  • 35.Evren B, Evren C, Guler MH. An open clinical trial of venlafaxine in the treatment of pain, depressive and anxiety symptoms in fibromyalgia. The Pain Clinic. 2006;18:167–73. [Google Scholar]
  • 36.Biederman J, Spencer TJ, Newcorn JH, Gao H, Milton DR, Feldman P, et al. Effect of comorbid symptoms of oppositional defiant disorder on responses to atomoxetine in children with ADHD: a meta-analysis of controlled clinical trial data. Psychopharmacology. 2007;190:31–41. doi: 10.1007/s00213-006-0565-2. [DOI] [PubMed] [Google Scholar]
  • 37.Caballero J, Nahata MC. Atomoxetine Hydrochloride for the treatment of Attention-Deficit/Hyperactivity Disorder. Clinical Therapeutics: The International Peer-Reviewed. Journal of Drug Therapy. 2003;25:3065–83. doi: 10.1016/s0149-2918(03)90092-0. [DOI] [PubMed] [Google Scholar]
  • 38.Bhatara VS, Aparasu RR. Pharmacotherapy with atomoxetine for US children and adolescents. Annals of Clinical Psychiatry. 2007;19:175–80. doi: 10.1080/10401230701465244. [DOI] [PubMed] [Google Scholar]
  • 39.Mazei-Robison MS, Couch RS, Shelton RC, Stein MA, Blakely RD. Sequence variation in the human dopamine transporter gene in children with attention deficit hyperactivity disorder. Neuropharmacology. 2005;49:724–36. doi: 10.1016/j.neuropharm.2005.08.003. [DOI] [PubMed] [Google Scholar]
  • 40.Balfour DJK, Ridley DL. The effects of nicotine on neural pathways in depression: A factor in nicotine addiction? Pharmacology, Biochemistry and Behavior. 2000;66:79–85. doi: 10.1016/s0091-3057(00)00205-7. [DOI] [PubMed] [Google Scholar]
  • 41.Hall SM, Humfleet GL, Reus VI, Muñoz RF, Hartz DT, Maude-Griffin R. Psychological intervention and antidepressant treatment in smoking cessation. Archives of General Psychiatry. 2002;59:929–37. doi: 10.1001/archpsyc.59.10.930. [DOI] [PubMed] [Google Scholar]
  • 42.Kinnunen T, Henning L, Nordstrom BL. Smoking cessation in individuals with depression: Recommendations for treatment. CNS Drugs. 1999;11:93–103. [Google Scholar]
  • 43.Iversen SD, Iversen LL. Dopamine: 50 years in perspective. Trends Neurosci. 2007;30:188–93. doi: 10.1016/j.tins.2007.03.002. [DOI] [PubMed] [Google Scholar]
  • 44.Ikemoto S, Panksepp J. The role of nucleus accumbens dopamine in motivated behavior: a unifying interpretation with special reference to reward-seeking. Brain Res Brain Res Rev. 1999;31:6–41. doi: 10.1016/s0165-0173(99)00023-5. [DOI] [PubMed] [Google Scholar]
  • 45.Pennartz CM, Groenewegen HJ, Lopes da Silva FH. The nucleus accumbens as a complex of functionally distinct neuronal ensembles: an integration of behavioural, electrophysiological and anatomical data. Prog Neurobiol. 1994;42:719–761. doi: 10.1016/0301-0082(94)90025-6. [DOI] [PubMed] [Google Scholar]
  • 46.Groenewegen HJ. The basal ganglia and motor control. Neural Plast. 2003;10:107–20. doi: 10.1155/NP.2003.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Graybiel AM. Habits, rituals, and the evaluative brain. Annu Rev Neurosci. 2008;31:359–87. doi: 10.1146/annurev.neuro.29.051605.112851. [DOI] [PubMed] [Google Scholar]
  • 48.Seamans JK, Yang CR. The principal features and mechanisms of dopamine modulation in the prefrontal cortex. Prog Neurobiol. 2004;74:1–58. doi: 10.1016/j.pneurobio.2004.05.006. [DOI] [PubMed] [Google Scholar]
  • 49.Ritz MC, Lamb RJ, Goldberg SR, Kuhar MJ. Cocaine receptors on dopamine transporters are related to self-administration of cocaine. Science. 1987;237:1219–23. doi: 10.1126/science.2820058. [DOI] [PubMed] [Google Scholar]
  • 50.Venton BJ, Seipel AT, Phillips PE, Wetsel WC, Gitler D, Greengard P, et al. Cocaine increases dopamine release by mobilization of a synapsin-dependent reserve pool. J Neurosci. 2006;26:3206–9. doi: 10.1523/JNEUROSCI.4901-04.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Volkow ND, Wang GJ, Fischman MW, Foltin R, Fowler JS, Franceschi D, et al. Effects of route of administration on cocaine induced dopamine transporter blockade in the human brain. Life Sci. 67:1507–15. doi: 10.1016/s0024-3205(00)00731-1. 200. [DOI] [PubMed] [Google Scholar]
  • 52.Chen N, Zhen J, Reith ME. Mutation of Trp84 and Asp313 of the dopamine transporter reveals similar mode of binding interaction for GBR12909 and benztropine as opposed to cocaine. J Neurochem. 2004;89:853–64. doi: 10.1111/j.1471-4159.2004.02386.x. [DOI] [PubMed] [Google Scholar]
  • 53.Dar DE, Mayo C, Uhl GR. The interaction of methylphenidate and benztropine with the dopamine transporter is different than other substrates and ligands. Biochem Pharmacol. 2005;70:461–9. doi: 10.1016/j.bcp.2005.04.032. [DOI] [PubMed] [Google Scholar]
  • 54.Loland CJ, Desai RI, Zou MF, Cao J, Grundt P, Gerstbrein K, et al. Relationship between conformational changes in the dopamine transporter and cocaine-like subjective effects of uptake inhibitors. Mol Pharmacol. 2008;73:813–23. doi: 10.1124/mol.107.039800. [DOI] [PubMed] [Google Scholar]
  • 55.Rothman RB, Baumann MH, Prisinzano TE, Newman AH. Dopamine transport inhibitors based on GBR12909 and benztropine as potential medications to treat cocaine addiction. Biochem Pharmacol. 2008;75:2–16. doi: 10.1016/j.bcp.2007.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Tanda G, Newman AH, Katz JL. Discovery of drugs to treat cocaine dependence: behavioral and neurochemical effects of atypical dopamine transport inhibitors. Adv Pharmacol. 2009;57:253–89. doi: 10.1016/S1054-3589(08)57007-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Meltzer PC, Kryatova O, Pham-Huu DP, Donovan P, Janowsky A. The synthesis of bivalent 2beta-carbomethoxy-3beta-(3,4-dichlorophenyl)-8-heterobicyclo[3.2.1]octane s as probes for proximal binding sites on the dopamine and serotonin transporters. Bioorg Med Chem. 2008;16:1832–1841. doi: 10.1016/j.bmc.2007.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Ferragud A, Velazquez-Sanchez C, Hernandez-Rabaza V, Nacher A, Merino V, Carda M, et al. A dopamine transport inhibitor with markedly low abuse liability suppresses cocaine self-administration in the rat. Psychopharmacology (Berl) 2009;207:281–289. doi: 10.1007/s00213-009-1653-x. [DOI] [PubMed] [Google Scholar]
  • 59.Velazquez-Sanchez C, Ferragud A, Hernandez-Rabaza V, Nacher A, Merino V, Carda M, et al. The dopamine uptake inhibitor 3 alpha-[bis(4′-fluorophenyl)metoxy]-tropane reduces cocaine-induced early-gene expression, locomotor activity, and conditioned reward. Neuropsychopharmacology. 2009;34:2497–507. doi: 10.1038/npp.2009.78. [DOI] [PubMed] [Google Scholar]
  • 60.Velazquez-Sanchez C, Ferragud A, Murga J, Carda M, Canales JJ. The high affinity dopamine uptake inhibitor, JHW 007, blocks cocaine-induced reward, locomotor stimulation and sensitization. Eur. Neuropsychopharmacol. 2010;20:501–8. doi: 10.1016/j.euroneuro.2010.03.005. [DOI] [PubMed] [Google Scholar]
  • 61.Moore RY, Halaris AE, Jones BE. Serotonin neurons of the midbrain raphe: ascending projections. J Comp Neurol. 1978;180:417–438. doi: 10.1002/cne.901800302. [DOI] [PubMed] [Google Scholar]
  • 62.Green AR. Neuropharmacology of 5-hydroxytryptamine. Br J Pharmacol. 2006;147:S145–S152. doi: 10.1038/sj.bjp.0706427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Fantegrossi WE, Murnane KS, Reissig CJ. The behavioral pharmacology of hallucinogens. Biochem Pharmacol. 2008;75:17–33. doi: 10.1016/j.bcp.2007.07.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Howell LL, Kimmel HL. Monoamine transporters and psychostimulant addiction. Biochem Pharmacol. 2008;75:196–217. doi: 10.1016/j.bcp.2007.08.003. [DOI] [PubMed] [Google Scholar]
  • 65.Papakostas GI. Tolerability of modern antidepressants. J Clin Psychiatry. 2008;69:8–13. [PubMed] [Google Scholar]
  • 66.Cryan JF, Valentino RJ, Lucki I. Assessing substrates underlying the behavioral effects of antidepressants using the modified rat forced swimming test. Neurosci Biobehav Rev. 2005;29:547–69. doi: 10.1016/j.neubiorev.2005.03.008. [DOI] [PubMed] [Google Scholar]
  • 67.Petty F, Davis LL, Kabel D, Kramer GL. Serotonin dysfunction disorders: a behavioral neurochemistry perspective. J Clin Psychiatry. 1996;57:11–16. [PubMed] [Google Scholar]
  • 68.Skolnick P, Popik P, Janowsky A, Beer B, Lippa AS. Antidepressant-like actions of DOV 21,947: a “triple” reuptake inhibitor. Eur J Pharmacol. 2003;461:99–104. doi: 10.1016/s0014-2999(03)01310-4. [DOI] [PubMed] [Google Scholar]
  • 69.Popik P, Krawczyk M, Golembiowska K, Nowak G, Janowsky A, Skolnick P, et al. Pharmacological profile of the “triple” monoamine neurotransmitter uptake inhibitor, DOV 102,677. Cell Mol Neurobiol. 2006;26:857–73. doi: 10.1007/s10571-006-9012-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Chen Z, Yang J, Tobak A. Designing new treatments for depression and anxiety. IDrugs. 2008;11:189–97. [PubMed] [Google Scholar]
  • 71.Daws LC. Unfaithful neurotransmitter transporters: focus on serotonin uptake and implications for antidepressant efficacy. Pharmacol Ther. 2009;121:89–99. doi: 10.1016/j.pharmthera.2008.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Spealman RD. Noradrenergic involvement in the discriminative stimulus effects of cocaine in squirrel monkeys. J Pharmacol Exp Ther. 1995;275:53–62. [PubMed] [Google Scholar]
  • 73.Platt DM, Rowlett JK, Spealman RD. Noradrenergic mechanisms in cocaine-induced reinstatement of drug seeking in squirrel monkeys. J Pharmacol Exp Ther. 2007;322:894–902. doi: 10.1124/jpet.107.121806. [DOI] [PubMed] [Google Scholar]
  • 74.Sofuoglu M, Poling J, Hill K, Kosten T. Atomoxetine attenuates dextroamphetamine effects in humans. Am J Drug Alcohol Abuse. 2009;35:412–416. doi: 10.3109/00952990903383961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Trigo JM, Renoir T, Lanfumey L, Hamon M, Lesch KP, Robledo P, et al. 3,4-methylenedioxymethamphetamine self-administration is abolished in serotonin transporter knockout mice. Biol. Psychiatry. 2007;62:669–79. doi: 10.1016/j.biopsych.2006.11.005. [DOI] [PubMed] [Google Scholar]
  • 76.Liechti ME, Baumann C, Gamma A, Vollenweider FX. Acute psychological effects of 3,4-methylenedioxymethamphetamine (MDMA, “Ecstasy”) are attenuated by the serotonin uptake inhibitor citalopram. Neuropsychopharmacology. 2000;22:513–21. doi: 10.1016/S0893-133X(99)00148-7. [DOI] [PubMed] [Google Scholar]
  • 77.Moore RY, Bloom FE. Central catecholamine neuron systems: anatomy and physiology of the norepinephrine and epinephrine systems. Annu Rev Neurosci. 1979;2:113–68. doi: 10.1146/annurev.ne.02.030179.000553. [DOI] [PubMed] [Google Scholar]
  • 78.Berridge CW, Waterhouse BD. The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res Brain Res Rev. 2003;42:33–84. doi: 10.1016/s0165-0173(03)00143-7. [DOI] [PubMed] [Google Scholar]
  • 79.Ressler KJ, Nemeroff CB. Role of norepinephrine in the pathophysiology and treatment of mood disorders. Biol Psychiatry. 1999;46:1219–33. doi: 10.1016/s0006-3223(99)00127-4. [DOI] [PubMed] [Google Scholar]
  • 80.Tsigos C, Chrousos GP. Hypothalamic-pituitary-adrenal axis, neuroendocrine factors and stress. J Psychosom Res. 2002;53:865–71. doi: 10.1016/s0022-3999(02)00429-4. [DOI] [PubMed] [Google Scholar]
  • 81.Chrousos GP. Stress and disorders of the stress system. Nat Rev Endocrinol. 2009;5:374–81. doi: 10.1038/nrendo.2009.106. [DOI] [PubMed] [Google Scholar]
  • 82.Dunn AJ, Swiergiel AH. The role of corticotropin-releasing factor and noradrenaline in stress-related responses, and the inter-relationships between the two systems. Eur J Pharmacol. 2008;583:186–93. doi: 10.1016/j.ejphar.2007.11.069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Andersen J, Kristensen AS, Bang-Andersen B, Stromgaard K. Recent advances in the understanding of the interaction of antidepressant drugs with serotonin and norepinephrine transporters. Chem Commun (Camb) 2009;7:3677–92. doi: 10.1039/b903035m. [DOI] [PubMed] [Google Scholar]
  • 84.Harkin A, Kelly JP, McNamara M, Connor TJ, Dredge K, Redmond A, Leonard BE. Activity and onset of action of reboxetine and effect of combination with sertraline in an animal model of depression. Eur J Pharmacol. 1999;364:123–32. doi: 10.1016/s0014-2999(98)00838-3. [DOI] [PubMed] [Google Scholar]
  • 85.Connor TJ, Kelliher P, Harkin A, Kelly JP, Leonard BE. Reboxetine attenuates forced swim test-induced behavioural and neurochemical alterations in the rat. Eur. J. Pharmacol. 1999;379:125–33. doi: 10.1016/s0014-2999(99)00492-6. [DOI] [PubMed] [Google Scholar]
  • 86.Cryan JF, Page ME, Lucki I. Noradrenergic lesions differentially alter the antidepressant-like effects of reboxetine in a modified forced swim test. Eur J Pharmacol. 2002;436:197–205. doi: 10.1016/s0014-2999(01)01628-4. [DOI] [PubMed] [Google Scholar]
  • 87.Rygula R, Abumaria N, Havemann-Reinecke U, Ruther E, Hiemke C, Zernig G, et al. Pharmacological validation of a chronic social stress model of depression in rats: effects of reboxetine, haloperidol and diazepam. Behav Pharmacol. 2008;19:183–196. doi: 10.1097/FBP.0b013e3282fe8871. (2008) [DOI] [PubMed] [Google Scholar]
  • 88.Hahn MK, Steele A, Couch RS, Stein MA, Krueger JJ. Novel and functional norepinephrine transporter protein variants identified in attention-deficit hyperactivity disorder. Neuropharmacology. 2009;57:694–701. doi: 10.1016/j.neuropharm.2009.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Liu YP, Lin YL, Chuang CH, Kao YC, Chang ST, Tung CS. Alpha adrenergic modulation on effects of norepinephrine transporter inhibitor reboxetine in five-choice serial reaction time task. J Biomed Sci. 2009;16:72. doi: 10.1186/1423-0127-16-72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Robinson ES, Eagle DM, Mar AC, Bari A, Banerjee G, Jiang X, et al. Similar effects of the selective noradrenaline reuptake inhibitor atomoxetine on three distinct forms of impulsivity in the rat. Neuropsychopharmacology. 2008;33:1028–37. doi: 10.1038/sj.npp.1301487. [DOI] [PubMed] [Google Scholar]
  • 91.Liu LL, Yang J, Lei GF, Wang GJ, Wang YW, Sun RP. Atomoxetine increases histamine release and improves learning deficits in an animal model of attention-deficit hyperactivity disorder: the spontaneously hypertensive rat. Basic Clin Pharmacol Toxicol. 2008;102:527–32. doi: 10.1111/j.1742-7843.2008.00230.x. [DOI] [PubMed] [Google Scholar]
  • 92.Vaughan B, Fegert J, Kratochvil CJ. Update on atomoxetine in the treatment of attention-deficit/hyperactivity disorder. Expert Opin Pharmacother. 2009;10:669–76. doi: 10.1517/14656560902762873. [DOI] [PubMed] [Google Scholar]
  • 93.Brown JM, Hanson GR, Fleckenstein AE. Regulation of the vesicular monoamine transporter-2: a novel mechanism for cocaine and other psychostimulants. J. Pharmacol. Exp. Ther. 2001;296:762–7. [PubMed] [Google Scholar]
  • 94.Riddle EL, Topham MK, Haycock JW, Hanson GR, Fleckenstein AE. Differential trafficking of the vesicular monoamine transporter-2 by methamphetamine and cocaine. Eur J Pharmacol. 2002;449:71–4. doi: 10.1016/s0014-2999(02)01985-4. [DOI] [PubMed] [Google Scholar]
  • 95.Farnsworth SJ, Volz TJ, Hanson GR, Fleckenstein AE. Cocaine alters vesicular dopamine sequestration and potassium-stimulated dopamine release: the role of D2 receptor activation. J Pharmacol Exp Ther. 2009;328:807–812. doi: 10.1124/jpet.108.146159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Sulzer D, Chen TK, Lau YY, Kristensen H, Rayport S, Ewing A. Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. J Neurosci. 1995;15:4102–8. doi: 10.1523/JNEUROSCI.15-05-04102.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Hansen JP, Riddle EL, Sandoval V, Brown JM, Gibb JW, Hanson GR, et al. Methylenedioxymethamphetamine decreases plasmalemmal and vesicular dopamine transport: mechanisms and implications for neurotoxicity. J Pharmacol Exp Ther. 2002;300:1093–100. doi: 10.1124/jpet.300.3.1093. [DOI] [PubMed] [Google Scholar]
  • 98.Bortolato M, Chen K, Shih JC. Monoamine oxidase inactivation: from pathophysiology to therapeutics. Adv Drug Deliv Rev. 2008;60:1527–33. doi: 10.1016/j.addr.2008.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Wimalasena K. Vesicular monoamine transporters: Structure-function, pharmacology, and medicinal chemistry. Med Res Rev. 2010 doi: 10.1002/med.20187. DOI: 10.1002/med.20187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Yamashita A, Singh SK, Kawate T, Jin Y, Gouaux E. Crystal structure of a bacterial homologue of Na+/Cl--dependent neurotransmitter transporters. Nature. 2005;437:215–23. doi: 10.1038/nature03978. [DOI] [PubMed] [Google Scholar]
  • 101.Guastella J, Nelson N, Nelson H, Czyzyk L, Keynan S, Miedel MC, et al. Cloning and expression of a rat brain GABA transporter. Science. 1990;249:1303–6. doi: 10.1126/science.1975955. [DOI] [PubMed] [Google Scholar]
  • 102.Nelson H, Mandiyan S, Nelson N. Cloning of the human brain GABA transporter. FEBS Lett. 1990;269:181–4. doi: 10.1016/0014-5793(90)81149-i. [DOI] [PubMed] [Google Scholar]
  • 103.Pacholczyk T, Blakely RD, Amara SG. Expression cloning of a cocaine- and antidepressant-sensitive human noradrenaline transporter. Nature. 1991;350:350–4. doi: 10.1038/350350a0. [DOI] [PubMed] [Google Scholar]
  • 104.Kilty JE, Lorang D, Amara SG. Cloning and expression of a cocaine-sensitive rat dopamine transporter. Science. 1991;254:578–9. doi: 10.1126/science.1948035. [DOI] [PubMed] [Google Scholar]
  • 105.Shimada S, Kitayama S, Lin CL, Patel A, Nanthakumar E, Gregor P, et al. Cloning and expression of a cocaine-sensitive dopamine transporter complementary DNA. Science. 1991;254:576–8. doi: 10.1126/science.1948034. [DOI] [PubMed] [Google Scholar]
  • 106.Hoffman BJ, Mezey E, Brownstein MJ. Cloning of a serotonin transporter affected by antidepressants. Science. 1991;254:579–80. doi: 10.1126/science.1948036. [DOI] [PubMed] [Google Scholar]
  • 107.Erickson JD, Eiden LE, Hoffman BJ. Expression cloning of a reserpine-sensitive vesicular monoamine transporter. Proc Natl Acad Sci USA. 1992;89:10993–7. doi: 10.1073/pnas.89.22.10993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Liu Y, Peter D, Roghani A, Schuldiner S, Privé GG, Eisenberg D, et al. A cDNA that suppresses MPP+ toxicity encodes a vesicular amine transporter. Cell. 1992;70:539–51. doi: 10.1016/0092-8674(92)90425-c. [DOI] [PubMed] [Google Scholar]
  • 109.Sarker S, Weissensteiner R, Steiner I, Sitte HH, Ecker GF, Freissmuth M, Sucic S. The high-affinity binding site for tricyclic antidepressants resides in the outer vestibule of the serotonin transporter. Mol Pharmacol. 2010 Sep 9; doi: 10.1124/mol.110.067538. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Wang CI, Lewis RJ. Emerging structure-function relationships defining monoamine NSS transporter substrate and ligand affinity. Biochem Pharmacol. 2010;79:1083–91. doi: 10.1016/j.bcp.2009.11.019. [DOI] [PubMed] [Google Scholar]
  • 111.Celik L, Sinning S, Severinsen K, Hansen CG, Møller MS, Bols M, et al. Binding of serotonin to the human serotonin transporter. Molecular modeling and experimental validation. J Am Chem Soc. 2008;130:3853–65. doi: 10.1021/ja076403h. [DOI] [PubMed] [Google Scholar]
  • 112.Ravna AW, Jaronczyk M, Sylte I. A homology model of SERT based on the LeuTAa template. Bioorg Med Chem Lett. 2006;16:5594–7. doi: 10.1016/j.bmcl.2006.08.028. [DOI] [PubMed] [Google Scholar]
  • 113.Zhou Z, Zhen J, Karpowich NK, Goetz RM, Law CJ, Reith ME, et al. LeuT-desipramine structure reveals how antidepressants block neurotransmitter reuptake. Science. 2007;317:1390–3. doi: 10.1126/science.1147614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Kaufmann KW, Dawson ES, Henry LK, Field JR, Blakely RD, Meiler J. Structural determinants of species-selective substrate recognition in human and Drosophila serotonin transporters revealed through computational docking studies. Proteins. 2009;74:630–42. doi: 10.1002/prot.22178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Indarte M, Madura JD, Surratt CK. Dopamine transporter comparative molecular modeling and binding site prediction using the LeuT(Aa) leucine transporter as a template. Proteins. 2008;70:1033–46. doi: 10.1002/prot.21598. [DOI] [PubMed] [Google Scholar]
  • 116.Gedeon PC, Indarte M, Surratt CK, Madura JD. Molecular dynamics of leucine and dopamine transporter proteins in a model cell membrane lipid bilayer. Proteins. 2010;78:797–811. doi: 10.1002/prot.22601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Paczkowski FA, Sharpe IA, Dutertre S, Lewis RJ. chi-Conotoxin and tricyclic antidepressant interactions at the norepinephrine transporter define a new transporter model. J Biol Chem. 2007;282:17837–44. doi: 10.1074/jbc.M610813200. [DOI] [PubMed] [Google Scholar]
  • 118.Henry LK, Adkins EM, Han Q, Blakely RD. Serotonin and cocaine-sensitive inactivation of human serotonin transporters by methanethiosulfonates targeted to transmembrane domain I. J Biol Chem. 2003;278:37052–63. doi: 10.1074/jbc.M305514200. [DOI] [PubMed] [Google Scholar]
  • 119.Henry LK, Field JR, Adkins EM, Parnas ML, Vaughan RA, Zou MF, Newman AH, Blakely RD. Tyr-95 and Ile-172 in transmembrane segments 1 and 3 of human serotonin transporters interact to establish high affinity recognition of antidepressants. J Biol Chem. 2006;281:2012–23. doi: 10.1074/jbc.M505055200. [DOI] [PubMed] [Google Scholar]
  • 120.Field JR, Henry LK, Blakely RD. Transmembrane domain 6 of the human serotonin transporter contributes to an aqueously accessible binding pocket for serotonin and the psychostimulant 3,4-methylene dioxymethamphetamine. J Biol Chem. 2010;285:11270–80. doi: 10.1074/jbc.M109.093658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Chen JG, Sachpatzidis A, Rudnick G. The third transmembrane domain of the serotonin transporter contains residues associated with substrate and cocaine binding. J Biol Chem. 1997;272:28321–7. doi: 10.1074/jbc.272.45.28321. [DOI] [PubMed] [Google Scholar]
  • 122.Zomot E, Bendahan A, Quick M, Zhao Y, Javitch JA, Kanner BI. Mechanism of chloride interaction with neurotransmitter:sodium symporters. Nature. 2007;449:726–30. doi: 10.1038/nature06133. [DOI] [PubMed] [Google Scholar]
  • 123.Forrest LR, Tavoulari S, Zhang YW, Rudnick G, Honig B. Identification of a chloride ion binding site in Na+/Cl -dependent transporters. Proc Natl Acad Sci USA. 2007;104:12761–6. doi: 10.1073/pnas.0705600104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Tavoulari S, Forrest LR, Rudnick G. Fluoxetine (Prozac) binding to serotonin transporter is modulated by chloride and conformational changes. J Neurosci. 2009;29:9635–43. doi: 10.1523/JNEUROSCI.0440-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Erreger K, Grewer C, Javitch JA, Galli A. Currents in response to rapid concentration jumps of amphetamine uncover novel aspects of human dopamine transporter function. J Neurosci. 2008;28:976–89. doi: 10.1523/JNEUROSCI.2796-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Milner HE, Beliveau R, Jarvis SM. The in situ size of the dopamine transporter is a tetramer as estimated by radiation inactivation. Biochim Biophys Acta. 1994;1190:185–7. doi: 10.1016/0005-2736(94)90051-5. [DOI] [PubMed] [Google Scholar]
  • 127.Schmid JA, Scholze P, Kudlacek O, Freissmuth M, Singer EA, Sitte HH. Oligomerization of the human serotonin transporter and of the rat GABA transporter 1 visualized by fluorescence resonance energy transfer microscopy in living cells. J Biol Chem. 2001;276:3805–10. doi: 10.1074/jbc.M007357200. [DOI] [PubMed] [Google Scholar]
  • 128.Hastrup H, Karlin A, Javitch JA. Symmetrical dimer of the human dopamine transporter revealed by cross-linking Cys-306 at the extracellular end of the sixth transmembrane segment. Proc Natl Acad Sci USA. 2001;98:10055–60. doi: 10.1073/pnas.181344298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Torres GE, Carneiro A, Seamans K, Fiorentini C, Sweeney A, Yao WD, et al. Oligomerization and trafficking of the human dopamine transporter. Mutational analysis identifies critical domains important for the functional expression of the transporter. J Biol Chem. 2003;278:2731–9. doi: 10.1074/jbc.M201926200. [DOI] [PubMed] [Google Scholar]
  • 130.Hastrup H, Sen N, Javitch JA. The human dopamine transporter forms a tetramer in the plasma membrane: cross-linking of a cysteine in the fourth transmembrane segment is sensitive to cocaine analogs. J Biol Chem. 2003;278:45045–8. doi: 10.1074/jbc.C300349200. [DOI] [PubMed] [Google Scholar]
  • 131.Sorkina T, Doolen S, Galperin E, Zahniser NR, Sorkin A. Oligomerization of dopamine transporters visualized in living cells by fluorescence resonance energy transfer microscopy. J Biol Chem. 2003;278:28274–83. doi: 10.1074/jbc.M210652200. [DOI] [PubMed] [Google Scholar]
  • 132.Sitte HH, Farhan H, Javitch JA. Sodium-dependent neurotransmitter transporters: oligomerization as a determinant of transporter function and trafficking. Mol Interv. 2004;4:38–47. doi: 10.1124/mi.4.1.38. [DOI] [PubMed] [Google Scholar]
  • 133.Seidel S, Singer EA, Just H, Farhan H, Scholze P, Kudlacek O, Holy M, Koppatz K, Krivanek P, Freissmuth M, Sitte HH. Amphetamines take two to tango: an oligomer-based counter-transport model of neurotransmitter transport explores the amphetamine action. Mol Pharmacol. 2005;67:140–51. doi: 10.1124/mol.67.1.. [DOI] [PubMed] [Google Scholar]
  • 134.Miner LH, Jedema HP, Moore FW, Blakely RD, Grace AA, Sesack SR. Chronic stress increases the plasmalemmal distribution of the norepinephrine transporter and the coexpression of tyrosine hydroxylase in norepinephrine axons in the prefrontal cortex. J Neurosci. 2006;26:1571–78. doi: 10.1523/JNEUROSCI.4450-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Nirenberg MJ, Vaughan RA, Uhl GR, Kuhar MJ, Pickel VM. The dopamine transporter is localized to dendritic and axonal plasma membranes of nigrostriatal dopaminergic neurons. J Neurosci. 1996;16:436–47. doi: 10.1523/JNEUROSCI.16-02-00436.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Zhou FC, Tao-Cheng JH, Segu L, Patel T, Wang Y. Serotonin transporters are located on the axons beyond the synaptic junctions: anatomical and functional evidence. Brain Res. 1998;805:241–254. doi: 10.1016/s0006-8993(98)00691-x. [DOI] [PubMed] [Google Scholar]
  • 137.Chen BT, Rice ME. Novel Ca2+ dependence and time course of somatodendritic dopamine release: substantia nigra versus striatum. J Neurosci. 2001;21:7841–7. doi: 10.1523/JNEUROSCI.21-19-07841.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Beckstead MJ, Grandy DK, Wickman K, Williams JT. Vesicular dopamine release elicits an inhibitory postsynaptic current in midbrain dopamine neurons. Neuron. 2004;42:939–46. doi: 10.1016/j.neuron.2004.05.019. [DOI] [PubMed] [Google Scholar]
  • 139.de Kock CP, Cornelisse LN, Burnashev N, Lodder JC, Timmerman AJ, Couey JJ, et al. NMDA receptors trigger neurosecretion of 5-HT within dorsal raphe nucleus of the rat in the absence of action potential firing. J Physiol. 2006;577:891–905. doi: 10.1113/jphysiol.2006.115311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Zahniser NR, Sorkin A. Trafficking of dopamine transporters in psychostimulant actions. Semin Cell Dev Biol. 2009;20:411–7. doi: 10.1016/j.semcdb.2009.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Daws LC, Callaghan PD, Moron JA, Kahlig KM, Shippenberg TS, Javitch JA, et al. Cocaine increases dopamine uptake and cell surface expression of dopamine transporters. Biochem Biophys Res Comm. 2002;290:1545–50. doi: 10.1006/bbrc.2002.6384. [DOI] [PubMed] [Google Scholar]
  • 142.Little KY, Elmer LW, Zhong H, Scheys JO, Zhang L. Cocaine induction of dopamine transporter trafficking to the plasma membrane. Mol Pharmacol. 2002;61:436–45. doi: 10.1124/mol.61.2.436. [DOI] [PubMed] [Google Scholar]
  • 143.Sulzer D, Sonders MS, Poulsen NW, Galli A. Mechanisms of neurotransmitter release by amphetamines: a review. Prog Neurobiol. 2005;75:406–33. doi: 10.1016/j.pneurobio.2005.04.003. [DOI] [PubMed] [Google Scholar]
  • 144.Fleckenstein AE, Volz TJ, Riddle EL, Gibb JW, Hanson GR. New insights into the mechanism of action of amphetamines. Annu Rev Pharmacol Toxicol. 2007;47:681–98. doi: 10.1146/annurev.pharmtox.47.120505.105140. [DOI] [PubMed] [Google Scholar]
  • 145.Khoshbouei H, Sen N, Guptaroy B, Johnson L, Lund D, Gnegy ME, et al. N-terminal phosphorylation of the dopamine transporter is required for amphetamine-induced efflux. Plos Biol. 2004;2:387–93. doi: 10.1371/journal.pbio.0020078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Johnson LA, Guptaroy B, Lund D, Shamban S, Gnegy ME. Regulation of amphetamine-stimulated dopamine efflux by protein kinase C {beta} J Biol Chem. 2005;280:10914–9. doi: 10.1074/jbc.M413887200. [DOI] [PubMed] [Google Scholar]
  • 147.Fog JU, Khoshbouei H, Holy M, Owens WA, Vaegter CB, Sen N, et al. Calmodulin kinase II interacts with the dopamine transporter C terminus to regulate amphetamine-induced reverse transport. Neuron. 2006;51:417–29. doi: 10.1016/j.neuron.2006.06.028. [DOI] [PubMed] [Google Scholar]
  • 148.Saunders C, Ferrer JV, Shi L, Chen J, Merrill G, Lamb ME, et al. Amphetamine-induced loss of human dopamine transporter activity: an internalization-dependent and cocaine-sensitive mechanism. Proc. Natl. Acad. Sci. USA. 2000;97:6850–5. doi: 10.1073/pnas.110035297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Furman CA, Chen R, Guptaroy B, Zhang M, Holz RW, Gnegy M. Dopamine and amphetamine rapidly increase dopamine transporter trafficking to the surface: live-cell imaging using total internal reflection fluorescence microscopy. J Neurosci. 2009;29:3328–36. doi: 10.1523/JNEUROSCI.5386-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Zahniser N, Doolen S. Chronic and acute regulation of Na+/Cl− dependent neurotransmitter transporters: drugs, substrates, presynaptic receptors, and signaling systems. Pharmacol Ther. 2001;92:21–55. doi: 10.1016/s0163-7258(01)00158-9. [DOI] [PubMed] [Google Scholar]
  • 151.Torres GE, Gainetdinov RR, Caron MG. Plasma membrane monoamine transporters: structure, regulation and function. Nat Rev Neurosci. 2003;4:13–25. doi: 10.1038/nrn1008. [DOI] [PubMed] [Google Scholar]
  • 152.Vaughan RA. Phosphorylation and regulation of psychostimulant-sensitive neurotransmitter transporters. J Pharmacol Exp Ther. 2004;310:1–7. doi: 10.1124/jpet.103.052423. [DOI] [PubMed] [Google Scholar]
  • 153.Melikian HE. Neurotransmitter transporter trafficking: endocytosis, recycling, and regulation. Pharmacol Ther. 2004;104:17–27. doi: 10.1016/j.pharmthera.2004.07.006. [DOI] [PubMed] [Google Scholar]
  • 154.Mortensen OV, Amara SG. Dynamic regulation of the dopamine transporter. Eur J Pharmacol. 2003;479:159–70. doi: 10.1016/j.ejphar.2003.08.066. [DOI] [PubMed] [Google Scholar]
  • 155.Loder MK, Melikian HE. The dopamine transporter constitutively and recycles in a protein kinase C-regulated manner in stably transfected PC12 cell lines. J Biol Chem. 2003;278:22168–74. doi: 10.1074/jbc.M301845200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Daniels GM, Amara SG. Regulated trafficking of the human dopamine transporter. Clathrin-mediated internalization and lysosomal degradation in response to phorbol esters. J Biol Chem. 1999;274:35794–801. doi: 10.1074/jbc.274.50.35794. [DOI] [PubMed] [Google Scholar]
  • 157.Granas C, Ferrer J, Loland CJ, Javitch JA, Gether U. N-terminal truncation of the dopamine transporter abolishes phorbol ester- and substance P receptor-stimulated phosphorylation without impairing transporter internalization. J Biol Chem. 2003;278:4990–5000. doi: 10.1074/jbc.M205058200. [DOI] [PubMed] [Google Scholar]
  • 158.Holton KL, Loder MK, Melikian HE. Nonclassical, distinct endocytic signals dictate constitutive and PKC-regulated neurotransmitter transporter internalization. Nat Neurosci. 8:881–8. doi: 10.1038/nn1478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Sorkina T, Hoover BR, Zahniser NR, Sorkin A. Constitutive and protein kinase C-induced internalization of the dopamine transporter is mediated by a clathrin-dependent mechanism. Traffic. 2005;6:157–70. doi: 10.1111/j.1600-0854.2005.00259.x. [DOI] [PubMed] [Google Scholar]
  • 160.Boudanova E, Navaroli DM, Stevens Z, Melikian HE. Dopamine transporter endocytic determinants: carboxy terminal residues critical for basal and PKC-stimulated internalization. Mol Cell Neurosci. 2008;39:211–7. doi: 10.1016/j.mcn.2008.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Sorkina T, Miranda M, Dionne KR, Hoover BR, Zahniser NR, Sorkin A. RNA interference screen reveals an essential role of Nedd4-2 in dopamine transporter ubiquitination and endocytosis. J Neurosci. 2006;26:8195–205. doi: 10.1523/JNEUROSCI.1301-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Sorkina T, Richards TL, Rao A, Zahniser NR, Sorkin A. Negative regulation of dopamine transporter endocytosis by membrane-proximal N-terminal residues. J. Neurosci. 2009;29:1361–74. doi: 10.1523/JNEUROSCI.3250-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Jiang H, Jiang Q, Feng J. Parkin increases dopamine uptake by enhancing the cell surface expression of dopamine transporter. J Biol Chem. 2004;279:54380–6. doi: 10.1074/jbc.M409282200. [DOI] [PubMed] [Google Scholar]
  • 164.Miranda M, Sorkin A. Regulation of receptors and transporters by ubiquitination: new insights into surprisingly similar mechanisms. Mol Interv. 2007;7:157–67. doi: 10.1124/mi.7.3.7. [DOI] [PubMed] [Google Scholar]
  • 165.Torres GE, Yao WD, Mohn AR, Quan H, Kim KM, Levey AI, et al. Functional interaction between monoamine plasma membrane transporters and the synaptic PDZ domain-containing protein PICK1. Neuron. 2001;30:121–4. doi: 10.1016/s0896-6273(01)00267-7. [DOI] [PubMed] [Google Scholar]
  • 166.Carneiro AM, Ingram SL, Beaulieu JM, Sweeney A, Amara SG, Thomas SM, et al. The multiple LIM domain-containing adaptor protein Hic-5 synaptically colocalizes and interacts with the dopamine transporter. J Neurosci. 2002;22:7045–54. doi: 10.1523/JNEUROSCI.22-16-07045.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Lee FJ, Liu F, Pristupa ZB, Niznik HB. Direct binding and functional coupling of alpha-synuclein to the dopamine transporters accelerate dopamine-induced apoptosis. FASEB J. 2001;15:916–26. doi: 10.1096/fj.00-0334com. [DOI] [PubMed] [Google Scholar]
  • 168.Wersinger C, Rusnak M, Sidhu A. Modulation of the trafficking of the human serotonin transporter by human alpha-synuclein. Eur J Neurosci. 2006;24:55–64. doi: 10.1111/j.1460-9568.2006.04900.x. [DOI] [PubMed] [Google Scholar]
  • 169.Wersinger C, Jeannotte A, Sidhu A. Attenuation of the norepinephrine transporter activity and trafficking via interactions with alpha-synuclein. Eur J Neurosci. 2006;24:3141–52. doi: 10.1111/j.1460-9568.2006.05181.x. [DOI] [PubMed] [Google Scholar]
  • 170.Bauman AL, Apparsundaram S, Ramamoorthy S, Wadzinski BE, Vaughan RA, Blakely RD. Cocaine and antidepressant-sensitive biogenic amine transporters exist in regulated complexes with protein phosphatase 2A. J Neurosci. 2000;20:7571–8. doi: 10.1523/JNEUROSCI.20-20-07571.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Vina-Vilaseca A, Sorkin A. Lysine 63-linked polyubiquitination of the dopamine transporter requires WW3 and WW4 domains of Nedd4-2 and UBE2D ubiquitin-conjugating enzymes. J Biol Chem. 2010;285:7645–56. doi: 10.1074/jbc.M109.058990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Marazziti D, Mandillo S, Di Pietro C, Golini E, Matteoni R, Tocchini-Valentini GP. GPR37 associates with the dopamine transporter to modulate dopamine uptake and behavioral responses to dopaminergic drugs. Proc Natl Acad Sci USA. 2007;104:9846–51. doi: 10.1073/pnas.0703368104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Moszczynska A, Saleh J, Zhang H, Vukusic B, Lee FJ, Liu F. Parkin disrupts the alpha-synuclein/dopamine transporter interaction: consequences toward dopamine-induced toxicity. J Mol Neurosci. 2007;32:217–27. doi: 10.1007/s12031-007-0037-0. [DOI] [PubMed] [Google Scholar]
  • 174.Lee FJ, Pei L, Moszczynska A, Vukusic B, Fletcher PJ, Liu F. Dopamine transporter cell surface localization facilitated by a direct interaction with the dopamine D2 receptor. EMBO J. 2007;26:2127–36. doi: 10.1038/sj.emboj.7601656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Lee FJ, Pei L, Liu F. Disruption of the dopamine transporter-dopamine D2 receptor interaction in schizophrenia. Synapse. 2009;63:710–2. doi: 10.1002/syn.20648. [DOI] [PubMed] [Google Scholar]
  • 176.Lee KH, Kim MY, Kim DH, Lee YS. Syntaxin 1A and receptor for activated C kinase interact with the N-terminal region of human dopamine transporter. Neurochem Res. 2004;29:1405–9. doi: 10.1023/b:nere.0000026404.08779.43. [DOI] [PubMed] [Google Scholar]
  • 177.Binda F, Dipace C, Bowton E, Robertson SD, Lute BJ, Fog JU, et al. Syntaxin 1A interaction with the dopamine transporter promotes amphetamine-induced dopamine efflux. Mol Pharmacol. 2008;74:1101–8. doi: 10.1124/mol.108.048447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Egaña LA, Cuevas RA, Baust TB, Parra LA, Leak RK, Hochendoner S, et al. Physical and functional interaction between the dopamine transporter and the synaptic vesicle protein synaptogyrin-3. J Neurosci. 2009;29:4592–604. doi: 10.1523/JNEUROSCI.4559-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Chen R, Furman CA, Zhang M, Kim MN, Gereau RW, Leitges M, et al. Protein kinase Cbeta is a critical regulator of dopamine transporter trafficking and regulates the behavioral response to amphetamine in mice. J Pharmacol Exp Ther. 2009;328:912–20. doi: 10.1124/jpet.108.147959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Ramamoorthy S, Blakely RD. Phosphorylation and sequestration of serotonin transporters differentially modulated by psychostimulants. Science. 1999;285:763–6. doi: 10.1126/science.285.5428.763. [DOI] [PubMed] [Google Scholar]
  • 181.Quick MW. Regulating the conducting states of a mammalian serotonintransporter. Neuron. 2003;40:537–49. doi: 10.1016/s0896-6273(03)00605-6. [DOI] [PubMed] [Google Scholar]
  • 182.Ciccone MA, Timmons M, Phillips A, Quick MW. Calcium/calmodulin-dependent kinase II regulates the interaction between the serotonin transporter and syntaxin 1A. Neuropharmacology. 2008;55:763–70. doi: 10.1016/j.neuropharm.2008.06.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Carneiro AM, Blakely RD. Serotonin-, protein kinase C-, and Hic-5-associated redistribution of the platelet serotonin transporter. J Biol Chem. 2006;281:24769–80. doi: 10.1074/jbc.M603877200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Müller HK, Wiborg O, Haase J. Subcellular redistribution of the serotonin transporter by secretory carrier membrane protein 2. J Biol Chem. 2006;281:28901–09. doi: 10.1074/jbc.M602848200. [DOI] [PubMed] [Google Scholar]
  • 185.Chanrion B, Mannoury la Cour C, Bertaso F, Lerner-Natoli M, Freissmuth M, Millan MJ, et al. Physical interaction between the serotonin transporter and neuronal nitric oxide synthase underlies reciprocal modulation of their activity. Proc Natl Acad Sci USA. 2007;104:8119–24. doi: 10.1073/pnas.0610964104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Steiner JA, Carneiro AM, Wright J, Matthies HJ, Prasad HC, Nicki CK, et al. cGMP-dependent protein kinase Ialpha associates with the antidepressant-sensitive serotonin transporter and dictates rapid modulation of serotonin uptake. Mol Brain. 2009;2:26. doi: 10.1186/1756-6606-2-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Sung U, Apparsundaram S, Galli A, Kahlig KM, Savchenko V, Schroeter S, et al. A regulated interaction of syntaxin 1A with the antidepressant-sensitive norepinephrine transporter establishes catecholamine clearance capacity. J Neurosci. 2003;23:1697–709. doi: 10.1523/JNEUROSCI.23-05-01697.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Sung U, Jennings JL, Link AJ, Blakely RD. Proteomic analysis of human norepinephrine transporter complexes reveals associations with protein phosphatase 2A anchoring subunit and 14-3-3 proteins. Biochem Biophys Res Commun. 2005;333:671–8. doi: 10.1016/j.bbrc.2005.05.165. [DOI] [PubMed] [Google Scholar]
  • 189.Cartier EA, Parra LA, Baust TB, Quiroz M, Salazar G, Faundez V, et al. A biochemical and functional protein complex involving dopamine synthesis and transport into synaptic vesicles. J Biol Chem. 2010;285:1957–66. doi: 10.1074/jbc.M109.054510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Sora I, Hall FS, Andrews AM, Itokawa M, Li XF, Wei HB, et al. Molecular mechanisms of cocaine reward: combined dopamine and serotonin transporter knockouts eliminate cocaine place preference. Proc Natl Acad Sci USA. 2001;98:5300–5. doi: 10.1073/pnas.091039298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Sora I, Wichems C, Takahashi N, Li XF, Zeng Z, Revay R, et al. Cocaine reward models: conditioned place preference can be established in dopamine- and in serotonin-transporter knockout mice. Proc Natl Acad Sci USA. 1998;95:7699–704. doi: 10.1073/pnas.95.13.7699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Gainetdinov RR, Jones SR, Caron MG. Functional hyperdopaminergia in dopamine transporter knock-out mice. Biol Psychiatry. 1999;46:303–11. doi: 10.1016/s0006-3223(99)00122-5. [DOI] [PubMed] [Google Scholar]
  • 193.Gainetdinov RR, Wetsel WC, Jones SR, Levin ED, Jaber M, Caron MG. Role of serotonin in the paradoxical calming effect of psychostimulants on hyperactivity. Science. 1999;283:397–401. doi: 10.1126/science.283.5400.397. [DOI] [PubMed] [Google Scholar]
  • 194.Wisor JP, Nishino S, Sora I, Uhl GH, Mignot E, Edgar DM. Dopaminergic role in stimulant-induced wakefulness. J Neurosci. 2001;21:1787–94. doi: 10.1523/JNEUROSCI.21-05-01787.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Mead AN, Rocha BA, Donovan DM, Katz JL. Intravenous cocaine induced-activity and behavioural sensitization in norepinephrine-, but not dopamine-transporter knockout mice. Eur J Neurosci. 2002;16:514–20. doi: 10.1046/j.1460-9568.2002.02104.x. [DOI] [PubMed] [Google Scholar]
  • 196.Cyr M, Beaulieu JM, Laakso A, Sotnikova TD, Yao WD, Bohn LM, et al. Sustained elevation of extracellular dopamine causes motor dysfunction and selective degeneration of striatal GABAergic neurons. Proc Natl Acad Sci USA. 2003;100:11035–40. doi: 10.1073/pnas.1831768100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Fernagut PO, Chalon S, Diguet E, Guilloteau D, Tison F, Jaber M. Motor behaviour deficits and their histopathological and functional correlates in the nigrostriatal system of dopamine transporter knockout mice. Neuroscience. 2003;116:1123–30. doi: 10.1016/s0306-4522(02)00778-9. [DOI] [PubMed] [Google Scholar]
  • 198.Hironaka N, Ikeda K, Sora I, Uhl GR, Niki H. Food-reinforced operant behavior in dopamine transporter knockout mice: enhanced resistance to extinction. Ann N Y Acad Sci. 2004;1025:140–5. doi: 10.1196/annals.1316.018. [DOI] [PubMed] [Google Scholar]
  • 199.Morice E, Billard JM, Denis C, Mathieu F, Betancur C, Epelbaum J, et al. Parallel loss of hippocampal LTD and cognitive flexibility in a genetic model of hyperdopaminergia. Neuropsychopharmacology. 2007;32:2108–16. doi: 10.1038/sj.npp.1301354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Thomsen M, Hall FS, Uhl GR, Caine SB. Dramatically decreased cocaine self-administration in dopamine but not serotonin transporter knock-out mice. J Neurosci. 2009;29:1087–92. doi: 10.1523/JNEUROSCI.4037-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zhuang X, Oosting RS, Jones SR, Gainetdinov RR, Miller GW, Caron MG, et al. Hyperactivity and impaired response habituation in hyperdopaminergic mice. Proc Natl Acad Sci USA. 2001;98:1982–7. doi: 10.1073/pnas.98.4.1982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Ralph-Williams RJ, Paulus MP, Zhuang X, Hen R, Geyer MA. Valproate attenuates hyperactive and perseverative behaviors in mutant mice with a dysregulated dopamine system. Biol Psychiatry. 2003;53:352–9. doi: 10.1016/s0006-3223(02)01489-0. [DOI] [PubMed] [Google Scholar]
  • 203.Berridge KC, Aldridge JW, Houchard KR, Zhuang X. Sequential super-stereotypy of an instinctive fixed action pattern in hyper-dopaminergic mutant mice: a model of obsessive compulsive disorder and Tourette’s. BMC Biol. 2005;3:4. doi: 10.1186/1741-7007-3-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Tilley MR, Cagniard B, Zhuang X, Han DD, Tiao N, Gu HH. Cocaine reward and locomotion stimulation in mice with reduced dopamine transporter expression. BMC Neurosci. 2007;8:42. doi: 10.1186/1471-2202-8-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Pecina S, Cagniard B, Berridge KC, Aldridge JW, Zhuang X. Hyperdopaminergic mutant mice have higher “wanting” but not “liking” for sweet rewards. J Neurosci. 2003;23:9395–402. doi: 10.1523/JNEUROSCI.23-28-09395.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Yin HH, Zhuang X, Balleine BW. Instrumental learning in hyperdopaminergic mice. Neurobiol Learn Mem. 2006;85:283–8. doi: 10.1016/j.nlm.2005.12.001. [DOI] [PubMed] [Google Scholar]
  • 207.Balci F, Ludvig EA, Abner R, Zhuang X, Poon P, Brunner D. Motivational effects on interval timing in dopamine transporter (DAT) knockdown mice. Brain Res. 2010;1325:89–99. doi: 10.1016/j.brainres.2010.02.034. [DOI] [PubMed] [Google Scholar]
  • 208.Donovan DM, Miner LL, Perry MP, Revay RS, Sharpe LG, Przedborski S, et al. Cocaine reward and MPTP toxicity: alteration by regional variant dopamine transporter overexpression. Brain Res Mol Brain Res. 1999;73:37–49. doi: 10.1016/s0169-328x(99)00235-1. [DOI] [PubMed] [Google Scholar]
  • 209.Salahpour A, Ramsey AJ, Medvedev IO, Kile B, Sotnikova TD, Holmstrand E, et al. Increased amphetamine-induced hyperactivity and reward in mice overexpressing the dopamine transporter. Proc Natl Acad Sci USA. 2008;105:4405–10. doi: 10.1073/pnas.0707646105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Thomsen M, Han DD, Gu HH, Caine SB. Lack of cocaine self-administration in mice expressing a cocaine-insensitive dopamine transporter. J Pharmacol Exp Ther. 2009;331:204–11. doi: 10.1124/jpet.109.156265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Chen R, Tilley MR, Wei H, Zhou F, Zhou FM, Ching S, Quan N, Stephens RL, Hill ER, Nottoli T, Han DD, Gu HH. Abolished cocaine reward in mice with a cocaine-insensitive dopamine transporter. Proc Natl Acad Sci USA. 2006;103:9333–8. doi: 10.1073/pnas.0600905103. 2006) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Tilley MR, Gu HH. Dopamine transporter inhibition is required for cocaine-induced stereotypy. Neuroreport. 2008;19:1137–40. doi: 10.1097/WNR.0b013e3283063183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Tilley MR, O’Neill B, Han DD, Gu HH. Cocaine does not produce reward in absence of dopamine transporter inhibition. Neuroreport. 2009;20:9–12. doi: 10.1097/WNR.0b013e32831b9ce4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Fox MA, Jensen CL, Gallagher PS, Murphy DL. Receptor mediation of exaggerated responses to serotonin-enhancing drugs in serotonin transporter (SERT)-deficient mice. Neuropharmacology. 2007;53:643–56. doi: 10.1016/j.neuropharm.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 215.Fox MA, Jensen CL, Murphy DL. Tramadol and another atypical opioid meperidine have exaggerated serotonin syndrome behavioural effects, but decreased analgesic effects, in genetically deficient serotonin transporter (SERT) mice. Int J Neuropsychopharmacol. 2009;12:1055–65. doi: 10.1017/S146114570900011X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Kalueff AV, Jensen CL, Murphy DL. Locomotory patterns, spatiotemporal organization of exploration and spatial memory in serotonin transporter knockout mice. Brain Res. 2007;1169:87–97. doi: 10.1016/j.brainres.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 217.Holmes A, Murphy DL, Crawley JN. Reduced aggression in mice lacking the serotonin transporter. Psychopharmacology (Berl) 2002;161:160–7. doi: 10.1007/s00213-002-1024-3. [DOI] [PubMed] [Google Scholar]
  • 218.Holmes A, Yang RJ, Murphy DL, Crawley JN. Evaluation of antidepressant-related behavioral responses in mice lacking the serotonin transporter. Neuropsychopharmacology. 2002;27:914–23. doi: 10.1016/S0893-133X(02)00374-3. [DOI] [PubMed] [Google Scholar]
  • 219.Holmes A, Yang RJ, Lesch KP, Crawley JN, Murphy DL. Mice lacking the serotonin transporter exhibit 5-HT(1A) receptor-mediated abnormalities in tests for anxiety-like behavior. Neuropsychopharmacology. 2003;28:2077–88. doi: 10.1038/sj.npp.1300266. [DOI] [PubMed] [Google Scholar]
  • 220.Zhao S, Edwards J, Carroll J, Wiedholz L, Millstein RA, Jaing C, et al. Insertion mutation at the C-terminus of the serotonin transporter disrupts brain serotonin function and emotion-related behaviors in mice. Neuroscience. 2006;140:321–34. doi: 10.1016/j.neuroscience.2006.01.049. [DOI] [PubMed] [Google Scholar]
  • 221.Kalueff AV, Fox MA, Gallagher PS, Murphy DL. Hypolocomotion, anxiety and serotonin syndrome-like behavior contribute to the complex phenotype of serotonin transporter knockout mice. Genes Brain Behav. 2007;6:389–400. doi: 10.1111/j.1601-183X.2006.00270.x. [DOI] [PubMed] [Google Scholar]
  • 222.Kalueff AV, Ren-Patterson RF, Murphy DL. The developing use of heterozygous mutant mouse models in brain monoamine transporter research. Trends Pharmacol Sci. 2007;28:122–7. doi: 10.1016/j.tips.2007.01.002. [DOI] [PubMed] [Google Scholar]
  • 223.Olivier JD, Jans LA, Korte-Bouws GA, Korte SM, Deen PM, Cools AR, et al. Acute tryptophan depletion dose dependently impairs object memory in serotonin transporter knockout rats. Psychopharmacology (Berl) 2008;200:243–54. doi: 10.1007/s00213-008-1201-0. [DOI] [PubMed] [Google Scholar]
  • 224.Perona MT, Waters S, Hall FS, Sora I, Lesch KP, Murphy DL, et al. Animal models of depression in dopamine, serotonin, and norepinephrine transporter knockout mice: prominent effects of dopamine transporter deletions. Behav Pharmacol. 2008;19:566–74. doi: 10.1097/FBP.0b013e32830cd80f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Homberg JR, van den Bos R, den Heijer E, Suer R, Cuppen E. Serotonin transporter dosage modulates long-term decision-making in rat and human. Neuropharmacology. 2008;55:80–4. doi: 10.1016/j.neuropharm.2008.04.016. [DOI] [PubMed] [Google Scholar]
  • 226.Homberg JR, Pattij T, Janssen MC, Ronken E, De Boer SF, Schoffelmeer AN, et al. Serotonin transporter deficiency in rats improves inhibitory control but not behavioural flexibility. Eur J Neurosci. 2007;26:2066–73. doi: 10.1111/j.1460-9568.2007.05839.x. [DOI] [PubMed] [Google Scholar]
  • 227.Homberg JR, Schiepers OJ, Schoffelmeer AN, Cuppen E, Vanderschuren LJ. Acute and constitutive increases in central serotonin levels reduce social play behaviour in peri-adolescent rats. Psychopharmacology (Berl) 2007;195:175–82. doi: 10.1007/s00213-007-0895-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Holmes A, Lit Q, Murphy DL, Gold E, Crawley JN. Abnormal anxiety-related behavior in serotonin transporter null mutant mice: the influence of genetic background. Genes Brain Behav. 2003;2:365–80. doi: 10.1046/j.1601-1848.2003.00050.x. [DOI] [PubMed] [Google Scholar]
  • 229.Wellman CL, Izquierdo A, Garrett JE, Martin KP, Carroll J, Millstein R, et al. Impaired stress-coping and fear extinction and abnormal corticolimbic morphology in serotonin transporter knock-out mice. J Neurosci. 2007;27:684–91. doi: 10.1523/JNEUROSCI.4595-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Homberg JR, De Boer SF, Raaso HS, Olivier JD, Verheul M, Ronken E, et al. Adaptations in pre- and postsynaptic 5-HT1A receptor function and cocaine supersensitivity in serotonin transporter knockout rats. Psychopharmacology (Berl) 2008;200:367–80. doi: 10.1007/s00213-008-1212-x. [DOI] [PubMed] [Google Scholar]
  • 231.Olivier JD, Jans LA, Blokland A, Broers NJ, Homberg JR, Ellenbroek BA, et al. Serotonin transporter deficiency in rats contributes to impaired object memory. Genes Brain Behav. 2009;8:829–34. doi: 10.1111/j.1601-183X.2009.00530.x. [DOI] [PubMed] [Google Scholar]
  • 232.Jennings KA, Loder MK, Sheward WJ, Pei Q, Deacon RM, Benson MA, et al. Increased expression of the 5-HT transporter confers a low-anxiety phenotype linked to decreased 5-HT transmission. J Neurosci. 2006;26:8955–64. doi: 10.1523/JNEUROSCI.5356-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Dawson N, Ferrington L, Olverman HJ, Harmar AJ, Kelly PA. Sex influences the effect of a lifelong increase in serotonin transporter function on cerebral metabolism. J Neurosci Res. 2009;87:2375–85. doi: 10.1002/jnr.22062. [DOI] [PubMed] [Google Scholar]
  • 234.Xu F, Gainetdinov RR, Wetsel WC, Jones SR, Bohn LM, Miller GW, et al. Mice lacking the norepinephrine transporter are supersensitive to psychostimulants. Nat Neurosci. 2000;3:465–71. doi: 10.1038/74839. [DOI] [PubMed] [Google Scholar]
  • 235.Dziedzicka-Wasylewska M, Faron-Gorecka A, Kusmider M, Drozdowska E, Rogoz Z, Siwanowicz J, et al. Effect of antidepressant drugs in mice lacking the norepinephrine transporter. Neuropsychopharmacology. 2006;31:2424–32. doi: 10.1038/sj.npp.1301064. [DOI] [PubMed] [Google Scholar]
  • 236.Haenisch B, Bilkei-Gorzo A, Caron MG, Bonisch H. Knockout of the norepinephrine transporter and pharmacologically diverse antidepressants prevent behavioral and brain neurotrophin alterations in two chronic stress models of depression. J Neurochem. 2009;111:403–16. doi: 10.1111/j.1471-4159.2009.06345.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Bohn LM, Xu F, Gainetdinov RR, Caron MG. Potentiated opioid analgesia in norepinephrine transporter knock-out mice. J Neurosci. 2000;20:9040–5. doi: 10.1523/JNEUROSCI.20-24-09040.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Rommelfanger KS, Weinshenker D, Miller GW. Reduced MPTP toxicity in noradrenaline transporter knockout mice. J Neurochem. 2004;91:1116–24. doi: 10.1111/j.1471-4159.2004.02785.x. [DOI] [PubMed] [Google Scholar]
  • 239.Kaminski RM, Shippenberg TS, Witkin JM, Rocha BA. Genetic deletion of the norepinephrine transporter decreases vulnerability to seizures. Neurosci Lett. 2005;382:51–5. doi: 10.1016/j.neulet.2005.02.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Ahern TH, Javors MA, Eagles DA, Martillotti J, Mitchell HA, Liles LC, et al. The effects of chronic norepinephrine transporter inactivation on seizure susceptibility in mice. Neuropsychopharmacology. 2006;31:730–8. doi: 10.1038/sj.npp.1300847. [DOI] [PubMed] [Google Scholar]
  • 241.Keller NR, Diedrich A, Appalsamy M, Tuntrakool S, Lonce S, Finney C, et al. Norepinephrine transporter-deficient mice exhibit excessive tachycardia and elevated blood pressure with wakefulness and activity. Circulation. 2004;110:1191–6. doi: 10.1161/01.CIR.0000141804.90845.E6. [DOI] [PubMed] [Google Scholar]
  • 242.Fukushima S, Shen H, Hata H, Ohara A, Ohmi K, Ikeda K, et al. Methamphetamine-induced locomotor activity and sensitization in dopamine transporter and vesicular monoamine transporter 2 double mutant mice. Psychopharmacology (Berl) 2007;193:55–62. doi: 10.1007/s00213-007-0749-4. [DOI] [PubMed] [Google Scholar]
  • 243.Fumagalli F, Gainetdinov RR, Wang YM, Valenzano KJ, Miller GW, Caron MG. Increased methamphetamine neurotoxicity in heterozygous vesicular monoamine transporter 2 knock-out mice. J Neurosci. 1999;19:2424–31. doi: 10.1523/JNEUROSCI.19-07-02424.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Savelieva KV, Caudle WM, Miller GW. Altered ethanol-associated behaviors in vesicular monoamine transporter heterozygote knockout mice. Alcohol. 2006;40:87–94. doi: 10.1016/j.alcohol.2006.09.030. [DOI] [PubMed] [Google Scholar]
  • 245.Gainetdinov RR, Fumagalli F, Wang YM, Jones SR, Levey AI, Miller GW, et al. Increased MPTP neurotoxicity in vesicular monoamine transporter 2 heterozygote knockout mice. J Neurochem. 1998;70:1973–8. doi: 10.1046/j.1471-4159.1998.70051973.x. [DOI] [PubMed] [Google Scholar]
  • 246.Larsen KE, Fon EA, Hastings TG, Edwards RH, Sulzer D. Methamphetamine-induced degeneration of dopaminergic neurons involves autophagy and upregulation of dopamine synthesis. J Neurosci. 2002;22:8951–60. doi: 10.1523/JNEUROSCI.22-20-08951.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Kariya S, Takahashi N, Hirano M, Ueno S. Increased vulnerability to L-DOPA toxicity in dopaminergic neurons From VMAT2 heterozygote knockout mice. J Mol Neurosci. 2005;27:277–79. doi: 10.1385/JMN:27:3:277. [DOI] [PubMed] [Google Scholar]
  • 248.Hall FS, Sora I, Uhl GR. Sex-dependent modulation of ethanol consumption in vesicular monoamine transporter 2 (VMAT2) and dopamine transporter (DAT) knockout mice. Neuropsychopharmacology. 2003;28:620–8. doi: 10.1038/sj.npp.1300070. [DOI] [PubMed] [Google Scholar]
  • 249.Fukui M, Rodriguiz RM, Zhou J, Jiang SX, Phillips LE, Caron MG, et al. Vmat2 heterozygous mutant mice display a depressive-like phenotype. J Neurosci. 2007;27:10520–9. doi: 10.1523/JNEUROSCI.4388-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Mooslehner KA, Chan PM, Xu W, Liu L, Smadja C, Humby T, et al. Mice with very low expression of the vesicular monoamine transporter 2 gene survive into adulthood: potential mouse model for parkinsonism. Mol Cell Biol. 2001;21:5321–31. doi: 10.1128/MCB.21.16.5321-5331.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Guillot TS, Shepherd KR, Richardson JR, Wang MZ, Li Y, Emson PC, et al. Reduced vesicular storage of dopamine exacerbates methamphetamine-induced neurodegeneration and astrogliosis. J Neurochem. 2008;106:2205–17. doi: 10.1111/j.1471-4159.2008.05568.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Colebrooke RE, Humby T, Lynch PJ, McGowan DP, Xia J, Emson PC. Age-related decline in striatal dopamine content and motor performance occurs in the absence of nigral cell loss in a genetic mouse model of Parkinson’s disease. Eur J Neurosci. 2006;24:2622–30. doi: 10.1111/j.1460-9568.2006.05143.x. [DOI] [PubMed] [Google Scholar]
  • 253.Caudle WM, Richardson JR, Wang MZ, Taylor TN, Guillot TS, McCormack AL, et al. Reduced vesicular storage of dopamine causes progressive nigrostriatal neurodegeneration. J Neurosci. 2007;27:8138–48. doi: 10.1523/JNEUROSCI.0319-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Taylor TN, Caudle WM, Shepherd KR, Noorian A, Jackson CR, Iuvone PM, et al. Nonmotor symptoms of Parkinson’s disease revealed in an animal model with reduced monoamine storage capacity. J Neurosci. 2009;29:8103–13. doi: 10.1523/JNEUROSCI.1495-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Lin Z, Madras BK. Human genetics and pharmacology of neurotransmitter transporters. Handb Exp Pharmacol. 2006;175:327–71. doi: 10.1007/3-540-29784-7_16. [DOI] [PubMed] [Google Scholar]
  • 256.Miller GM, Madras BK. Polymorphisms in the 3′-untranslated region of human and monkey dopamine transporter genes affect reporter gene expression. Mol Psychiatry. 2002;7:44–55. doi: 10.1038/sj.mp.4000921. [DOI] [PubMed] [Google Scholar]
  • 257.Jacobsen LK, Staley JK, Zoghbi SS, Seibyl JP, Kosten TR, Innis RB, et al. Prediction of dopamine transporter binding availability by genotype: a preliminary report. Am J Psychiatry. 2000;157:1700–3. doi: 10.1176/appi.ajp.157.10.1700. [DOI] [PubMed] [Google Scholar]
  • 258.van de Giessen EM, de Win MM, Tanck MW, van den Brink W, Baas F, Booij J. Striatal dopamine transporter availability associated with polymorphisms in the dopamine transporter gene SLC6A3. J Nucl Med. 2009;50:45–52. doi: 10.2967/jnumed.108.053652. [DOI] [PubMed] [Google Scholar]
  • 259.Greenwood TA, Kelsoe JR. Promoter and intronic variants affect the transcriptional regulation of the human dopamine transporter gene. Genomics. 2003;82:511–20. doi: 10.1016/s0888-7543(03)00142-3. [DOI] [PubMed] [Google Scholar]
  • 260.Hill M, Anney RJ, Gill M, Hawi Z. Functional analysis of intron 8 and 3′ UTR variable number of tandem repeats of SLC6A3: differential activity of intron 8 variants. Pharmacogenomics J. 2009 Dec 22; doi: 10.1038/tpj.2009.66. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 261.Althaus M, Groen Y, Wijers AA, Minderaa RB, Kema IP, Dijck JD, et al. Variants of the SLC6A3 (DAT1) polymorphism affect performance monitoring-related cortical evoked potentials that are associated with ADHD. Biol Psychol. 2010 May 5; doi: 10.1016/j.biopsycho.2010.04.007. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 262.Durston S, Fossella JA, Mulder MJ, Casey BJ, Ziermans TB, Vessaz MN, et al. Dopamine transporter genotype conveys familial risk of attention-deficit/hyperactivity disorder through striatal activation. J Am Acad Child Adolesc Psychiatry. 2008;47:61–7. doi: 10.1097/chi.0b013e31815a5f17. [DOI] [PubMed] [Google Scholar]
  • 263.Brown AB, Biederman J, Valera EM, Doyle AE, Bush G, Spencer T, et al. Effect of dopamine transporter gene (SLC6A3) variation on dorsal anterior cingulate function in attention-deficit/hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:365–75. doi: 10.1002/ajmg.b.31022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Forbes EE, Brown SM, Kimak M, Ferrell RE, Manuck SB, Hariri AR. Genetic variation in components of dopamine neurotransmission impacts ventral striatal reactivity associated with impulsivity. Mol Psychiatry. 2009;14:60–70. doi: 10.1038/sj.mp.4002086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Friedel S, Saar K, Sauer S, Dempfle A, Walitza S, Renner T, et al. Association and linkage of allelic variants of the dopamine transporter gene in ADHD. Mol Psychiatry. 2007;12:923–33. doi: 10.1038/sj.mp.4001986. [DOI] [PubMed] [Google Scholar]
  • 266.Bobb AJ, Castellanos FX, Addington AM, Rapoport JL. Molecular genetic studies of ADHD: 1991 to 2004. Am J Med Genet B Neuropsychiatr Genet. 2005;132:109–25. [PubMed] [Google Scholar]
  • 267.Kim CH, Hahn MK, Joung Y, Anderson SL, Steele AH, Mazei-Robinson MS, et al. A polymorphism in the norepinephrine transporter gene alters promoter activity and is associated with attention-deficit hyperactivity disorder. Proc Natl Acad Sci USA. 2006;103:19164–9. doi: 10.1073/pnas.0510836103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Courtet P, Baud P, Abbar M, Boulenger JP, Castelnau D, Mouthon D, et al. Association between violent suicidal behavior and the low activity allele of the serotonin transporter gene. Mol Psychiatry. 2001;6:338–41. doi: 10.1038/sj.mp.4000856. [DOI] [PubMed] [Google Scholar]
  • 269.Ueno S, Nakamura M, Mikami M, Kondoh K, Ishiguro H, Arinami T, et al. Identification of a novel polymorphism of the human dopamine transporter (DAT1) gene and the significant association with alcoholism. Mol Psychiatry. 1999;4:552–557. doi: 10.1038/sj.mp.4000562. [DOI] [PubMed] [Google Scholar]
  • 270.Köhnke MD, Batra A, Kolb W, Köhnke AM, Lutz U, Schick S, et al. Association of the dopamine transporter gene with alcoholism. Alcohol Alcohol. 2005;40:339–42. doi: 10.1093/alcalc/agh179. [DOI] [PubMed] [Google Scholar]
  • 271.Seneviratne C, Huang W, Ait-Daoud N, Li MD, Johnson BA. Characterization of a functional polymorphism in the 3′ UTR of SLC6A4 and its association with drinking intensity. Alcohol Clin Exp Res. 2009;33:332–9. doi: 10.1111/j.1530-0277.2008.00837.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Lin Z, Walther D, Yu XY, Li S, Drgon T, Uhl GR. SLC18A2 promoter haplotypes and identification of a novel protective factor against alcoholism. Hum Mol Genet. 2005;14:1393–404. doi: 10.1093/hmg/ddi148. [DOI] [PubMed] [Google Scholar]
  • 273.Schwab SG, Franke PE, Hoefgen B, Guttenthaler V, Lichtermann D, Trixler M, et al. Association of DNA polymorphisms in the synaptic vesicular amine transporter gene (SLC18A2) with alcohol and nicotine dependence. Neuropsychopharmacology. 2005;30:2263–8. doi: 10.1038/sj.npp.1300809. [DOI] [PubMed] [Google Scholar]
  • 274.Quaranta D, Bizzarro A, Marra C, Vita MG, Seripa D, Pilotto A, et al. Psychotic symptoms in Alzheimer’s disease and 5-HTTLPR polymorphism of the serotonin transporter gene: evidence for an association. J Alzheimers Dis. 2009;16:173–80. doi: 10.3233/JAD-2009-0950. [DOI] [PubMed] [Google Scholar]
  • 275.Joyce PR, McHugh PC, Light KJ, Rowe S, Miller AL, Kennedy MA. Relationships between angry-impulsive personality traits and genetic polymorphisms of the dopamine transporter. Biol Psychiatry. 2009;66:717–21. doi: 10.1016/j.biopsych.2009.03.005. [DOI] [PubMed] [Google Scholar]
  • 276.Urwin RE, Bennetts B, Wilcken B, Lampropoulos B, Beumont P, Clarke S, et al. Anorexia nervosa (restrictive subtype) is associated with a polymorphism in the novel norepinephrine transporter gene promoter polymorphic region. Mol Psychiatry. 2002;7:652–7. doi: 10.1038/sj.mp.4001080. [DOI] [PubMed] [Google Scholar]
  • 277.Matsushita S, Suzuki K, Murayama M, Nishiguchi N, Hishimoto A, Takeda A, et al. Serotonin transporter regulatory region polymorphism is associated with anorexia nervosa. Am J Med Genet B Neuropsychiatr Genet. 2004;128:114–7. doi: 10.1002/ajmg.b.30022. [DOI] [PubMed] [Google Scholar]
  • 278.Wray NR, James MR, Gordon SD, Dumenil T, Ryan L, Coventry WL, et al. Accurate, Large-Scale Genotyping of 5HTTLPR and Flanking Single Nucleotide Polymorphisms in an Association Study of Depression, Anxiety, and Personality Measures. Biol Psychiatry. 2009;66:468–76. doi: 10.1016/j.biopsych.2009.04.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Wassink TH, Hazlett HC, Epping EA, Arndt S, Dager SR, Schellenberg GD, et al. Cerebral cortical gray matter overgrowth and functional variation of the serotonin transporter gene in autism. Arch Gen Psychiatry. 2007;64:709–17. doi: 10.1001/archpsyc.64.6.709. [DOI] [PubMed] [Google Scholar]
  • 280.Keikhaee MR, Fadai F, Sargolzaee MR, Javanbakht A, Najmabadi H, Ohadi M. Association analysis of the dopamine transporter (DAT1)-67A/T polymorphism in bipolar disorder. Am J Med Genet B Neuropsychiatr Genet. 2005;135:47–9. doi: 10.1002/ajmg.b.30174. [DOI] [PubMed] [Google Scholar]
  • 281.van Munster BC, de Rooij SE, Yazdanpanah M, Tienari PJ, Pitkälä KH, Osse RJ, et al. The association of the dopamine transporter gene and the dopamine receptor 2 gene with delirium, a meta-analysis. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:648–55. doi: 10.1002/ajmg.b.31034. [DOI] [PubMed] [Google Scholar]
  • 282.Dong C, Wong ML, Licinio J. Sequence variations of ABCB1, SLC6A2, SLC6A3, SLC6A4, CREB1, CRHR1 and NTRK2: association with major depression and antidepressant response in Mexican-Americans. Mol Psychiatry. 2009;14:1105–18. doi: 10.1038/mp.2009.92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.López-León S, Janssens AC, González-Zuloeta Ladd AM, Del-Favero J, Claes SJ, et al. Meta-analyses of genetic studies on major depressive disorder. Mol Psychiatry. 2008;13:772–85. doi: 10.1038/sj.mp.4002088. [DOI] [PubMed] [Google Scholar]
  • 284.Pattarachotanant N, Sritharathikhun T, Suttirat S, Tencomnao T. Association of C/T polymorphism in intron 14 of the dopamine transporter gene (rs40184) with major depression in a northeastern Thai population. Genet Mol Res. 2010;9:565–72. doi: 10.4238/vol9-1gmr757. [DOI] [PubMed] [Google Scholar]
  • 285.Min W, Li T, Ma X, Li Z, Yu T, Gao D, et al. Monoamine transporter gene polymorphisms affect susceptibility to depression and predict antidepressant response. Psychopharmacology (Berl) 2009;205:409–17. doi: 10.1007/s00213-009-1550-3. [DOI] [PubMed] [Google Scholar]
  • 286.Christiansen L, Tan Q, Iachina M, Bathum L, Kruse TA, McGue M, et al. Candidate gene polymorphisms in the serotonergic pathway: influence on depression symptomatology in an elderly population. Biol Psychiatry. 2007;61:223–30. doi: 10.1016/j.biopsych.2006.03.046. [DOI] [PubMed] [Google Scholar]
  • 287.Guindalini C, Howard M, Haddley K, Laranjeira R, Collier D, Ammar N, et al. A dopamine transporter gene functional variant associated with cocaine abuse in a Brazilian sample. Proc Natl Acad Sci USA. 2006;103:4552–7. doi: 10.1073/pnas.0504789103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Gelernter J, Kranzler HR, Satel SL, Rao PA. Genetic association between dopamine transporter protein alleles and cocaine-induced paranoia. Neuropsychopharmacology. 1994;11:195–200. doi: 10.1038/sj.npp.1380106. [DOI] [PubMed] [Google Scholar]
  • 289.Dlugos AM, Hamidovic A, Palmer AA, de Wit H. Further evidence of association between amphetamine response and SLC6A2 gene variants. Psychopharmacology (Berl) 2009;206:501–11. doi: 10.1007/s00213-009-1628-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Sikander A, Rana SV, Sinha SK, Prasad KK, Arora SK, Sharma SK, et al. Serotonin transporter promoter variant: Analysis in Indian IBS patients and control population. J Clin Gastroenterol. 2009;43:957–61. doi: 10.1097/MCG.0b013e3181b37e8c. [DOI] [PubMed] [Google Scholar]
  • 291.Ulrich S, Hersberger M, Fischler M, Nussbaumer-Ochsner Y, Treder U, Russi EW, et al. Genetic polymorphisms of the serotonin transporter, but not the 2a receptor or nitric oxide synthetase, are associated with pulmonary hypertension in chronic obstructive pulmonary disease. Respiration. 2010;79:288–95. doi: 10.1159/000226243. [DOI] [PubMed] [Google Scholar]
  • 292.Ksiazek P, Buraczynska K, Buraczynska M. Norepinephrine transporter gene (NET) polymorphism in patients with type 2 diabetes. Kidney Blood Press Res. 2006;29:338–43. doi: 10.1159/000097356. [DOI] [PubMed] [Google Scholar]
  • 293.Deuschle M, Schredl M, Schilling C, Wüst S, Frank J, Witt SH, et al. Association between a serotonin transporter length polymorphism and primary insomnia. Sleep. 2010;33:343–7. doi: 10.1093/sleep/33.3.343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Sen S, Burmeister M, Ghosh D. Meta-analysis of the association between a serotonin transporter promoter polymorphism (5-HTTLPR) and anxiety-related personality traits. Am J Med Genet B Neuropsychiatr Genet. 2004;127:85–9. doi: 10.1002/ajmg.b.20158. [DOI] [PubMed] [Google Scholar]
  • 295.Voyiaziakis E, Evgrafov O, Li D, Yoon HJ, Tabares P, Samuels J, et al. Association of SLC6A4 variants with obsessive-compulsive disorder in a large multicenter US family study. Mol Psychiatry. 2009 Oct 6; doi: 10.1038/mp.2009.100. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Shannon JR, Flattem NL, Jordan J, Jacob G, Black BK, Biaggioni I, Blakely RD, Robertson D. Orthostatic intolerance and tachycardia associated with norepinephrine-transporter deficiency. N Engl J Med. 2000;342:541–9. doi: 10.1056/NEJM200002243420803. [DOI] [PubMed] [Google Scholar]
  • 297.Treister R, Pud D, Ebstein RP, Laiba E, Gershon E, Haddad M, et al. Associations between polymorphisms in dopamine neurotransmitter pathway genes and pain response in healthy humans. Pain. 2009;147:187–93. doi: 10.1016/j.pain.2009.09.001. [DOI] [PubMed] [Google Scholar]
  • 298.Kelada SN, Checkoway H, Kardia SL, Carlson CS, Costa-Mallen P, Eaton DL, et al. 5′ and 3′ region variability in the dopamine transporter gene (SLC6A3), pesticide exposure and Parkinson’s disease risk: a hypothesis-generating study. Hum Mol Genet. 2006;15:3055–62. doi: 10.1093/hmg/ddl247. [DOI] [PubMed] [Google Scholar]
  • 299.Albani D, Vittori A, Batelli S, Polito L, De Mauro S, Galimberti D, et al. Serotonin transporter gene polymorphic element 5-HTTLPR increases the risk of sporadic Parkinson’s disease in Italy. Eur Neurol. 2009;62:120–3. doi: 10.1159/000222784. [DOI] [PubMed] [Google Scholar]
  • 300.Glatt CE, Wahner AD, White DJ, Ruiz-Linares A, Ritz B. Gain-of-function haplotypes in the vesicular monoamine transporter promoter are protective for Parkinson disease in women. Hum Mol Genet. 2006;15:299–305. doi: 10.1093/hmg/ddi445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Grabe HJ, Spitzer C, Schwahn C, Marcinek A, Frahnow A, Barnow S, et al. Serotonin transporter gene (SLC6A4) promoter polymorphisms and the susceptibility to posttraumatic stress disorder in the general population. Am J Psychiatry. 2009;166:926–33. doi: 10.1176/appi.ajp.2009.08101542. [DOI] [PubMed] [Google Scholar]
  • 302.Talkowski ME, Kirov G, Bamne M, Georgieva L, Torres G, Mansour H, et al. A network of dopaminergic gene variations implicated as risk factors for schizophrenia. Hum Mol Genet. 2008;17:747–58. doi: 10.1093/hmg/ddm347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Allen NC, Bagade S, McQueen MB, Ioannidis JP, Kavvoura FK, Khoury MJ, et al. Systematic meta-analyses and field synopsis of genetic association studies in schizophrenia: the SzGene database. Nat Genet. 2008;40:827–34. doi: 10.1038/ng.171. [DOI] [PubMed] [Google Scholar]
  • 304.Vandenbergh DJ, Bennett CJ, Grant MD, Strasser AA, O’Connor R, Stauffer RL, et al. Smoking status and the human dopamine transporter variable number of tandem repeats (VNTR) polymorphism: failure to replicate and finding that never-smokers may be different. Nicotine Tob Res. 2002;4:333–40. doi: 10.1080/14622200210142689. [DOI] [PubMed] [Google Scholar]
  • 305.Lin PY, Tsai G. Association between serotonin transporter gene promoter polymorphism and suicide: results of a meta-analysis. Biol Psychiatry. 2004;55:1023–30. doi: 10.1016/j.biopsych.2004.02.006. [DOI] [PubMed] [Google Scholar]
  • 306.Munafo MR, Clark T, Flint J. Does measurement instrument moderate the association between the serotonin transporter gene and anxiety-related personality traits? A meta-analysis. Mol Psychiatry. 2005;10:415–9. doi: 10.1038/sj.mp.4001627. [DOI] [PubMed] [Google Scholar]
  • 307.Strug LJ, Suresh R, Fyer AJ, Talati A, Adams PB, Li W, et al. Panic disorder is associated with the serotonin transporter gene (SLC6A4) but not the promoter region (5-HTTLPR) Mol Psychiatry. 2010;15:166–76. doi: 10.1038/mp.2008.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Noskova TG, Kazantseva AV, Gareeva AE, Gaĭsina DA, Tuktarova SU, Khusnutdinova EK. [Association of several polymorphic loci of serotoninergic genes with unipolar depression] Genetika. 2009;45:842–8. [PubMed] [Google Scholar]
  • 309.Lohoff FW, Weller AE, Bloch PJ, Buono RJ, Doyle GA, Ferraro TN, et al. Association between polymorphisms in the vesicular monoamine transporter 1 gene (VMAT1/SLC18A1) on chromosome 8p and schizophrenia. Neuropsychobiology. 2008;57:55–60. doi: 10.1159/000129668. [DOI] [PubMed] [Google Scholar]
  • 310.Richards M, Iijima Y, Kondo H, Shizuno T, Hori H, Arima K, et al. Association study of the vesicular monoamine transporter 1 (VMAT1) gene with schizophrenia in a Japanese population. Behav Brain Funct. 2006;2:39. doi: 10.1186/1744-9081-2-39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Lohoff FW, Dahl JP, Ferraro TN, Arnold SE, Gallinat J, Sander T, et al. Variations in the vesicular monoamine transporter 1 gene (VMAT1/SLC18A1) are associated with bipolar i disorder. Neuropsychopharmacology. 2006;31:2739–47. doi: 10.1038/sj.npp.1301196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Kooij JS, Boonstra AM, Vermeulen SH, Heister AG, Burger H, Buitelaar JK, et al. Response to methylphenidate in adults with ADHD is associated with a polymorphism in SLC6A3 (DAT1) Am J Med Genet B Neuropsychiatr Genet. 2008;147B:201–8. doi: 10.1002/ajmg.b.30586. [DOI] [PubMed] [Google Scholar]
  • 313.Joober R, Grizenko N, Sengupta S, Amor LB, Schmitz N, Schwartz G, et al. Dopamine transporter 3′-UTR VNTR genotype and ADHD: a pharmaco-behavioural genetic study with methylphenidate. Neuropsychopharmacology. 2007;32:1370–6. doi: 10.1038/sj.npp.1301240. [DOI] [PubMed] [Google Scholar]
  • 314.Kirchheiner J, Nickchen K, Sasse J, Bauer M, Roots I, Brockmöller J. A 40-basepair VNTR polymorphism in the dopamine transporter (DAT1) gene and the rapid response to antidepressant treatment. Pharmacogenomics J. 2007;7:48–55. doi: 10.1038/sj.tpj.6500398. [DOI] [PubMed] [Google Scholar]
  • 315.Stapleton JA, Sutherland G, O’Gara C. Association between dopamine transporter genotypes and smoking cessation: a meta-analysis. Addict Biol. 2007;12:221–6. doi: 10.1111/j.1369-1600.2007.00058.x. [DOI] [PubMed] [Google Scholar]
  • 316.O’Gara C, Stapleton J, Sutherland G, Guindalini C, Neale B, Breen G, et al. Dopamine transporter polymorphisms are associated with short-term response to smoking cessation treatment. Pharmacogenet Genomics. 2007;17:61–7. doi: 10.1097/01.fpc.0000236328.18928.4c. [DOI] [PubMed] [Google Scholar]
  • 317.Xu M, Xing Q, Li S, Zheng Y, Wu S, Gao R, et al. Pharmacogenetic effects of dopamine transporter gene polymorphisms on response to chlorpromazine and clozapine and on extrapyramidal syndrome in schizophrenia. Prog Neuropsychopharmacol Biol Psychiatry. 2010 May 24; doi: 10.1016/j.pnpbp.2010.05.017. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 318.Franke B, Vasquez AA, Johansson S, Hoogman M, Romanos J, Boreatti-Hümmer A, et al. Multicenter analysis of the SLC6A3/DAT1 VNTR haplotype in persistent ADHD suggests differential involvement of the gene in childhood and persistent ADHD. Neuropsychopharmacology. 2010;35:656–64. doi: 10.1038/npp.2009.170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Pal P, Mihanović M, Molnar S, Xi H, Sun G, Guha S, et al. Association of tagging single nucleotide polymorphisms on 8 candidate genes in dopaminergic pathway with schizophrenia in Croatian population. Croat Med J. 2009;50:361–9. doi: 10.3325/cmj.2009.50.361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Sieminska A, Buczkowski K, Jassem E, Niedoszytko M, Tkacz E. Influences of polymorphic variants of DRD2 and SLC6A3 genes, and their combinations on smoking in Polish population. BMC Med Genet. 2009;10:92. doi: 10.1186/1471-2350-10-92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Sáiz PA, García-Portilla MP, Arango C, Morales B, Arias B, Corcoran P, et al. Genetic polymorphisms in the dopamine-2 receptor (DRD2), dopamine-3 receptor (DRD3), and dopamine transporter (SLC6A3) genes in schizophrenia: Data from an association study. Prog Neuropsychopharmacol Biol Psychiatry. 2010;34:26–31. doi: 10.1016/j.pnpbp.2009.09.008. [DOI] [PubMed] [Google Scholar]
  • 322.Bergen AW, Conti DV, Van Den Berg D, Lee W, Liu J, Li D, et al. Dopamine genes and nicotine dependence in treatment-seeking and community smokers. Neuropsychopharmacology. 2009;34:2252–64. doi: 10.1038/npp.2009.52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Lee SS, Chronis-Tuscano A, Keenan K, Pelham WE, Loney J, Van Hulle CA, et al. Association of maternal dopamine transporter genotype with negative parenting: evidence for gene x environment interaction with child disruptive behavior. Mol Psychiatry. 2010;15:548–58. doi: 10.1038/mp.2008.102. [DOI] [PubMed] [Google Scholar]
  • 324.Neuman RJ, Lobos E, Reich W, Henderson CA, Sun LW, Todd RD. Prenatal smoking exposure and dopaminergic genotypes interact to cause a severe ADHD subtype. Biol Psychiatry. 2007;61:1320–8. doi: 10.1016/j.biopsych.2006.08.049. [DOI] [PubMed] [Google Scholar]
  • 325.Frankle WG, Narendran R, Huang Y, Hwang DR, Lombardo I, Cangiano C, et al. Serotonin Transporter Availability in Patients with Schizophrenia: A Positron Emission Tomography Imaging Study with [(11)C]DASB. Biol Psychiatry. 2005;57:1510–6. doi: 10.1016/j.biopsych.2005.02.028. [DOI] [PubMed] [Google Scholar]
  • 326.Little KY, McLaughlin DP, Zhang L, Livermore CS, Dalack GW, McFinton PR, et al. Cocaine, ethanol, and genotype effects on human midbrain serotonin transporter binding sites and mRNA levels. Am J Psychiatry. 1998;155:207–13. doi: 10.1176/ajp.155.2.207. [DOI] [PubMed] [Google Scholar]
  • 327.Heils A, Teufel A, Petri S, Stober G, Riederer P, Bengel D, et al. Allelic variation of human serotonin transporter gene expression. J Neurochem. 1996;66:2621–4. doi: 10.1046/j.1471-4159.1996.66062621.x. [DOI] [PubMed] [Google Scholar]
  • 328.Eddahibi S, Humbert M, Fadel E, Raffestin B, Darmon M, Capron F, et al. Serotonin transporter overexpression is responsible for pulmonary artery smooth muscle hyperplasia in primary pulmonary hypertension. J Clin Invest. 2001;108:1141–50. doi: 10.1172/JCI12805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Lesch KP, Bengel D, Heils A, Sabol SZ, Greenberg BD, Petri S, et al. Association of anxiety-related traits with a polymorphism in the serotonin transporter gene regulatory region. Science. 1996;274:1527–31. doi: 10.1126/science.274.5292.1527. [DOI] [PubMed] [Google Scholar]
  • 330.Bonvicini C, Minelli A, Scassellati C, Bortolomasi M, Segala M, Sartori R, et al. Serotonin transporter gene polymorphisms and treatment-resistant depression. Prog Neuropsychopharmacol Biol Psychiatry. 2010 May 5; doi: 10.1016/j.pnpbp.2010.04.020. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 331.Min W, Li T, Ma X, Li Z, Yu T, Gao D, et al. Monoamine transporter gene polymorphisms affect susceptibility to depression and predict antidepressant response. Psychopharmacology (Berl) 2009;205:409–17. doi: 10.1007/s00213-009-1550-3. [DOI] [PubMed] [Google Scholar]
  • 332.Koenen KC, Aiello AE, Bakshis E, Amstadter AB, Ruggiero KJ, Acierno R, et al. Modification of the association between serotonin transporter genotype and risk of posttraumatic stress disorder in adults by county-level social environment. Am J Epidemiol. 2009;169:704–11. doi: 10.1093/aje/kwn397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Mrazek DA, Rush AJ, Biernacka JM, O’Kane DJ, Cunningham JM, Wieben ED, et al. SLC6A4 variation and citalopram response. Am J Med Genet B Neuropsychiatr Genet. 2009;150B:341–51. doi: 10.1002/ajmg.b.30816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Hu XZ, Rush AJ, Charney D, Wilson AF, Sorant AJ, Papanicolaou GJ, et al. Association between a functional serotonin transporter promoter polymorphism and citalopram treatment in adult outpatients with major depression. Arch Gen Psychiatry. 2007;64:783–92. doi: 10.1001/archpsyc.64.7.783. [DOI] [PubMed] [Google Scholar]
  • 335.Wilkie MJ, Smith G, Day RK, Matthews K, Smith D, Blackwood D, et al. Polymorphisms in the SLC6A4 and HTR2A genes influence treatment outcome following antidepressant therapy. Pharmacogenomics J. 2009;9:61–70. doi: 10.1038/sj.tpj.6500491. [DOI] [PubMed] [Google Scholar]
  • 336.Otte C, McCaffery J, Ali S, Whooley MA. Association of a serotonin transporter polymorphism (5-HTTLPR) with depression, perceived stress, and norepinephrine in patients with coronary disease: the Heart and Soul Study. Am J Psychiatry. 2007;164:1379–84. doi: 10.1176/appi.ajp.2007.06101617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Mössner R, Henneberg A, Schmitt A, Syagailo YV, Grässle M, Hennig T, et al. Allelic variation of serotonin transporter expression is associated with depression in Parkinson’s disease. Mol Psychiatry. 2001;6:350–2. doi: 10.1038/sj.mp.4000849. [DOI] [PubMed] [Google Scholar]
  • 338.Lasky-Su J, Neale BM, Franke B, Anney RJ, Zhou K, Maller JB, et al. Genome-wide association scan of quantitative traits for attention deficit hyperactivity disorder identifies novel associations and confirms candidate gene associations. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1345–54. doi: 10.1002/ajmg.b.30867. [DOI] [PubMed] [Google Scholar]
  • 339.Mick E, Neale B, Middleton FA, McGough JJ, Faraone SV. Genome-wide association study of response to methylphenidate in 187 children with attention-deficit/hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1412–8. doi: 10.1002/ajmg.b.30865. [DOI] [PubMed] [Google Scholar]
  • 340.Little KY, Krolewski DM, Zhang L, Cassin BJ. Loss of striatal vesicular monoamine transporter protein (VMAT2) in human cocaine users. Am J Psychiatry. 2003;160:47–55. doi: 10.1176/appi.ajp.160.1.47. [DOI] [PubMed] [Google Scholar]
  • 341.Gilman S, Koeppe RA, Adams KM, Junck L, Kluin KJ, Johnson-Greene D, et al. Decreased striatal monoaminergic terminals in severe chronic alcoholism demonstrated with (+)[11C]dihydrotetrabenazine and positron emission tomography. Ann Neurol. 1998;44:326–33. doi: 10.1002/ana.410440307. [DOI] [PubMed] [Google Scholar]
  • 342.Hansson SR, Hoffman BJ, Mezey E. Ontogeny of vesicular monoamine transporter mRNAs VMAT1 and VMAT2. I. The developing rat central nervous system. Brain Res Dev Brain Res. 1998;110:135–58. doi: 10.1016/s0165-3806(98)00104-7. [DOI] [PubMed] [Google Scholar]
  • 343.Buckley NA, McManus PR. Fatal toxicity of serotoninergic and other antidepressant drugs: analysis of United Kingdom mortality data. British Medical Journal. 2003;325:1332–3. doi: 10.1136/bmj.325.7376.1332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 344.Kulkarni SK, Dhir A. Current investigational drugs for Major Depression. Expert Opinion on Investigationa. 2009;18:767–88. doi: 10.1517/13543780902880850. [DOI] [PubMed] [Google Scholar]
  • 345.Schatzberg AF. Pharmacological principles of antidepressant efficacy. Human Psychopharmacology. 2002;17:S17–22. doi: 10.1002/hup.399. [DOI] [PubMed] [Google Scholar]
  • 346.Youdim MB, Weinstock M. Therapeutic applications of selective and non-selective inhibitors of monoamine oxidase A and B that do not cause significant tyramine potentiation. Neurotoxicology. 2004;25:243–50. doi: 10.1016/S0161-813X(03)00103-7. [DOI] [PubMed] [Google Scholar]
  • 347.Kessing LV, Hansen MG, Andersen PK, Angst J. The predictive effect of episodes on the risk of recurrence in depressive and bipolar disorders-a life-long perspective. Acta Psychiatria Scandinavian. 2004;109:339–44. doi: 10.1046/j.1600-0447.2003.00266.x. [DOI] [PubMed] [Google Scholar]
  • 348.Baldessarini RJ, Ghaemi SN, Viguera AC. Tolerance in antidepressant treatment. Psychotherapy and Psychosomatics. 2002;71:177–9. doi: 10.1159/000063641. [DOI] [PubMed] [Google Scholar]
  • 349.Mischoulon D, Opitz G, Kelly K, Fava M, Rosenbaum J. A Preliminary open study of the tolerability and effectiveness of Nefazodone in Major Depressive Disorder: Comparing patients who recently discontinued an SSRI with those on no recent antidepressant treatment. Depression and Anxiety. 2004;19:43–50. doi: 10.1002/da.10127. [DOI] [PubMed] [Google Scholar]
  • 350.Ruhé HG, Huyser J, Swinkels JA, Schene AH. Switching antidepressants after a first selective serotonin reuptake inhibitor in major depressive disorder: A systematic review. Journal of Clinical Psychiatry. 2006;67:1836–1855. doi: 10.4088/jcp.v67n1203. [DOI] [PubMed] [Google Scholar]
  • 351.Pundiak TM, Case BG, Peselow ED, Mulcare L. Discontinuation of maintenance selective serotonin reuptake inhibitor monotherapy after 5 years of stable response: A naturalistic study. Journal of Clinical Psychiatry. 2008;69:1811–7. doi: 10.4088/jcp.v69n1117. [DOI] [PubMed] [Google Scholar]
  • 352.Pato MT, Hill JL, Murphy DL. A clomipramine dosage reduction study in the course of long-term treatment of obsessive-compulsive disorder patients. Psychopharmacology Bulletin. 1990;26:211–4. [PubMed] [Google Scholar]
  • 353.Runyon SP, Carroll FI. Dopamine transporter ligands: recent developments and therapeutic potential. Curr Top Med Chem. 2006;6:1825–43. doi: 10.2174/156802606778249775. [DOI] [PubMed] [Google Scholar]

RESOURCES