14.01.2013 Views

Regulation of the dopamine transporter - Addiction Research ...

Regulation of the dopamine transporter - Addiction Research ...

Regulation of the dopamine transporter - Addiction Research ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

ANNALS OF THE NEW YORK ACADEMY OF SCIENCES<br />

Issue: <strong>Addiction</strong> Reviews 2<br />

<strong>Regulation</strong> <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong><br />

Aspects relevant to psychostimulant drugs <strong>of</strong> abuse<br />

Kyle C. Schmitt 1 and Maarten E. A. Reith 1,2<br />

Ann. N.Y. Acad. Sci. ISSN 0077-8923<br />

1 2 Department <strong>of</strong> Pharmacology, New York University School <strong>of</strong> Medicine, New York, New York, USA. Department <strong>of</strong><br />

Psychiatry, New York University School <strong>of</strong> Medicine, New York, New York, USA<br />

Address for correspondence: Maarten E. A. Reith, New York University School <strong>of</strong> Medicine, 550 First Ave., New York, NY<br />

10016. maarten.reith@med.nyu.edu<br />

Dopaminergic signaling in <strong>the</strong> brain is primarily modulated by <strong>dopamine</strong> <strong>transporter</strong>s (DATs), which actively<br />

translocate extraneuronal <strong>dopamine</strong> back into <strong>dopamine</strong>rgic neurons. Transporter proteins are highly dynamic,<br />

continuously trafficking between plasmalemmal and endosomal membranes. Changes in DAT membrane trafficking<br />

kinetics can rapidly regulate <strong>dopamine</strong>rgic tone by altering <strong>the</strong> number <strong>of</strong> <strong>transporter</strong>s present at <strong>the</strong> cell surface.<br />

Various psychostimulant DAT ligands—acting ei<strong>the</strong>r as amphetamine-like substrates or cocaine-like nontranslocated<br />

inhibitors—affect <strong>transporter</strong> trafficking, triggering rapid insertion or removal <strong>of</strong> plasmalemmal DATs. In this<br />

review, we focus on <strong>the</strong> effects <strong>of</strong> psychostimulants <strong>of</strong> addiction (particularly d-methamphetamine and cocaine)<br />

on DAT regulation and membrane trafficking, with an emphasis on how <strong>the</strong>se drugs may influence intracellular<br />

signaling cascades and <strong>transporter</strong>-associated scaffolding proteins to affect DAT regulation. In addition, we consider<br />

involvement <strong>of</strong> presynaptic receptors for <strong>dopamine</strong> and o<strong>the</strong>r ligands in DAT regulation. Finally, we discuss possible<br />

implications <strong>of</strong> <strong>transporter</strong> regulation to <strong>the</strong> putative toxicity <strong>of</strong> several substituted amphetamine derivatives<br />

commonly used as recreational drugs, as well as to <strong>the</strong> design <strong>of</strong> <strong>the</strong>rapeutics for cocaine addiction.<br />

Keywords: <strong>dopamine</strong> <strong>transporter</strong>; trafficking; regulation; amphetaminergic substrates; internalization; cocaine;<br />

methamphetamine; MDMA<br />

Introduction<br />

The neuronal <strong>dopamine</strong> <strong>transporter</strong> (DAT)—a<br />

member <strong>of</strong> <strong>the</strong> neurotransmitter sodium symporter<br />

protein superfamily (<strong>the</strong> SLC6 gene family)—<br />

regulates <strong>dopamine</strong>rgic neurotransmission in <strong>the</strong><br />

brain by actively clearing extraneuronal <strong>dopamine</strong>,<br />

using <strong>the</strong> energy <strong>of</strong> cellular ionic electrochemical<br />

gradients. 1 Members <strong>of</strong> <strong>the</strong> neurotransmitter<br />

sodium symporter protein family are glycoproteins<br />

with 12 transmembrane-spanning domains, with<br />

both <strong>the</strong> amino and carboxyl termini located on<br />

<strong>the</strong> inner face <strong>of</strong> <strong>the</strong> membrane. 2 The DAT plays<br />

an integral role in cognition, affect, behavioral reinforcement,<br />

and motor function, and DAT pathologies<br />

are suspected to contribute to disorders, such as<br />

depression, attention deficit–hyperactivity disorder,<br />

anhedonia, Parkinson’s disease, and addiction. 2,3<br />

The DAT is a target <strong>of</strong> several clinically used drugs,<br />

such as <strong>the</strong> psychostimulants methylphenidate, d-<br />

amphetamine, and modafinil, as well as <strong>the</strong> antidepressant<br />

bupropion. In addition, <strong>the</strong> reinforcing and<br />

euphoric effects <strong>of</strong> <strong>the</strong> powerfully addictive psychostimulants<br />

cocaine and d-methamphetamine<br />

(“crystal meth,” a vastly more potent analogue <strong>of</strong><br />

amphetamine, <strong>of</strong>ten administered in large doses by<br />

vaporization) are primarily mediated by interaction<br />

with <strong>the</strong> DAT. 4 <strong>Research</strong>ers in several disciplines<br />

have empirically demonstrated <strong>the</strong> integral nature <strong>of</strong><br />

<strong>the</strong> DAT protein in both <strong>the</strong> acute and chronic effects<br />

<strong>of</strong> cocaine and in cocaine addiction (see, e.g., Mash 5<br />

for review). A major recent finding is that DATknockout<br />

mice show an attenuated response to cocaine<br />

and amphetamines and a reduced preference<br />

for cocaine under self-administration paradigms. 6<br />

These mice, however, still self-administer cocaine—<br />

although more sessions were needed to meet<br />

self-administration criteria 7 —indicating that developmentally<br />

compensatory non<strong>dopamine</strong>rgic mechanisms<br />

can mediate cocaine-taking behavior in<br />

doi: 10.1111/j.1749-6632.2009.05148.x<br />

316 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

DAT-lacking animals. In particular, cocaine’s interaction<br />

with <strong>the</strong> serotonin <strong>transporter</strong> (SERT)<br />

has been hypo<strong>the</strong>sized to contribute to continued<br />

self-administration in DAT-knockout animals. 7,8<br />

Never<strong>the</strong>less, convincing preclinical evidence implicating<br />

specific activity at <strong>the</strong> DAT in cocaine dependence<br />

comes from <strong>the</strong> observation that <strong>the</strong> reinforcing<br />

effect <strong>of</strong> cocaine is lost in transgenic mice<br />

expressing a triple point–mutated DAT with preserved<br />

(albeit somewhat reduced) substrate translocation<br />

but little appreciable affinity for cocaine. 9<br />

It is conceivable that elevated basal <strong>dopamine</strong>rgic<br />

tone in <strong>the</strong>se mice (a result <strong>of</strong> <strong>the</strong>ir reducedfunction<br />

DAT is<strong>of</strong>orm) causes adaptive changes,<br />

altering <strong>the</strong> response to cocaine; however, knockdown<br />

mutant mice with a ∼90% reduction in<br />

DAT expression—but with functionally unmodified<br />

DATs and elevated basal <strong>dopamine</strong>—still exhibit robust,<br />

wild-type–like preference for cocaine, indicating<br />

that interaction with <strong>the</strong> DAT is necessary for<br />

<strong>the</strong> reinforcing effects <strong>of</strong> cocaine in animals that<br />

carry <strong>the</strong> DAT. 10 Moreover, a recent investigation<br />

by Thomsen et al. showed that a different strain<br />

<strong>of</strong> DAT –/– (DAT knockout) mice generally fail to<br />

acquire intravenous cocaine self-administration behavior.<br />

11 Importantly, in <strong>the</strong> minority <strong>of</strong> mice that<br />

do initially show self-administration, this behavior<br />

is quickly extinguished as <strong>the</strong> response requirement<br />

(<strong>the</strong> “work” required to obtain a dose <strong>of</strong> cocaine) is<br />

increased. In contrast, both wild-type and SERT –/–<br />

mice maintain consistent self-administration, even<br />

when <strong>the</strong> response requirement for a single dose <strong>of</strong><br />

intravenous cocaine is increased 50- to 100-fold. Unlike<br />

cocaine, o<strong>the</strong>r <strong>dopamine</strong>rgic compounds (i.e.,<br />

<strong>the</strong> <strong>dopamine</strong> D1-like receptor agonist SKF82958)<br />

and food rewards are comparably administered by<br />

DAT –/– and wild-type mice, indicating that elimination<br />

<strong>of</strong> <strong>the</strong> DAT does not globally hinder behavioral<br />

reinforcement. These data strongly suggest that cocaine<br />

is not a reliable reinforcer in <strong>the</strong> absence <strong>of</strong> <strong>the</strong><br />

DAT, despite some intersubject differences in initial<br />

responses among DAT –/– mice. 11<br />

The DAT governs both <strong>the</strong> duration and magnitude<br />

<strong>of</strong> <strong>dopamine</strong>rgic neurotransmission by actively<br />

translocating <strong>dopamine</strong> from <strong>the</strong> extracellular space<br />

into presynaptic neurons. Extracellular <strong>dopamine</strong> is<br />

also subject to enzymatic catabolism—thus terminating<br />

its signaling potential—however, studies <strong>of</strong><br />

DAT-knockout mice 6,12 have demonstrated that <strong>the</strong><br />

DAT is <strong>the</strong> chief arbiter <strong>of</strong> <strong>dopamine</strong>rgic signal-<br />

ing flux (for review, see Gainetdinov and Caron 3 ).<br />

Dopaminergic transmission is volume transmission<br />

with an extrasynaptic predominance <strong>of</strong> <strong>dopamine</strong><br />

receptors and <strong>transporter</strong>s; <strong>the</strong> function <strong>of</strong> <strong>the</strong> DAT<br />

is not to remove <strong>dopamine</strong> from <strong>the</strong> synapse but<br />

ra<strong>the</strong>r to regulate <strong>dopamine</strong>’s action at extrasynaptic<br />

<strong>dopamine</strong> receptors (see Rice and Cragg 13 for<br />

review and highly recommended, updated cartoon<br />

<strong>of</strong> a <strong>dopamine</strong> synapse). The overall rate <strong>of</strong> de novo<br />

DAT protein syn<strong>the</strong>sis is a relatively slow process—<br />

<strong>the</strong> half-life <strong>of</strong> striatal <strong>transporter</strong> proteins, for example,<br />

is approximately 2–3 days. 14,15 Hence, it is<br />

not surprising that cells use an assortment <strong>of</strong> posttranslational<br />

regulatory strategies to rapidly alter<br />

DAT function—without requiring de novo protein<br />

syn<strong>the</strong>sis or transcriptional-level modifications—to<br />

dynamically shift <strong>the</strong> tone <strong>of</strong> <strong>dopamine</strong>rgic activity<br />

in response to various endogenous and exogenous<br />

stimuli. There are two conceivable manners<br />

in which a cell could modulate DAT function “on<br />

<strong>the</strong> fly” (on a time scale <strong>of</strong> seconds to minutes) in<br />

response to changing environmental conditions—<br />

ei<strong>the</strong>r via direct modification <strong>of</strong> single-<strong>transporter</strong><br />

parameters, such as intrinsic substrate permeation<br />

efficacy, or by affecting <strong>the</strong> trafficking <strong>of</strong> <strong>the</strong> <strong>transporter</strong>,<br />

redistributing DATs between <strong>the</strong> plasma<br />

membrane and intracellular endosomal compartments.<br />

4 Experimental evidence has demonstrated<br />

that both processes may play some role in acute DAT<br />

regulation, but that redistribution <strong>of</strong> <strong>transporter</strong>s<br />

to and from <strong>the</strong> cell surface is <strong>the</strong> predominant<br />

mechanism. 16,17<br />

This regulatory method—rapid insertion or removal<br />

<strong>of</strong> plasmalemmal <strong>transporter</strong> proteins—may<br />

seem like a cumbersome way <strong>of</strong> acutely controlling<br />

neurotransmitter signaling; however, it makes logistical<br />

sense when we consider <strong>the</strong> kinetics <strong>of</strong> <strong>the</strong><br />

substrate translocation cycle. That is, <strong>transporter</strong>s<br />

must undergo a conformational shift between <strong>the</strong><br />

discrete phases <strong>of</strong> extracellular substrate binding<br />

(outward facing) and cytosolic substrate release (inward<br />

facing), limiting <strong>the</strong> potential maximum uptake<br />

velocity to around one substrate molecule per<br />

second per <strong>transporter</strong>. 4 Ion channels, by contrast,<br />

can achieve flux velocities on <strong>the</strong> order <strong>of</strong> millions <strong>of</strong><br />

ions per second per channel. 18 Because <strong>the</strong>y <strong>of</strong>fer a<br />

far wider dynamic range in “substrate” flux rate, ion<br />

channels are more amenable to regulation strategies<br />

involving posttranslational modification at <strong>the</strong> individual<br />

protein level than <strong>transporter</strong>s, such as <strong>the</strong><br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 317


DAT <strong>Regulation</strong> Schmitt & Reith<br />

Figure 1. The many parallel factors influencing DAT trafficking and membrane distribution. (A) The DAT undergoes<br />

continuous constitutive recycling between <strong>the</strong> plasma membrane and early endosomal compartments. (B) and (C)<br />

Substrates, such as amphetamine trigger internalization <strong>of</strong> plasmalemmal DATs, whereas cocaine-like inhibitors<br />

upregulate surface DAT expression. (D) and (E) Alterations in <strong>the</strong> phosphorylation state <strong>of</strong> DAT and associated<br />

scaffolding proteins by intracellular signaling cascades modulates <strong>transporter</strong> trafficking—although direct phosphorylation<br />

<strong>of</strong> DAT proteins is probably not caused by PKC, this kinase is an integral participant in DAT dynamics. (F)<br />

Physical coupling <strong>of</strong> <strong>the</strong> DAT to <strong>dopamine</strong> D2 receptors encourages recruitment <strong>of</strong> DAT to <strong>the</strong> plasma membrane.<br />

(G) PKC is also involved in ubiquitin-mediated DAT regulation, with monoubiquitination hypo<strong>the</strong>sized to target<br />

internalized DAT to <strong>the</strong> lysosomal degradation pathway. (In color in Annals online.)<br />

DAT. Moreover, as Deken et al. have proposed, <strong>the</strong><br />

ability to rapidly alter <strong>transporter</strong> trafficking may<br />

allow cells to regulate <strong>the</strong> exocytic release <strong>of</strong> neurotransmitters<br />

and <strong>the</strong> insertion <strong>of</strong> surface <strong>transporter</strong>s<br />

in parallel. 19 This hypo<strong>the</strong>sis was prompted<br />

by <strong>the</strong> finding that <strong>the</strong> GABA <strong>transporter</strong> (GAT1)<br />

is trafficked to and from <strong>the</strong> plasma membrane in<br />

recycling vesicles similar in size and morphology<br />

to synaptic vesicles (but lacking synaptophysin and<br />

<strong>the</strong> vesicular GABA <strong>transporter</strong>) and that <strong>the</strong> rate<br />

<strong>of</strong> recycling depends on intracellular Ca 2+ concentration,<br />

indicating that <strong>transporter</strong> trafficking may<br />

indeed be coupled to transmitter release. 19<br />

<strong>Regulation</strong><strong>of</strong>DATactivitycanbetriggeredby<br />

a myriad <strong>of</strong> exogenous factors and cellular events,<br />

such as (1) small-molecule ligands targeting ei<strong>the</strong>r<br />

<strong>the</strong> DAT itself (i.e., substrates and inhibitors) or<br />

presynaptic G protein-coupled receptors (GPCRs)<br />

that affect DAT, (2) enzymatic modification (e.g.,<br />

phosphorylation) involving intracellular secondmessenger<br />

cascades, and (3) protein–protein<br />

interactions between <strong>the</strong> DAT and o<strong>the</strong>r transmembrane<br />

or cytoskeletal scaffolding proteins. A graphical<br />

overview <strong>of</strong> <strong>the</strong> putative factors influencing<br />

DAT membrane trafficking is given in Figure 1. Our<br />

primary focus in this review will be on short-term<br />

DAT regulation in response to acute exposure to psychostimulants,<br />

such as cocaine- and amphetaminelike<br />

drugs. Although <strong>the</strong> complete signaling mechanisms<br />

underlying substrate- and inhibitor-mediated<br />

318 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

DAT regulation have yet to be elucidated, we<br />

will attempt to highlight instances where DATaffecting<br />

drugs may directly influence intracellular<br />

signaling cascades (which can, in turn, affect<br />

DAT regulation). We will also briefly discuss <strong>the</strong> involvement<br />

<strong>of</strong> presynaptic GPCRs and <strong>transporter</strong>associated<br />

scaffolding proteins in DAT regulation.<br />

Finally, <strong>the</strong> possible implications <strong>of</strong> <strong>transporter</strong><br />

protein regulation to <strong>the</strong> putative toxicity <strong>of</strong> several<br />

substituted amphetamine derivatives will be<br />

discussed.<br />

Direct regulation <strong>of</strong> <strong>transporter</strong> function by<br />

DAT ligands<br />

Effects <strong>of</strong> <strong>transporter</strong> substrates<br />

Substrates that are actively translocated by <strong>the</strong><br />

DAT (o<strong>the</strong>r than <strong>dopamine</strong> itself) include endogenous<br />

trace amines, such as tyramine and �phenethylamine,<br />

<strong>the</strong> neurotoxic mitochondrial poison<br />

1-methyl-4-phenylpyridinium (MPP + ), and<br />

<strong>the</strong> amphetamines, such as <strong>the</strong> clinically used<br />

stimulant d-amphetamine and its more potent<br />

congener d-methamphetamine, a highly-abused<br />

addictive drug. 20 Extracellular substrates can produce<br />

ei<strong>the</strong>r transient upregulation or significant<br />

downregulation <strong>of</strong> DAT activity, depending on <strong>the</strong><br />

duration <strong>of</strong> substrate exposure. Evidence that substrates<br />

for <strong>the</strong> DAT can induce rapid alterations<br />

in <strong>transporter</strong> activity was first obtained in striatal<br />

synaptosomes prepared from rats 1 h after treatment<br />

with high-dose d-methamphetamine (15 mg/kg <strong>of</strong><br />

body weight). At this dose, acute methamphetamine<br />

administration resulted in a 65% decrease in synaptosomal<br />

[ 3H]<strong>dopamine</strong> uptake compared with<br />

vehicle. 21 Fur<strong>the</strong>r analysis showed that this<br />

downregulation is transient—it was observed in<br />

synaptosomes prepared 1 h after, but not 24 h after,<br />

methamphetamine treatment—and is due to<br />

a treatment-associated drop in transport velocity<br />

(V max), not an alteration in <strong>the</strong> <strong>transporter</strong>’s affinity<br />

(Km) for <strong>dopamine</strong> and is not associated with<br />

changes in overall DAT protein concentration. 22<br />

These data are consistent with <strong>the</strong> hypo<strong>the</strong>sis<br />

that methamphetamine rapidly and transiently diminishes<br />

<strong>the</strong> activity <strong>of</strong> <strong>the</strong> DAT at <strong>the</strong> cell surface;<br />

however, <strong>the</strong>y do not distinguish between a<br />

trafficking-dependent mechanism (internalization<br />

<strong>of</strong> plasmalemmal <strong>transporter</strong>s) and a posttranslational<br />

regulatory effect at <strong>the</strong> individual <strong>transporter</strong><br />

level. Support for a trafficking-dependent mechanism<br />

has come from fur<strong>the</strong>r research employing<br />

fluorescent microscopy to visualize tagged human<br />

DAT (hDAT) fusion proteins expressed in cultured<br />

cells. For example, Saunders et al. demonstrated that<br />

exposing HEK293 cells transfected with fluorescenttagged<br />

hDAT to d-amphetamine triggers migration<br />

<strong>of</strong> <strong>the</strong> fluorescent DAT from <strong>the</strong> cell surface<br />

to <strong>the</strong> intracellular compartment, with visible DAT<br />

accumulation in punctate intracellular vesicles. 23<br />

In line with <strong>the</strong> previous findings for methamphetamine,<br />

preincubation <strong>of</strong> <strong>the</strong> HEK-hDAT cells<br />

with amphetamine for 1 h significantly reduced<br />

cellular [ 3 H]<strong>dopamine</strong> uptake, affecting <strong>the</strong> V max<br />

transport parameter but not <strong>the</strong> Km for <strong>dopamine</strong>.<br />

Amphetamine-induced <strong>transporter</strong> internalization<br />

was apparent after as little as 20 min (maximal at 1-h<br />

exposure time) and was found to be dependent on<br />

internalization <strong>of</strong> <strong>the</strong> DAT by clathrin-coated vesicles,<br />

as coexpression <strong>of</strong> a dominant-negative form<br />

<strong>of</strong> dynamin I—a GTPase responsible for cleaving<br />

nascent clathrin-coated vesicles budding from <strong>the</strong><br />

plasma membrane—in <strong>the</strong> HEK-hDAT cells prevented<br />

<strong>the</strong> internalizing effect <strong>of</strong> amphetamine. 23<br />

The authors also noted that <strong>dopamine</strong> itself causes<br />

a similar redistribution <strong>of</strong> plasmalemmal DAT to<br />

intracellular vesicles (albeit at a higher dosage than<br />

amphetamine), implying that regulation <strong>of</strong> surface<br />

DAT expression is a general effect <strong>of</strong> substratelike<br />

compounds. Extending this qualitative observation,<br />

Chi and Reith found that pretreatment with<br />

<strong>dopamine</strong> for 1 h decreases surface DAT expression<br />

by approximately 30% in both HEK-hDAT cells<br />

and rat striatal synaptosomes, an effect that cleavable<br />

biotinylation assays suggest is due to enhanced<br />

endocytosis <strong>of</strong> <strong>the</strong> <strong>transporter</strong> protein. 24 Fur<strong>the</strong>r<br />

corroboration <strong>of</strong> increased internalization and sequestration<br />

<strong>of</strong> surface-localized <strong>transporter</strong>s in response<br />

to substrates is provided by <strong>the</strong> fluorescent<br />

microscopy studies <strong>of</strong> Sorkina et al., who visualized<br />

amphetamine-mediated DAT endocytosis in<br />

living cells and later demonstrated that internalized<br />

DATs colocalize with well-characterized markers <strong>of</strong><br />

early and recycling endosomes, such as Rab5 and<br />

EEA1. 25,26<br />

In what quaternary form is DAT internalized by<br />

<strong>the</strong> action <strong>of</strong> amphetamine? Is <strong>the</strong>re a relationship<br />

between <strong>the</strong> oligomerization <strong>of</strong> DAT and its internalization?<br />

We found that in hDAT-expressing HEK-<br />

293 cells, d-amphetamine shifted <strong>the</strong> distribution <strong>of</strong><br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 319


DAT <strong>Regulation</strong> Schmitt & Reith<br />

surface DAT toward a smaller ratio <strong>of</strong> oligomers to<br />

monomers in what appears to reflect a dissociation<br />

<strong>of</strong> DAT oligomers. 27 Along with <strong>the</strong> reduction <strong>of</strong> <strong>the</strong><br />

fraction <strong>of</strong> oligomerized DAT, <strong>the</strong> amount <strong>of</strong> surface<br />

DAT was decreased, and blocking endocytosis with<br />

phenylarsine oxide or sucrose counteracted <strong>the</strong>se<br />

effects <strong>of</strong> amphetamine. 27 We speculated that DAT<br />

at <strong>the</strong> cell surface is distributed between oligomers<br />

and monomers and that monomers are internalized;<br />

d-amphetamine could promote <strong>the</strong> formation <strong>of</strong><br />

monomers, which <strong>the</strong>n become internalized, reducing<br />

<strong>the</strong> DAT’s presence at <strong>the</strong> surface. When endocytosisispreventedbyanexogenouslyaddedblocker,<br />

monomeric DAT formed by amphetamine remains<br />

at <strong>the</strong> surface and reassociates to form oligomers. In<br />

this scenario, oligomerization and internalization<br />

are linked in <strong>the</strong> effects <strong>of</strong> amphetamine. One can<br />

grasp <strong>the</strong> functional relevance <strong>of</strong> a changed distribution<br />

between oligomerized and monomeric DAT<br />

at <strong>the</strong> surface only once <strong>the</strong> functional properties <strong>of</strong><br />

<strong>the</strong>se forms <strong>of</strong> <strong>the</strong> <strong>transporter</strong> are known.<br />

Substrate-mediated downregulation <strong>of</strong> surface<br />

DAT expression has also been demonstrated with<br />

electrophysiological techniques. In heterologous expression<br />

systems, DAT function is assessed by measuring<br />

<strong>transporter</strong>-associated currents with twoelectrode<br />

voltage clamp recording, whereas in vivo,<br />

microamperometry is used to electrochemically<br />

monitor <strong>dopamine</strong> clearance in real-time. In hDATexpressing<br />

Xenopus oocytes, brief repeated bath perfusion<br />

with <strong>the</strong> substrates <strong>dopamine</strong>, amphetamine<br />

and tyramine (1 min <strong>of</strong> substrate exposure every<br />

5 min, for 1 h in total) gradually reduces DAT<br />

function, as shown by progressive decline <strong>of</strong> substrate<br />

translocation-associated currents. 28 Of <strong>the</strong><br />

three substrates, amphetamine was found to be<br />

<strong>the</strong> most efficacious, producing a significant drop<br />

in current magnitude after <strong>the</strong> fewest repeat perfusions.<br />

This gradual decline in DAT function is<br />

accompanied by a decrease in specific [ 3 H]CFT<br />

binding (Bmax) to intact oocytes, indicating that<br />

<strong>the</strong> reduction in <strong>transporter</strong> currents originates<br />

from loss <strong>of</strong> surface-localized DATs ra<strong>the</strong>r than<br />

modification <strong>of</strong> <strong>the</strong> electrophysiological parameters<br />

<strong>of</strong> individual <strong>transporter</strong>s. Kahlig et al. provided<br />

fur<strong>the</strong>r substantiation <strong>of</strong> this finding—using<br />

a combination <strong>of</strong> fluorescence microscopy and<br />

patch-clamp recording, <strong>the</strong> authors demonstrated<br />

that amphetamine-induced internalization <strong>of</strong> <strong>the</strong><br />

DAT is not paralleled by alteration <strong>of</strong> single-<br />

<strong>transporter</strong> current dynamics, reinforcing <strong>the</strong> idea<br />

that substrate-mediated DAT regulation is <strong>the</strong> result<br />

<strong>of</strong> endocytosis <strong>of</strong> plasmalemmal DAT. 29 In line<br />

with <strong>the</strong> findings in heterologous cells, recurrent<br />

infusion <strong>of</strong> <strong>dopamine</strong> in vivo by reverse microdialysis<br />

(one infusion every 2 min) results in a robust<br />

reduction <strong>of</strong> <strong>dopamine</strong> clearance in rat striatum;<br />

however, no significant change was observed<br />

in <strong>the</strong> nucleus accumbens. 28 This observation implies<br />

that DAT regulation might be an anatomically<br />

specific phenomenon, differing between brain regions.<br />

A similar anatomical discrepancy has been<br />

shown for d-methamphetamine, which decreases<br />

[ 3 H]<strong>dopamine</strong> uptake in striatal synaptosomes but<br />

not in those prepared from <strong>the</strong> nucleus accumbens.<br />

22 A recent study by Richards and Zahniser<br />

also highlights <strong>the</strong> difference between <strong>the</strong> striatum<br />

and <strong>the</strong> nucleus accumbens in terms <strong>of</strong> substrateinduced<br />

DAT regulation: brief (15 min) preincubation<br />

with 20 �M amphetamine reduces <strong>the</strong> V max <strong>of</strong><br />

[ 3 H]<strong>dopamine</strong> uptake in rat striatal synaptosomes<br />

but does not significantly alter <strong>the</strong> V max in synaptosomes<br />

from <strong>the</strong> nucleus accumbens. 30 Curiously,<br />

however, <strong>the</strong> authors found that synaptosomes from<br />

both regions were significantly affected if prepared<br />

from rats treated systemically with 2-mg/kg amphetamine<br />

for 45 min. Fur<strong>the</strong>r inquiry is needed<br />

to delineate potential region-specific differences in<br />

DAT regulation and between in vivo and in vitro<br />

approaches.<br />

Many studies have now conclusively shown<br />

that several substrates, such as <strong>dopamine</strong>, damphetamine,<br />

and d-methamphetamine, can trigger<br />

cytosolic redistribution <strong>of</strong> plasmalemmal DAT.<br />

Nearly all <strong>the</strong>se studies used relatively long substrate<br />

exposure times—anywhere from 15 min to a few<br />

hours. Far less is known about <strong>the</strong> immediate—on<br />

a scale <strong>of</strong> seconds to minutes—effects <strong>of</strong> substrates<br />

on DAT trafficking. A study from <strong>the</strong> Gnegy lab 31<br />

suggests that substrates, such as amphetamine,<br />

may have time-dependent biphasic effects on DAT<br />

membrane trafficking. Using confocal microscopy,<br />

reversible biotinylation, and intact-cell [ 3 H]CFT<br />

(2�-carbomethoxy-3�-(4-fluorophenyl)tropane)<br />

binding, <strong>the</strong> authors demonstrated that acute<br />

exposure to 3 �M d-amphetamine causes an extremely<br />

rapid upregulation in plasmalemmal DAT<br />

expression: surface DAT levels were significantly<br />

increased (on <strong>the</strong> order <strong>of</strong> 60–70%) within 30 s <strong>of</strong><br />

amphetamine treatment and remained elevated for<br />

320 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

around 1 min, before quickly dropping back to baseline<br />

levels within 3 min. In accordance with most<br />

substrate-induced trafficking studies, DAT surface<br />

expression began to drop below vehicle levels<br />

after 20 min <strong>of</strong> amphetamine exposure. Curiously,<br />

<strong>dopamine</strong> itself was not found to induce <strong>the</strong> same<br />

immediate upregulation <strong>of</strong> surface DAT expression<br />

as amphetamine, indicating that this biphasic<br />

trafficking pattern might not extend to all DAT<br />

substrates and may contribute to amphetamine’s<br />

ability to stimulate <strong>dopamine</strong> efflux via <strong>the</strong> DAT. 31<br />

However, a more recent study from this group—<br />

using total internal reflection fluorescent microscopy<br />

to visualize hDAT trafficking dynamics in<br />

N2Aneuroblastomacellswithreal-timetemporal<br />

resolution—shows evidence that both <strong>dopamine</strong><br />

(10 �M) and d-amphetamine (5 �M) induce<br />

changes in surface DAT expression within seconds,<br />

but <strong>the</strong> effect <strong>of</strong> amphetamine persisted longer after<br />

substrate washout. 32<br />

Effects <strong>of</strong> DAT inhibitors<br />

Compared with <strong>the</strong> robust literature on DAT regulation<br />

by phenethylamine substrates, <strong>the</strong>re is a paucity<br />

<strong>of</strong> studies investigating <strong>the</strong> acute regulatory effects<br />

<strong>of</strong> DAT inhibitors. The few in vitro regulatory studies<br />

using heterologous cell systems, as well as <strong>the</strong> bulk<br />

<strong>of</strong> in vivo and clinical studies, have focused solely<br />

on <strong>the</strong> classical DAT inhibitor cocaine, doubtless a<br />

result <strong>of</strong> its notorious reputation as one <strong>of</strong> <strong>the</strong> most<br />

addictive compounds known to humans. 33 Atypical<br />

DAT inhibitors lacking cocaine-like abuse potential<br />

34,35 remain almost entirely uninvestigated;<br />

thus, it is not yet possible to make global statements<br />

regarding <strong>the</strong> regulatory effects <strong>of</strong> inhibitors<br />

o<strong>the</strong>r than cocaine. However, at least for cocaine,<br />

it appears that nontranslocated DAT inhibitors not<br />

only block <strong>the</strong> acute regulatory effects <strong>of</strong> substrates,<br />

<strong>the</strong>y may exert <strong>the</strong> opposite effect: insertion <strong>of</strong> DATs<br />

from <strong>the</strong> endosomic recycling pool into <strong>the</strong> plasma<br />

membrane. 16,36<br />

Chronic administration <strong>of</strong> cocaine upregulates<br />

striatal DAT expression in rhesus monkeys, an<br />

effect that persists for more than 30 days after<br />

cocaine withdrawal. 37 Increased DAT expression<br />

has also been shown in postmortem analyses <strong>of</strong><br />

brain tissue from human cocaine addicts 38 and<br />

synaptosomes prepared from this tissue exhibit<br />

greater [ 3 H]<strong>dopamine</strong> uptake than synaptosomes<br />

from age-matched cocaine-naïve individuals. 39 In<br />

addition, cyclic voltammetry studies <strong>of</strong> rats that<br />

exhibit a self-administration preference for high<br />

doses <strong>of</strong> cocaine reveal a specific increase in <strong>the</strong><br />

V max <strong>of</strong> <strong>dopamine</strong> uptake with no effect on <strong>the</strong><br />

Km for <strong>dopamine</strong>, suggesting an upregulation <strong>of</strong><br />

DAT expression. 40 Interestingly, <strong>the</strong>se results are<br />

exactly inverted compared with <strong>the</strong> findings for<br />

d-methamphetamine, indicating that <strong>the</strong> two psychostimulants<br />

possess opposing effects on DAT trafficking,<br />

despite <strong>the</strong>ir comparably high abuse potentials.<br />

Different effects on plasmalemmal DAT<br />

expression may play a role in <strong>the</strong> observed efficacy<br />

<strong>of</strong> d-amphetamine for <strong>the</strong> treatment <strong>of</strong> cocaine<br />

dependence in preclinical and early clinical<br />

trials. 41–44<br />

In studies <strong>of</strong> hDAT-expressing cultured cells,<br />

preincubation with cocaine has been routinely<br />

demonstrated to prevent <strong>the</strong> DAT-internalizing effects<br />

<strong>of</strong> substrates (e.g., Refs. 23, 24, and 45), as well<br />

as <strong>the</strong> putative amphetamine-induced brief, transient<br />

upregulation <strong>of</strong> surface DAT levels. 31,32 The<br />

acute effects <strong>of</strong> cocaine on DAT trafficking in <strong>the</strong><br />

absence <strong>of</strong> substrates are less consistent: whereas<br />

some research groups have shown that cocaine elicits<br />

a rapid increase in plasmalemmal DAT expression,<br />

46,47 o<strong>the</strong>rs have found no effect <strong>of</strong> cocaine pretreatment<br />

on DAT activity or surface expression. 24,48<br />

In <strong>the</strong> study by Little et al., treatment <strong>of</strong> N2A-hDAT<br />

cells with 1 �M cocaine for 24 h increased cell<br />

surface DAT expression by approximately 30%. 47<br />

Remarkably, this same magnitude <strong>of</strong> increase was<br />

detected using a wide array <strong>of</strong> techniques, including<br />

biotinylation, intact-cell [ 3 H]CFT binding, and<br />

visualization <strong>of</strong> plasma membrane anti-DAT immun<strong>of</strong>luorescence<br />

with confocal microscopy. 47 The<br />

long incubation time used in this study (24 h) may<br />

underlie some <strong>of</strong> <strong>the</strong> discrepancy with o<strong>the</strong>r reports<br />

showing no effect <strong>of</strong> cocaine, because most studies<br />

have used incubation times on <strong>the</strong> order <strong>of</strong> 30 min to<br />

1 h—curiously, however, Daws et al.demonstrateda<br />

cocaine-mediated increase in surface expressed DAT<br />

in HEK-hDAT cells after only 1 h <strong>of</strong> treatment. 46<br />

Cellular signaling cascades,<br />

phosphorylation, and DAT regulation<br />

DAT phosphorylation and trafficking<br />

Members <strong>of</strong> <strong>the</strong> neurotransmitter sodium symporter<br />

superfamily are large proteins—<strong>the</strong> hDAT<br />

possesses 620 amino acid residues—containing<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 321


DAT <strong>Regulation</strong> Schmitt & Reith<br />

many cytoplasmic-facing consensus sites for phosphorylation,<br />

many <strong>of</strong> which lie on <strong>the</strong> N terminus<br />

<strong>of</strong> <strong>the</strong> DAT protein (for a review <strong>of</strong> functional<br />

moieties in <strong>the</strong> DAT amino acid sequence,<br />

see Volz and Schenk 2 ). Protein phosphorylation is a<br />

de facto posttranslational strategy for regulating protein<br />

function, altering protein–protein interactions<br />

and transducing exogenous stimuli. Because phosphorylation<br />

<strong>of</strong> a particular residue in a protein can<br />

both drastically change its three-dimensional shape<br />

and serve as a signaling marker, is not surprising that<br />

<strong>the</strong> phosphorylation state <strong>of</strong> <strong>the</strong> DAT has a pr<strong>of</strong>ound<br />

influence on both <strong>the</strong> intrinsic activity (i.e., affinity<br />

for and responsiveness to various ligands) and membrane<br />

distribution <strong>of</strong> <strong>the</strong> <strong>transporter</strong>. 4 Several protein<br />

kinases and phosphatases involved in key signaling<br />

cascades can affect DAT function, including<br />

members <strong>of</strong> <strong>the</strong> protein kinase C (PKC), protein kinase<br />

A, phosphatidylinositol-3-kinase (PI3K), protein<br />

tyrosine kinase, Ca 2+ /calmodulin kinase, protein<br />

phosphatase 1, and mitogen-activated protein<br />

kinase (MAPK) families. 49–52 Of <strong>the</strong>se, PKC has been<br />

<strong>the</strong> most thoroughly investigated; however, a direct<br />

phosphotransferase reaction between PKC and <strong>the</strong><br />

DAT has not been demonstrated, suggesting that<br />

o<strong>the</strong>r (currently undefined) downstream kinases<br />

may be responsible for <strong>the</strong> direct phosphorylation<br />

<strong>of</strong> <strong>the</strong> DAT. 4,51 In addition to possible direct effects<br />

on <strong>the</strong> DAT and on DAT membrane trafficking, activation<br />

(or, by <strong>the</strong> same token, inhibition) <strong>of</strong> <strong>the</strong>se<br />

phosphorylation pathways can affect <strong>the</strong> regulatory<br />

and transmitter-releasing action <strong>of</strong> substrates, such<br />

as amphetamine. As we discuss in <strong>the</strong> following,<br />

both substrate exposure and PKC activation can<br />

result in DAT downregulation; however, despite a<br />

myriad <strong>of</strong> studies demonstrating <strong>the</strong> importance<br />

<strong>of</strong> protein kinases and phosphatases in substratemediated<br />

regulation <strong>of</strong> plasmalemmal DAT expression,<br />

explicit details regarding <strong>the</strong> mechanistic link<br />

between <strong>the</strong> two processes have been elusive. 50 In<br />

contrast with substrates, classical DAT inhibitors,<br />

such as cocaine, �-CFT, and methylphenidate, have<br />

little effect on DAT phosphorylation. 48<br />

Protein kinase C<br />

Functional regulation <strong>of</strong> <strong>the</strong> DAT in response to<br />

PKC activation was first demonstrated in vitro by<br />

using phorbol-ester kinase modulators, such as 4�phorbol-12-myristate-13-acetate<br />

(�-PMA). Incubation<br />

with �-PMA (or o<strong>the</strong>r nonphorbol PKC<br />

activators) dramatically increases levels <strong>of</strong> 32 PO4labeled<br />

immunoprecipitated DAT in both synaptosomes<br />

53 and cells heterologously expressing DAT. 54<br />

Increased phosphorylation was not observed in<br />

<strong>the</strong> presence <strong>of</strong> <strong>the</strong> inactive phorbol analogue 4�phorbol-12,13-didecanoate,<br />

and all effects <strong>of</strong> �-<br />

PMA were abolished by cotreatment with <strong>the</strong> PKC<br />

inhibitor bisindoylmaleimide I, highlighting <strong>the</strong> involvement<br />

<strong>of</strong> PKC. 53 Analogous to <strong>the</strong> findings with<br />

DAT substrates, pretreatment with PKC activators<br />

acutely and rapidly reduces [ 3 H]<strong>dopamine</strong> uptake<br />

by lowering <strong>the</strong> V max for uptake without affecting<br />

<strong>the</strong> Km valueandreduces<strong>the</strong>Bmax <strong>of</strong> [ 3 H]mazindol<br />

binding to intact Xenopus oocytes, 55 suggesting a<br />

role for phosphorylation in promoting <strong>the</strong> internalization<br />

<strong>of</strong> plasmalemmal DAT. A follow-up study by<br />

Melikian and Buckley substantiated this hypo<strong>the</strong>sis:<br />

<strong>the</strong> authors found that induction <strong>of</strong> PKC with<br />

�-PMA results in intracellular redistribution <strong>of</strong> <strong>the</strong><br />

DAT from <strong>the</strong> plasma membrane to recycling endosomal<br />

compartments in hDAT-expressing PC12<br />

cells. 56 Endocytic trafficking <strong>of</strong> <strong>the</strong> DAT to early<br />

and recycling endosomes in response to PKC activation<br />

has since been demonstrated in many different<br />

cell models and has been visualized in live cells,<br />

using various fluorescent-tagged DATs. 25,57 PKCmediated<br />

loss <strong>of</strong> plasmalemmal DAT is probably due<br />

to a combination <strong>of</strong> increased clathrin-dependent<br />

endocytosis and decreased recycling from endosomal<br />

compartments. 58 Proteolytic digestion studies<br />

<strong>of</strong> 32 PO4-labeled DAT indicate that serine residues<br />

in <strong>the</strong> N terminus are major target sites <strong>of</strong> phosphorylation<br />

after exogenous pharmacological PKC activation.<br />

59 In both constitutive and PKC-mediated<br />

DAT internalization, an endocytic motif has been<br />

shown to be required in <strong>the</strong> DAT C terminus, that<br />

is, residues 587–596; <strong>the</strong> results do not rule out <strong>the</strong><br />

possibility that o<strong>the</strong>r DAT sequences play a role in<br />

<strong>the</strong> interaction between <strong>the</strong> <strong>transporter</strong> and <strong>the</strong> endocytic<br />

machinery, with such sequences being sensitive<br />

to <strong>the</strong> local environment within <strong>the</strong> 587–596<br />

motif. 60 Essential elements <strong>of</strong> <strong>the</strong> DAT C-terminal<br />

endocytic motif are conserved within <strong>the</strong> SLC6<br />

<strong>transporter</strong> family but not in any o<strong>the</strong>r protein, indicating<br />

that DAT endocytosis may involve a clathrinindependent<br />

component. 60 Recent studies by <strong>the</strong><br />

group <strong>of</strong> Vaughan 61 does not indicate clathrinindependent<br />

endocytosis <strong>of</strong> DAT, but <strong>the</strong>y do<br />

suggest a role for cholesterol-sensitive raftassociated<br />

DAT that responds to PKC with reduced<br />

322 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

transport, in addition to non–raft-associated DATs<br />

responding to PKC by a classical clathrin-dependent<br />

mechanism.<br />

On <strong>the</strong> basis <strong>of</strong> <strong>the</strong> preceding studies, it would<br />

be tempting to posit that trafficking <strong>of</strong> <strong>the</strong> DAT is<br />

mediated (at least in part) by PKC, with increased<br />

phosphorylation signaling that a given DAT protein<br />

is ready to be internalized. Recent work by<br />

Cervinski et al. 45 shows that d-methamphetamine<br />

increases DAT phosphorylation in addition to triggering<br />

internalization—both <strong>of</strong> which are prevented<br />

by PKC inhibitors. Taken toge<strong>the</strong>r, <strong>the</strong>se findings<br />

prompt a tantalizing hypo<strong>the</strong>sis positing a link<br />

between substrate interaction, <strong>transporter</strong> phosphorylation<br />

by PKC, and subsequent membrane<br />

redistribution. Unfortunately, several lines <strong>of</strong> evidence<br />

suggest that <strong>the</strong> relationship between PKC,<br />

<strong>transporter</strong> phosphorylation, and membrane trafficking<br />

is far more complex. For example, removal<br />

<strong>of</strong> three classical PKC consensus sites from <strong>the</strong> DAT<br />

protein by mutagenesis <strong>of</strong> <strong>the</strong> target residues (Ser-<br />

262, Ser-586, and Thr-613) to glycine prevents <strong>the</strong><br />

phosphorylation normally observed after PKC activation<br />

with �-PMA but fails to prevent �-PMA–<br />

induced internalization and endosomal trafficking<br />

<strong>of</strong> <strong>the</strong> DAT. 62 Similarly, truncation <strong>of</strong> <strong>the</strong> <strong>transporter</strong><br />

distal N terminus—which bears several serine<br />

residues implicated in phosphorylation after PKC<br />

activation—also eliminates �-PMA–induced DAT<br />

phosphorylation without hindering <strong>the</strong> typical endocytic<br />

response in HEK cells. 63 Fur<strong>the</strong>rmore, although<br />

removal <strong>of</strong> <strong>the</strong> serine-rich distal N terminus<br />

prevents methamphetamine-induced DAT phosphorylation,<br />

it does not affect internalization <strong>of</strong> <strong>the</strong><br />

DAT in response to <strong>the</strong> substrate. 45 That is, although<br />

PKC clearly plays a role in <strong>the</strong> mechanism underlying<br />

<strong>the</strong> trafficking effects <strong>of</strong> amphetaminergic substrates,<br />

direct phosphorylation <strong>of</strong> <strong>the</strong> DAT is not part<br />

<strong>of</strong> that role. It is possible that activation <strong>of</strong> PKC results<br />

in DAT phosphorylation via a more circuitous<br />

route, involving thus far unidentified downstream<br />

kinases; however, <strong>the</strong> observation that substrateinduced<br />

trafficking can occur without appreciable<br />

phosphorylation <strong>of</strong> <strong>the</strong> DAT protein begs <strong>the</strong><br />

question <strong>of</strong> whe<strong>the</strong>r DAT phosphorylation actually<br />

serves as a “proendocytosis” signal. Instead, it may<br />

be that PKC regulates <strong>the</strong> action <strong>of</strong> a DAT-associated<br />

scaffolding or cytoskeletal protein that is ultimately<br />

responsible for determining <strong>the</strong> trafficking fate <strong>of</strong><br />

a given DAT. 49,51 Data suggest that scaffolding<br />

proteins can control <strong>the</strong> trafficking dynamics <strong>of</strong> <strong>the</strong><br />

noradrenaline <strong>transporter</strong> (NET): <strong>the</strong> cytosolic scaffolding<br />

protein syntaxin 1A—a well-known mediator<br />

<strong>of</strong> plasmalemmal vesicle docking and fusion—<br />

directly interacts with <strong>the</strong> NET and is involved with<br />

regulation <strong>of</strong> surface NET expression levels. 64 Moreover,<br />

exposure to amphetamine results in cytosolic<br />

redistribution <strong>of</strong> plasmalemmal NETs with a concomitant<br />

increase in <strong>the</strong> association <strong>of</strong> <strong>the</strong> NET protein<br />

and syntaxin 1A at <strong>the</strong> plasma membrane. 65 It<br />

is not clear whe<strong>the</strong>r syntaxin 1A promotes <strong>the</strong> internalization<br />

<strong>of</strong> <strong>the</strong> NET or stabilizes its presence<br />

at <strong>the</strong> plasma membrane (classically, syntaxin 1A is<br />

considered an exocytosis-promoting vesicle fusion<br />

protein), but <strong>the</strong> mere interaction between <strong>the</strong> two<br />

indicates that scaffolding proteins involved in regulated<br />

vesicular neurotransmitter release may also<br />

play a role in regulating transmitter <strong>transporter</strong>s.<br />

Although explicit syntaxin 1A-dependence in<br />

amphetamine-mediated trafficking has not yet been<br />

shown with <strong>the</strong> DAT, syntaxin 1A does interact with<br />

<strong>the</strong> N terminus <strong>of</strong> <strong>the</strong> DAT, 66 and this interaction<br />

has been recently shown to promote <strong>the</strong> efflux <strong>of</strong><br />

intracellular <strong>dopamine</strong> by amphetamine. 67 Interestingly,<br />

PKC (<strong>the</strong> classical is<strong>of</strong>orm, PKC-�) is known<br />

to promote amphetamine-stimulated <strong>dopamine</strong><br />

efflux via interaction with a DAT-associated 68<br />

protein complex; PKC-� is also necessary for <strong>the</strong><br />

rapid short-term increase in surface DAT levels<br />

upon substrate exposure. 32 In <strong>the</strong> syntaxin study by<br />

Lee et al., 66 <strong>the</strong> authors also noted an interaction<br />

between <strong>the</strong> N terminus <strong>of</strong> <strong>the</strong> DAT and a protein<br />

known as <strong>the</strong> receptor for activated C kinases<br />

(RACK1). RACK1 can bind to activated PKC<br />

molecules and o<strong>the</strong>r intracellular signaling kinases<br />

and hence may help transduce increased intracellular<br />

PKC activity into a DAT proendocytosis signal<br />

without <strong>the</strong> need for a direct phosphotransferase<br />

interaction between <strong>the</strong> DAT and PKC. This<br />

protein–protein interaction might also contribute<br />

to amphetamine-induced <strong>dopamine</strong> efflux by<br />

attracting activated PKC to <strong>the</strong> <strong>transporter</strong>,<br />

promoting an inward-facing, “efflux-favoring”<br />

conformational state. 68,69 Fur<strong>the</strong>rmore, <strong>the</strong><br />

conformational state <strong>of</strong> <strong>the</strong> DAT protein itself may<br />

influence its internalization rate, because disruption<br />

<strong>of</strong> <strong>the</strong> outward-facing conformational state by mutation<br />

<strong>of</strong> membrane-proximal N-terminal residues<br />

Arg-60 or Trp-63 to alanine results in increased<br />

DAT endocytosis. 70 Substrates could encourage an<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 323


DAT <strong>Regulation</strong> Schmitt & Reith<br />

inward-facing <strong>transporter</strong> conformation via <strong>the</strong><br />

process <strong>of</strong> translocation, suggesting a potential<br />

link between high concentrations <strong>of</strong> substrate and<br />

eventual DAT endocytosis.<br />

Thus, even if <strong>the</strong> identity <strong>of</strong> <strong>the</strong> terminal protein<br />

mediator <strong>of</strong> substrate-mediated DAT internalization<br />

is obfuscated, PKC still appears to be a requisite<br />

middleman in <strong>the</strong> endocytic signaling cascade.<br />

However, ano<strong>the</strong>r significant question remains: how<br />

might amphetaminergic substrates activate PKC in<br />

<strong>the</strong> first place? PKC activity is regulated by Ca 2+<br />

and diacylglycerol, a phospholipid metabolite generated<br />

in concert with inositol triphosphate by <strong>the</strong><br />

enzyme phospholipase C (PLC), which is also Ca 2+<br />

dependent. Stimulation <strong>of</strong> PKC activity by diacylglycerol<br />

can be amplified by arachidonic acid (for<br />

references, see Zhang and Reith 71 ), which has been<br />

shown to affect DAT function (e.g., Refs. 71–73);<br />

however, <strong>the</strong> effect <strong>of</strong> amphetamine on endogenous<br />

arachidonic acid—potentially through regulation<br />

<strong>of</strong> phospholipase A2, releasing arachidonic<br />

acid—is not known. Usually, PLC is activated by<br />

agonist binding at GPCRs coupled to <strong>the</strong> Gq/11type<br />

�-subunit; however, compounds, such as amphetamine<br />

and methamphetamine, lack significant<br />

affinity for GPCRs o<strong>the</strong>r than <strong>the</strong> trace amine–<br />

associated receptor 1 (TAAR1), a Gs-coupled receptor.<br />

74 It is conceivable that in certain experimental<br />

conditions and cell types, endogenous <strong>dopamine</strong><br />

released by amphetamines could activate <strong>dopamine</strong><br />

receptors, which can affect DAT function (see section<br />

on GPCR-mediated DAT regulation). However,<br />

Cervinski et al. 45 observed no differences in<br />

methamphetamine-induced DAT phosphorylation<br />

or internalization when cells were simultaneously<br />

pretreated with D1- andD2-like receptor antagonists.<br />

How, <strong>the</strong>n, could amphetamine induce PLC<br />

activity? Although <strong>the</strong>re have been few investigations<br />

into <strong>the</strong> non–<strong>transporter</strong>-mediated actions<br />

<strong>of</strong> amphetamines, research by Giambalvo indicates<br />

that treatment with 10 �M amphetamine enhances<br />

PLC activity in rat striatal synaptosomes and that<br />

pharmacological inhibition <strong>of</strong> PLC attenuates <strong>the</strong><br />

usual activation <strong>of</strong> PKC by amphetamine. 75 Interestingly,<br />

inhibition <strong>of</strong> <strong>the</strong> Na + /Ca 2+ antiporter with<br />

amiloride also blocked amphetamine-induced PKC<br />

activation, suggesting that <strong>the</strong> ionic effects <strong>of</strong> DAT<br />

substrate translocation can directly influence activation<br />

<strong>of</strong> intracellular signaling cascades. When substrates<br />

are transported across <strong>the</strong> plasma membrane<br />

via <strong>the</strong> DAT, sodium ions are cotransported, and<br />

this increase in intracellular sodium concentration<br />

may cause Na + /Ca 2+ antiporters at mitochondrial<br />

and endoplasmic reticulum membranes to operate<br />

in reverse, favoring a net flux <strong>of</strong> Ca 2+ into <strong>the</strong> cytosolic<br />

compartment. Because PLC is also a Ca 2+ -<br />

dependent enzyme, <strong>the</strong> increase in intracellular calcium<br />

may catalyze PKC activation via stimulation<br />

<strong>of</strong> PLC in addition to its direct effect on PKC. 75<br />

PI3K, MAPK, and o<strong>the</strong>r protein kinases<br />

Protein kinases o<strong>the</strong>r than PKC are also involved in<br />

DAT phosphorylation, but again, evidence <strong>of</strong> definitive<br />

phosphotransferase reactions has proven elusive.<br />

76 As with PKC, both <strong>the</strong> PI3K and MAPK signaling<br />

pathways influence phosphorylation <strong>of</strong> DAT<br />

serine and threonine residues, 49 but <strong>the</strong>ir regulatory<br />

effects on DAT function appear to be opposite that<br />

<strong>of</strong> PKC, with kinase inhibition resulting in downregulation<br />

<strong>of</strong> and kinase activation resulting in upregulation<br />

<strong>of</strong> DAT activity, respectively. PI3K can be activated<br />

by GPCRs and by receptor tyrosine kinases,<br />

such as <strong>the</strong> insulin receptor. Although <strong>the</strong> classical<br />

is<strong>of</strong>orm <strong>of</strong> PI3K itself is not considered a serine/threonine<br />

protein kinase, it is directly upstream<br />

<strong>of</strong> <strong>the</strong> serine/threonine kinase Akt (also known as<br />

protein kinase B). Inhibition <strong>of</strong> PI3K decreases <strong>the</strong><br />

V max <strong>of</strong> <strong>dopamine</strong> uptake and causes loss <strong>of</strong> surface<br />

expressed DAT, which—like PKC- and substrateinduced<br />

downregulation—is dependent on endocytosis<br />

by clathrin-coated vesicles. 77,78 Conversely,<br />

treatment with insulin or expression <strong>of</strong> constitutively<br />

active forms <strong>of</strong> PI3K increases [ 3 H]<strong>dopamine</strong><br />

uptake, a sign <strong>of</strong> putative surface DAT upregulation.<br />

77 Unlike inhibition <strong>of</strong> PKC, inhibition <strong>of</strong> PI3K<br />

has been recently demonstrated to attenuate <strong>the</strong><br />

<strong>dopamine</strong>-releasing effects <strong>of</strong> amphetamine; however,<br />

this is most likely due to reduction in cell<br />

surface DAT expression ra<strong>the</strong>r than a direct effect<br />

<strong>of</strong> PI3K on <strong>transporter</strong> function. 79 The MAPK<br />

pathway is a complex signaling network consisting<br />

<strong>of</strong> a three-tiered series <strong>of</strong> various serine/threonine<br />

protein kinases: primary members <strong>of</strong> <strong>the</strong> MAPK<br />

protein family (such as extracellular signal–<br />

regulated kinases-1 and -2 [ERK1/2]) are activated<br />

by secondary upstream kinases known as<br />

MAP/ERK kinases (MEKs) which are in turn activated<br />

by a tertiary group <strong>of</strong> fur<strong>the</strong>r upstream kinases<br />

known as MEK kinases. 80 Reducing ERK1 and ERK2<br />

324 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

phosphorylation by using small-molecule inhibitors<br />

<strong>of</strong> MEK affects DAT function and trafficking similarly<br />

to PI3K inhibition, decreasing <strong>the</strong> V max <strong>of</strong><br />

<strong>dopamine</strong> uptake without altering <strong>the</strong> Km and<br />

inducing cytosolic redistribution <strong>of</strong> plasmalemmal<br />

DAT into endosomal compartments. 78,81 The<br />

MAPK and PI3K signaling cascades are probably<br />

involved in <strong>the</strong> regulation <strong>of</strong> DAT activity by various<br />

classes <strong>of</strong> transmembrane receptors, such as<br />

GPCRs (discussed in <strong>the</strong> following section) and receptor<br />

tyrosine kinases. For example, in addition<br />

to insulin—which has been shown to increase plasmalemmal<br />

DAT function—activation <strong>of</strong> <strong>the</strong> neurotrophin<br />

receptor tyrosine kinase TrkB with brainderived<br />

neurotrophic factor also upregulates DAT<br />

activity, an effect that hinges on both <strong>the</strong> MAPK and<br />

PI3K pathways. 52 MAPK family members ERK1 and<br />

ERK2 have also been implicated in cocaine addiction<br />

and relapse after cocaine withdrawal, as central<br />

mediators <strong>of</strong> long-term sensitization to cocaine associated<br />

cues (for review, see Lu et al. 82 ).<br />

Ubiquitination and o<strong>the</strong>r direct<br />

posttranslational modifications<br />

Covalent attachment <strong>of</strong> <strong>the</strong> small soluble protein<br />

ubiquitin to <strong>the</strong> ε-amino moiety <strong>of</strong> lysine residues<br />

in target cellular proteins is known as ubiquitination.<br />

The ubiquitination reaction is catalyzed by a<br />

multistep enzymatic system, with <strong>the</strong> final covalent<br />

attachment mediated by members <strong>of</strong> a family <strong>of</strong> proteins<br />

known as E3 ubiquitin ligases (<strong>the</strong> multifarious<br />

functions <strong>of</strong> ubiquitin are reviewed in Welchman<br />

et al. 83 ). The most widely recognized function <strong>of</strong><br />

ubiquitination is polyubiquitination, which occurs<br />

when several ubiquitin subunits are attached to one<br />

lysine, forming a chain. Polyubiquitin chains serve<br />

as a signaling motif, indicating that a target protein<br />

is to be trafficked to <strong>the</strong> 26S proteasome for proteolytic<br />

destruction. However, attachment <strong>of</strong> one<br />

ubiquitin subunit to a target protein (monoubiquitination)<br />

is also a signaling motif and is thought<br />

to encourage retention <strong>of</strong> target proteins in sorting<br />

endosomes for eventual trafficking to <strong>the</strong> lysosomal<br />

degradation pathway. 84 Ubiquitination can<br />

fulfill roles o<strong>the</strong>r than those directly serving protein<br />

breakdown. For example, ubiquitination <strong>of</strong> <strong>the</strong><br />

yeast �-factor receptor Ste2 promotes endocytosis<br />

<strong>of</strong> <strong>the</strong> receptor–ligand complex prior to degradation<br />

in <strong>the</strong> vacuole. 85 The yeast multidrug <strong>transporter</strong><br />

Pdr5, 86 and <strong>the</strong> Ste6 �-factor pheromone<br />

<strong>transporter</strong>, 87 both members <strong>of</strong> <strong>the</strong> ATP-binding<br />

cassette multidrug <strong>transporter</strong> family, appear to be<br />

prepared by ubiquitination for endocytic delivery to<br />

<strong>the</strong> vacuole for proteolytic turnover. In <strong>the</strong>se cases,<br />

ubiquitination may serve as a signal for protein trafficking<br />

ra<strong>the</strong>r than a signal for protein degradation<br />

itself. Work by Hicke’s group 88 has provided<br />

more information on ubiquitin as a signal for endocytosis<br />

<strong>of</strong> <strong>the</strong> plasma membrane protein Ste2p,<br />

<strong>the</strong> mating pheromone �-factor receptor. Unlike<br />

<strong>the</strong> ubiquitin proteasome recognition signal, <strong>the</strong><br />

internalization signal does not require polyubiquitin<br />

formation through Lys-18 but ra<strong>the</strong>r relies on<br />

monoubiquitination <strong>of</strong> one lysine residue <strong>of</strong> <strong>the</strong><br />

Ste2p.<br />

There is evidence that, like phosphorylation,<br />

ubiquitination <strong>of</strong> <strong>the</strong> DAT affects cell surface expression<br />

and membrane trafficking <strong>of</strong> <strong>the</strong> <strong>transporter</strong>.<br />

Interestingly, many <strong>of</strong> <strong>the</strong> same intracellular signaling<br />

kinases involved in DAT phosphorylation may<br />

also be responsible for signaling via ubiquitination.<br />

For example, Miranda et al. demonstrated that activation<br />

<strong>of</strong> PKC by treatment with �-PMA increases<br />

ubiquitination <strong>of</strong> <strong>the</strong> DAT and that <strong>the</strong> major proportion<br />

<strong>of</strong> ubiquitin-conjugated DAT was present<br />

as <strong>the</strong> monoubiquitinated species. 89 Using Förster<br />

resonance energy transfer, <strong>the</strong> authors also showed<br />

that fluorescent-tagged hDAT and ubiquitin are associated<br />

in late endosomes and multivesicular bodies<br />

destined for lysosomal degradation. Much like<br />

PKC-associated <strong>transporter</strong> phosphorylation, ubiquitination<br />

<strong>of</strong> <strong>the</strong> DAT depends upon residues residing<br />

in <strong>the</strong> N terminus <strong>of</strong> <strong>the</strong> <strong>transporter</strong>, because<br />

simultaneous mutation <strong>of</strong> three amino-terminal lysine<br />

residues (Lys-19, Lys-27, and Lys-35) inhibits<br />

PKC-mediated ubiquitination. 90 Ubiquitin conjugation<br />

to <strong>the</strong> DAT is catalyzed by <strong>the</strong> ubiquitin ligase<br />

Nedd4-2. 91 Although it is clear that ubiquitination<br />

is involved in <strong>the</strong> regulated endocytosis <strong>of</strong> <strong>the</strong> DAT, a<br />

relationship between substrate-induced <strong>transporter</strong><br />

trafficking and ubiquitination has not yet been<br />

established.<br />

Presynaptic G protein-coupled receptors<br />

affecting DAT function<br />

Because DAT protein function and membrane distribution<br />

are acutely regulated via several different<br />

intracellular signaling cascades, it is not surprising<br />

that activation <strong>of</strong> various types <strong>of</strong> GPCRs expressed<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 325


DAT <strong>Regulation</strong> Schmitt & Reith<br />

by <strong>dopamine</strong>rgic neurons can affect <strong>the</strong> DAT. Of <strong>the</strong><br />

possible GPCR subtypes that can affect DAT function,<br />

<strong>dopamine</strong> D2-like autoreceptors are <strong>the</strong> most<br />

logical candidates, with <strong>the</strong>ir presynaptic localization<br />

and clearly defined role in regulatory inhibition<br />

<strong>of</strong> both tyrosine hydroxylase 92 and vesicular<br />

<strong>dopamine</strong> release. 93 However, although D2-like autoreceptors<br />

are <strong>the</strong> most thoroughly investigated,<br />

o<strong>the</strong>r GPCRs also appear to modulate <strong>dopamine</strong>rgic<br />

neurotransmission by altering DAT function.<br />

In particular, activation <strong>of</strong> both �-opioid receptors<br />

and TAAR1 receptors (a member <strong>of</strong> a recently discovered<br />

family <strong>of</strong> receptors for endogenous trace<br />

amines) has been shown to elicit changes in DAT<br />

function.<br />

Dopamine D2 and D3 autoreceptors<br />

The D2-like family <strong>of</strong> <strong>dopamine</strong> receptors comprises<br />

D2,D3,andD4 receptor subtypes. In addition<br />

to <strong>the</strong>ir role as classical postsynaptic receptors, <strong>the</strong><br />

D2 and D3 subtypes are expressed by <strong>dopamine</strong>rgic<br />

neurons <strong>the</strong>mselves, serving as presynaptic autoreceptors.<br />

Activation <strong>of</strong> D2-like autoreceptors attenuates<br />

<strong>dopamine</strong>rgic neurotransmission via several<br />

parallel mechanisms: in PC12 cells, for example,<br />

treatment with <strong>the</strong> D2/D3 agonist quinpirole causes<br />

feedback inhibition <strong>of</strong> tyrosine hydroxylase—<strong>the</strong><br />

rate-determining enzyme in <strong>dopamine</strong> syn<strong>the</strong>sis—<br />

and decreases K + -evoked <strong>dopamine</strong> release. 93 In<br />

addition to reducing stimulated <strong>dopamine</strong> release,<br />

recent evidence suggests that D2 and D3 receptor<br />

activation also reduces extracellular <strong>dopamine</strong> concentration<br />

via acute upregulation <strong>of</strong> DAT function.<br />

This effect was first demonstrated in <strong>the</strong> early 1990s<br />

by using rotating disk electrode voltammetry to<br />

measure <strong>the</strong> rate <strong>of</strong> extracellular <strong>dopamine</strong> clearance.<br />

That is, Meiergerd et al. demonstrated 94 that<br />

<strong>the</strong> D2/D3 agonist quinpirole increases <strong>the</strong> V max <strong>of</strong><br />

<strong>dopamine</strong> transport into rat striatal synaptosomes,<br />

an effect that is reversed by concomitant administration<br />

<strong>of</strong> <strong>the</strong> antagonist sulpiride. The authors also<br />

showed in vivo evidence <strong>of</strong> <strong>the</strong> inverse situation: a<br />

reductionin<strong>the</strong>V max <strong>of</strong> <strong>dopamine</strong> clearance in rats<br />

chronically treated with <strong>the</strong> potent D2 antagonist<br />

haloperidol. In support <strong>of</strong> this finding, localized intrastriatal<br />

application <strong>of</strong> <strong>the</strong> selective D2 antagonist<br />

raclopride has also been shown to decrease clearance<br />

<strong>of</strong> exogenous <strong>dopamine</strong> applied locally via a<br />

second micropipette. 95 Studies using selective D3<br />

receptor ligands paint a similar picture: <strong>the</strong> selective<br />

D3 agonist PD128907 increases electrochemically<br />

measured <strong>dopamine</strong> clearance in rat nucleus accumbens<br />

slices, producing an increase in transport V max<br />

with no effect on <strong>the</strong> K m <strong>of</strong> <strong>dopamine</strong> after a 10min<br />

preincubation. 96 In contrast, <strong>the</strong> D3 receptor<br />

antagonist GR103691 decreased <strong>dopamine</strong> significantly<br />

below vehicle levels. In Xenopus oocytes coexpressing<br />

hDAT and D2R transcripts, activation <strong>of</strong><br />

D2 receptors with apomorphine for as little as 5 min<br />

results in a 40% increase <strong>of</strong> whole-cell [ 3 H]CFT<br />

binding (Bmax) with no alteration in <strong>the</strong> Ki value<br />

for CFT, suggesting that activation <strong>of</strong> D2 receptors<br />

triggers rapid trafficking <strong>of</strong> intracellular DATs to <strong>the</strong><br />

plasma membrane. 97 Upregulation <strong>of</strong> surface DAT<br />

expression by D2 receptor agonists requires Gi/Go<br />

protein coupling, because pretreatment with pertussis<br />

toxin (PTX) blocked <strong>the</strong> increase in [ 3 H]CFT<br />

binding. In our collaborative study with <strong>the</strong> group<br />

<strong>of</strong> Garris, in vivo voltammetry was used to monitor<br />

<strong>dopamine</strong> uptake and release in rat striatum<br />

and nucleus accumbens. 98 We found that D2R antagonism<br />

reduced uptake at all frequencies used to<br />

stimulate <strong>the</strong> medial forebrain bundle, whereas release<br />

was enhanced only at lower frequencies. Be<br />

that as it may, both processes—D2R regulation <strong>of</strong><br />

<strong>dopamine</strong> uptake as well as release—serve as parallel<br />

feedback mechanisms aimed at reducing extracellular<br />

<strong>dopamine</strong> levels when high <strong>dopamine</strong> levels<br />

activate D2 receptors.<br />

Most investigations into <strong>the</strong> effects <strong>of</strong> D2-like receptor<br />

ligands on DAT function in vitro have used<br />

electrochemical methods to assess <strong>dopamine</strong> uptake<br />

in animal tissues, in lieu <strong>of</strong> more commonly<br />

used measurement <strong>of</strong> [ 3 H]<strong>dopamine</strong> uptake by heterologous<br />

cells. The reason for this methodological<br />

preference is purely technical: <strong>dopamine</strong> is a<br />

high-affinity agonist at D2 and D3 receptors, creating<br />

an obvious confound because <strong>the</strong> radioligand<br />

will activate <strong>the</strong> receptors as well. Although electrochemical<br />

detection has greater temporal resolution<br />

than is possible with cell models, 16 it does not allow<br />

direct observation <strong>of</strong> effects on DAT trafficking<br />

in concert with kinetic measures <strong>of</strong> DAT function.<br />

Recent studies have overcome this issue by using a<br />

novel syn<strong>the</strong>tic DAT substrate that is easily detected<br />

yet exhibits negligible affinity for <strong>dopamine</strong> receptors.<br />

The substrate 4-(4-(dimethylamino)styryl)-<br />

N-methylpyridinium (ASP + )—a styryl analogue<br />

<strong>of</strong> <strong>the</strong> neurotoxic substrate MPP + —has <strong>the</strong> added<br />

benefit <strong>of</strong> possessing a fluorescent �-conjugated<br />

326 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

diaryle<strong>the</strong>ne moiety (with excitation and emission<br />

spectra <strong>of</strong> 488 nm and 609 nm, respectively 99 ).<br />

Hence, unlike o<strong>the</strong>r potential surrogate DAT substrates<br />

(e.g., [ 3 H]MPP + and [ 3 H]amphetamine), it<br />

is possible to measure <strong>the</strong> cellular uptake and accumulation<br />

<strong>of</strong> ASP + in real time by using fluorescent<br />

microscopy. In EM4 cells coexpressing hDAT and<br />

D2R transcripts, receptor activation with quinpirole<br />

increases <strong>the</strong> rate <strong>of</strong> ASP + uptake compared<br />

with vehicle treatment—this effect was abolished<br />

in cells treated with PTX, corroborating <strong>the</strong> necessity<br />

<strong>of</strong> Gi/Go protein–coupled signaling for functional<br />

DAT upregulation. 100 Moreover, inhibition <strong>of</strong><br />

MEK (which, in turn, prevents ERK1/2 activation)<br />

blocked <strong>the</strong> effects <strong>of</strong> quinpirole, whereas PI3K inhibition<br />

had no effect. Treatment with quinpirole<br />

also increases surface expressed DAT (with a concomitant<br />

decrease in intracellular-localized DAT),<br />

as measured by biotinylation. Similarly, quinpirole<br />

also increases plasmalemmal DAT expression and<br />

ASP + uptake in EM4 cells coexpressing hDAT and<br />

D3 receptors. 101 Like <strong>the</strong> DAT upregulation mediated<br />

by D2 receptor activation, this effect is PTX<br />

sensitive; however, both <strong>the</strong> MAPK and PI3K pathways<br />

are required for <strong>the</strong> quinpirole-induced increase<br />

in DAT function. Interestingly, much like <strong>the</strong><br />

observations <strong>of</strong> Johnson et al. 31 for amphetaminergic<br />

substrates, D3 activation results in a rapid increase<br />

in surface DAT expression that eventually<br />

gives way to significantly reduced plasmalemmal<br />

<strong>transporter</strong> expression upon prolonged D3 agonist<br />

exposure. 101 As a tritiated alternative to fluorescent<br />

ASP + , it also appears possible to use tracer amounts<br />

<strong>of</strong> [ 3 H]tyramine for monitoring DAT function in<br />

experiments disentangling <strong>transporter</strong> and receptor<br />

phenomena, because tyramine has been shown to<br />

display ra<strong>the</strong>r low potency (micromolar) in activating<br />

D2 <strong>dopamine</strong> receptors. 100 It will be important<br />

to show that <strong>the</strong> tritiated ligand does not appreciably<br />

bind to <strong>dopamine</strong> receptors; caution is required<br />

because <strong>the</strong> thorough study <strong>of</strong> Zapata et al. 101<br />

shows considerable binding <strong>of</strong> [ 3 H]<strong>dopamine</strong> to D3<br />

<strong>dopamine</strong> receptors that can be displaced by 0.1 �M<br />

spiroperidol in cells that do not express hDAT compared<br />

with cells subjected to <strong>the</strong> same conditions<br />

with hDAT on board.<br />

One unexpected finding in this series <strong>of</strong> receptormediated<br />

ASP + -uptake experiments was <strong>the</strong><br />

observation <strong>of</strong> positive resonance energy transfer<br />

between fluorescent-tagged hDAT and a biolumi-<br />

nescent D2R–luciferase construct, suggesting that<br />

<strong>the</strong> DAT and D2R proteins are in proximity (


DAT <strong>Regulation</strong> Schmitt & Reith<br />

that <strong>the</strong>y are bona fide neuromodulators. �-<br />

Phenethylamine (�-PEA), p-tyramine, and DMT<br />

act as full agonists at <strong>the</strong> trace amine receptor<br />

TAAR1, triggering <strong>the</strong> production cAMP. 106<br />

Interestingly, like �-PEA, <strong>the</strong> DAT substrates damphetamine<br />

and d-methamphetamine are also<br />

full agonists at TAAR1. 74 In HEK cells expressing<br />

hDAT and TAAR1, exposure to �-PEA causes a<br />

rapid reduction in [ 3 H]<strong>dopamine</strong> uptake in both<br />

HEK cells and mouse synaptosomes—importantly,<br />

this effect does not appear to be due to generalized<br />

DAT substrate activity <strong>of</strong> �-PEA, because <strong>the</strong> effect<br />

is absent in synaptosomes prepared from TAAR1knockout<br />

mice. 107 Later analysis <strong>of</strong> transfected HEK<br />

cells demonstrated that TAAR1 and <strong>dopamine</strong> D2<br />

receptors may have opposing effects on DAT regulation:<br />

cells expressing DAT and D2 receptors show <strong>the</strong><br />

expected increase in [ 3 H]<strong>dopamine</strong> uptake, whereas<br />

cells expressing DAT and TAAR1 show a decrease<br />

in [ 3 H]<strong>dopamine</strong> uptake under identical conditions.<br />

108 However, <strong>the</strong> relevance <strong>of</strong> this finding is<br />

difficult to interpret, because <strong>dopamine</strong> was used as<br />

both <strong>the</strong> D2R and TAAR1 agonist, but as discussed in<br />

<strong>the</strong> foregoing, radiolabeled [ 3 H]<strong>dopamine</strong> used to<br />

measure uptake would also probably activate receptors.<br />

Because <strong>the</strong> neuromodulatory roles <strong>of</strong> <strong>the</strong>se<br />

once-mysterious trace amines are just now beginning<br />

to be uncovered, fur<strong>the</strong>r investigations into<br />

<strong>the</strong> effects <strong>of</strong> TAAR1 receptor activation on <strong>transporter</strong><br />

trafficking are clearly needed. Subtle modulation<br />

<strong>of</strong> monoaminergic neurotransmission by<br />

TAAR1 agonists has, for example, been speculated<br />

to underlie <strong>the</strong> anxiolytic effects <strong>of</strong> small doses <strong>of</strong><br />

d-amphetamine and DMT in humans. 109<br />

Downregulation <strong>of</strong> DAT activity by κ-opioid<br />

receptors<br />

Interest in <strong>the</strong> effects <strong>of</strong> �-opioid receptor activation<br />

on DAT function stems from initial observations<br />

that �-opioid agonists attenuate <strong>the</strong> dramatic<br />

rise in extracellular <strong>dopamine</strong> levels after cocaine<br />

administration. In <strong>the</strong> rat nucleus accumbens, for<br />

example, a single high-dose injection <strong>of</strong> cocaine<br />

(20 mg/kg) causes a 10-fold increase in extracellular<br />

<strong>dopamine</strong> concentration—pretreatment with<br />

<strong>the</strong> �-agonist U50488 decreases this effect by more<br />

than 50%. 110 Using voltammetry and in vivo microdialysis,<br />

Thompson et al. demonstrated that <strong>the</strong><br />

DATmodulatoryeffects<strong>of</strong>�-opioid agonists differ<br />

between acute and chronic agonist administra-<br />

tion. 111 That is, whereas acute administration <strong>of</strong> <strong>the</strong><br />

�-agonist U69593 increases <strong>dopamine</strong> uptake, repeated<br />

administration results in a decrease in uptake<br />

with a concomitant decrease in surface DAT<br />

labeling by <strong>the</strong> tropane radioligand [ 125 I]CIT (2�carbomethoxy-3�-(4-iodophenyl)tropane)<br />

but no<br />

loss in total DAT protein. Importantly, <strong>the</strong> effect<br />

<strong>of</strong> repeated cocaine treatment was opposite that <strong>of</strong><br />

U69593 (a significant acceleration <strong>of</strong> <strong>dopamine</strong> uptake,<br />

indicating an upregulation <strong>of</strong> surface DATs),<br />

but <strong>the</strong> effect <strong>of</strong> both drugs in combination was not<br />

significantly different from control. 111 The mechanism<br />

underlying �-opioid–induced DAT regulation<br />

is unknown, but it is hypo<strong>the</strong>sized that activation<br />

<strong>of</strong> PI3K by �-receptor agonists plays a role. 16 The<br />

DAT-modulating effects <strong>of</strong> �-opioids prompt consideration<br />

<strong>of</strong> <strong>the</strong>ir receptors as potential targets for<br />

novel cocaine addiction <strong>the</strong>rapeutics. For example,<br />

�-opioid receptor agonists attenuate <strong>the</strong> acute reinforcing<br />

effects <strong>of</strong> cocaine, and repeated �-opioid<br />

agonist treatment may oppose some <strong>of</strong> <strong>the</strong> alterations<br />

in <strong>dopamine</strong>rgic transmission observed after<br />

withdrawal from chronic cocaine. 111<br />

Possible implications <strong>of</strong> DAT regulation to<br />

<strong>the</strong> putative toxicity <strong>of</strong> several substituted<br />

amphetamines<br />

Historically, neurotransmitter <strong>transporter</strong>s, such as<br />

<strong>the</strong> DAT, were regarded as relatively static entities:<br />

“molecular vacuums” expressed at a constant level<br />

at presynaptic terminals. 16 This preconceived notion<br />

has changed dramatically over <strong>the</strong> last decade,<br />

because many studies have indicated that—akin to<br />

classically defined transmembrane receptors—<strong>the</strong><br />

activity and membrane trafficking <strong>of</strong> <strong>transporter</strong><br />

proteins is rapidly and differentially regulated by<br />

interaction with chemically distinct ligands. In fact,<br />

current research suggests that DAT trafficking is<br />

<strong>the</strong> primary mechanism by which cells modulate<br />

extracellular <strong>dopamine</strong>rgic tone. That is, dynamic<br />

change in <strong>the</strong> rate <strong>of</strong> <strong>dopamine</strong> uptake is achieved<br />

by shifting <strong>the</strong> distribution <strong>of</strong> DAT proteins between<br />

<strong>the</strong> plasmalemmal and endosomal compartments,<br />

rapidly altering <strong>the</strong> number <strong>of</strong> DATs at <strong>the</strong> cell surface.<br />

As discussed earlier, <strong>the</strong> most robust example<br />

<strong>of</strong> ligand-mediated <strong>transporter</strong> redistribution is <strong>the</strong><br />

internalization response triggered by amphetaminergic<br />

substrates. Long before it was known that<br />

<strong>transporter</strong> proteins undergo dynamic trafficking<br />

328 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

and that this trafficking can be rapidly regulated,<br />

findings <strong>of</strong> in vitro amphetamine-induced reduction<br />

in <strong>transporter</strong> radioligand binding were interpreted<br />

as evidence <strong>of</strong> <strong>the</strong> in vivo neurotoxicity <strong>of</strong><br />

<strong>the</strong>se compounds. With <strong>the</strong> new knowledge on DAT<br />

trafficking phenomena, such findings will need to be<br />

reinterpreted. In <strong>the</strong> following, we will discuss <strong>the</strong><br />

possible implications <strong>of</strong> <strong>transporter</strong> protein regulation<br />

to <strong>the</strong> putative toxicity <strong>of</strong> several substituted<br />

amphetamine derivatives that are <strong>of</strong>ten encountered<br />

as drugs <strong>of</strong> addiction (<strong>the</strong> prototypical example being<br />

methamphetamine) and/or recreational drugs.<br />

We will also discuss <strong>the</strong> potential utility <strong>of</strong> agents<br />

that reduce plasmalemmal DAT expression in <strong>the</strong><br />

rectification <strong>of</strong> <strong>dopamine</strong>rgic neuroadaptation that<br />

occurs in chronic cocaine addiction. 37,40<br />

The amphetamines are an immensely diverse class<br />

<strong>of</strong> compounds and <strong>the</strong> structure-activity relationship<br />

<strong>of</strong> <strong>the</strong> amphetaminergic structural framework<br />

has been thoroughly investigated. 112 Amphetamine<br />

derivatives can elicit a wide range <strong>of</strong> phenomenological<br />

and neurochemical effects; however, for this<br />

discussion, we will focus solely on stimulant-like<br />

amphetamine derivatives that act as monoamine<br />

<strong>transporter</strong> substrates and will not consider any <strong>of</strong><br />

<strong>the</strong> psychedelic amphetamine derivatives, which act<br />

primarily at serotonin 5-HT2 receptors. 113 Whereas<br />

amphetamine itself exhibits roughly equipotent<br />

substrate activity at <strong>the</strong> DAT and NET with little<br />

meaningful activity at <strong>the</strong> SERT—amphetamine inhibits<br />

[ 3 H]DA and [ 3 H]NE uptake with 112- and 98fold<br />

greater potency than [ 3 H]5-HT, respectively—<br />

certain amphetamine analogues bearing substitutions<br />

on <strong>the</strong> phenyl ring and amine nitrogen have<br />

unique activity pr<strong>of</strong>iles at <strong>the</strong> three monoamine<br />

<strong>transporter</strong>s. 114 For example, aromatic substitution<br />

with a bulky, electron-rich moiety (such as a<br />

methoxy group or a halogen larger than fluorine) at<br />

<strong>the</strong> 3- or 4-position usually decreases substrate activity<br />

at <strong>the</strong> DAT but concomitantly increases affinity<br />

toward <strong>the</strong> SERT. 115,116 Hence, amphetamine<br />

derivatives, such as 4-chloroamphetamine (PCA),<br />

4-methoxyamphetamine (4-MA; more traditionally<br />

known as PMA, but we shall refer to it as 4-MA<br />

to prevent confusion with <strong>the</strong> phorbol ester �-<br />

PMA) and 3-trifluoromethyl-N-ethylamphetamine<br />

(fenfluramine) are <strong>of</strong>ten labeled as “serotonergic<br />

amphetamines” (although PCA still possesses<br />

significant affinity for both <strong>the</strong> DAT and<br />

NET 116 ). Alkylation <strong>of</strong> <strong>the</strong> amphetamine nitro-<br />

gen to <strong>the</strong> secondary amine generally increases<br />

SERT affinity without drastically altering potency<br />

at ei<strong>the</strong>r <strong>the</strong> DAT or NET; however, alkyl<br />

moieties longer than N-ethyl impede substratelike<br />

activity, with d-N-butylamphetamine being<br />

virtually inactive in behavioral assays. 117<br />

Therefore, d-methamphetamine exhibits roughly<br />

2.5-fold greater potency than d-amphetamine<br />

as a SERT substrate. 115 The empathogen 3,4methylenedioxyamphetamine<br />

(MDA) and its Nmethyl<br />

analogue (MDMA)—both widely used<br />

recreational drugs, <strong>of</strong>ten sold (with any number <strong>of</strong><br />

admixes) under <strong>the</strong> name “ecstasy”—have affinity<br />

for all three <strong>of</strong> <strong>the</strong> monoamine <strong>transporter</strong>s, with<br />

slightly greater substrate activity at <strong>the</strong> SERT and<br />

NET than at <strong>the</strong> DAT. 118–120 SERT-affecting substituted<br />

amphetamine analogues decrease plasmalemmal<br />

SERT expression, paralleling <strong>the</strong> endocytic effect<br />

that amphetamine itself has on <strong>the</strong> DAT. Acute<br />

administration <strong>of</strong> high doses <strong>of</strong> d-fenfluramine or<br />

PCA to rats decreases <strong>the</strong> Bmax for SERT radioligand<br />

binding by 30–60% (depending upon <strong>the</strong> interval<br />

after drug administration), without significant<br />

changes in whole-cell SERT protein levels, 121 as<br />

measured by Western blot analysis (but also see Xie<br />

et al. 122 ). MDMA has also been frequently shown<br />

to cause loss <strong>of</strong> both radiolabeled SERT and DAT<br />

binding sites in experimental animals. 123–125<br />

Because monoaminergic neurotransmission is so<br />

intimately involved in cognition, affect, behavioral<br />

reinforcement, and motor function, 2,126 it is not surprising<br />

that <strong>the</strong> implications <strong>of</strong> substrate-induced<br />

monoamine <strong>transporter</strong> downregulation on human<br />

physiological and psychical function are highly debated<br />

topics. The amphetaminergic substrate that<br />

has received <strong>the</strong> most attention in this debate is<br />

MDMA. Because <strong>of</strong> its widespread popularity as a<br />

recreational drug and nearly ubiquitous presence in<br />

<strong>the</strong> party scene, ascertaining <strong>the</strong> potential <strong>of</strong> recreational<br />

doses <strong>of</strong> MDMA to elicit neurotoxicity in<br />

humans is an important public health question.<br />

Fur<strong>the</strong>rmore, elucidation <strong>of</strong> <strong>the</strong> mechanism underlying<br />

MDMAergic neurotoxicity will also provide<br />

insight into strategies to minimize potential damage<br />

dealt by recreational MDMA use.<br />

Initial neurochemical studies in rats revealed a<br />

dramatic and protracted reduction in <strong>the</strong> density <strong>of</strong><br />

SERT expression and in total brain 5-HT content<br />

after high doses <strong>of</strong> ei<strong>the</strong>r MDA or MDMA. 123 Similar<br />

losses <strong>of</strong> SERT expression were observed after<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 329


DAT <strong>Regulation</strong> Schmitt & Reith<br />

treatment with <strong>the</strong> well-established serotonergic<br />

neurotoxin 5,7-dihydroxytryptamine (5,7-DHT).<br />

Hence, authors <strong>of</strong> <strong>the</strong>se early studies concluded<br />

that MDA-like drugs produce neurotoxic insult to<br />

serotonergic neurons and suggested that <strong>the</strong> magnitude<br />

and duration <strong>of</strong> <strong>the</strong> decline in SERT expression<br />

constitutes a method <strong>of</strong> quantifying <strong>the</strong><br />

extent <strong>of</strong> this damage. The reduction in SERT expression<br />

that occurs after MDMA administration<br />

is quantitatively similar to reductions caused by<br />

4-substituted amphetamines, such as 4-MA and<br />

PCA. 127,128 In light <strong>of</strong> <strong>the</strong> fact that membrane trafficking<br />

<strong>of</strong> monoamine <strong>transporter</strong> proteins is altered<br />

upon exposure to amphetaminergic substrates—<br />

and that such changes persist beyond <strong>the</strong> duration<br />

<strong>of</strong> acute drug exposure—<strong>the</strong> question is<br />

thus whe<strong>the</strong>r or not reductions in SERT expression<br />

represent mere changes in <strong>transporter</strong> trafficking<br />

and membrane distribution or true damage to<br />

serotonergic neuronal function. Evidence in support<br />

<strong>of</strong> both conclusions exists. Experiments by<br />

Wang et al. 125,129 showed that MDMA administration<br />

reduces radioligand binding to plasmalemmal<br />

SERTs in rats; but unlike <strong>the</strong> specific serotonergic<br />

toxin 5,7-DHT, it nei<strong>the</strong>r alters total SERT<br />

protein expression nor induces microglial activation.<br />

In contrast, <strong>the</strong> results <strong>of</strong> Xie et al. 122 —who<br />

used a different, more specific antibody for immunochemical<br />

detection <strong>of</strong> <strong>the</strong> SERT protein—<br />

do indicate substantial loss <strong>of</strong> SERT protein after<br />

treatment with ei<strong>the</strong>r 5,7-DHT or <strong>the</strong> serotonergic<br />

amphetamines MDMA, PCA, and fenfluramine,<br />

suggesting functional damage to serotonergic nerve<br />

terminals (Ref. 122, Figs. 8–10). However, some<br />

<strong>of</strong> <strong>the</strong> differences seen by <strong>the</strong>se two groups may<br />

be related to respective MDMA dosing regimens:<br />

22.5 mg/kg total MDMA (3 × 7.5 mg/kg, each 2 h<br />

apart) was given in <strong>the</strong> studies by Wang et al., 125,129<br />

whereas a total dose <strong>of</strong> 45 mg/kg (3 × 15 mg/kg,<br />

each 1.5 h apart) was used in <strong>the</strong> Xie et al.<br />

study. 122 Callaghan et al. compared <strong>the</strong> effects <strong>of</strong> repeated<br />

administration <strong>of</strong> ei<strong>the</strong>r MDMA or 4-MA on<br />

in vitro measures <strong>of</strong> neurodegeneration (such as<br />

plasmalemmal SERT binding) with in vivo assessment<br />

<strong>of</strong> 5-HT clearance by using real-time microamperometry.<br />

128 Here, <strong>the</strong> authors found that<br />

[ 3 H]cyanoimipramine binding to <strong>the</strong> SERT was<br />

significantly reduced 2 weeks after repeated administration<br />

<strong>of</strong> ei<strong>the</strong>r high-dose 4-MA or MDMA<br />

(15 mg/kg, twice daily for 4 days); however, clear-<br />

ance <strong>of</strong> locally applied 5-HT was reduced only in<br />

rats treated with a high dose <strong>of</strong> 4-MA. The observation<br />

that 5-HT clearance in vivo was unaltered by<br />

even a high-dose MDMA treatment regimen suggests<br />

that in vitro measures, such as plasmalemmal<br />

SERT radioligand binding, do not necessarily predict<br />

<strong>the</strong> functional state <strong>of</strong> serotonergic neurons in<br />

vivo. Moreover, <strong>the</strong>se findings highlight that <strong>the</strong> toxicity<br />

<strong>of</strong> substituted amphetamines can vary greatly<br />

depending upon <strong>the</strong>ir individual chemical structure.<br />

Here 4-MA appears to have a greater potential<br />

than MDMA to elicit toxicity in vivo. This finding<br />

is certainly consistent with epidemiological data, 130<br />

which show a high incidence <strong>of</strong> mortality after consumption<br />

<strong>of</strong> clandestine pressed pills containing 4-<br />

MA sold as “ecstasy.” By contrast, fatalities caused<br />

by consumption <strong>of</strong> moderate doses <strong>of</strong> unadulterated<br />

MDMA (as opposed to ecstasy pills) in <strong>the</strong> absence<br />

<strong>of</strong> o<strong>the</strong>r drugs are exceedingly rare.<br />

Paradoxically, localized direct administration <strong>of</strong><br />

MDMA in rats by in vivo reverse microdialysis—<br />

at concentrations equivalent to those attained after<br />

high-dose systemic administration—does not result<br />

in <strong>the</strong> typical indicators <strong>of</strong> serotonergic neurotoxicity,<br />

despite eliciting dramatically greater increases<br />

in extracellular monoamine levels. 131 Surprisingly,<br />

MDMA was demonstrated to exert neuroprotective<br />

effects in cultured rat fetal neurons, indicating that<br />

direct exposure to <strong>the</strong> chemical itself is not necessarily<br />

cytotoxic. 132 In addition, whereas systemically<br />

administered MDA and MDMA acutely increase<br />

<strong>the</strong> concentration <strong>of</strong> radical reactive oxygen<br />

species (ROS) and produce oxidative stress in <strong>the</strong><br />

brain, direct intracerebroventricular administration<br />

is devoid <strong>of</strong> such effects. 131,133<br />

The difference in toxicological pr<strong>of</strong>ile between<br />

direct microdialysis and systemic routes <strong>of</strong> administration<br />

suggests that much <strong>of</strong> <strong>the</strong> toxicity <strong>of</strong><br />

MDA and its N-alkylated cousins is <strong>the</strong> result <strong>of</strong><br />

metabolic transformation <strong>of</strong> parent compounds<br />

into more potent neurotoxins. 134 In humans,<br />

MDA and MDMA are largely metabolized via<br />

demethylenation by <strong>the</strong> hepatic CYP2D6 enzyme<br />

to <strong>the</strong> respective 3,4-dihydroxyamphetamine<br />

species (�-methyl<strong>dopamine</strong> and �,Ndimethyl<strong>dopamine</strong>).<br />

135 In recent years, several<br />

investigators have revealed that—much like<br />

<strong>dopamine</strong>—<strong>the</strong>se catechol species readily undergo<br />

oxidation to semiquinone and quinone species<br />

that can form neurotoxic thioe<strong>the</strong>r compounds<br />

330 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

Figure 2. Metabolic reaction pathway <strong>of</strong> MDA-like compounds leading to <strong>the</strong> formation <strong>of</strong> neurotoxic prooxidant<br />

thioe<strong>the</strong>r species and <strong>the</strong> relative toxicity <strong>of</strong> metabolites. The parent compounds MDA and MDMA are not directly<br />

cytotoxic. 131,132 However, demethylenation <strong>of</strong> <strong>the</strong> 3,4-methylenedioxy moiety by <strong>the</strong> hepatic CYP2D6 enzyme 135<br />

results in <strong>the</strong> formation <strong>of</strong> 3,4-dihydroxyamphetamines, which are readily oxidized to quinones. 134 These quinones<br />

can form conjugates with GSH, greatly potentiating neurotoxicity. 138 Fur<strong>the</strong>r metabolism in <strong>the</strong> brain via <strong>the</strong><br />

mercapturic acid pathway results in toxic NAC-conjugated species. Although <strong>the</strong> quinones, GSH-conjugates, and<br />

NAC-conjugates all promote <strong>the</strong> formation <strong>of</strong> reactive oxygen species, NAC-substituted compounds appear to possess<br />

<strong>the</strong> greatest relativistic potential for toxicity owing to <strong>the</strong>ir protracted rate <strong>of</strong> elimination from <strong>the</strong> brain. 137 (In color<br />

in Annals online.)<br />

in vivo. 136,137 Quinones are strong electrophiles:<br />

<strong>the</strong>y can serve to propagate ROS via a one-electron<br />

redox reaction to a stable semiquinone radical<br />

or form conjugates with reducing agents via<br />

two-electron nucleophilic attack. 138 In particular,<br />

<strong>the</strong> dopaminoquinone-like metabolite <strong>of</strong><br />

MDA/MDMA can undergo conjugation with<br />

<strong>the</strong> cysteinyl thiol moiety <strong>of</strong> <strong>the</strong> endogenous<br />

reductant glutathione (GSH) to form <strong>the</strong> thioe<strong>the</strong>r<br />

5-(glutathionyl)-�-methyl<strong>dopamine</strong> (5-(GSH)-<br />

�MeDA) or its N-methyl analogue. 136 In <strong>the</strong> central<br />

nervous system, <strong>the</strong> 5-(glutathionyl)-thioe<strong>the</strong>rs<br />

are ultimately metabolized via <strong>the</strong> mercapturic<br />

acid pathway to form 5-(N-acetyl-cysteinyl)-<br />

�-methyl<strong>dopamine</strong> (5-(NAC)-�MeDA) and its<br />

N-methyl analogue (5-(NAC)-�,N-diMeDA).<br />

The structural formulae <strong>of</strong> <strong>the</strong>se metabolites<br />

and <strong>the</strong>ir relative potencies as neurotoxins are<br />

displayed in Figure 2. The thioe<strong>the</strong>r metabolites are<br />

potent serotonergic neurotoxins that can induce<br />

a dose-dependent increase in ROS formation and<br />

cause caspase-3–mediated apoptosis in cultured<br />

cortical neurons. 137 Moreover, direct intrastriatal<br />

administration <strong>of</strong> pure 5-(NAC)-�,N-diMeDA in<br />

rats fully recapitulates <strong>the</strong> serotonergic toxicity<br />

observed with systemic high-dose MDMA. 136<br />

These thioe<strong>the</strong>rs are detectable in <strong>the</strong> rat brain<br />

after systemic administration <strong>of</strong> a high-dose<br />

neurotoxic regimen <strong>of</strong> MDMA—repeated dosing<br />

leads to significant accumulation due to a<br />

nonlinear increase in elimination half-life. 139<br />

Interestingly, cotreatment with <strong>the</strong> antioxidant<br />

GSH precursor N-acetylcysteine (NAC) successfully<br />

protects cultured cortical neurons from<br />

thioe<strong>the</strong>r-induced apoptosis and attenuates <strong>the</strong><br />

formation <strong>of</strong> ROS, even though NAC could<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 331


DAT <strong>Regulation</strong> Schmitt & Reith<br />

obviously contribute to <strong>the</strong> formation <strong>of</strong> <strong>the</strong><br />

5-(NAC)-thioe<strong>the</strong>r metabolites. 137 Pretreatment<br />

with <strong>the</strong> antioxidant compound acetyl-L-carnitine<br />

was also shown to prevent oxidative damage to<br />

neuronal mitochondria in vivo after high-dose<br />

MDMA administration in rats. 140 These data<br />

strongly indicate that <strong>the</strong> toxicity <strong>of</strong> MDMA and its<br />

analogues is due chiefly to oxidative stress induced<br />

by <strong>the</strong> potent and selective serotonergic neurotoxins<br />

formed during metabolism <strong>of</strong> <strong>the</strong> parent<br />

compound. Moreover, <strong>the</strong>se findings suggest that<br />

avoidance <strong>of</strong> repeated dosing and concomitant use<br />

<strong>of</strong> central nervous system–penetrating antioxidant<br />

compounds are prudent harm reduction strategies<br />

for preventing or limiting <strong>the</strong> damage inherent to<br />

recreational MDMA use.<br />

If <strong>the</strong> neurotoxic effects <strong>of</strong> <strong>the</strong> MDA-type ring<br />

substituted amphetamines is due primarily to <strong>the</strong><br />

formation <strong>of</strong> prooxidant quinone metabolites, what<br />

can be said <strong>of</strong> amphetamine derivatives that are<br />

unlikely to form such unstable metabolites, such<br />

as d-methamphetamine? Despite being similar to<br />

amphetamine itself in structure, pharmacokinetics,<br />

and metabolic fate, methamphetamine exhibits increased<br />

toxicity and significantly greater abuse potential<br />

than <strong>the</strong> parent compound. 20,141 Toxic insult<br />

to both serotonergic and <strong>dopamine</strong>rgic neuronal<br />

systems after high doses <strong>of</strong> d-methamphetamine has<br />

been routinely demonstrated in experimental animals<br />

(see Cadet et al. 142 for a review). In humans,<br />

chronic high-dose methamphetamine addicts show<br />

moderate deficits in cognitive function, and in vivo<br />

neuroimaging techniques reveal decreased surface<br />

DAT concentration 143 and increased microglial activation<br />

compared with control subjects. 144 Eventual<br />

recovery from <strong>the</strong>se effects appears possible,<br />

but normalization <strong>of</strong> <strong>dopamine</strong>rgic function is an<br />

extremely slow process, taking months to years. 145<br />

Similarly to neurotoxicity associated with MDA-like<br />

substituted amphetamines, <strong>the</strong> toxic response to<br />

methamphetamine—particularly in <strong>dopamine</strong>rgic<br />

neurons—also involves oxidative stress by quinonemediated<br />

formation <strong>of</strong> ROS and subsequent activation<br />

<strong>of</strong> proapoptotic signaling cascades. 142<br />

Findings that antioxidant compounds can reduce<br />

neurotoxic insult by methamphetamine support <strong>of</strong><br />

this <strong>the</strong>ory: for example, high doses <strong>of</strong> <strong>the</strong> aforementioned<br />

antioxidant NAC ameliorate <strong>the</strong> longterm<br />

loss <strong>of</strong> striatal DAT in moneys after repeated<br />

injections <strong>of</strong> methamphetamine. 146 However, un-<br />

like with MDA/MDMA, here <strong>the</strong> likely culprit is<br />

not a metabolic product <strong>of</strong> <strong>the</strong> parent stimulant but<br />

ra<strong>the</strong>r <strong>dopamine</strong> itself.<br />

Free-floating <strong>dopamine</strong> (that is, existing outside<br />

synaptic vesicles) can be spontaneously oxidized<br />

to dopaminoquinone and successively converted to<br />

5-cysteinyl-catechol thioe<strong>the</strong>rs in a manner identical<br />

to <strong>the</strong> reaction <strong>of</strong> <strong>the</strong> MDA metabolite 3,4dihydroxyamphetamine<br />

shown in Figure 2. 147 If sufficiently<br />

large concentrations <strong>of</strong> soluble <strong>dopamine</strong><br />

are present, mechanisms to rapidly clear or safely<br />

inactivate <strong>dopamine</strong>—namely, <strong>the</strong> DAT and <strong>the</strong><br />

catabolic enzyme COMT—will presumably be<br />

saturated. Excess <strong>dopamine</strong> will thus be nonspecifically<br />

oxidized, resulting in dose-dependent formation<br />

<strong>of</strong> dopaminoquinone. How might this oxidative<br />

mechanism explain <strong>the</strong> increased toxicity <strong>of</strong><br />

methamphetamine compared with amphetamine<br />

itself? A series <strong>of</strong> experiments by Goodwin et al.<br />

systematically highlighted, for <strong>the</strong> first time, some<br />

vital differences between <strong>the</strong> two compounds in<br />

<strong>the</strong>ir respective effects on <strong>the</strong> DAT and <strong>dopamine</strong>rgic<br />

function. 20 Notably, <strong>the</strong> authors demonstrated<br />

that d-methamphetamine is fivefold more effective<br />

at stimulating <strong>the</strong> DAT to release <strong>dopamine</strong> in<br />

hDAT-expressing HEK cells than d-amphetamine at<br />

identical concentrations. Also, in vivo chronoamperometry<br />

in <strong>the</strong> rat nucleus accumbens after<br />

administration <strong>of</strong> an equivalent dose <strong>of</strong> ei<strong>the</strong>r<br />

methamphetamine or amphetamine (5 mg/kg) revealed<br />

that methamphetamine significantly prolonged<br />

<strong>dopamine</strong> clearance time compared with<br />

amphetamine, resulting in greater sustained external<br />

<strong>dopamine</strong> levels. However, no differences between<br />

<strong>the</strong> two were detected at a 1-mg/kg dose (Ref.<br />

20, Fig. 5).<br />

Thesedatasuggestthatathigh(yetequivalent)<br />

doses, methamphetamine has a far greater propensity<br />

than amphetamine to produce accumulation<br />

<strong>of</strong> excess external <strong>dopamine</strong>. The higher resultant<br />

concentrations <strong>of</strong> external <strong>dopamine</strong> conceivably<br />

contribute to <strong>the</strong> greater addictive potential<br />

<strong>of</strong> methamphetamine than with amphetamine and<br />

may also render methamphetamine more likely to<br />

cause oxidative neurodegeneration. Although <strong>the</strong><br />

specific DAT-mediated pharmacology <strong>of</strong> methamphetamine<br />

contributes to its particularly high toxicity<br />

and abuse potential, ano<strong>the</strong>r basic contributing<br />

factor is <strong>the</strong> dose and administration pattern typical<br />

<strong>of</strong> human methamphetamine addicts. Recreational<br />

332 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

methamphetamine users typically take upward <strong>of</strong><br />

100 mg (a “point”) per dose session—for those<br />

that administer it via inhalation (as smoked “crystal<br />

meth” or “ice”), this can be even higher—<strong>of</strong>ten<br />

taking multiple subsequent doses in a “binge” pattern<br />

(every few hours); chronic methamphetamine<br />

addicts can easily consume doses on <strong>the</strong> order <strong>of</strong><br />

grams per day. 145 Intranasal administration <strong>of</strong> a 50mg<br />

dose <strong>of</strong> d-methamphetamine results in an average<br />

plasma concentration <strong>of</strong> 758 nM in human<br />

subjects. 148 Sequential 100-mg doses could thus result<br />

in prolonged plasma drug concentrations ranging<br />

into <strong>the</strong> micromolar levels shown in animal<br />

and in vitro studies to elicit pr<strong>of</strong>ound <strong>dopamine</strong><br />

release, cytotoxicity, and long-lasting downregulation<br />

<strong>of</strong> DAT function. Methamphetamine addicts<br />

that consume gram quantities per day can<br />

achieve steady-state plasma concentrations exceeding<br />

13 �M. 74 Hence, it is likely that <strong>the</strong> transient reduction<br />

in plasmalemmal DAT expression observed<br />

after a single high-level dose <strong>of</strong> methamphetamine<br />

and <strong>the</strong> prolonged reduction seen in chronic<br />

high-dose methamphetamine users are two entirely<br />

different phenomena: <strong>the</strong> former representing a<br />

neuroadaptive feedback mechanism and <strong>the</strong> latter<br />

representing accumulated apoptotic oxidative injury<br />

to <strong>dopamine</strong>rgic neurons.<br />

Conclusion<br />

Neuronal DATs are <strong>the</strong> primary regulators <strong>of</strong><br />

extracellular <strong>dopamine</strong> in <strong>the</strong> brain and are thus responsible<br />

for both <strong>the</strong> duration and <strong>the</strong> magnitude<br />

<strong>of</strong> interneuronal <strong>dopamine</strong>rgic signals. Biogenic<br />

amine <strong>transporter</strong> proteins, such as <strong>the</strong> DAT, are<br />

highly dynamic entities that rapidly cycle between<br />

<strong>the</strong> plasma membrane and early endosomal compartments,<br />

probably in specialized vesicles roughly<br />

similar to synaptic vesicles in size (e.g., Ref. 19). By<br />

altering <strong>the</strong> pattern <strong>of</strong> DAT trafficking, cells can adjust<br />

<strong>the</strong> number <strong>of</strong> active (surface localized) DATs<br />

in real time, allowing for condition-specific adaptive<br />

changes in <strong>dopamine</strong>rgic signaling. DAT trafficking<br />

is regulated by several intracellular signaling cascades<br />

involving protein kinases and phosphatases.<br />

Activation or inhibition <strong>of</strong> PKC, PI3K, and members<br />

<strong>of</strong> <strong>the</strong> MAPK family can trigger <strong>transporter</strong><br />

membrane redistribution; however, <strong>the</strong> identity <strong>of</strong><br />

<strong>the</strong> kinase(s) ultimately responsible for direct phosphorylation<br />

<strong>of</strong> <strong>the</strong> DAT is currently unknown. With<br />

<strong>the</strong> recent evidence that syntaxin 1A and o<strong>the</strong>r scaffolding<br />

proteins interact with <strong>the</strong> DAT, it is certainly<br />

conceivable that central signaling kinases, such as<br />

PKC, affect DAT activity via DAT-associated scaffolding<br />

or cytoskeletal proteins that more directly<br />

control <strong>the</strong> <strong>transporter</strong>’s trafficking fate. In addition,<br />

ligands for certain presynaptic GPCRs found<br />

on <strong>dopamine</strong>rgic neurons (such as <strong>dopamine</strong> D2like<br />

autoreceptors and �-opioid receptors) can also<br />

acutely regulate DAT surface presence, probably by<br />

modulating <strong>the</strong> activity <strong>of</strong> <strong>the</strong>se cell signal transduction<br />

pathways.<br />

Finally, DAT trafficking is acutely regulated by ligands<br />

that directly interact with <strong>the</strong> DAT itself (substrates<br />

and inhibitors). Acute exposure to high doses<br />

<strong>of</strong> substrates (including several amphetaminergic<br />

compounds and <strong>dopamine</strong> itself) elicits a biphasic<br />

effect on DAT trafficking: initially, a rapid yet<br />

transient (beginning within seconds and lasting for<br />

several minutes) increase in surface DAT occurs,<br />

followed by significant <strong>transporter</strong> endocytosis that<br />

begins after approximately 15 min <strong>of</strong> substrate exposure.<br />

In direct contrast with <strong>the</strong> internalizing effects<br />

<strong>of</strong> substrates, cocaine-like DAT inhibitors appear<br />

to induce upregulation <strong>of</strong> surface DAT levels. Upregulation<br />

<strong>of</strong> surface DAT concentration might be<br />

responsible for much <strong>of</strong> <strong>the</strong> anhedonic, depressionlike<br />

symptoms that chronic cocaine addicts experience<br />

during withdrawal—doubtless triggering<br />

cravings for more cocaine. The DAT-internalizing<br />

effect <strong>of</strong> substrates, such as d-amphetamine, might<br />

thus be useful in ameliorating symptoms <strong>of</strong> cocaine<br />

withdrawal. In limited preclinical trials,<br />

d-amphetamine has shown efficacy in reducing cocaine<br />

consumption and craving. 43 Amphetaminergic<br />

prodrugs would be desirable for this purpose,<br />

because <strong>the</strong>ir requisite metabolic activation after<br />

oral administration and slow onset <strong>of</strong> action would<br />

prevent <strong>the</strong> possibility <strong>of</strong> parenteral abuse, a clear<br />

concern in psychostimulant-addicted patients. The<br />

atypical psychostimulant compound modafinil also<br />

shows preclinical promise as a potential <strong>the</strong>rapeutic<br />

for stimulant addiction. Modafinil decreases cocaine<br />

craving during withdrawal 149 and attenuates<br />

<strong>the</strong> effect <strong>of</strong> concomitantly administered cocaine 150<br />

or methamphetamine, 151 warranting investigation<br />

into its effects on DAT trafficking. There is also evidence<br />

that withdrawal from psychostimulant use<br />

is associated with deficits in not only <strong>dopamine</strong>rgic<br />

but also serotonergic function. Moreover, agents<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 333


DAT <strong>Regulation</strong> Schmitt & Reith<br />

that increase serotonergic activity can attenuate reinforcing<br />

effects mediated by excessive <strong>dopamine</strong><br />

release. 152 Therefore, novel amphetaminergic substrates<br />

functioning as dual <strong>dopamine</strong>/serotonin<br />

releasers—but lacking <strong>the</strong> prooxidant metabolites<br />

<strong>of</strong> MDA-like compounds or <strong>the</strong> 5-HT2B receptor agonist<br />

activity <strong>of</strong> fenfluramine and its metabolites—<br />

may be beneficial as treatment agents. 153<br />

Conflicts <strong>of</strong> interest<br />

The authors declare no conflicts <strong>of</strong> interest.<br />

References<br />

1. Ge<strong>the</strong>r, U., P.H. Andersen, O.M. Larsson & A. Schousboe.<br />

2006. Neurotransmitter <strong>transporter</strong>s: molecular<br />

function <strong>of</strong> important drug targets. Trends Pharmacol.<br />

Sci. 27: 375–383.<br />

2. Volz, T.J. & J.O. Schenk. 2005. A comprehensive atlas <strong>of</strong><br />

<strong>the</strong> topography <strong>of</strong> functional groups <strong>of</strong> <strong>the</strong> <strong>dopamine</strong><br />

<strong>transporter</strong>. Synapse 58: 72–94.<br />

3. Gainetdinov, R.R. & M.G. Caron. 2003. Monoamine<br />

<strong>transporter</strong>s: from genes to behavior. Annu. Rev. Pharmacol.<br />

Toxicol. 43: 261–284.<br />

4. Mortensen, O. & S. Amara. 2003. Dynamic regulation<br />

<strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong>. Eur. J. Pharmacol. 479:<br />

159–170.<br />

5. Mash, D.C. 2008. Dopamine <strong>transporter</strong>, disease states<br />

and pathology. In Dopamine Transporters: Chemistry,<br />

Biology and Pharmacology. M.L.Trudell&S.Izenwasser,<br />

Eds.: 29–47. Wiley. Hoboken, NJ.<br />

6. Giros,B.,M.Jaber,S.R.Jones,et al. 1996. Hyperlocomotion<br />

and indifference to cocaine and amphetamine<br />

in mice lacking <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong>. Nature 379:<br />

606–612.<br />

7. Rocha, B.A., F. Fumagalli, R.R. Gainetdinov, et al. 1998.<br />

Cocaine self-administration in <strong>dopamine</strong>-<strong>transporter</strong><br />

knockout mice. Nat. Neurosci. 1: 132–137.<br />

8. Mateo, Y., E.A. Budygin, C.E. John & S.R. Jones. 2004.<br />

Role <strong>of</strong> serotonin in cocaine effects in mice with reduced<br />

<strong>dopamine</strong> <strong>transporter</strong> function. Proc. Natl. Acad. Sci.<br />

USA 101: 372–377.<br />

9. Chen, R., M.R. Tilley, H. Wei, et al. 2006. Abolished<br />

cocaine reward in mice with a cocaine-insensitive<br />

<strong>dopamine</strong> <strong>transporter</strong>. Proc. Natl. Acad. Sci. USA 103:<br />

9333–9338.<br />

10. Tilley, M.R., B. Cagniard, X. Zhuang, et al. 2007. Cocaine<br />

reward and locomotion stimulation in mice with<br />

reduced <strong>dopamine</strong> <strong>transporter</strong> expression. BMC Neurosci.<br />

8: 42. (online).<br />

11. Thomsen, M., F.S. Hall, G.R. Uhl & S.B. Caine. 2009.<br />

Dramatically decreased cocaine self-administration in<br />

<strong>dopamine</strong> but not serotonin <strong>transporter</strong> knock-out<br />

mice. J. Neurosci. 29: 1087–1092.<br />

12. Jones, S.R., R.R. Gainetdinov, M. Jaber, et al. 1998.<br />

Pr<strong>of</strong>ound neuronal plasticity in response to inactivation<br />

<strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong>. Proc. Natl. Acad. Sci. USA<br />

95: 4029–4034.<br />

13. Rice, M.E. & S.J. Cragg. 2008. Dopamine spillover after<br />

quantal release: rethinking <strong>dopamine</strong> transmission<br />

in <strong>the</strong> nigrostriatal pathway. Brain Res. Rev. 58: 303–<br />

313.<br />

14. Kimmel, H.L., F.I. Carroll & M.J. Kuhar. 2000.<br />

Dopamine <strong>transporter</strong> syn<strong>the</strong>sis and degradation rate<br />

in rat striatum and nucleus accumbens using RTI-76.<br />

Neuropharmacology 39: 578–585.<br />

15. Kimmel, H.L., F.I. Carroll & M.J. Kuhar. 2003. Withdrawal<br />

from repeated cocaine alters <strong>dopamine</strong> <strong>transporter</strong>proteinturnoverin<strong>the</strong>ratstriatum.J.Pharmacol.<br />

Exp. Ther. 304: 15–21.<br />

16. Gulley, J.M. & N.R. Zahniser. 2003. Rapid regulation<br />

<strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong> function by substrates, blockers<br />

and presynaptic receptor ligands. Eur. J. Pharmacol.<br />

479: 139–152.<br />

17. Zahniser, N.R. & A. Sorkin. 2004. Rapid regulation <strong>of</strong><br />

<strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong>: role in stimulant addiction?<br />

Neuropharmacology 47: 80–91.<br />

18. Hille, B. 2001. Ion Channels <strong>of</strong> Excitable Membranes,3rd<br />

edn. Sinauer Associates Publishing. Sunderland, MA.<br />

19. Deken, S.L., D. Wang & M. Quick. 2003. Plasma membrane<br />

GABA <strong>transporter</strong>s reside on distinct vesicles<br />

and undergo rapid regulated recycling. J. Neurosci. 23:<br />

1563–1568.<br />

20. Goodwin, J.S., G.A. Larson, J. Swant, et al. 2009. Amphetamine<br />

and methamphetamine differentially affect<br />

<strong>dopamine</strong> <strong>transporter</strong>s in vitro and in vivo. J. Biol.<br />

Chem. 284: 2978–2989.<br />

21. Fleckenstein, A.E., R.R. Metzger, D.G. Wilkins, et al.<br />

1997. Rapid and reversible effects <strong>of</strong> methamphetamine<br />

on <strong>dopamine</strong> <strong>transporter</strong>s. J. Pharmacol. Exp. Ther. 282:<br />

834–838.<br />

22. Kokoshka, J.M., R.A. Vaughan, G.R. Hanson & A.E.<br />

Fleckenstein. 1998. Nature <strong>of</strong> methamphetamineinduced<br />

rapid and reversible changes in <strong>dopamine</strong><br />

<strong>transporter</strong>s. Eur. J. Pharmacol. 361: 269–275.<br />

23. Saunders, C., J.V. Ferrer, L. Shi, et al. 2000.<br />

Amphetamine-induced loss <strong>of</strong> human <strong>dopamine</strong> <strong>transporter</strong><br />

activity: an internalization-dependent and<br />

cocaine-sensitive mechanism. Proc.Natl.Acad.Sci.USA<br />

97: 6850–6855.<br />

334 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

24. Chi, L. & M.E. Reith. 2003. Substrate-induced trafficking<br />

<strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> in heterologously<br />

expressing cells and in rat striatal synaptosomal preparations.<br />

J. Pharmacol. Exp. Ther. 307: 729–736.<br />

25. Sorkina, T., S. Doolen, E. Galperin, et al. 2003.<br />

Oligomerization <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s visualized<br />

in living cells by fluorescence resonance energy transfer<br />

microscopy. J. Biol. Chem. 278: 28274–28283.<br />

26. Sorkina, T., B.R. Hoover, N.R. Zahniser & A. Sorkin.<br />

2005. Constitutive and protein kinase C-induced internalization<br />

<strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> is mediated<br />

by a clathrin-dependent mechanism. Traffic 6: 157–<br />

170.<br />

27. Chen, N. & M.E. Reith. 2008. Substrates dissociate<br />

<strong>dopamine</strong> <strong>transporter</strong> oligomers. J. Neurochem. 105:<br />

910–920.<br />

28. Gulley, J.M., S. Doolen & N.R. Zahniser. 2002.<br />

Brief, repeated exposure to substrates down-regulates<br />

<strong>dopamine</strong> <strong>transporter</strong> function in Xenopus oocytes in<br />

vitro and rat dorsal striatum in vivo. J. Neurochem. 83:<br />

400–411.<br />

29. Kahlig, K.M., J. Javitch & A. Galli. 2004. Amphetamine<br />

regulation <strong>of</strong> <strong>dopamine</strong> transport. Combined measurements<br />

<strong>of</strong> <strong>transporter</strong> currents and <strong>transporter</strong> imaging<br />

support <strong>the</strong> endocytosis <strong>of</strong> an active carrier. J. Biol.<br />

Chem. 279: 8966–8975.<br />

30. Richards, T.L. & N.R. Zahniser. 2009. Rapid substrateinduced<br />

down-regulation in function and surface localization<br />

<strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s: rat dorsal striatum<br />

versus nucleus accumbens. J. Neurochem. 108: 1575–<br />

1584.<br />

31.Johnson,L.A.,C.A.Furman,M.Zhang,et al. 2005.<br />

Rapid delivery <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> to <strong>the</strong> plasmalemmal<br />

membrane upon amphetamine stimulation.<br />

Neuropharmacology 49: 750–758.<br />

32. Furman, C.A., R. Chen, B. Guptaroy, et al.<br />

2009. Dopamine and amphetamine rapidly increase<br />

<strong>dopamine</strong> <strong>transporter</strong> trafficking to <strong>the</strong> surface: livecell<br />

imaging using total internal reflection fluorescence<br />

microscopy. J. Neurosci. 29: 3328–3336.<br />

33. Das, G. 1993. Cocaine abuse in North America: a milestone<br />

in history. J. Clin. Pharmacol. 33: 296–310.<br />

34. Loland, C., R.I. Desai, M.F. Zou, et al. 2008. Relationship<br />

between conformational changes in <strong>the</strong> <strong>dopamine</strong><br />

<strong>transporter</strong> and cocaine-like subjective effects <strong>of</strong> uptake<br />

inhibitors. Mol. Pharmacol. 73: 813–823.<br />

35. Schmitt, K.C., J. Zhen, P. Kharkar, et al. 2008. Interaction<br />

<strong>of</strong> cocaine-, benztropine- and GBR12909like<br />

compounds with wild-type and mutant human<br />

<strong>dopamine</strong> <strong>transporter</strong>s: molecular features that dif-<br />

ferentially determine antagonist-binding properties. J.<br />

Neurochem. 107: 928–940.<br />

36. Kahlig, K.M. & A. Galli. 2003. <strong>Regulation</strong> <strong>of</strong> <strong>dopamine</strong><br />

<strong>transporter</strong> function and plasma membrane expression<br />

by <strong>dopamine</strong>, amphetamine, and cocaine. Eur. J. Pharmacol.<br />

479: 153–158.<br />

37. Beveridge, T.J., H.R. Smith, M.A. Nader & L.J. Porrino.<br />

2009. Abstinence from chronic cocaine selfadministration<br />

alters striatal <strong>dopamine</strong> systems in<br />

rhesus monkeys. Neuropsychopharmacology 34: 1162–<br />

1171.<br />

38. Little,K.Y.,L.Zhang,T.Desmond,et al. 1999. Striatal<br />

<strong>dopamine</strong>rgic abnormalities in human cocaine users.<br />

Am.J.Psychiatry156: 238–245.<br />

39. Mash, D.C., J. Pablo, Q. Ouyang, et al. 2002. Dopamine<br />

transport function is elevated in cocaine users. J. Neurochem.<br />

81: 292–300.<br />

40. Oleson, E.B., S. Talluri, S.R. Childers, et al. 2009.<br />

Dopamine uptake changes associated with cocaine selfadministration.<br />

Neuropsychopharmacology 34: 1174–<br />

1184.<br />

41. Negus, S.S. & N.K. Mello. 2003. Effects <strong>of</strong> chronic<br />

d-amphetamine treatment on cocaine- and foodmaintained<br />

responding under a second-order schedule<br />

in rhesus monkeys. Drug. Alcohol Dep. 70: 39–52.<br />

42. Grabowski, J., J. Shearer, J. Merrill & S.S. Negus. 2004.<br />

Agonist-like, replacement pharmaco<strong>the</strong>rapy for stimulant<br />

abuse and dependence. Addict. Behav. 29: 1439–<br />

1464.<br />

43. Castells, X., M. Casas, X. Vidal, et al. 2007. Efficacy <strong>of</strong><br />

central nervous system stimulant treatment for cocaine<br />

dependence: a systematic review and meta-analysis <strong>of</strong><br />

randomized controlled clinical trials. <strong>Addiction</strong> 102:<br />

1871–1887.<br />

44. Rush, C.R., W.W. Stoops & L.R. Hays. 2009. Cocaine<br />

effects during d-amphetamine maintenance: A. Human<br />

laboratory analysis <strong>of</strong> safety, tolerability and efficacy.<br />

Drug. Alcohol Dep. 99: 261–271.<br />

45. Cervinski, M.A., J.D. Foster & R.A. Vaughan. 2005. Psychoactive<br />

substrates stimulate <strong>dopamine</strong> <strong>transporter</strong><br />

phosphorylation and down-regulation by cocainesensitive<br />

and protein kinase C-dependent mechanisms.<br />

J. Biol. Chem. 49: 40442–40449.<br />

46. Daws, L.C., P.D. Callaghan, J.A. Morón, et al. 2002.<br />

Cocaine increases <strong>dopamine</strong> uptake and cell surface<br />

expression <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s. Biochem. Biophys.<br />

Res. Commun. 290: 1545–1550.<br />

47. Little, K.Y., L.W. Elmer, H. Zhong, et al. 2002. Cocaine<br />

induction <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong> trafficking to <strong>the</strong><br />

plasma membrane. Mol. Pharmacol. 61: 436–445.<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 335


DAT <strong>Regulation</strong> Schmitt & Reith<br />

48. Gorentla, B.K. & R.A. Vaughan. 2005. Differential effects<br />

<strong>of</strong> <strong>dopamine</strong> and psychoactive drugs on <strong>dopamine</strong><br />

<strong>transporter</strong> phosphorylation and regulation. Neuropharmacology<br />

49: 759–768.<br />

49. Vaughan, R.A. 2004. Phosphorylation and regulation<br />

<strong>of</strong> psychostimulant-sensitive neurotransmitter <strong>transporter</strong>s.<br />

J. Pharmacol. Exp. Ther. 310: 1–7.<br />

50. Melikian, H.E. 2004. Neurotransmitter <strong>transporter</strong><br />

trafficking: endocytosis, recycling, and regulation.<br />

Pharmacol. Ther. 104: 17–27.<br />

51. Foster, J.D., M.A. Cervinski, B.K. Gorentla & R.A.<br />

Vaughan. 2006. <strong>Regulation</strong> <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong><br />

by phosphorylation. Handbook Exp. Pharmacol. 175:<br />

197–214.<br />

52. Hoover, B.R., C.V. Everett, A. Sorkin & N.R. Zahniser.<br />

2007. Rapid regulation <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s by<br />

tyrosine kinases in rat neuronal preparations. J. Neurochem.<br />

101: 1258–1271.<br />

53. Vaughan, R.A., R.A. Huff, G.R. Uhl & M.J. Kuhar. 1997.<br />

Protein kinase C–mediated phosphorylation and functional<br />

regulation <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s in striatal<br />

synaptosomes. J. Biol. Chem. 272: 15541–15546.<br />

54. Huff, R.A., R.A. Vaughan, M.J. Kuhar & G.R. Uhl.<br />

1997. Phorbol esters increase <strong>dopamine</strong> <strong>transporter</strong><br />

phosphorylation and decrease transport Vmax. J. Neurochem.<br />

68: 225–232.<br />

55. Zhu, S.J., M.P. Kavanaugh, M.S. Sonders, et al. 1997.<br />

Activation <strong>of</strong> protein kinase C inhibits uptake, currents<br />

and binding associated with <strong>the</strong> human <strong>dopamine</strong><br />

<strong>transporter</strong> expressed in Xenopus oocytes. J. Pharmacol.<br />

Exp. Ther. 282: 1358–1365.<br />

56. Melikian, H.E. & K.M. Buckley. 1999. Membrane trafficking<br />

regulates <strong>the</strong> activity <strong>of</strong> <strong>the</strong> human <strong>dopamine</strong><br />

<strong>transporter</strong>. J. Neurosci. 19: 7699–7710.<br />

57. Daniels, G.M. & S.G. Amara. 1999. Regulated trafficking<br />

<strong>of</strong> <strong>the</strong> human <strong>dopamine</strong> <strong>transporter</strong>. Clathrinmediated<br />

internalization and lysosomal degradation in<br />

response to phorbol esters. J. Biol. Chem. 274: 35794–<br />

35801.<br />

58. Loder, M.K. & H.E. Melikian. 2003. The <strong>dopamine</strong><br />

<strong>transporter</strong> constitutively internalizes and recycles in<br />

a protein kinase C-regulated manner in stably transfected<br />

PC12 cell lines. J. Biol. Chem. 278: 22168–<br />

22174.<br />

59. Foster, J.D., B. Pananusorn & R.A. Vaughan. 2002.<br />

Dopamine <strong>transporter</strong>s are phosphorylated on Nterminal<br />

serines in rat striatum. J. Biol. Chem. 277:<br />

25178–25186.<br />

60. Holton, K.L., M.K. Loder & H.E. Melikian. 2005. Nonclassical,<br />

distinct endocytic signals dictate constitutive<br />

and PKC-regulated neurotransmitter <strong>transporter</strong> internalization.<br />

Nat. Neurosci. 8: 881–888.<br />

61. Foster, J.D., S.D. Adkins, J.R. Lever & R.A. Vaughan.<br />

2008. Phorbol ester induced trafficking-independent<br />

regulation and enhanced phosphorylation <strong>of</strong> <strong>the</strong><br />

<strong>dopamine</strong> <strong>transporter</strong> associated with membrane<br />

rafts and cholesterol. J. Neurochem. 105: 1683–<br />

1699.<br />

62. Chang, M.Y., S.H. Lee, J.H. Kim, et al. 2001. Protein<br />

kinase C-mediated functional regulation <strong>of</strong> <strong>dopamine</strong><br />

<strong>transporter</strong> is not achieved by direct phosphorylation<br />

<strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> protein. J. Neurochem. 77:<br />

754–761.<br />

63. Granäs,C.,J.Ferrer,C.J.Loland,et al. 2003. N-terminal<br />

truncation <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> abolishes phorbol<br />

ester- and substance P receptor-stimulated phosphorylation<br />

without impairing <strong>transporter</strong> internalization.<br />

J. Biol. Chem. 278: 4990–5000.<br />

64. Sung, U., S. Apparsundaram, A. Galli, et al. 2003.<br />

A regulated interaction <strong>of</strong> syntaxin 1A with <strong>the</strong><br />

antidepressant-sensitive norepinephrine <strong>transporter</strong><br />

establishes catecholamine clearance capacity. J. Neurosci.<br />

23: 1697–1709.<br />

65. Dipace, C., U. Sung, F. Binda, et al. 2007. Amphetamine<br />

induces a calcium/calmodulin-dependent protein kinase<br />

II-dependent reduction in norepinephrine <strong>transporter</strong><br />

surface expression linked to changes in syntaxin<br />

1A/<strong>transporter</strong> complexes. Mol. Pharmacol. 71: 230–<br />

239.<br />

66. Lee, K.H., M.Y. Kim, D.H. Kim & Y.S. Lee. 2004. Syntaxin<br />

1A and receptor for activated C kinase interact<br />

with <strong>the</strong> N-terminal region <strong>of</strong> human <strong>dopamine</strong> <strong>transporter</strong>.<br />

Neurochem. Res. 29: 1405–1409.<br />

67. Binda, F., C. Dipace, E. Bowton, et al. 2008. Syntaxin 1A<br />

interaction with <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> promotes<br />

amphetamine-induced <strong>dopamine</strong> efflux. Mol. Pharmacol.<br />

74: 1101–1108.<br />

68. Johnson, L.A., B. Guptaroy, D. Lund, et al. 2005. <strong>Regulation</strong><br />

<strong>of</strong> amphetamine-stimulated <strong>dopamine</strong> efflux by<br />

protein kinase C beta. J. Biol. Chem. 280: 10914–10919.<br />

69. Khoshbouei, H., N. Sen, B. Guptaroy, et al. 2004. Nterminal<br />

phosphorylation <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong><br />

is required for amphetamine-induced efflux. PLoS Biol.<br />

2: 387–393.<br />

70. Sorkina, T., T.L. Richards, A. Rao, et al. 2009. Negative<br />

regulation <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong> endocytosis by<br />

membrane-proximal N-terminal residues. J. Neurosci.<br />

29: 1361–1374.<br />

71. Zhang, L. & M.E. Reith. 1996. <strong>Regulation</strong> <strong>of</strong> <strong>the</strong> functional<br />

activity <strong>of</strong> <strong>the</strong> human <strong>dopamine</strong> <strong>transporter</strong> by<br />

336 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

<strong>the</strong> arachidonic acid pathway. Eur. J. Pharmacol. 315:<br />

345–354.<br />

72. Ingram, S.L. & S.G. Amara. 2000. Arachidonic acid<br />

stimulates a novel cocaine-sensitive cation conductance<br />

associated with <strong>the</strong> human <strong>dopamine</strong> <strong>transporter</strong>. J.<br />

Neurosci. 20: 550–557.<br />

73. Chen, N., M. Appell, J.L. Berfield & M.E. Reith. 2003.<br />

Inhibition by arachidonic acid and o<strong>the</strong>r fatty acids <strong>of</strong><br />

<strong>dopamine</strong> uptake at <strong>the</strong> human <strong>dopamine</strong> <strong>transporter</strong>.<br />

Eur. J. Pharmacol. 478: 89–95.<br />

74. Reese, E.A., J.R. Bunzow, S. Arttamangkul, et<br />

al. 2007. Trace amine-associated receptor 1 displays<br />

species-dependent stereoselectivity for isomers<br />

<strong>of</strong> methamphetamine, amphetamine, and parahydroxyamphetamine.<br />

J. Pharmacol. Exp. Ther. 321:<br />

178–186.<br />

75. Giambalvo, C.T. 2004. Mechanisms underlying <strong>the</strong><br />

effects <strong>of</strong> amphetamine on particulate PKC activity.<br />

Synapse 51: 128–139.<br />

76. Parnas, M.L. & R.A. Vaughan. 2008. Molecular structure<br />

and composition <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s. In<br />

Dopamine Transporters: Chemistry, Biology and Pharmacology.<br />

M.L. Trudell & S. Izenwasser, Eds.: 73–95.<br />

Wiley. Hoboken, NJ.<br />

77. Carvelli, L., J.A. Morón, K.M. Kahlig, et al. 2002. PI3kinase<br />

regulation <strong>of</strong> <strong>dopamine</strong> uptake. J. Neurochem.<br />

81: 859–869.<br />

78. Lin, Z., P.W. Zhang, X. Zhu, et al. 2003. Phosphatidylinositol<br />

3-kinase, protein kinase C, and MEK1/2 kinase<br />

regulation <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s (DAT) require Nterminal<br />

DAT phosphoacceptor sites. J. Biol. Chem. 278:<br />

20162–20170.<br />

79. Lute,B.J.,H.Khoshbouei,C.Saunders,et al. 2008. PI3K<br />

signaling supports amphetamine-induced <strong>dopamine</strong><br />

efflux. Biochem. Biophys. Res. Commun. 372: 656–661.<br />

80. Pearson, G., F. Robinson, T. Beers Gibson, et al. 2001.<br />

Mitogen-activated protein (MAP) kinase pathways:<br />

regulation and physiological functions. Endocr. Rev. 22:<br />

153–183.<br />

81. Morón, J.A., I. Zakharova, J.V. Ferrer, et al. 2003.<br />

Mitogen-activated protein kinase regulates <strong>dopamine</strong><br />

<strong>transporter</strong> surface expression and <strong>dopamine</strong> transport<br />

capacity. J. Neurosci. 23: 8480–8488.<br />

82. Lu, L., E. Koya, H. Zhai, et al. 2006. Role <strong>of</strong> ERK in<br />

cocaine addiction. Trends Neurosci. 29: 695–703.<br />

83. Welchman, R., C. Gordon & R.J. Mayer. 2005. Ubiquitin<br />

and ubiquitin-like proteins as multifunctional signals.<br />

Nat. Rev. Mol. Cell Biol. 6: 599–609.<br />

84. Hicke, L. 2001. Protein regulation by monoubiquitin.<br />

Nat. Rev. Mol. Cell Biol. 2: 195–201.<br />

85. Hicke, L. & H. Riezman. 1996. Ubiquitination <strong>of</strong> a<br />

yeast plasma membrane receptor signals its ligandstimulated<br />

endocytosis. Cell 84: 277–287.<br />

86. Egner, R. & K. Kuchler. 1996. The yeast multidrug <strong>transporter</strong><br />

Pdr5 <strong>of</strong> <strong>the</strong> plasma membrane is ubiquitinated<br />

prior to endocytosis and degradation in <strong>the</strong> vacuole.<br />

FEBS Lett. 378: 177–181.<br />

87. Kolling, R. & S. Losko. 1997. The linker region <strong>of</strong> <strong>the</strong><br />

ABC-<strong>transporter</strong> Ste6 mediates ubiquitination and fast<br />

turnover <strong>of</strong> <strong>the</strong> protein. EMBO J. 16: 2251–2261.<br />

88. Hicke, L. 1997. Ubiquitin-dependent internalization<br />

and down-regulation <strong>of</strong> plasma membrane proteins.<br />

FASEB J. 11: 1215–1226.<br />

89. Miranda, M., C.C. Wu, T. Sorkina, et al. 2005. Enhanced<br />

ubiquitylation and accelerated degradation <strong>of</strong><br />

<strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> mediated by protein kinase<br />

C. J. Biol. Chem. 280: 35617–35624.<br />

90. Miranda, M., K.R. Dionne, T. Sorkina & A. Sorkin.<br />

2007. Three ubiquitin conjugation sites in <strong>the</strong> amino<br />

terminus <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong> mediate protein<br />

kinase C-dependent endocytosis <strong>of</strong> <strong>the</strong> <strong>transporter</strong>.<br />

Mol. Biol. Cell 18: 313–323.<br />

91. Sorkina, T., M. Miranda, K.R. Dionne, et al. 2006. RNA<br />

interference screen reveals an essential role <strong>of</strong> Nedd4–2<br />

in <strong>dopamine</strong> <strong>transporter</strong> ubiquitination and endocytosis.<br />

J. Neurosci. 26: 8195–8205.<br />

92. Strait, K.A. & R. Kuczenski. 1986. Dopamine autoreceptor<br />

regulation <strong>of</strong> <strong>the</strong> kinetic state <strong>of</strong> striatal tyrosine<br />

hydroxylase. Mol. Pharmacol. 29: 561–569.<br />

93. Pothos, E.N., S. Przedborski, V. Davila, et al. 1998. D2-<br />

Like <strong>dopamine</strong> autoreceptor activation reduces quantal<br />

size in PC12 cells. J. Neurosci. 18: 5575–5585.<br />

94. Meiergerd, S.M., T.A. Patterson & J.O. Schenk. 1993.<br />

D2 receptors may modulate <strong>the</strong> function <strong>of</strong> <strong>the</strong> striatal<br />

<strong>transporter</strong> for <strong>dopamine</strong>: kinetic evidence from<br />

studies in vitro and in vivo. J. Neurochem. 61: 764–<br />

767.<br />

95. Cass, W.A. & G.A. Gerhardt. 1994. Direct in vivo<br />

evidence that D2 <strong>dopamine</strong> receptors can modulate<br />

<strong>dopamine</strong> uptake. Neurosci. Lett. 176: 259–263.<br />

96. Zapata, A. & T.S. Shippenberg. 2002. D(3) receptor<br />

ligands modulate extracellular <strong>dopamine</strong> clearance in<br />

<strong>the</strong> nucleus accumbens. J. Neurochem. 81: 1035–1042.<br />

97. Mayfield, R.D. & N.R. Zahniser. 2001. Dopamine<br />

D2 receptor regulation <strong>of</strong> <strong>the</strong> <strong>dopamine</strong> <strong>transporter</strong><br />

expressed in Xenopus laevis oocytes is voltageindependent.<br />

Mol. Pharmacol. 59: 113–121.<br />

98. Wu, Q., M.E. Reith, Q.D. Walker, et al. 2002. Concurrent<br />

autoreceptor-mediated control <strong>of</strong> <strong>dopamine</strong><br />

release and uptake during neurotransmission: an<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 337


DAT <strong>Regulation</strong> Schmitt & Reith<br />

in vivo voltammetric study. J. Neurosci. 22: 6272–<br />

6281.<br />

99. Mason, J.N., H. Farmer, I.D. Tomlinson, et al. 2005.<br />

Novel fluorescence-based approaches for <strong>the</strong> study <strong>of</strong><br />

biogenic amine <strong>transporter</strong> localization, activity, and<br />

regulation. J. Neurosci. Methods 143: 3–25.<br />

100. Bolan, E.A., B. Kivell, V. Jaligam, et al. 2007. D2 receptors<br />

regulate <strong>dopamine</strong> <strong>transporter</strong> function via an extracellular<br />

signal-regulated kinases 1 and 2-dependent<br />

and phosphoinositide 3 kinase-independent mechanism.<br />

Mol. Pharmacol. 71: 1222–1232.<br />

101. Zapata, A., B. Kivell, Y. Han, et al. 2007. <strong>Regulation</strong><br />

<strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong> function and cell surface expression<br />

by D3 <strong>dopamine</strong> receptors. J. Biol. Chem. 282:<br />

35842–35854.<br />

102. Lee, F.J., L. Pei, A. Moszczynska, et al. 2007. Dopamine<br />

<strong>transporter</strong> cell surface localization facilitated by a direct<br />

interaction with <strong>the</strong> <strong>dopamine</strong> D2 receptor. EMBO<br />

J. 26: 2127–2136.<br />

103. Bertolino, A., L. Fazio, A. Di Giorgio, et al. 2009. Genetically<br />

determined interaction between <strong>the</strong> <strong>dopamine</strong><br />

<strong>transporter</strong> and <strong>the</strong> D2 receptor on prefronto-striatal<br />

activity and volume in humans. J. Neurosci. 29: 1224–<br />

1234.<br />

104. Borowsky, B., N. Adham, K.A. Jones, et al. 2001. Trace<br />

amines: identification <strong>of</strong> a family <strong>of</strong> mammalian G<br />

protein-coupled receptors. Proc. Natl. Acad. Sci. USA<br />

98: 8966–8971.<br />

105. Burchett, S.A. & T.P. Hicks. 2006. The mysterious trace<br />

amines: protean neuromodulators <strong>of</strong> synaptic transmission<br />

in mammalian brain. Prog. Neurobiol. 79: 223–<br />

246.<br />

106. Bunzow, J.R., M.S. Sonders, S. Arttamangkul,<br />

et al. 2001. Amphetamine, 3,4-methylenedioxymethamphetamine,<br />

lysergic acid diethylamide, and<br />

metabolites <strong>of</strong> <strong>the</strong> catecholamine neurotransmitters are<br />

agonists <strong>of</strong> a rat trace amine receptor. Mol. Pharmacol.<br />

60: 1181–1188.<br />

107. Xie, Z. & G.M. Miller. 2008. Beta-phenylethylamine alters<br />

monoamine <strong>transporter</strong> function via trace amineassociated<br />

receptor 1: implication for modulatory roles<br />

<strong>of</strong> trace amines in brain. J. Pharmacol. Exp. Ther. 325:<br />

617–628.<br />

108. Xie, Z., S.V. Westmoreland & G.M. Miller. 2008. Modulation<br />

<strong>of</strong> monoamine <strong>transporter</strong>s by common biogenic<br />

amines via trace amine-associated receptor 1 and<br />

monoamine autoreceptors in human embryonic kidney<br />

293 cells and brain synaptosomes. J. Pharmacol.<br />

Exp. Ther. 325: 629–640.<br />

109. Jacob, M.S. & D.E. Presti. 2005. Endogenous psy-<br />

choactive tryptamines reconsidered: an anxiolytic role<br />

for dimethyltryptamine. Med. Hypo<strong>the</strong>ses 64: 930–<br />

937.<br />

110. Maisonneuve, I.M., S. Archer & S.D. Glick. 1994.<br />

U50,488, a kappa opioid receptor agonist, attenuates<br />

cocaine-induced increases in extracellular <strong>dopamine</strong> in<br />

<strong>the</strong> nucleus accumbens <strong>of</strong> rats. Neurosci. Lett. 181: 57–<br />

60.<br />

111. Thompson, A.C., A. Zapata, J.B. Justice, et al. 2000.<br />

Kappa-opioid receptor activation modifies <strong>dopamine</strong><br />

uptake in <strong>the</strong> nucleus accumbens and opposes <strong>the</strong> effects<br />

<strong>of</strong> cocaine. J. Neurosci. 20: 9333–9340.<br />

112. Shuglin, A. & A. Shuglin. 1991. Phenethylamines I Have<br />

Known and Loved.TransformPress.Berkeley,CA.<br />

113. Nichols, D.E. 2004. Hallucinogens. Pharmacol. Ther.<br />

101: 131–181.<br />

114. Rothman, R.B., M.H. Baumann, C.M. Dersch, et al.<br />

2001. Amphetamine-type central nervous system stimulants<br />

release norepinephrine more potently than <strong>the</strong>y<br />

release <strong>dopamine</strong> and serotonin. Synapse 39: 32–41.<br />

115. Rothman, R.B. & M.H. Baumann. 2003. Monoamine<br />

<strong>transporter</strong>s and psychostimulant drugs. Eur J. Pharmacol.<br />

479: 23–40.<br />

116. Blough, B.E. 2008. Dopamine-releasing agents. In<br />

Dopamine Transporters: Chemistry, Biology and Pharmacology.<br />

M.L. Trudell & S. Izenwasser, Eds.: 305–318.<br />

Wiley. Hoboken, NJ.<br />

117. Woolverton, W.L., G. Shybut & C.E. Johanson. 1980.<br />

Structure-activity relationships among some d-Nalkylated<br />

amphetamines. Pharmacol. Biochem. Behav.<br />

13: 869–876.<br />

118. Parker, M.A., D. Marona-Lewicka, D. Kurrasch, et<br />

al. 1998. Syn<strong>the</strong>sis and pharmacological evaluation<br />

<strong>of</strong> ring-methylated derivatives <strong>of</strong> 3,4-methylenedioxyamphetamine<br />

(MDA). J. Med. Chem. 41: 1001–1005.<br />

119. Setola, V., S.J. Hufeisen, K.J. Grande-Allen, et al. 2003.<br />

3,4-methylenedioxymethamphetamine (MDMA, “Ecstasy”)<br />

induces fenfluramine-like proliferative actions<br />

on human cardiac valvular interstitial cells in vitro. Mol.<br />

Pharmacol. 63: 1223–1229.<br />

120. Baumann, M.H., X. Wang & R.B. Rothman. 2007.<br />

3,4-Methylenedioxymethamphetamine (MDMA) neurotoxicity<br />

in rats: a reappraisal <strong>of</strong> past and present findings.<br />

Psychopharmacology 189: 407–424.<br />

121. Rothman, R.B., S. Jayanthi, X. Wang, et al. 2003. Highdose<br />

fenfluramine administration decreases serotonin<br />

<strong>transporter</strong> binding, but not serotonin <strong>transporter</strong> protein<br />

levels, in rat forebrain. Synapse 50: 233–239.<br />

122. Xie, T., L. Tong, M.W. McLane, et al. 2006. Loss <strong>of</strong><br />

serotonin <strong>transporter</strong> protein after MDMA and o<strong>the</strong>r<br />

338 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.


Schmitt & Reith DAT <strong>Regulation</strong><br />

ring-substituted amphetamines. Neuropsychopharmacology<br />

31: 2639–2651.<br />

123. Battaglia, G., S.Y. Yeh, O’E. Hearn, et al.<br />

1987. 3,4-Methylenedioxymethamphetamine and 3,4methylenedioxyamphetamine<br />

destroy serotonin terminals<br />

in rat brain: quantification <strong>of</strong> neurodegeneration<br />

by measurement <strong>of</strong> [ 3 H]paroxetine-labeled serotonin<br />

uptake sites. J. Pharmacol. Exp. Ther. 242: 911–916.<br />

124. Fleckenstein, A.E., H.M. Haughey, R.R. Metzger, et al.<br />

1999. Differential effects <strong>of</strong> psychostimulants and related<br />

agents on <strong>dopamine</strong>rgic and serotonergic <strong>transporter</strong><br />

function. Eur J. Pharmacol. 382: 45–49.<br />

125. Wang, X., M.H. Baumann, H. Xu & R.B. Rothman.<br />

2004. 3,4-methylenedioxy-methamphetamine<br />

(MDMA) administration to rats decreases brain tissue<br />

serotonin but not serotonin <strong>transporter</strong> protein and<br />

glial fibrillary acidic protein. Synapse 53: 240–248.<br />

126. Hahn, M.K. & R.D. Blakely. 2002. Monoamine <strong>transporter</strong><br />

gene structure and polymorphisms in relation to<br />

psychiatric and o<strong>the</strong>r complex disorders. Pharmacogenomics<br />

J. 2: 217–235.<br />

127. Sanders-Bush, E., J.A. Bushing & F. Sulser. 1975. Longterm<br />

effects <strong>of</strong> p-chloroamphetamine and related drugs<br />

on central serotonergic mechanisms. J. Pharmacol. Exp.<br />

Ther. 192: 33–41.<br />

128. Callaghan, P.D., W.A. Owens, M. Javors, et al.<br />

2007. In vivo analysis <strong>of</strong> serotonin clearance in<br />

rat hippocampus reveals that repeated administration<br />

<strong>of</strong> p-methoxyamphetamine (PMA), but not 3,4methylenedioxymethamphetamine<br />

(MDMA), leads to<br />

long-lasting deficits in serotonin <strong>transporter</strong> function.<br />

J. Neurochem. 100: 617–627.<br />

129. Wang, X., M.H. Baumann, H. Xu, et al. 2005. (+/−)-<br />

3,4-Methylenedioxymethamphetamine administration<br />

to rats does not decrease levels <strong>of</strong> <strong>the</strong> serotonin <strong>transporter</strong><br />

protein or alter its distribution between endosomes<br />

and <strong>the</strong> plasma membrane. J. Pharmacol. Exp.<br />

Ther. 314: 1002–1012.<br />

130. Kraner, J.C., D.J. McCoy, M.A. Evans, et al. 2001.<br />

Fatalities caused by <strong>the</strong> MDMA-related drug paramethoxyamphetamine<br />

(PMA). J. Anal. Toxicol. 25: 645–<br />

648.<br />

131. Esteban, B., E. O’Shea, J. Camarero, et al.<br />

2001. 3,4-Methylenedioxymethamphetamine induces<br />

monoamine release, but not toxicity, when administered<br />

centrally at a concentration occurring following<br />

a peripherally injected neurotoxic dose. Psychopharmacology<br />

154: 251–260.<br />

132. Lipton, J.W., E.G. Tolod, V.B. Thompson, et al. 2008.<br />

3,4-Methylenedioxymethamphetamine (ecstasy) pro-<br />

motes <strong>the</strong> survival <strong>of</strong> fetal <strong>dopamine</strong> neurons in culture.<br />

Neuropharmacology 55: 851–859.<br />

133. Paris, J.M. & K.A. Cunningham. 1992. Lack <strong>of</strong> serotonin<br />

neurotoxicity after intraraphe microinjection <strong>of</strong><br />

(+)-3,4-methylenedioxymethamphetamine (MDMA).<br />

Brain Res. Bull. 28: 115–119.<br />

134. Bai, F., S.S. Lau & T.J. Monks. 1999. Glutathione and<br />

N-acetylcysteine conjugates <strong>of</strong> alpha-methyl<strong>dopamine</strong><br />

produce serotonergic neurotoxicity: possible role in<br />

methylenedioxyamphetamine-mediated neurotoxicity.<br />

Chem. Res. Toxicol. 12: 1150–1157.<br />

135. De la Torre, R., M. Farré, P.N. Roset, et al. 2004.<br />

Human pharmacology <strong>of</strong> MDMA: pharmacokinetics,<br />

metabolism, and disposition. Ther. Drug Mon. 26: 137–<br />

144.<br />

136. Jones, D.C., C. Duvauchelle, A. Ikegami, et al. 2005.<br />

Serotonergic neurotoxic metabolites <strong>of</strong> ecstasy identified<br />

in rat brain. J. Pharmacol. Exp. Ther. 313: 422–431.<br />

137. Capela, J.P., C. Macedo, P.S. Branco, et al. 2007. Neurotoxicity<br />

mechanisms <strong>of</strong> thioe<strong>the</strong>r ecstasy metabolites.<br />

Neuroscience 146: 1743–1757.<br />

138. Jameson, G.N., J. Zhang, R.F. Jameson & W. Linert.<br />

2004. Kinetic evidence that cysteine reacts with<br />

dopaminoquinone via reversible adduct formation to<br />

yield 5-cysteinyl-<strong>dopamine</strong>: an important precursor <strong>of</strong><br />

neuromelanin. Org. Biomol. Chem. 2: 777–782.<br />

139. Erives, G.V., S.S. Lau & T.J. Monks. 2008. Accumulation<br />

<strong>of</strong> neurotoxic thioe<strong>the</strong>r metabolites <strong>of</strong> 3,4-<br />

(+/−)-methylenedioxymethamphetamine in rat brain.<br />

J. Pharmacol. Exp. Ther. 324: 284–291.<br />

140. Alves, E., Z. Binienda, F. Carvalho, et al. 2009.<br />

Acetyl-l-carnitine provides effective in vivo neuroprotection<br />

over 3,4-methylenedioximethamphetamineinduced<br />

mitochondrial neurotoxicity in <strong>the</strong> adolescent<br />

rat brain. Neuroscience 158: 514–523.<br />

141. Shoblock, J., E.B. Sullivan, I. Maisonneuve & S. Glick.<br />

2003. Neurochemical and behavioral differences between<br />

d-methamphetamine and d-amphetamine in<br />

rats. Psychopharmacology 165: 359–369.<br />

142. Cadet, J.L., S. Jayanthi & X. Deng. 2003. Speed kills:<br />

cellular and molecular bases <strong>of</strong> methamphetamineinduced<br />

nerve terminal degeneration and neuronal<br />

apoptosis. FASEB J. 17: 1775–1788.<br />

143. McCann, U.D., H. Kuwabara, A. Kumar, et al. 2008.<br />

Persistent cognitive and <strong>dopamine</strong> <strong>transporter</strong> deficits<br />

in abstinent methamphetamine users. Synapse 62: 91–<br />

100.<br />

144. Sekine, Y., Y. Ouchi, G. Sugihara, et al. 2008. Methamphetamine<br />

causes microglial activation in <strong>the</strong> brains <strong>of</strong><br />

human abusers. J. Neurosci. 28: 5756–5761.<br />

Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences. 339


DAT <strong>Regulation</strong> Schmitt & Reith<br />

145. Davidson, C., A.J. Gow, T.H. Lee & E.H. Ellinwood.<br />

2001. Methamphetamine neurotoxicity: necrotic and<br />

apoptotic mechanisms and relevance to human abuse<br />

and treatment. Brain Res. Brain Res. Rev. 36: 1–22<br />

146. Hashimoto, K., H. Tsukada, S. Nishiyama, et al. 2004.<br />

Protective effects <strong>of</strong> N-acetyl-l-cysteine on <strong>the</strong> reduction<br />

<strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s in <strong>the</strong> striatum <strong>of</strong><br />

monkeys treated with methamphetamine. Neuropsychopharmacology<br />

29: 2018–2023.<br />

147. Miyazaki, I. & M. Asanuma. 2009. Approaches to prevent<br />

<strong>dopamine</strong> quinone-induced neurotoxicity. Neurochem.<br />

Res. 34: 698–706.<br />

148. Harris, D.S., H. Boxenbaum, E.T. Everhart, et al.<br />

2003. The bioavailability <strong>of</strong> intranasal and smoked<br />

methamphetamine. Clin. Pharmacol. Ther. 74: 475–<br />

486.<br />

149. Dackis, C.A., K.M. Kampman, K.G. Lynch, et al.<br />

A double-blind, placebo-controlled trial <strong>of</strong> modafinil<br />

for cocaine dependence. Neuropsychopharmacology 30:<br />

205–211.<br />

150. Hart, C.L., M. Haney, S.K. Vosburg, et al. 2008. Smoked<br />

cocaine self-administration is decreased by modafinil.<br />

Neuropsychopharmacology 33: 761–768.<br />

151. Zolkowska, D., R. Jain, R. Rothman, et al. 2009. Evidence<br />

for <strong>the</strong> involvement <strong>of</strong> <strong>dopamine</strong> <strong>transporter</strong>s in<br />

behavioral stimulant effects <strong>of</strong> modafinil. J. Pharmacol.<br />

Exp. Ther. 329: 738–746.<br />

152. Rothman, R.B. & M.H. Baumann. 2006. Balance between<br />

<strong>dopamine</strong> and serotonin release modulates behavioral<br />

effects <strong>of</strong> amphetamine-type drugs. Ann. N. Y.<br />

Acad. Sci. 1074: 245–260.<br />

153. Rothman, R.B., B.E. Blough & M.H. Baumann. 2008.<br />

Dopamine/serotonin releasers as medications for stimulant<br />

addictions. Prog. Brain Res. 172: 385–406.<br />

340 Ann. N.Y. Acad. Sci. 1187 (2010) 316–340 c○ 2010 New York Academy <strong>of</strong> Sciences.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!