You are on page 1of 8

Pharmacology & Therapeutics 125 (2010) 394–401

Contents lists available at ScienceDirect

Pharmacology & Therapeutics


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / p h a r m t h e r a

Associate editor: H. Bönisch

Neuronal and non-neuronal GABA transporters as targets for antiepileptic drugs


Karsten K. Madsen a, H. Steve White b, Arne Schousboe a,⁎
a
Department of Pharmacology and Pharmacotherapy, Faculty of Pharmaceutical Sciences, University of Copenhagen, Copenhagen, Denmark
b
Anticonvulsant Drug Development Program, Department of Pharmacology and Toxicology, University of Utah, Salt Lake City, Utah, United States

a r t i c l e i n f o a b s t r a c t

Keywords: Epileptic seizure activity is associated with an imbalance between excitatory and inhibitory synaptic
Tiagabine activities. The latter is mediated by GABA, and several currently used antiepileptic drugs target entities of the
SNAP-5114 GABAergic synapse such as the receptors or the inactivation mechanism consisting of transmembrane
EF 1502
transport and enzymatic degradation. The development of tiagabine selectively inhibiting the GABA
Extrasynaptic GABA receptors
transporter GAT1 constitutes a proof of concept that the GABA transporters are interesting drug targets in the
Betaine/GABA transporter-1
context of antiepileptic drugs. The review provides a detailed analysis of the role of such transporters
pointing in particular to an interesting role of the transporters located extrasynaptically. It is suggested that
the betaine–GABA transporter BGT1 should receive particular interest in this context as the GABA analogue
EF 1502 (N-[4,4-bis(3-methyl-2-thienyl)-3-butenyl]-4-(methylamino)-4,5,6,7-tetrahydrobenzo[d]isoxazol-
3-ol) has been shown to possess a novel anticonvulsant profile in animal models of epilepsy, involving the
ability to inhibit GABA transport mediated by GAT1 and BGT1 at the same time.
© 2009 Elsevier Inc. All rights reserved.

Contents

1. Epilepsy: a heterogeneous neurological disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 394


2. GABAergic neurotransmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
3. Characterization of neuronal and non-neuronal GABA transporters . . . . . . . . . . . . . . . . . . . . . . 395
4. The era of cloning: impact on GAT pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
5. Localization of the GATs in healthy tissue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
6. Introducing lipophilic aromatic side chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
7. GAT inhibitors and the control of neuronal excitability . . . . . . . . . . . . . . . . . . . . . . . . . . 398
8. Anticonvulsant activity of GAT inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
9. Therapeutic aspects of GAT inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399

1. Epilepsy: a heterogeneous neurological disorder patients about 20% are either inadequately treated or treated at the
expense of severe side effects leaving about 30% of the patients with
Epilepsy is a heterogeneous neurological disorder characterized by refractory epilepsy (White, 1999; Madsen et al., 2009). A common
the onset of spontaneous convulsive and non-convulsive seizures. feature in epilepsy seems to involve hyperexcitable neurons which
Worldwide over 50 million people are affected and it is estimated that discharge in a highly synchronized manner to produce a seizure. It has
only 50% of the patients are adequately treated for their symptoms with been suggested that seizure activity is a consequence of an imbalance
the currently available antiepileptic drugs (AEDs). Of the remaining between the inhibitory and excitatory neurotransmission (Dalby &
Mody, 2001). The current available AEDs can be divided into three
groups based on their mechanism of action which includes: modulation
of voltage-dependent ion-channels, attenuation of excitatory neuro-
⁎ Corresponding author. Department of Pharmacology and Pharmacotherapy,
Faculty of Pharmaceutical Sciences, University of Copenhagen, Universitetsparken 2,
transmission, and enhancement of the inhibitory neurotransmission
2100 Copenhagen, Denmark. Tel.: +45 35 33 63 30. (Kwan et al., 2001). Inhibiting GABA transaminase (GABA-T), the
E-mail address: as@farma.ku.dk (A. Schousboe). catabolic enzyme involved in the degradation of GABA, has been shown

0163-7258/$ – see front matter © 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.pharmthera.2009.11.007
K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401 395

to increase the GABA content in nerve terminals thereby facilitating the astrocytes is metabolized and lost thereby draining the presynaptic
inhibitory GABAergic neurotransmission which mediates its anticon- neuron of inhibitory neurotransmitter (Schousboe, 1990, 2000; White
vulsant activity (Gale & Iadarola, 1980). Moreover, tiagabine, an et al., 2002).
inhibitor of GABA transport which increases the extracellular levels of The initial characterization of GATs has to a large extent been made
GABA has been shown to mediate seizure protection (Nielsen et al., possible due to a naturally occurring compound from the fly agaric
1991). However, increasing the amount of extracellular GABA does not mushroom, muscimol (Fig. 1). Muscimol is a potent GABAA receptor
alleviate all seizure types. Actually, tiagabine is mainly used as an add-on agonist with weak inhibitory effect on GAT and is a substrate for the
antiepileptic drug for treatment of partial seizures and unfortunately GABA-T. The pharmacological profile of muscimol was separated into
tiagabine is not without severe adverse side effects (Schousboe & White, the rigid GABA analogue, THIP (Gaboxadol, (4,5,6,7-tetrahydroisox-
2009). In fact absence seizures commonly seen in children can be azolo[5,4-c]pyridin-3-ol)) which is a selective extrasynaptic GABAA
exacerbated by non-selectively increasing GABA levels (Inaba et al., receptor agonist (Wafford & Ebert, 2006; Cremers & Ebert, 2007) and
2009). the rigid β-alanine analogue THPO ((4,5,6,7-tetrahydroisoxazolo[4,5-
c]pyridin-3-ol)) which retained GAT activity (Krogsgaard-Larsen &
2. GABAergic neurotransmission Johnston, 1975). As can be seen from Table 1, β-alanine and THPO
display a two-fold selectivity for astrocytic GAT compared to neuronal
After the independent discovery of GABA in 1950 by Awapara et al. GAT. However, the conversion of the isoxazolol group of THPO to the
(1950), Roberts and Frankel (1950) and Udenfriend (1950) much carboxyl group resulted in the generation of two compounds, nipecotic
attention was given to this small flexible molecule. What has come to acid and guvacine which are very potent inhibitors but with a
be widely accepted today is that GABA is the principal inhibitory concomitant loss of selectivity for astrocytic GAT (Schousboe et al.,
neurotransmitter in the mammalian brain; although this notion was 1979, 1981). Nipecotic acid and guvacine have subsequently been used
highly controversial at first (Krnjevic, 2004). The major route of GABA as lead structures in the efforts to synthesize potent GAT inhibitors.
biosynthesis involves decarboxylation of glutamate yielding GABA The glial selectivity of THPO prompted the synthesis of a series of
and CO2 via the enzyme glutamate decarboxylase (GAD) (Roberts & compounds where the secondary amino group of THPO was moved to
Kuriyama, 1968). GAD exists in two isoforms GAD65 and GAD67 both an exocyclic position becoming a primary amino group (exo-THPO)
activated by pyridoxal 5′-phosphate with the former isoform being in (Fig. 1). The pharmacological action of selected GAT inhibitors from
a readily inducible state pending neuronal activity and the latter this series is summarized in Table 1. Of special interest is its mono-
isoform being almost fully activated. GAD67 is found ubiquitously in methylated derivative (N-methyl-exo-THPO) which is the most
GABAergic neurons, whereas GAD65 is preferentially located in the selective inhibitor for astrocytic versus neuronal GAT seen until now
nerve terminals. Based on these findings it has been suggested that (Falch et al., 1999).
GAD65 is specialized to readily synthesize GABA under short-term
demand (Martin & Rimvall, 1993). In the nerve terminals GABA can 4. The era of cloning: impact on GAT pharmacology
be released to the synaptic cleft by two different pathways; i.e., via a
Ca2+ dependent vesicular release or Ca2+ independent release via The existence of pharmacologically different GAT populations on
transporter reversal (Belhage et al., 1993). Upon release to the neurons and astrocytes was further elaborated as cloning revealed
synaptic cleft GABA interacts with GABA receptors located pre and four subtypes of high affinity GATs. In 1986 an 80 kiloDalton (kDa)
postsynaptically. Presynaptic GABA receptors metabotropically me- glycoprotein was isolated from the rat brain with an apparent Km for
diate a negative feedback mechanism on GABA release whereas, GABA transport of 3 μM and it showed a dependence for Na+ and Cl−
postsynaptic GABA receptors can be both metabotropic (GABAB) and for transport. In comparison, the apparent Km for GABA for neuronal
ionotropic (GABAA) and mediate usually hyperpolarization of the cell and astrocytic GAT is 8 and 32 μM, respectively. Antibodies raised
(Watanabe et al., 2002). After dissociation from the receptor complex against GAT failed to cross-react with GABA receptors suggesting little
GABA is transported back into the presynaptic nerve terminal or into homology between these proteins (Radian et al., 1986). Four years
surrounding astrocytes via a high affinity GABA transport system later Guastella et al. (1990) succeeded in cloning this transporter and
thereby terminating GABA's inhibitory action (Iversen & Neal, 1968). designated it GAT-1 which consists of 599 amino acids with a Km for
It has been estimated that the neuronal GABA transport system is GABA of 7 μM. Inhibitors of neuronal and glial GATs suggested that
three- to six-fold more efficient than the astrocytic GABA transport GAT-1 displayed a similar pharmacology as the neuronal GAT
mechanisms which could indicate a reutilization of GABA taken up in (Guastella et al., 1990). That same year the human GAT-1 was cloned.
the neuron (Hertz & Schousboe, 1987). The degradation of GABA is It was found to consist of 599 amino acids (Nelson et al., 1990). From
brought about via the enzymes GABA transaminase (GABA-T) and the mouse, four GATs have been cloned and named GAT1, GAT2,
succinic semialdehyde dehydrogenase yielding succinate. GABA-T is GAT3, and GAT4 encoding proteins of 598, 614, 602, and 627 amino
located in both neurons and astrocytes with the highest activity in the acids, respectively with a Km for GABA of 7, 79, 18, and 0.7 μM,
latter cell type (Schousboe et al., 1980). respectively. GAT2 also shows affinity for betaine with a Km of 398 μM
(Liu et al., 1992, 1993). From canine, human and rat, a betaine–GABA
3. Characterization of neuronal and non-neuronal GABA transporters transporter (BGT-1) has been cloned which is homologous to mouse
GAT2, encoding proteins of 614 amino acids for the two former
Ten years before Iversen and Neal (1968) proved the existence of species and 628 amino acids for the latter species, and like mouse
high affinity GABA transporters (GATs) in neurons and astrocytes, GAT2 they show a greater affinity for GABA than for betaine
Elliott and van Gelder (1958), demonstrated that GABA in the (Yamauchi et al., 1992; Borden et al., 1995; Burnham et al., 1996).
incubation medium could accumulate in slices of cerebral cortices. The rat GAT-2 and GAT-3 clones are homologues of mouse GAT3 and
The accumulation of GABA in neurons and astrocytes can be blocked GAT4 encoding proteins of 602 and 627 amino acids, respectively with
by the GABA analogue diaminobutyric acid (DABA) and β-alanine, a Km for GABA of 8 and 12 μM. Both were sensitive to β-alanine
respectively, revealing a very important difference in the pharmacol- suggesting a greater similarity to glial GAT (Borden et al., 1992;
ogy of neuronal and astrocytic GAT's (Iversen & Neal, 1968; Iversen & Christiansen et al., 2007). The human GAT-2 and GAT-3 clones are also
Kelly, 1975). It has been hypothesized that selective inhibition of homologous to mouse GAT3 and GAT4 and encode proteins of 602 and
astrocytic GABA transport could possess superior anticonvulsant 632 amino acids (Borden et al., 1994; Christiansen et al., 2007).
properties compared to inhibition of neuronal GABA transport, since Unfortunately, the cloning of the four GATs has resulted in a rather
DABA has been shown to be proconvulsant and GABA taken up in confusing nomenclature. Table 2 summarizes the GAT nomenclature
396 K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401

Fig. 1. Chemical structures of muscimol, THIP (Gaboxadol) and selected of GABA uptake inhibitors used for the characterization of neuronal and non-neuronal GABA transporters.

categorized in columns for species homology. Throughout the et al., 2005). The expression of GAT2 in the brain is very limited and it
remainder of this review the nomenclature as proposed by the is found in the leptomeninges and the neonatal brain and maybe at
HUGO Gene Nomenclature Committee will be used. It seems that very low levels throughout the brain at extrasynaptic sites on both
GAT1 behaves as the neuronal GAT whereas GAT2 and GAT3 resemble neurons and astrocytes (Liu et al., 1993; Borden, 1996; Conti et al.,
the astrocytic GAT when transport of GABA is inhibited with DABA 1999). BGT1 has been found in the hippocampus and cortex,
and β-alanine, respectively. surprisingly not located closely to GABAergic synapses but in an
extrasynaptic region involving to a large extent astrocytes (Borden
5. Localization of the GATs in healthy tissue et al., 1995; Zhu & Ong, 2004) which could indicate a specialized
pharmacological action different from the other GATs. Both GAT2 and
GAT1 and GAT3 have been found exclusively in the brain, whereas BGT1 are found on the basolateral membrane in polarized MDCK cells
GAT2 and BGT1 have been found in multiple other organs. GAT1 is after transfection (Ahn et al., 1996) strikingly opposite to the
widely distributed throughout the entire brain (Durkin et al., 1995) localization of GAT1 and GAT3. In the developing mouse cerebral
and its expression closely follows that of GABAergic pathways (Radian cortex, the onset of GAT1 expression in the axon terminals is closely
et al., 1990). GAT1 is primarily found on presynaptic neurons in the related to synapse formation. GAT3 expression follows a couple of
synapse and to a minor extent on distal astrocytic processes in close days later with astrocyte sealing of the synapse (Takayama & Inoue,
proximity to axon terminals forming symmetric synapses (Borden, 2005). Fig. 2 is a graphical representation of the primary localization
1996; Conti et al., 1998). In polarized Madin Darby Canine Kidney of GAT1, BGT1 and GAT3 summarized in one synapse.
(MDCK) cells GAT1 was after transfection into the cells found
exclusively on the apical surface suggesting an axonal localization 6. Introducing lipophilic aromatic side chains
(Pietrini et al., 1994). The localization of GAT3 is more restricted when
compared to GAT1 showing strong intensity in retina, olfactory bulb, The subtype selectivity of THPO and the exocyclic derivatives
brainstem, diencephalon and low levels in hippocampus and cortex. showed an interesting feature; i.e., they were all GAT1 selective but the
GAT3 is predominantly expressed on distal astrocytic processes in inhibitory effect was greater on astrocytic GAT rather than neuronal
direct contact with GABAergic neurons (Durkin et al., 1995; Minelli GAT. This was somewhat unexpected given the localization of GAT1.
et al., 1996). Expression in MDCK cells places GAT3 on the apical Direct intracerebroventricular injections of N-methyl-exo-THPO, N-
surface (Ahn et al., 1996; Borden, 1996). The finding of GAT3 on ethyl-exo-THPO, exo-THPO, and tiagabine were evaluated for anti-
neurons in human tissue has been investigated further and it is convulsant activity in the audiogenic seizure (AGS) susceptible Frings
debated whether this localization is an artifact introduced after the mouse, a reflex seizure model that is non-discriminatory with respect
tissue has been removed in response to metabolic stress (Melone to seizure type. The IC50 value of the drugs obtained in neurons and
K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401 397

Table 1
Inhibition of GABA uptake by selected GABA analogues.
Data summarized from Schousboe (1979), Larsson et al. (1981, 1983, 1986, 1988), Suzdak
et al. (1992), Borden (1996), Thomsen et al. (1997), Bolvig et al. (1999), Falch et al. (1999),
White et al. (2002) and Sarup et al. (2003).

Compound GABA uptake inhibition IC50 or ⁎Km/i (μM)

Neurons Astrocytes GAT1 GAT2 GAT3 GAT4

GABA 8a 32a 17a 51a 15a 17a


Nipecotic acid 12 16 24 >1000 113 159
Guvacine 32 29 39 >1000 228 378
DABA 1000 >5000 128 528c 300 710
ACHC 200 700 132 1070c >1000 >10,000
β-Alanine 1666b 843b 2920 1100c 66 110
THPO 501b 262b 1300 3000 800 5000
Exo-THPO 780 250 1000 3000 >3000 >3000
N-methyl-exo-THPO 405 48 450 >3000 >3000 >3000
N-ethyl-exo-THPO 390 301 320 >1000 >1000 >1000
N-2-hydroxyethyl- 300 200 >500 >500 >500 >500
exo-THPO
N-4-phenylbutyl- 100 15 7 >500 >1000 >1000
exo-THPO
N-acetyloxyethyl- 200 18 550 >1000 >1000 >1000 Fig. 2. A graphical representation of the primary subcellular localization of GAT1, BGT1, and
exo-THPO GAT3. GAT1 is primarily located presynaptically. GAT3 is primarily located on distal astrocytic
(R/S)-EF1502 2 2 7 26 >300 >300 processes in close proximity to the synapse. BGT1 is located in the extrasynaptic region.
(R)-EF1502 1.5 0.65 4 22 >150 >150
(S)-EF1502 >100 >100 120 34 >150 >150
N-DPB-THPO 38b 26 30 200 >300 >1000
Overcoming this obstacle to produce systemically active com-
N-DPB-Nipecotic acid 1.3b 2.0b 0.64 7210c 550 4390
N-DPB-guvacine 4.9b 4.2b – – – – pounds was initiated by introducing lipophilic aromatic side chains
N-DPB-exo-THPO 1.4 0.6 6 100 >100 >100 such as 4,4-diphenylbut-3-en-1-yl (DPB) to the parent compounds
N-DPB-N-methyl- 5 2 2 200 >100 >100 such as R-nipecotic acid and guvacine yielding SKF89976A and
exo-THPO SKF100330A, respectively. These SKF compounds were approximately
NNC 05-2090 – – 19b 1.4b 41b 15b
20 times more potent than the parent structures thereby producing
SNAP-5114c – – 388 140 21 5
NO-711 1.24 0.64 – – – – inhibitors with an IC50 in the mid nanomolar range. Moreover, the two
Tiagabine 0.45 0.18 0.11 >100 >100 800 compounds showed no inhibitory effect towards 3H-muscimol
DABA: diaminobutyric acid; ACHC: amino-cyclohexane-carboxylic acid. binding in the nanomolar range confirming GAT selectivity (Yunger
a
Km. et al., 1984; Ali et al., 1985).
b
Ki. Based on this discovery, guvacine, R- and S-nipecotic acid were
c
Human homologue of BGT1. used as the GABA mimetic moiety in the synthesis of more potent GAT
inhibitors in which the length and electro-negativity of the linker was
astrocytes correlated extremely well with their anticonvulsant activity altered along with the di-aromatic region at the end of the linker
in the AGS susceptible Frings mouse. Furthermore, the anticonvulsant (Fig. 3) (Braestrup et al., 1990; Andersen et al., 1993, 1994, 1999;
activity correlated (r2 = 0.988) nicely with the IC50 obtained in the Knutsen et al., 1999; Andersen et al., 2001a,b). One compound which
astrocytes contrary to the IC50 in neurons (r2 = 0.5336) suggesting that gained particular interest was tiagabine (Gabitril®), a derivative of R-
selective inhibition of astrocytic GAT has therapeutic potential (White nipecotic acid. Tiagabine is approved for the treatment of partial
et al., 2002). epileptic seizures and is the only GAT inhibitor currently available for
The compounds described so far possess IC50 values in the low the treatment of epilepsy. Recently, selected compounds from the
micromolar range (Table 1); however, the therapeutic potential of above mentioned series were reevaluated for inhibitory effect on the
these molecules is limited due to their inability to penetrate the BBB at four cloned GATs in order to supplement the inhibitory effect
physiological pH. This inability to penetrate the blood–brain barrier is previously measured on synaptosomes. Not surprisingly these
due to the zwitterion characteristic of the drugs which is governed by compounds were found to preferentially inhibit GAT1; however,
the ionized/unionized (I/U) ratio of the compounds (Krogsgaard- some effect was also seen on the other subtypes. It seems that by
Larsen et al., 2000). increasing the length of the carbohydrate linker of diphenylamine
derivates a greater inhibitory effect was obtained on the non-GAT1
subtypes along with an almost unchanged affinity for GAT1; i.e., they
Table 2 lost their GAT1 preference (Clausen et al., 2006).
GABA transporter nomenclature across species.
Very few compounds developed possess selectivity towards non-
Species Nomenclature GAT1 subtypes. SNAP-5114 (Fig. 3) was synthesized and tested on the
SLC6 gene SLC6a1 SLC6a12 SLC6a13 SLC6a11
human GAT clones and found to display an 80-fold selectivity towards
a b c
GAT3 over GAT1 (Borden et al., 1994). This selectivity is not as
Rat GAT-1 BGT-1 GAT-2 GAT-3c
pronounced in mice where the selectivity between mouse GAT3 and
Human GAT-1d BGT-1e GAT-2f GAT-3g
Mouse GAT1h GAT2h GAT3h GAT4h GAT1 is about 30-fold (Kragler et al., 2005). A series of compounds
HUGO GAT1 BGT1 GAT2 GAT3 have recently been synthesized using proline and pyrrolidine-2-acetic
a
Guastella et al. (1990). acid as the GABA mimetic moiety substituted with three lipophilic side
b
Burnham et al. (1996). chains DPB, the side chain from tiagabine (4,4-[di(3-methylthiophen-
c
Borden et al. (1992). 2-yl)]phenylbut-3-en-1-yl), and the side chain from SNAP-5114 (2-
d
Nelson et al. (1990). [tris(4-methoxyphenyl)methoxy]ethyl). The results obtained with
e
Borden et al. (1995).
f
Christiansen et al. (2007).
these molecules suggest that the SNAP-5114 side chain conveys
g
Borden et al. (1994). selectivity towards GAT3 which is independent of the affinity of the
h
Liu et al. (1993). GABA mimetic moiety towards GAT1 and GAT3 (Fulep et al., 2006).
398 K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401

Fig. 3. General structure of GAT inhibitors. The vertical dashed lines cut the structure into the diaryl-region, the linker, and the amino acid moiety. DPB is presented as an example of
an aromatic side chain. Tiagabine, EF1502, SNAP-5114 and Lu-32-176B are the compounds used for anticonvulsant testing as described in this review.

Unfortunately, a drug with a better selectivity profile than SNAP-5114 receptors. Synaptically located GABAA receptors mediate a phasic
has not been found in either study, thereby leaving SNAP-5114 as one inhibition in response to high concentrations (low micromolar) of
of the best commercially available pharmacological tools for evaluat- intermittently released GABA from the presynaptic neuron. Extra-
ing the impact of GAT2/3 inhibition. synaptically located GABAA receptors respond to ambient levels of
Recently, the development of EF1502 ((RS)-4-[N-[1,1-bis(3-methyl- GABA that have diffused away from the synapse and mediate a
2-thienyl)but-1-en-4-yl]-N-methylamino]-4,5,6,7-tetrahydrobenzo[d] continuous tonic inhibition and show no, or very little, desensitization
isoxaxol-3-ol) in which N-methyl-exo-THPO serves as GABA mimetic in contrast to their synaptic counterparts (Saxena & MacDonald, 1994;
moiety substituted with the lipophilic side chain from tiagabine (Fig. 3) Brown et al., 2002).
was reported to be a GAT1/BGT1 selective inhibitor. Interestingly, GAT1 From adult rat hippocampal slices the tonic conductance in
selectivity was confined to R-EF1502 whereas the BGT1 selectivity was granule cells of the dentate gyrus is about 4 times larger than
independent of the stereochemistry of the GABA mimetic moiety spontaneous inhibitory postsynaptic currents (IPSC) when activated
(Table 1) (Clausen et al., 2005). In summary, tiagabine, SNAP-5114, and only by ambient GABA. Furthermore, by inhibiting GAT1 the tonic
EF1502 represent the three currently available drugs for investigation of inhibition was increased by 330% without affecting spontaneous IPSC
the therapeutic potential of the different GAT subtypes. (sIPSC) and zolpidem enhanced the phasic conductance by 66%
In addition to the subtype selectivity of the amino acid based without producing any effect on the tonic current (Nusser & Mody,
inhibitors like DABA and β-alanine and the lipophilic aromatic inhibitors 2002). From 3 to 4 week old guinea pig hippocampal slices a tonic
like tiagabine, EF1502, and SNAP-5114, these compounds also display conductance was seen in interneurons but not in pyramidal cells
different kinetics regarding inhibition of GABA uptake and some act as under basal conditions. The tonic current was pharmacologically
substrates for transport. While the simple amino acid related inhibitors different from phasic currents and much larger. Inhibition of GAT1
(Fig. 1) have proven to be competitive substrates for the transporter, the produced a comparable tonic inhibition in pyramidal cells as that
lipophilic derivatives (Fig. 3) display multiple forms of kinetics and are observed in interneurons under basal conditions without affecting the
not substrates (Larsson et al., 1988). From the recently reevaluated tonic conductance in interneurons. Reducing the tonic conductance in
lipophilic compounds described above, the individual inhibitors showed interneurons increases the inhibitory drive to pyramidal cells
differences in kinetic profiles between the GAT subtypes (Clausen et al., (Semyanov et al., 2003). Microdialysis and electrophysiological
2006). The therapeutic implication of different kinetics of GAT inhibitors studies on GAT1 knockout mice versus wildtype mice show that
has not been investigated; however, theoretical considerations can be KCl-evoked release of GABA was significantly increased along with a
made. Competitive inhibitors work at low concentrations of extracel- tonic conductance in layer II/III pyramidal cells. Moreover, the decay
lular GABA and loose effect due to competition as the concentration of time of evoked IPSC (eIPSC) was prolonged significantly contrary to
GABA increases. Non-competitive inhibitors, on the other hand, are spontaneous IPSC. Immunohistochemistry confirms that GAT2 or
independent of the extracellular concentration of GABA if the occupancy GAT3 expression was not altered; however, an increase in GAD65/67
of the available transporters is nearly complete. However, if the expression was observed. These results demonstrate that GAT1
occupancy is low as may be the situation in-vivo, the non-competitive regulates both phasic and tonic GABAA receptor mediated inhibition
inhibitor will behave as a competitive inhibitor, i.e. its effect will be in the cerebral cortex of mice (Bragina et al., 2008). In monkey globus
concentration dependent. pallidus the inhibition of either GAT1 or GAT3 increased the ambient
Why have we failed in producing non-GAT1 selective drugs? The GABA concentration resulting in a decreased neuronal firing (Galvan
reason for this could be a consequence of the parent structures used to et al., 2005). Recordings of GABA currents from rat neocortical slices
synthesize new analogues and the fact that the in-vitro assays (P17–22) were evaluated in the presence of either NO711 (GAT1
employed express primarily GAT1, thereby missing lead structures for selective inhibitor) or SNAP-5114 (GAT2/3 selective inhibitor) or a
non-GAT1 inhibitors. combination of both. NO711 decreased the amplitude and increased
the decay time of eIPSC, whereas SNAP-5114 was without effect.
7. GAT inhibitors and the control of neuronal excitability However, co-application synergistically decreased the amplitude and
increased the decay time. Neither NO711 nor SNAP-5114 alone
Neuronal excitability is controlled by two different inhibitory increased the tonic inhibitory current but co-application substantially
mechanisms mediated through synaptic and extrasynaptic GABA increased it. Taken together these results indicate that GAT1 regulates
K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401 399

the phasic inhibitory currents in the synapse whereas GAT1 and extrasynaptically as a result of inhibition of BGT1 which then interacts
GAT2/3 in combination can regulate the extrasynaptic tonic inhibitory with a GABA receptor population that is different from those GABA
currents (Keros & Hablitz, 2005). SNAP-5114 has also been shown to receptors activated by inhibition of GAT1 and GAT3. Therefore it could
increase GABA levels in the thalamus to 247 ± 27% of basal conditions be argued that in order to achieve synergism one should use GAT
but had no effect on GABA levels in the hippocampus, whereas inhibitors that produce a different regional increase in GABA levels
tiagabine was able to increase GABA levels both in the thalamus and and thus activate different GABA receptor populations; e.g., synaptic
hippocampus (Dalby, 2000). Neuronal excitability can be fine tuned and extrasynaptic. On this basis it may be argued that development of
by the use of GAT inhibitors. However, whether phasic or tonic inhibitors of extrasynaptically located GABA transporters such as
inhibitory currents are activated seems to depend on the neuronal BGT1 may represent a novel avenue to obtain antiepileptic drugs with
network being investigated, which undoubtedly is a consequence of a therapeutic potential different from that of tiagabine.
both GABA transporter and receptor density and subtype expression.
9. Therapeutic aspects of GAT inhibitors
8. Anticonvulsant activity of GAT inhibitors
Tight regulation of the synthesis, release, and removal of
The therapeutic potential of GAT inhibition has been confirmed synaptically released GABA is of fundamental importance for the
with the successful development of the GAT1 selective drug tiagabine maintenance of brain function at all levels. In addition to their value in
(Nielsen et al., 1991), albeit tiagabine has limited efficacy in epilepsy controlling seizure activity in patients with epilepsy, it is highly likely
management and exhibits numerous side effects such as agitation, that drugs acting on GABA transporters may represent potential
sedation and psychotic-like episodes in patients predisposed to therapeutic candidates for the treatment of other neurologic and
psychiatric illness (Schousboe & White, 2009). Unfortunately, the psychiatric conditions associated with dysfunction of the GABA
development of selective non-GAT1 inhibitors has been disappoint- system; e.g., anxiety, sleep disorders, chronic pain, post-traumatic
ing. However, the GAT1/BGT1 inhibitor EF1502 and the GAT2/GAT3 stress disorder, migraine, and others. Actually, several clinically
inhibitor SNAP-5114 do allow us to evaluate the therapeutic potential effective drugs used for the treatment of these disorders do in fact
of non-GAT1 inhibitors although the drugs are only semi-selective and act through an interaction with GABA receptors; i.e., the benzodia-
the results must be conservatively interpreted. An interesting zepines for the treatment of anxiety and sleep disorders. In this
question is: to what extent does the individual GATs contribute to context, inhibitors of GABA transporters exert their beneficial effects
the control of neuronal excitability? by making more GABA available at synaptic and extrasynaptic GABA
The electrophysiological data presented above suggests different receptors. Although the development of THIP (Gaboxadol) for the
mechanisms of action of GAT1 inhibitors depending on the circuitry treatment of sleep disorders has been halted, it certainly demon-
being investigated, and it is conceivable that the same fact holds for the strates the therapeutic potential whereby increasing extrasynaptic
other GAT subtypes. In-vivo studies in AGS susceptible Frings mice have GABA levels could play an important role in the control of CNS
attempted to address this question. From an isobologram study function (Krogsgaard-Larsen et al., 2004). In this context, it is
conducted by White et al. (2005) we gained several important pieces becoming increasingly clear that the anatomical distribution, regula-
of information about the combination between GAT1 and BGT1 tion and trafficking of GABA transporters may have important
inhibitors. The two selective GAT1 inhibitors tiagabine and LU-32- implications for the treatment of a number of neurological disorders
176B in combination resulted in an additive effect at three fixed dose and diseases. To this point, a greater understanding of the molecular
mixtures indicating that both drugs indirectly interacted with the same biology, distribution, and factors that regulate their function will be
GAT population. Interestingly, the combination of the GAT1/BGT1 critical for developing a new class of therapeutic agents that target
selective inhibitor EF1502 with either tiagabine or LU-32-176B (GAT1 this important regulator of the CNS function.
selective) produced a synergistic anticonvulsant interaction at all dose
mixtures, strongly indicating that the additional inhibition of BGT1 was a Acknowledgment
requirement for the synergism. The combination between tiagabine and
EF1502 was also reevaluated in a model of partial epilepsy, the corneal The secretarial assistance of Ms. Hanne Danø is cordially
kindling model described by Matagne and Klitgaard (1998) and the acknowledged. The experimental work has been supported by the
synergistic effect was seen at one dose mixture (Madsen et al., 2009). The Lundbeck Foundation (R19-A2199).
same study examined the effect of combining SNAP-5114, the GAT2/3
inhibitor with tiagabine. This resulted in an additive anticonvulsant References
effect at all dose mixtures, whereas SNAP-5114 in combination with
EF1502 produced a synergistic interaction in one dose mixture and an Ahn, J., Mundigl, O., Muth, T. R., Rudnick, G., & Caplan, M. J. (1996). Polarized expression
additive interaction in the two others (Madsen et al., 2009). of GABA transporters in Madin–Darby canine kidney cells and cultured hippocam-
GAT inhibitors time dependently increase the extracellular level of pal neurons. J Biol Chem 271, 6917–6924.
Ali, F. E., Bondinell, W. E., Dandridge, P. A., Frazee, J. S., Garvey, E., Girard, G. R., et al.
GABA which then interacts with different GABA receptor populations (1985). Orally active and potent inhibitors of gamma-aminobutyric acid uptake.
thereby facilitating inhibitory neurotransmission. Thus, GAT inhibi- J Med Chem 28, 653–660.
tors only produce an indirect inhibition of neuronal excitability. It is Andersen, K. E., Begtrup, M., Chorghade, M. S., Lee, E. C., Lau, J., Lundt, B. F., et al. (1994).
The synthesis of novel GABA uptake inhibitors. 2. Synthesis of 5-hydroxytiagabine,
known that tiagabine, EF1502 and SNAP-5114 possess anticonvulsant a human metabolite of the GABA reuptake inhibitor tiagabine. Tetrahedron 50,
activity in several established models of epilepsy (Nielsen et al., 1991; 8699–8710.
Dalby, 2000; White et al., 2005). The isobologram studies suggest that Andersen, K. E., Braestrup, C., Grønwald, F. C., Jørgensen, A. S., Nielsen, E. B., Sonnewald,
U., et al. (1993). The synthesis of novel GABA uptake inhibitors. 1. Elucidation of the
in order to achieve synergism, BGT1 needs to be inhibited along with structure–activity studies leading to the choice of (R)-1-[4,4-bis(3-methyl-2-
either GAT1 or GAT3. A possible explanation for this can be found in thienyl)-3-butenyl]-3-piperidinecarboxylic acid (tiagabine) as an anticonvulsant
the localization of the GATs (see Fig. 2). As described previously, GAT1 drug candidate. J Med Chem 36, 1716–1725.
Andersen, K. E., Lau, J., Lundt, B. F., Petersen, H., Huusfeldt, P. O., Suzdak, P. D., et al.
is localized to the presynaptic neuron and perisynaptic distal
(2001). Synthesis of novel GABA uptake inhibitors. Part 6: preparation and
astrocytic processes (Borden, 1996), GAT3 is located exclusively on evaluation of N-Omega asymmetrically substituted nipecotic acid derivatives.
distal astrocytic processes perisynaptically (Durkin et al., 1995; Bioorg Med Chem 9, 2773–2785.
Minelli et al., 1996), and BGT1 is found exclusively extrasynaptically Andersen, K. E., Sørensen, J. L., Huusfeldt, P. O., Knutsen, L. J., Lau, J., Lundt, B. F., et al.
(1999). Synthesis of novel GABA uptake inhibitors. 4. Bioisosteric transformation
at asymmetric synapses (Zhu & Ong, 2004). It is very likely that the and successive optimization of known GABA uptake inhibitors leading to a series of
synergism obtained using EF1502 arises because GABA levels rise potent anticonvulsant drug candidates. J Med Chem 42, 4281–4291.
400 K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401

Andersen, K. E., Sørensen, J. L., Lau, J., Lundt, B. F., Petersen, H., Huusfeldt, P. O., et al. Inaba, Y., D'Antuono, M., Bertazzoni, G., Biagini, G., & Avoli, M. (2009). Diminished
(2001). Synthesis of novel gamma-aminobutyric acid (GABA) uptake inhibitors. presynaptic GABA(B) receptor function in the neocortex of a genetic model of
5. (1) Preparation and structure–activity studies of tricyclic analogues of known absence epilepsy. Neurosignals 17, 121–131.
GABA uptake inhibitors. J Med Chem 44, 2152–2163. Iversen, L. L., & Kelly, J. S. (1975). Uptake and metabolism of gamma-aminobutyric acid
Awapara, J., Landua, A. J., Fuerst, R., & Seale, B. (1950). Free gamma-aminobutyric acid in by neurones and glial cells. Biochem Pharmacol 24, 933–938.
brain. J Biol Chem 187, 35–39. Iversen, L. L., & Neal, M. J. (1968). The uptake of [3H]GABA by slices of rat cerebral
Belhage, B., Hansen, G. H., & Schousboe, A. (1993). Depolarization by K+ and glutamate cortex. J Neurochem 15, 1141–1149.
activates different neurotransmitter release mechanisms in GABAergic neurons: Keros, S., & Hablitz, J. J. (2005). Subtype-specific GABA transporter antagonists
vesicular versus non-vesicular release of GABA. Neuroscience 54, 1019–1034. synergistically modulate phasic and tonic GABAA conductances in rat neocortex.
Bolvig, T., Larsson, O. M., Pickering, D. S., Nelson, N., Falch, E., Krogsgaard-Larsen, P., et al. J Neurophysiol 94, 2073–2085.
(1999). Action of bicyclic isoxazole GABA analogues on GABA transporters and its Knutsen, L. J., Andersen, K. E., Lau, J., Lundt, B. F., Henry, R. F., Morton, H. E., et al. (1999).
relation to anticonvulsant activity. Eur J Pharmacol 375, 367–374. Synthesis of novel GABA uptake inhibitors. 3. Diaryloxime and diarylvinyl ether
Borden, L. A. (1996). GABA transporter heterogeneity: pharmacology and cellular derivatives of nipecotic acid and guvacine as anticonvulsant agents. J Med Chem 42,
localization. Neurochem Int 29, 335–356. 3447–3462.
Borden, L. A., Dhar, T. G., Smith, K. E., Branchek, T. A., Gluchowski, C., & Weinshank, R. L. (1994). Kragler, A., Hofner, G., & Wanner, K. T. (2005). Novel parent structures for inhibitors of
Cloning of the human homologue of the GABA transporter GAT-3 and identification of a the murine GABA transporters mGAT3 and mGAT4. Eur J Pharmacol 519, 43–47.
novel inhibitor with selectivity for this site. Receptors Channels 2, 207–213. Krnjevic, K. (2004). How does a little acronym become a big transmitter? Biochem
Borden, L. A., Smith, K. E., Gustafson, E. L., Branchek, T. A., & Weinshank, R. L. (1995). Pharmacol 68, 1549–1555.
Cloning and expression of a betaine/GABA transporter from human brain. J Neurochem Krogsgaard-Larsen, P., Frolund, B., & Frydenvang, K. (2000). GABA uptake inhibitors. Design,
64, 977–984. molecular pharmacology and therapeutic aspects. Curr Pharm Des 6, 1193–1209.
Borden, L. A., Smith, K. E., Hartig, P. R., Branchek, T. A., & Weinshank, R. L. (1992). Krogsgaard-Larsen, P., Frolund, B., Liljefors, T., & Ebert, B. (2004). GABA(A) agonists and
Molecular heterogeneity of the gamma-aminobutyric acid (GABA) transport partial agonists: THIP (Gaboxadol) as a non-opioid analgesic and a novel type of
system. Cloning of two novel high affinity GABA transporters from rat brain. J Biol hypnotic. Biochem Pharmacol 68, 1573–1580.
Chem 267, 21098–21104. Krogsgaard-Larsen, P., & Johnston, G. A. (1975). Inhibition of GABA uptake in rat brain
Braestrup, C., Nielsen, E. B., Sonnewald, U., Knutsen, L. J., Andersen, K. E., Jansen, J. A., et al. slices by nipecotic acid, various isoxazoles and related compounds. J Neurochem 25,
(1990). (R)-N-[4, 4-bis(3-methyl-2-thienyl)but-3-en-1-yl]nipecotic acid binds with 797–802.
high affinity to the brain gamma-aminobutyric acid uptake carrier. J Neurochem 54, Kwan, P., Sills, G. J., & Brodie, M. J. (2001). The mechanisms of action of commonly used
639–647. antiepileptic drugs. Pharmacol Ther 90, 21–34.
Bragina, L., Marchionni, I., Omrani, A., Cozzi, A., Pellegrini-Giampietro, D. E., Cherubini, Larsson, O. M., Thorbek, P., Krogsgaard-Larsen, P., & Schousboe, A. (1981). Effect of
E., et al. (2008). GAT-1 regulates both tonic and phasic GABA(A) receptor-mediated homo-∃-proline and other heterocyclic GABA analogues on GABA uptake in neurons
inhibition in the cerebral cortex. J Neurochem 105, 1781–1793. and astroglial cells and on GABA receptor binding. J Neurochem 37, 1509–1516.
Brown, N., Kerby, J., Bonnert, T. P., Whiting, P. J., & Wafford, K. A. (2002). Larsson, O. M., Johnston, G. A. R., & Schousboe, A. (1983). Differences in uptake kinetics
Pharmacological characterization of a novel cell line expressing human alpha(4) of cis-3-aminocyclohexane carboxylic acid into neurons and astrocytes in primary
beta(3)delta GABA(A) receptors. Br J Pharmacol 136, 965–974. cultures. Brain Res 260, 279–285.
Burnham, C. E., Buerk, B., Schmidt, C., & Bucuvalas, J. C. (1996). A liver-specific isoform Larsson, O. M., Griffiths, R., Allen, I. C., & Schousboe, A. (1986). Mutual inhibition kinetic
of the betaine/GABA transporter in the rat: cDNA sequence and organ distribution. analysis of (γ-aminobutyric acid, taurine, taurine and β-alanine high affinity
Biochim Biophys Acta 1284, 4–8. transport into neurons and astrocytes: evidence for similarity between the taurine
Christiansen, B., Meinild, A. K., Jensen, A. A., & Brauner-Osborne, H. (2007). Cloning and and β-alanine carriers in both cell types. J Neurochem 47, 426–432.
characterization of a functional human gamma-aminobutyric acid (GABA) Larsson, O. M., Falch, E., Krogsgaard-Larsen, P., & Schousboe, A. (1988). Kinetic
transporter, human GAT-2. J Biol Chem 282, 19331–19341. characterization of inhibition of gamma-aminobutyric acid uptake into cultured
Clausen, R. P., Madsen, K., Larsson, O. M., Frolund, B., Krogsgaard-Larsen, P., & Schousboe, A. neurons and astrocytes by 4, 4-diphenyl-3-butenyl derivatives of nipecotic acid
(2006). Structure–activity relationship and pharmacology of gamma-aminobutyric and guvacine. J Neurochem 50, 818–823.
acid (GABA) transport inhibitors. Adv Pharmacol 54, 265–284. Liu, Q. R., Lopez-Corcuera, B., Mandiyan, S., Nelson, H., & Nelson, N. (1993). Molecular
Clausen, R. P., Moltzen, E. K., Perregaard, J., Lenz, S. M., Sanchez, C., Falch, E., et al. (2005). characterization of four pharmacologically distinct gamma-aminobutyric acid
Selective inhibitors of GABA uptake: synthesis and molecular pharmacology of 4-N- transporters in mouse brain [corrected]. J Biol Chem 268, 2106–2112.
methylamino-4, 5, 6, 7-tetrahydrobenzo[d]isoxazol-3-ol analogues. Bioorg Med Liu, Q. R., Mandiyan, S., Nelson, H., & Nelson, N. (1992). A family of genes encoding
Chem 13, 895–908. neurotransmitter transporters. Proc Natl Acad Sci U S A 89, 6639–6643.
Conti, F., Melone, M., De Biasi, S., Minelli, A., Brecha, N. C., & Ducati, A. (1998). Neuronal Madsen, K. K., Clausen, R. P., Larsson, O. M., Krogsgaard-Larsen, P., Schousboe, A., &
and glial localization of GAT-1, a high-affinity gamma-aminobutyric acid plasma White, H. S. (2009). Synaptic and extrasynaptic GABA transporters as targets for
membrane transporter, in human cerebral cortex: with a note on its distribution in anti-epileptic drugs. J Neurochem 109(Suppl 1), 139–144.
monkey cortex. J Comp Neurol 396, 51–63. Martin, D. L., & Rimvall, K. (1993). Regulation of gamma-aminobutyric acid synthesis in
Conti, F., Zuccarello, L. V., Barbaresi, P., Minelli, A., Brecha, N. C., & Melone, M. (1999). the brain. J Neurochem 60, 395–407.
Neuronal, glial, and epithelial localization of gamma-aminobutyric acid transporter Matagne, A., & Klitgaard, H. (1998). Validation of corneally kindled mice: a sensitive
2, a high-affinity gamma-aminobutyric acid plasma membrane transporter, in the screening model for partial epilepsy in man. Epilepsy Res 31, 59–71.
cerebral cortex and neighboring structures. J Comp Neurol 409, 482–494. Melone, M., Barbaresi, P., Fattorini, G., & Conti, F. (2005). Neuronal localization of the
Cremers, T., & Ebert, B. (2007). Plasma and CNS concentrations of gaboxadol in rats GABA transporter GAT-3 in human cerebral cortex: a procedural artifact? J Chem
following subcutaneous administration. Eur J Pharmacol 562, 47–52. Neuroanat 30, 45–54.
Dalby, N. O. (2000). GABA-level increasing and anticonvulsant effects of three different Minelli, A., DeBiasi, S., Brecha, N. C., Zuccarello, L. V., & Conti, F. (1996). GAT-3, a high-
GABA uptake inhibitors. Neuropharmacology 39, 2399–2407. affinity GABA plasma membrane transporter, is localized to astrocytic processes,
Dalby, N. O., & Mody, I. (2001). The process of epileptogenesis: a pathophysiological and it is not confined to the vicinity of GABAergic synapses in the cerebral cortex.
approach. Curr Opin Neurol 14, 187–192. J Neurosci 16, 6255–6264.
Durkin, M. M., Smith, K. E., Borden, L. A., Weinshank, R. L., Branchek, T. A., & Gustafson, E. L. Nelson, H., Mandiyan, S., & Nelson, N. (1990). Cloning of the human brain GABA
(1995). Localization of messenger RNAs encoding three GABA transporters in rat transporter. FEBS Lett 269, 181–184.
brain: an in situ hybridization study. Brain Res Mol Brain Res 33, 7–21. Nielsen, E. B., Suzdak, P. D., Andersen, K. E., Knutsen, L. J., Sonnewald, U., & Braestrup, C.
Elliott, K. A., & van Gelder, N. M. (1958). Occlusion and metabolism of gamma- (1991). Characterization of tiagabine (NO-328), a new potent and selective GABA
aminobutyric acid by brain tissue. J Neurochem 3, 28–40. uptake inhibitor. Eur J Pharmacol 196, 257–266.
Falch, E., Perregaard, J., Frølund, B., SLkilde, B., Buur, A., Hansen, L. M., et al. (1999). Nusser, Z., & Mody, I. (2002). Selective modulation of tonic and phasic inhibitions in
Selective inhibitors of glial GABA uptake: synthesis, absolute stereochemistry, and dentate gyrus granule cells. J Neurophysiol 87, 2624–2628.
pharmacology of the enantiomers of 3-hydroxy-4-amino-4, 5, 6, 7-tetrahydro-1, 2- Pietrini, G., Suh, Y. J., Edelmann, L., Rudnick, G., & Caplan, M. J. (1994). The axonal
benzisoxazole (exo-THPO) and analogues. J Med Chem 42, 5402–5414. gamma-aminobutyric acid transporter GAT-1 is sorted to the apical membranes of
Fulep, G. H., Hoesl, C. E., Hofner, G., & Wanner, K. T. (2006). New highly potent GABA polarized epithelial cells. J Biol Chem 269, 4668–4674.
uptake inhibitors selective for GAT-1 and GAT-3 derived from (R)- and (S)-proline Radian, R., Bendahan, A., & Kanner, B. I. (1986). Purification and identification of the
and homologous pyrrolidine-2-alkanoic acids. Eur J Med Chem 41, 809–824. functional sodium- and chloride-coupled gamma-aminobutyric acid transport
Gale, K., & Iadarola, M. J. (1980). Seizure protection and increased nerve-terminal glycoprotein from rat brain. J Biol Chem 261, 15437–15441.
GABA: delayed effects of GABA transaminase inhibition. Science 208, 288–291. Radian, R., Ottersen, O. P., Storm-Mathisen, J., Castel, M., & Kanner, B. I. (1990).
Galvan, A., Villalba, R. M., West, S. M., Maidment, N. T., Ackerson, L. C., Smith, Y., et al. Immunocytochemical localization of the GABA transporter in rat brain. J Neurosci
(2005). GABAergic modulation of the activity of globus pallidus neurons in 10, 1319–1330.
primates: in vivo analysis of the functions of GABA receptors and GABA Roberts, E., & Frankel, S. (1950). γ-Aminobutyric acid in brain: its formation from
transporters. J Neurophysiol 94, 990–1000. glutamic acid. J Biol Chem 187, 55–63.
Guastella, J., Nelson, N., Nelson, H., Czyzyk, L., Keynan, S., Miedel, M. C., et al. (1990). Roberts, E., & Kuriyama, K. (1968). Biochemical-physiological correlations in studies of
Cloning and expression of a rat brain GABA transporter. Science 249, 1303–1306. the gamma-aminobutyric acid system. Brain Res 8, 1–35.
Hertz, L., & Schousboe, A. (1987). Primary cultures of GABAergic and glutamatergic Sarup, A., Larsson, O. M., Bolvig, T., Frølund, B., Krogsgaard-Larsen, P., & Schousboe, A.
neurons as model systems to study neurotransmitter functions. I Differentiated (2003). Effects of 3-hydroxy-4-amino-4, 5, 6, 7-tetrahydro-1, 2-benzisoxazol (exo-
cells. In A. Vernadakis, A. Privat, J. M. Lauder, P. S. Timiras, & E. Giacobini (Eds.), THPO) and its N-substituted analogs on GABA transport in cultured neurons and
Model systems of development and aging of the nervous system (pp. 19–32). Boston: astrocytes and by the four cloned mouse GABA transporters. Neurochem Int 43,
Martinus Nijhoff Publishing. 445–451.
K.K. Madsen et al. / Pharmacology & Therapeutics 125 (2010) 394–401 401

Saxena, N. C., & MacDonald, R. L. (1994). Assembly of GABAA receptor subunits: role of Udenfriend, S. (1950). Identification of gamma-aminobutyric acid in brain by the
the delta subunit. J Neurosci 14, 7077–7086. isotope derivative method. J Biol Chem 187, 65–69.
Schousboe, A. (1979). Effects of GABA analogues on the high-affinity uptake of GABA in Wafford, K. A., & Ebert, B. (2006). Gaboxadol—a new awakening in sleep. Curr Opin
astrocytes in primary cultures. In P. Mandel, & F. V. De Feudis (Eds.), GABA- Pharmacol 6, 30–36.
Biochemistry and CNS Function (pp. 219–237). New York: Plenum Publishing Corp. Watanabe, M., Maemura, K., Kanbara, K., Tamayama, T., & Hayasaki, H. (2002). GABA
Schousboe, A. (1990). Neurochemical alterations associated with epilepsy or seizure and GABA receptors in the central nervous system and other organs. In K. W. Jeon
activity. In M. Dam, & L. Gram (Eds.), Comprehensive epileptology (pp. 1–16). New (Ed.), A survey of cell biology (pp. 1–47). San Diego: Academic Press, Inc.
York: Raven Press, Ltd. White, H. S. (1999). Comparative anticonvulsant and mechanistic profile of the
Schousboe, A. (2000). Pharmacological and functional characterization of astrocytic established and newer antiepileptic drugs. Epilepsia 40(Suppl 5), S2–S10.
GABA transport: a short review. Neurochem Res 25, 1241–1244. White, H. S., Sarup, A., Bolvig, T., Kristensen, A. S., Petersen, G., Nelson, N., et al. (2002).
Schousboe, A., & White, H. S. (2009). In P. Schwartzkroin (Ed.), Modulation of excitability Correlation between anticonvulsant activity and inhibitory action on glial gamma-
via glutamate and GABA transportersEncyclopedia of basic epilepsy research Vol. 1. aminobutyric acid uptake of the highly selective mouse gamma-aminobutyric acid
(pp. 397–401) Oxford, UK: Elsevier Ltd. transporter 1 inhibitor 3-hydroxy-4-amino-4,5,6,7-tetrahydro-1,2-benzisoxazole
Schousboe, A., Larsson, O. M., Hertz, L., & Krogsgaard-Larsen, P. (1981). Heterocyclic and its N-alkylated analogs. J Pharmacol Exp Ther 302, 636–644.
GABA analogues as new selective inhibitors of astroglial GABA transport. Drug Dev White, H. S., Watson, W. P., Hansen, S. L., Slough, S., Perregaard, J., Sarup, A., et al.
Res 1, 115–127. (2005). First demonstration of a functional role for central nervous system betaine/
Schousboe, A., Saito, K., & Wu, J. Y. (1980). Characterization and cellular and subcellular {gamma}-aminobutyric acid transporter (mGAT2) based on synergistic anticon-
localization of GABA–transaminase. Brain Res Bull 5, 71–76. vulsant action among inhibitors of mGAT1 and mGAT2. J Pharmacol Exp Ther 312,
Schousboe, A., Thorbek, P., Hertz, L., & Krogsgaard-Larsen, P. (1979). Effects of GABA 866–874.
analogues of restricted conformation on GABA transport in astrocytes and brain Yamauchi, A., Uchida, S., Kwon, H. M., Preston, A. S., Robey, R. B., Garcia-Perez, A., et al.
cortex slices and on GABA receptor binding. J Neurochem 33, 181–189. (1992). Cloning of a Na(+)- and Cl(−)-dependent betaine transporter that is
Semyanov, A., Walker, M. C., & Kullmann, D. M. (2003). GABA uptake regulates cortical regulated by hypertonicity. J Biol Chem 267, 649–652.
excitability via cell type-specific tonic inhibition. Nat Neurosci 6, 484–490. Yunger, L. M., Fowler, P. J., Zarevics, P., & Setler, P. E. (1984). Novel inhibitors of gamma-
Suzdak, P. D., Frederiksen, K., Andersen, K. E., Sørensen, P. O., Knutsen, L. J., & Nielsen, E. B. aminobutyric acid (GABA) uptake: anticonvulsant actions in rats and mice. J Pharmacol
(1992). NNC-711, a novel potent and selective gamma-aminobutyric acid uptake Exp Ther 228, 109–115.
inhibitor: pharmacological characterization. Eur J Pharmacol 224, 189–198. Zhu, X. M., & Ong, W. Y. (2004). A light and electron microscopic study of betaine/GABA
Takayama, C., & Inoue, Y. (2005). Developmental expression of GABA transporter-1 and transporter distribution in the monkey cerebral neocortex and hippocampus.
3 during formation of the GABAergic synapses in the mouse cerebellar cortex. Brain J Neurocytol 33, 233–240.
Res Dev Brain Res 158, 41–49.
Thomsen, C., Sørensen, P. O., & Egebjerg, J. (1997). 1-(3-(9H-carbazol-9-yl)-1-propyl)-
4-(2-methoxyphenyl)-4-piperidinol, a novel subtype selective inhibitor of the
mouse type II GABA-transporter. Br J Pharmacol 120, 983–985.

You might also like