Sunteți pe pagina 1din 14

The mechanisms of action of commonly used antiepileptic drugs

Patrick Kwan, Graeme J. Sills*, Martin J. Brodie


Epilepsy Unit, University Department of Medicine and Therapeutics, Western Infirmary, Glasgow G11 6NT, Scotland, UK
Abstract
In the past decade, nine new drugs have been licensed for the treatment of epilepsy. With limited clinical experience of these agents, the
mechanisms of action of antiepileptic drugs may be an important criterion in the selection of the most suitable treatment regimens for
individual patients. At the cellular level, three basic mechanisms are recognised: modulation of voltage-dependent ion channels, enhancement
of inhibitory neurotransmission, and attenuation of excitatory transmission. In this review, we will attempt to introduce the concepts of ion
channel and neurotransmitter modulation and, thereafter, group currently used antiepileptic drugs according to their principal mechanisms of
action. D 2001 Elsevier Science Inc. All rights reserved.
Keywords: Antiepileptic drugs; Mechanisms of action; Epilepsy
Abbreviations: AED, antiepileptic drug; AMPA, a-amino-3-hydroxy-5-methyl-isoxazole-4-propionic acid; BGT, betaine g-aminobutyric acid transporter;
BZD, benzodiazepine; CBZ, carbamazepine; ESM, ethosuximide; FBM, felbamate; GABA, g-aminobutyric acid; GABA-T, g-aminobutyric acid transaminase;
GAD, glutamic acid decarboxylase; GAT, g-aminobutyric acid transporter; GBP, gabapentin; LEV, levetiracetam; LTG, lamotrigine; NMDA, N-methyl-D-
aspartate; OXC, oxcarbazepine; PB, phenobarbital; PHT, phenytoin; PRM, primidone; TGB, tiagabine; TPM, topiramate; VGB, vigabatrin; VPA, sodium
valproate; ZNS, zonisamide.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.1. Classification of mechanisms of action . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2. Targets for antiepileptic drug action. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1. Ion channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.1. Na
+
channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.2. Ca
2+
channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.3. K
+
channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2. g-Aminobutyric acid-mediated inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3. Glutamate-mediated excitation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3. Modulation of ion channels by antiepileptic drugs . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1. Phenytoin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2. Carbamazepine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3. Lamotrigine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.4. Oxcarbazepine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5. Ethosuximide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.6. Zonisamide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4. Potentiation of g-aminobutyric acid by antiepileptic drugs . . . . . . . . . . . . . . . . . . . . . 26
4.1. Phenobarbital . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2. Benzodiazepines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3. Vigabatrin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4. Tiagabine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
0163-7258/01/$ see front matter D 2001 Elsevier Science Inc. All rights reserved.
PII: S0163- 7258( 01) 00122- X
* Corresponding author. Tel.: +44-141-211-2770; fax: +44-141-334-9329.
E-mail address: g.j.sills@clinmed.gla.ac.uk (G.J. Sills).
Pharmacology & Therapeutics 90 (2001) 2134
5. Antiepileptic drugs with multiple mechanisms of action . . . . . . . . . . . . . . . . . . . . . . 28
5.1. Sodium valproate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2. Gabapentin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3. Felbamate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.4. Topiramate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
6. Antiepileptic drugs with unknown mechanisms of action. . . . . . . . . . . . . . . . . . . . . . 30
6.1. Levetiracetam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1. Introduction
The serendipitous discovery of phenobarbital (PB) in
1912 marked the beginning of the modern pharmacotherapy
of epilepsy (Brodie & Dichter, 1996). In the ensuing 70
years, phenytoin (PHT), carbamazepine (CBZ), ethosuxi-
mide (ESM), sodium valproate (VPA), and a range of
benzodiazepines (BZDs) became available and can be
regarded as established antiepileptic drugs (AEDs; Bro-
die & Dichter, 1996). In the past decade, nine new agents
have been licensed as add-on treatment for difficult-to-
control epilepsy (Table 1) and still more are under evalua-
tion (Dichter & Brodie, 1996).
The undoubtedly beneficial expansion of the pharmaco-
logical armamentarium for the treatment of epilepsy does,
however, complicate selection of the most suitable AED (or
combination of AEDs) for individual patients. With limited
clinical experience of the new agents, the mechanisms of
action of individual AEDs may be an important criterion in
this decision-making process (Brodie, 1999; Brodie &
Kwan, 2000).
1.1. Classification of mechanisms of action
Practical approaches to drug treatment have resorted
primarily to symptom control, i.e., suppression of seiz-
ures. Although the mechanisms of action of the currently
marketed AEDs are still not completely understood, they
ultimately involve alteration of the balance between neu-
ronal excitation and inhibition (White, 1999). At the
cellular level, three basic mechanisms are recognised:
modulation of voltage-dependent ion channels (Na
+
,
Ca
2 +
, K
+
), enhancement of g-aminobutyric acid
(GABA)-mediated inhibitory neurotransmission, and
attenuation of excitatory (particularly glutamate-mediated)
transmission (Meldrum, 1996).
In this review, we will attempt to introduce the concepts
of ion channel and neurotransmitter modulation and, there-
after, group the currently used AEDs according to their
principal mechanisms of action (Table 1). It is, however,
increasingly recognised that many AEDs possess multiple
primary mechanisms. In addition, many less well-charac-
terised mechanisms may also contribute to the anticonvul-
sant effect of any given compound. Finally, to complicate
the issue further, the primary mode of action of some AEDs
remains to be discovered.
2. Targets for antiepileptic drug action
2.1. Ion channels
2.1.1. Na
+
channels
In the nervous system, voltage-gated ion channels control
the flow of cations across surface and internal cell mem-
branes (Barchi, 1998). Of these, the Na
+
channel is argu-
ably of principal importance. Voltage-dependent Na
+
channels are responsible for the upstroke of the neuronal
action potential, and ultimately control the intrinsic excit-
ability of the nervous system (Porter & Rogawski, 1992).
The neuronal Na
+
channel has a multi-subunit structure that
forms a Na
+
-selective, voltage-gated pore through the
plasma membrane. The protein structure undergoes confor-
mational alterations in response to changes in membrane
potential, regulating conductance through the intrinsic pore
(Ragsdale & Avoli, 1998).
Table 1
Proposed mechanisms of antiepileptic drug action
# Na
+
channels
# Ca
2 +
channels
" K
+
channels
" Inhibitory
transmission
# Excitatory
transmission
Established AEDs
PHT ++ +
CBZ ++ +
ESM + ++
PB + + ++ +
BZDs + ++
VPA + + + + +
New AEDs
LTG ++ + +
OXC ++ + + +
ZNS ++ ++
VGB + ++
TGB + ++
GBP + + + +
FBM ++ ++ + + + +
TPM ++ ++ + + + +
LEV + + +
+++, primary action; ++, probable action; +, possible action.
Data from Upton (1994), Schachter (1995), Macdonald & Kelly (1995),
Meldrum (1996), Coulter (1997), and White (1999).
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 22
The main structural component of the neuronal Na
+
channel is the a-subunit, which forms the ion-conducting
pore and confers voltage dependency (Catterall, 1992). In
the mammalian brain, the a-subunit associates with two
auxiliary subunits designated b
1
and b
2
. The b-subunits are
not required for basic Na
+
channel activity, but they
modulate the expression and function of individual channels
(Ragsdale & Avoli, 1998).
At normal membrane potentials, most Na
+
channels
exist in a closed, resting state. Upon depolarisation, the
channel activates, facilitating ion flux. Thereafter, the Na
+
channel enters an inactivated state, from which it is not
readily re-activated. Repolarisation of the neuronal mem-
brane rapidly converts the channel back to a resting state,
from which it can respond to subsequent depolarisations
(Catterall, 1992; Ragsdale & Avoli, 1998). Neuronal Na
+
channels can cycle through these functional states within a
few milliseconds. This characteristic is essential for sustain-
ing the rapid bursts of action potentials necessary for some
normal brain functions, and is implicated in the production
of epileptic discharges. The neuronal Na
+
channel repre-
sents one of the most important targets for AED action
(Upton, 1994; Macdonald & Kelly, 1995; Meldrum, 1996;
White, 1999).
2.1.2. Ca
2+
channels
Voltage-dependent Ca
2 +
channels share key structural
elements and sequence homology with their Na
+
channel
counterparts (Barchi, 1998). The a
1
-subunit of the Ca
2 +
channel is the homologue of the a-subunit of the Na
+
channel. It forms the Ca
2 +
-sensitive channel pore and
confers voltage dependency (Catterall, 1995). In the mam-
malian brain, the a
1
-subunit heterogeneously associates
with other subunits designated b, g, and d.
Voltage-sensitive Ca
2 +
channels can be broadly classi-
fied into low or high threshold, according to the membrane
potential at which they are activated (Hofmann et al., 1994).
The low-threshold T-type Ca
2 +
channel is expressed pre-
dominantly in thalamocortical relay neurones, where it is
believed to be instrumental in the generation of the rhythmic
3-Hz spike-and-wave discharge that is characteristic of
generalised absence seizures (Coulter et al., 1989a). High-
threshold Ca
2 +
channels are subclassified by their pharma-
cological properties into L-, N-, P-, Q-, and R-types (Hof-
mann et al., 1994; Catterall, 1995; Dolphin, 1995). These
channels are distributed throughout the nervous system on
dendrites, cell bodies, and nerve terminals. The N-, P-, and
Q-type channels, in particular, have been implicated in the
control of neurotransmitter release at the synapse (Stefani
et al., 1997).
Interest in Ca
2 +
channels has heightened in recent years,
following the identification of subunit-specific genetic
mutations that can alter channel structure and/or function
and that have been implicated in several human neurological
diseases (Ophoff et al., 1998). Several AEDs have been
reported to block voltage-sensitive Ca
2 +
channels in a
subtype-specific manner, an effect that may contribute to
their antiepileptic actions (Stefani et al., 1997).
2.1.3. K
+
channels
Neuronal K
+
channels are large protein complexes that
form tetrameric structures, the monomers of which are
structurally and genetically related to the a- and a
1
-subunits
of the Na
+
and Ca
2 +
channel, respectively (Barchi, 1998).
The association of four subunits (monomers) in the neuronal
membrane is required for the formation of a K
+
-sensitive
pore and, therefore, channel function (Pongs, 1999). More
than 40 distinct K
+
channel subunits have been identified,
together with several auxiliary subunits. Given heterologous
arrangement, it is possible that countless populations of K
+
channels, with individual functions and distributions, are
expressed in the mammalian brain (Pongs, 1999).
At the neuronal level, K
+
channels are intimately
involved in excitability. They are responsible for the action
potential downstroke or, more specifically, repolarisation of
the plasma membrane in the aftermath of Na
+
channel
activation (Pongs, 1999). Direct activation of voltage-
dependent K
+
channels hyperpolarises the neuronal mem-
brane and limits action potential firing (Porter & Rogaw-
ski, 1992). Accordingly, K
+
channel activators have
anticonvulsant effects in some experimental seizure models
(Gandolfo et al., 1989; Rostock et al., 1996), whereas K
+
channel blockers precipitate seizures (Yamaguchi &
Rogawski, 1992).
Potentiation of voltage-sensitive K
+
channel currents
may prove to be an important target for future AED
development. The novel antiepileptic agent retigabine, cur-
rently undergoing Phase II clinical trial, is believed to exert
its effects, at least in part, by activation of the KCNQ2/
KCNQ3 K
+
channels (Rundfeldt & Netzer, 2000). Muta-
tions in the KCNQ2/KCNQ3 channels have been reported
in benign neonatal familial convulsions, a generalised epi-
lepsy syndrome (Rogawski, 2000).
2.2. g-Aminobutyric acid-mediated inhibition
GABA is the predominant inhibitory neurotransmitter in
the mammalian CNS, where it is released at up to 40% of all
synapses (Olsen & Avoli, 1997). Impairment of GABA
function is widely recognised to provoke seizures, whereas
facilitation has an anticonvulsant effect (Loscher, 1999).
GABA is synthesised from glutamate, exclusively in
GABAergic neurones, by the action of the enzyme glutamic
acid decarboxylase (GAD; Loscher, 1999). Upon synaptic
release, GABA acts on its three specific receptors, GABA
A
,
GABA
B
, and the newly characterised GABA
C
. GABA
receptors are distinguished by their pharmacology and
function (Johnston, 1996). The GABA
A
receptor belongs
to the ligand-gated ion channel superfamily, and responds to
GABA binding by increasing Cl

conductance, resulting in
neuronal hyperpolarisation (Rabow et al., 1995). GABA
B
receptors are G-protein-linked, activation of which leads to
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 23
an increase in K
+
conductance (Olsen & Avoli, 1997). It
has recently been proposed that GABA
A
and GABA
B
receptors may have evolved from the GABA
C
receptor,
which is comparatively simpler in structure and pharmacol-
ogy (Johnston, 1996).
Of the three receptor subtypes, the GABA
A
receptor,
with its pentameric subunit array and central Cl

ion pore,
is perhaps the best understood (Rabow et al., 1995).
GABA
A
receptors are composed of various combinations
of 5 subunits, a
(16)
, b
(13)
, g
(13)
, d, and e. Receptor
physiology, pharmacology, and distribution differs, depend-
ing on subunit composition (Mody, 1998). Theoretically,
tens of thousands of different GABA
A
receptors could
exist, but naturally only 10 or fewer combinations of
subtypes are encountered.
Following receptor activation, GABA is removed from
the synaptic cleft into localised nerve terminals and glial
cells, by specific membrane-bound transport molecules.
Currently, four active transport systems, GABA transporter
(GAT)-1, GAT-2, GAT-3, and betaine GAT (BGT)-1, have
been described (Borden et al., 1992). GABA has a variable
affinity for these transporters, and only GAT-1, predomi-
nantly located in the cerebral cortex and hippocampus, has
GABA as its principal substrate (Guastella et al., 1990).
After removal from the synapse, GABA is either recycled to
the readily releasable neurotransmitter pool (GABAergic
nerve terminals only) or metabolised (neurones and glial
cells) to the inactive molecule succinic acid semialdehyde
by the action of the mitochondrial enzyme GABA-trans-
aminase (GABA-T; Meldrum, 1995).
Several AEDs exert their effects, at least in part, by
actions on the GABAergic system. Increased GABA syn-
thesis, increased release, allosteric receptor facilitation, and
reduced inactivation have all been implicated in the mech-
anisms of action of commonly used agents (Sills et al.,
1999). The GABA system also represents the most success-
ful target for the rational design of novel antiepileptic
compounds (Loscher, 1998).
2.3. Glutamate-mediated excitation
Glutamate is the principal excitatory neurotransmitter in
the mammalian brain (Meldrum, 2000). Focal injection of
glutamate induces seizures in animals, and over-activation
of glutamatergic transmission or abnormal glutamate recep-
tor properties are observed in certain experimental seizure
models and human epilepsy syndromes (Meldrum, 1995).
Inhibition of the neuronal release of glutamate and blockade
of its receptors have received considerable attention in the
search for novel AEDs (Meldrum, 2000).
Glutamate is synthesised from glutamine by the action of
the enzyme glutaminase in glutamatergic neurones (Daikhin
& Yudkoff, 2000). Following synaptic release, glutamate
exerts its pharmacological effects on several receptors,
classified into ionotropic and metabotropic families. Gluta-
mate is removed from the synaptic cleft into nerve terminals
and glial cells by the action of several specific transporters
(Meldrum et al., 1999). Glial glutamate uptake is of princi-
pal importance. Glial cells convert glutamate into glutamine
by the action of the enzyme glutamine synthetase. Gluta-
mine is subsequently transferred to glutamatergic neurones,
completing the cycle (Daikhin & Yudkoff, 2000).
Like GABA receptors, ionotropic glutamate receptors are
comprised of various combinations of subunits forming
tetrameric and pentameric arrays. They are classified into
three specific subtypes, a-amino-3-hydroxy-5-methyl-iso-
xazole-4-propionic acid (AMPA), kainate, and N-methyl-D-
aspartate (NMDA), which form ligand-gated ion channels,
permeable to Na
+
and, depending on subtype and subunit
composition, Ca
2 +
ions (Trist, 2000). The NMDA receptor
is further distinguished by having glycine as a co-agonist.
The AMPA and kainate subtypes of the glutamate receptor
are implicated in fast excitatory neurotransmission, whereas
the NMDA receptor, quiescent at resting membrane poten-
tial, is recruited during periods of prolonged depolarisation
(Meldrum, 2000). The metabotropic family of glutamate
receptors, also classified into three distinct subtypes (Groups
I, II, and III), are G-protein linked and predominantly
presynaptic, possibly controlling neurotransmitter release
(Meldrum, 2000).
Although none of the commonly used AEDs exert their
pharmacological effects solely by an action on the glutamate
system, blockade of ionotropic glutamate receptors is
believed to contribute to the antiepileptic activity of several
compounds (Upton, 1994; Macdonald & Kelly, 1995; Mel-
drum, 1996; White, 1999). In addition, several AEDs have
been reported to reduce glutamate release, although this
effect may be more indicative of their actions on neuronal
Ca
2 +
channels than a direct effect on the glutamate system
(Stefani et al., 1997).
3. Modulation of ion channels by antiepileptic drugs
3.1. Phenytoin
PHT was discovered following a search to identify a
nonsedative analogue of PB (Fig. 1; Merritt & Putnam,
1938). It has become a first-line treatment for partial and
generalised tonic-clonic seizures (Brodie & Dichter, 1996).
PHT is believed to exert its anticonvulsant effect pri-
marily by an action on voltage-dependent Na
+
channels
(Tunnicliff, 1996), binding to the fast inactivated state of
the channel and reducing the frequency of sustained repet-
itive firing of action potentials, without affecting their
amplitude or duration (McLean & Macdonald, 1983).
PHT inhibits high-frequency repetitive firing in a voltage-
dependent manner, with limitation of firing increased after
depolarisation and removed by hyperpolarisation (Schwartz
& Grigat, 1989). Na
+
channel blockade with PHT is also
frequency-dependent, being more effective at higher fre-
quencies of neuronal stimulation. An effect on persistent
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 24
Na
+
currents has also been suggested (Lampl et al., 1998).
The precise binding site for PHT on the Na
+
channel is
unclear (Ragsdale et al., 1996), and there may be a
preferential effect on different subtypes of Na
+
channels
(Song et al., 1996).
PHT has also been reported to block high voltage-
activated Ca
2 +
channels (Schumacher et al., 1998), to
attenuate post-ictal glutamate release (Rowley et al., 1995)
and, paradoxically, to reduce K
+
currents (Nobile &
Vercellino, 1997). There is further unsubstantiated evi-
dence to suggest that PHT potentiates the action of GABA
at specific molecular subtypes of the GABA
A
receptor
(Granger et al., 1995).
3.2. Carbamazepine
CBZ is chemically related to the tricyclic antidepressants
(Fig. 1). First introduced in 1963, it is widely used in the
treatment of partial and generalised tonic-clonic seizures
(Brodie & French, 2000). CBZ has been reported to stabilise
the inactive form of the Na
+
channel in a voltage-,
frequency-, and time-dependent fashion (Courtney & Etter,
1983). Although similar to PHT in this respect, subtle
differences in the mechanisms of action of the two drugs
may exist. Accordingly, CBZ has a greater binding rate
constant, but lower affinity, for the inactivated Na
+
channel
(Kuo et al., 1997), whereas PHT produces a more pro-
nounced slowing of recovery from the fast inactivated state
(Schwartz & Grigat, 1989).
Inhibition of glutamatergic neurotransmission has also
been implicated in the mechanism of CBZ action. Recent
evidence suggests that it inhibits the rise in intracellular free
Ca
2 +
induced by NMDA and glycine in rat cerebellar
granule cells (Hough et al., 1996) and blocks veratrine-
induced release of endogenous glutamate (Waldmeier et al.,
1995). Unlike PHT, there is no evidence that CBZ directly
interacts with Ca
2 +
channels or potentiates the actions of
GABA. However, effects on the serotonin (Dailey et al.,
1997a, 1997b) and adenosine (Marangos et al., 1983) sys-
tems have been reported. Whether these additional actions
contribute to the anticonvulsant effects of the drug is unclear.
3.3. Lamotrigine
Lamotrigine (LTG; Fig. 2) is a new AED that was
developed as a result of a once-presumed link between
anticonvulsant and antifolate properties (Reynolds et al.,
1966). It has proved to have a broad spectrum of activity,
with efficacy for partial, absence, myoclonic, and tonic-
clonic seizures (Leach & Brodie, 1995). There is no
evidence that the anticonvulsant activity of LTG is related
to its weak antifolate effect (Macdonald & Kelly, 1995).
Instead, like PHT and CBZ, it inhibits sustained repetitive
firing of action potentials (Cheung et al., 1992; Wang et al.,
1993) by blocking Na
+
channels in a voltage- and use-
dependent manner (Cheung et al., 1992; Lang et al., 1993;
Zona & Avoli, 1997).
The broad clinical profile of LTG suggests that its effects
on the Na
+
channel may differ from those observed with
PHT and CBZ. Unlike PHT, LTG acts principally on the
slow inactivated state of the channel (Kuo & Lu, 1997). In
addition, LTG may exhibit differential sensitivity for the
Fig. 1. Molecular structures of the established AEDs.
Fig. 2. Molecular structures of some new AEDs.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 25
various a-subunits of the Na
+
channel (Coulter, 1997).
These subunits have a markedly different regional distribu-
tion in the brain (Catterall, 1992). Indeed, it has been
suggested that LTG may selectively target Na
+
channels
on neurones that synthesise glutamate and aspartate (Leach
et al., 1986).
In addition to Na
+
channel effects, LTG reduces whole-
cell Ca
2 +
currents in rat amygdalar neurones, possibly via
the N- and P-type channels that have been implicated in
neurotransmitter release (Stefani et al., 1996b, 1997; Wang
et al., 1996). This effect might explain the inhibition of
electrically stimulated glutamate release from rat spinal
dorsal horn slices observed with the drug (Teoh et al.,
1995). It remains possible, however, that LTG possesses
additional unidentified mechanisms that confer its relatively
broader clinical spectrum when compared with other Na
+
channel-blocking agents.
3.4. Oxcarbazepine
Oxcarbazepine (OXC; Fig. 3) is a relatively novel AED,
with widespread geographical approval for clinical use
(Tecoma, 1999). It is closely related to CBZ in structure.
The keto substitutions at the 10 and 11 positions of the
dibenzazepine nucleus do not affect the therapeutic profile
of the drug when compared with CBZ, but result in altered
biotransformation and better tolerability (White, 1999). The
structural modifications circumvent the 10,11-epoxide
metabolite of CBZ that is believed to be responsible for
many of its side effects and its ability to induce cytochrome
P450-dependent hepatic metabolism (Tecoma, 1999). OXC
is essentially a pro-drug, and is rapidly and completely
reduced in the liver to its active metabolite, the monohy-
droxy derivative (10,11-dihydro-10-hydroxy CBZ; Edito-
rial, 1989).
In terms of mechanisms of action, OXC appears to exert
its pharmacological effects by blockade of voltage-depend-
ent Na
+
channels in a manner similar to that reported for
PHT and CBZ (McLean et al., 1994). It also reduces
presynaptic glutamate release, possibly by blocking high-
threshold Ca
2 +
currents (Calabresi et al., 1995; Stefani et
al., 1995, 1997). Interestingly, unlike any other licensed
AED, OXC may additionally increase K
+
channel conduc-
tance (McLean et al., 1994).
3.5. Ethosuximide
ESM has been used in the treatment of generalised
absence seizures for over 30 years (Fig. 1; Brodie & Dichter,
1997). It has no consistent efficacy for any other seizure
type. ESM exerts its anti-absence effects by reducing T-type
Ca
2 +
currents in thalamocortical relay neurones (Coulter et
al., 1989b, 1989c). As discussed in Section 2.1.2, the low-
threshold T-type Ca
2 +
channel predominates in these neuro-
nes, where it is believed to play a fundamental role in the
generation of the characteristic 3-Hz spike-and-wave dis-
charge of absence epilepsy (Coulter et al., 1989a). ESM
blocks the T-type Ca
2 +
channel and prevents synchronised
firing. It has no other known mechanisms of action (Rogaw-
ski & Porter, 1990).
3.6. Zonisamide
The development of zonisamide (ZNS; Fig. 3) was, until
recently, suspended following observations linking the drug
to an increased incidence of renal calculi (Leppik et al.,
1993; Leppik, 1999). However, long-term experience in
Japan, where the drug is licensed, failed to substantiate
these concerns. Clinical evidence to date suggests that ZNS
is effective against partial and generalised seizures, and has
particular efficacy in the progressive myoclonic epilepsies
that are often resistant to AED treatment (Dichter & Brodie,
1996; Kyllerman & Ben-Menachem, 1998).
The principal pharmacological action of ZNS involves
modulation of voltage-dependent ion channels. Like LTG,
ZNS enhances slow Na
+
channel inactivation (Schauf,
1987) and reduces sustained repetitive firing in spinal cord
neurones (Rock et al., 1989). It also blocks low-threshold T-
type Ca
2 +
currents, which may account for its anti-absence
effects (Suzuki et al., 1992).
ZNS also inhibits carbonic anhydrase, although this
action is believed to be too weak to contribute to its
antiepileptic effect (Leppik, 1999; Rho & Sankar, 1999).
It has also been shown to inhibit ligand binding to GABA
A
receptors and the associated BZD recognition site (Mimaki
et al., 1988). Other proposed mechanisms include inhibition
of monoamine release (Kawata et al., 1999) and metabolism
(Okada et al., 1995).
4. Potentiation of g-aminobutyric acid by
antiepileptic drugs
4.1. Phenobarbital
The barbiturates have been used since the early 1900s for
their sedative, anaesthetic, and anticonvulsant properties. Fig. 3. Molecular structures of remaining new AEDs.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 26
PB (Fig. 1) is still commonly prescribed worldwide for
epilepsy, although its cognitive and behavioural side effects
have limited its use, particularly in the developed world
(Mattson et al., 1985; Brodie & Dichter, 1997). The use of
primidone (PRM; Fig. 1) has been similarly restricted. It is
largely metabolised to PB, although there is evidence to
suggest that it possesses an additional active metabolite.
Variations in the experimental anticonvulsant profiles of PB
and PRM (Bourgeois et al., 1983), and the relatively greater
toxicity of PRM (Mattson et al., 1985), may be conferred by
this metabolite.
PB exerts its pharmacological effects by allosteric acti-
vation of the GABA
A
receptor, increasing the duration of
Cl

channel opening, without affecting the frequency of


opening or channel conductance (Macdonald et al., 1989;
Twyman et al., 1989). The barbiturates can also activate the
GABA
A
receptor directly, in the absence of GABA, an
effect that may underlie their sedative properties (Rho et al.,
1996; White, 1999). Additional mechanisms of barbiturate
action have been reported, including blockade of high-
voltage-activated Ca
2 +
channels (Rogawski & Porter,
1990) and an inhibitory effect on the AMPA/kainate subtype
of glutamate receptor (Davies, 1995).
4.2. Benzodiazepines
More than 50 chemically distinct BZDs are marketed
worldwide. Diazepam, lorazepam, clobazam, and clonaze-
pam are those most commonly used as AEDs (Fig. 4; Brodie
& Dichter, 1996; Dichter & Brodie, 1996). The antiepileptic
BZDs have a broad spectrum of clinical activity, with
efficacy in the partial and idiopathic generalised epilepsies
(Dichter & Brodie, 1996) and for the acute treatment of
status epilepticus (Treiman et al., 1998). The development
of tolerance to their pharmacological effects has, however,
restricted their use in chronic treatment regimens (Brodie &
Dichter, 1996).
The BZDs bind to a specific site in the brain (Squires &
Braestrup, 1977), now recognised as the a-subunit of the
GABA
A
receptor (Macdonald & Kelly, 1995). Binding
results in allosteric activation of the receptor, increasing
the frequency of Cl

channel opening, without affecting


open duration or channel conductance (Study & Barker,
1981; Twyman et al., 1989). Unlike the barbiturates, the
BZDs are unable to directly activate the GABA
A
receptor in
the absence of GABA (White, 1999).
Augmentation of GABAergic inhibition in the thalamus
can result in the de-inactivation of T-type Ca
2 +
channels,
triggering a strong low-threshold burst and enhancing
development of the thalamocortical rhythmicity that is
characteristic of absence seizures (Coulter, 1997). Some
GABAergic AEDs accordingly exacerbate absence seizures;
others are without effect, while the BZDs are notable by
their efficacy (Dichter & Brodie, 1996).
This phenomenon is explained by selective augmentation
of GABA
A
-mediated inhibition within the nucleus reticula-
ris thalami, but not the thalamus itself, reducing the syn-
chronising input to the thalamocortical circuits (Steriade &
Llinas, 1988; Coulter, 1997). A differential distribution of
a-subunits of the GABA
A
receptor between the nucleus
reticularis thalami and other thalamic nuclei may underlie
this apparent contradiction (Coulter, 1997).
4.3. Vigabatrin
In 1989, vigabatrin (VGB; Fig. 2) became the first of the
new generation of AEDs to be licensed in the United
Kingdom. It was initially approved as adjunctive therapy
for partial seizures with or without secondary generalisation
(Dichter & Brodie, 1996). It has subsequently demonstrated
particular efficacy in the treatment of infantile spasms
(Appleton et al., 1999). Recent evidence of an association
between VGB treatment and concentric visual field con-
striction will, however, limit the future use of the drug
(Kalviainen et al., 1999).
VGB is an ethyl analogue of GABA that is widely
recognised to exert its pharmacological effects by inhibition
of GABA-T, the enzyme responsible for the catabolism of
GABA (Jung et al., 1977). As a consequence, it elevates
GABA levels, potentiating inhibitory neurotransmission
throughout the brain (Schechter et al., 1977). VGB is classed
as a suicide inhibitor. It is transformed by GABA-T to an
active metabolite, which, thereafter, irreversibly binds to the
active site of the enzyme (Lippert et al., 1977).
A single dose of VGB reduces mouse brain GABA-T
activity to 20% of control levels and produces a 4-fold
increase in whole brain GABA concentrations. This effect
persists for over 24 hr, with GABA-T activity and GABA
concentrations only returning to normal upon the synthesis
of new enzyme protein over a period of 45 days Fig. 4. Structures of BDZ compounds commonly used in epilepsy.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 27
(Schechter et al., 1977). Similar effects are also observed in
humans (Petroff et al., 1996b; Petroff & Rothman, 1998).
Preliminary evidence suggests that, in addition to an effect
on GABA-T, VGB may block the uptake of GABA into
glial cells (Leach et al., 1996).
4.4. Tiagabine
Tiagabine (TGB) is a novel AED, recently licensed
widely for the adjunctive treatment of partial seizures with
or without secondary generalisation (Fig. 3; Leach &
Brodie, 1998). It is an analogue of nipecotic acid, a proto-
typic GABA uptake blocker, which is widely recognised to
prevent GABA transport into both neurones and glial cells
(Krogsgaard-Larsen et al., 1987). Nipecotic acid, however,
fails to cross the blood-brain barrier following systemic
administration. This problem is overcome by linking it to a
lipophilic anchor to form TGB, which is able to cross the
blood-brain barrier more readily (Brodie, 1995).
TGB inhibits GABA uptake into synaptosomal mem-
branes, neurones, and glial cells. It has a greater affinity
(2.5-fold) for glial than for neuronal uptake (Braestrup et al.,
1990). TGB has a selective action on the GAT-1 GABA
transporter, with little or no activity on GAT-2, GAT-3, or
BGT-1 (Borden et al., 1994). As a result, the pharmaco-
logical effect of TGB reflects the regional distribution of
GAT-1, and is mainly restricted to the cerebral cortex and
hippocampus (Meldrum & Chapman, 1999).
Unlike VGB, which elevates whole brain GABA con-
centrations, TGB temporarily prolongs the presence of
GABA in the synaptic cleft (Sills et al., 1999). This
important distinction has been confirmed by in vivo micro-
dialysis studies in rats (Fink-Jensen et al., 1992) and in the
hippocampus of a patient with refractory partial epilepsy
(During et al., 1992). In rat hippocampal slices, TGB pro-
longs the duration, but not the magnitude, of the peak
inhibitory postsynaptic current, consistent with delayed
clearance of GABA from the synapse (Roepstorff & Lam-
bert, 1992). Inhibition of GABA uptake is the only known
mechanism of TGB action (Brodie, 1995).
5. Antiepileptic drugs with multiple mechanisms
of action
5.1. Sodium valproate
The antiepileptic properties of VPA were discovered
serendipitously when valproic acid was employed in animal
studies as a solvent for drugs under formal investigation
(Fig. 1; Meunier et al., 1963). VPA has since proved to be
an extremely useful AED, with a broad spectrum of activity
and particular efficacy in the generalised epilepsies (Edito-
rial, 1988; Brodie & Dichter, 1996). However, the precise
mechanisms by which it exerts its antiepileptic effects
remain to be conclusively determined.
VPA has been reported to block voltage-dependent Na
+
channels. It reduces sustained repetitive firing of mouse
neurones in culture (McLean & Macdonald, 1986), inhibits
Na
+
channels in Xenopus leavis myelinated neurones (van
Dongen et al., 1986), and reduces Na
+
currents in neo-
cortical neurones (Zona & Avoli, 1990). However, rat
hippocampal slice studies suggest that, unlike PHT and
CBZ, VPA has no effect on the recovery of Na
+
channels
from the inactivated state (Albus & Williamson, 1998).
VPA may also block T-type Ca
2 +
channels in a manner
similar to that reported for ESM. Such an effect would
explain its efficacy against generalised absence seizures.
However, the reduction of T-type Ca
2 +
currents observed
with VPA in rat primary afferent neurones is modest and
requires relatively high drug concentrations (Kelly et al.,
1990). In addition, VPA appears to have no effect on Ca
2 +
channel conductance in rat thalamic neurones (Coulter
et al., 1989c).
There is evidence to suggest that VPA elevates whole
brain GABA levels and potentiates GABA responses,
possibly by enhancing GAD activity and inhibiting
GABA degradation (Loscher, 1999). Anecdotal reports
suggest that the drug also augments GABA release
(Rowley et al., 1995) and blocks GABA uptake (Sills et
al., 1996). The reproducibility of these effects has, how-
ever, been questioned (Rogawski & Porter, 1990). It is
suggested that the GABAergic effects of VPA exhibit a
degree of regional specificity within the brain and that
inconsistent results reflect the resolution of individual
studies (Loscher, 1999).
Single doses of VPA decrease brain levels of the
excitatory amino acid aspartate, without influencing those
of glutamate or GABA (Schechter et al., 1978). Decreases
in aspartate concentration have been shown to correspond
with the period of anticonvulsant activity in animal models
(Chapman et al., 1983). The relative anticonvulsant poten-
cies of a series of valproate analogues also correlates more
closely with their ability to reduce brain aspartate levels
than with their effects on GABA concentration (Chapman
et al., 1984). The potential contribution of any of the
above effects to the clinical activity of VPA remains to
be determined.
5.2. Gabapentin
Gabapentin (GBP) is a novel compound, structurally
related to GABA, which is effective in the adjunctive
treatment of partial seizures, with or without secondary
generalisation (Fig. 2; Dichter & Brodie, 1996). It was
originally designed as a GABAmimetic that could freely
cross the blood-brain barrier. However, subsequent studies
have shown that GBP does not directly interact with GABA
receptors (Taylor et al., 1998) or transporters (Su et al.,
1995; Macdonald & Greenfield, 1997).
There is evidence that GBP may increase the synthesis
(Taylor et al., 1992) and nonvesicular release of GABA
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 28
(Gotz et al., 1993), and may prevent its metabolism (Leach
et al., 1997). Using
1
H NMR spectroscopy, GBP has been
shown to elevate GABA concentrations in the occipital
cortex of epileptic patients (Petroff et al., 1996a). Whether
this observation is the result of enhanced synthesis,
increased release, or reduced metabolism of GABA remains
to be determined.
Early efforts to identify the mechanism of action of GBP
proposed an interaction with the L-amino acid transport
system, resulting in alterations in the cytosolic and extrac-
ellular concentrations of several amino acids, including L-
leucine, L-valine, and L-phenylalanine (Su et al., 1995).
Inhibition of branched chain amino acid transferase and
augmentation of glutamate dehydrogenase activities have
also been mooted (Goldlust et al., 1995). Investigations of
Na
+
channel function suggest that GBP does not affect
sustained repetitive firing in spinal cord neurones (Taylor et
al., 1988), although prolonged exposure reduces high-fre-
quency action potential firing in central neurones (Wamil &
McLean, 1994).
The identification of a specific binding site for GBP in
the mammalian brain, and its subsequent unveiling as the
a
2
d-subunit of the L-type voltage-dependent Ca
2 +
channel
(Gee et al., 1996), suggested another potential pharmaco-
logical mechanism. The implications of these findings
remain to be fully investigated, but the lack of effect of
GBP on whole cell Ca
2 +
currents in human dentate granule
cells acutely isolated from patients with temporal lobe
epilepsy (Schumacher et al., 1998) questions the pharmaco-
logical relevance of this finding.
5.3. Felbamate
Felbamate (FBM; Fig. 2) was licensed in the United
States in 1993 as monotherapy and add-on treatment for
partial-onset and primary generalised tonic-clonic seizures
in adults and for children with LennoxGastaut syndrome
(Dichter & Brodie, 1996). Its initial clinical success was
tempered by the significant incidence of aplastic anaemia
and hepatotoxicity revealed in postmarketing surveillance
(Pellock & Brodie, 1997).
FBM is believed to be the first effective AED with a
direct action on the NMDA subtype of glutamate receptor. It
inhibits NMDA/glycine-stimulated increases in intracellular
Ca
2 +
(Taylor et al., 1995), reduces inward currents evoked
by NMDA application to striatal neurones (Pisani et al.,
1995), and blocks NMDA receptor-mediated excitatory
postsynaptic potentials (Pugliese & Corradetti, 1996). Con-
siderable evidence suggests that FBM interacts with the
strychnine-insensitive glycine recognition site on the
NMDA receptor complex (White et al., 1995b). FBM
inhibits the binding of high-affinity glycine antagonists at
this site (McCabe et al., 1993), and its anticonvulsant effects
in several experimental models are blocked by glycine and
synthetic glycine site compounds (Coffin et al., 1994; White
et al., 1995b).
FBM facilitates the effects of diazepam against exper-
imentally induced seizures, implying an additional action at
the GABA
A
receptor (Gordon et al., 1991). Early studies
failed to demonstrate an effect of FBM on ligand binding to
the GABA, BZD, or picrotoxin recognition sites on the
receptor complex (Ticku et al., 1991). However, subsequent
investigations suggested that, in addition to inhibiting
NMDA responses, FBM potentiated GABA
A
receptor cur-
rents (Rho et al., 1994). The precise binding site for FBM on
the GABA
A
receptor remains to be identified, although a
weak interaction with the barbiturate recognition site has
been mooted (Rho et al., 1997).
Additional actions of FBM on voltage-dependent ion
channels have been reported. It reduces voltage-dependent
Na
+
currents in striatal neurones (Pisani et al., 1995) and
stabilises the inactive state of the a-subunit of Na
+
channels transiently expressed in Xenopus oocytes (Taglia-
latela et al., 1996). In cortical and neostriatal neurones,
FBM reduces high-voltage-activated Ca
2 +
channels, an
effect reversed by the L-type channel antagonist nifedipine,
but not by N- or P-type channel blockers (Stefani et al.,
1996a). The effects of FBM on voltage-activated ion
channels may explain the decrease in veratridine-induced
release of glutamate observed with the drug (Srinivasan
et al., 1996).
5.4. Topiramate
The sulfamate derivative topiramate (TPM; Fig. 2) is
active against partial-onset and generalised seizures in
humans (Wilson & Brodie, 1996; Brodie & French,
2000). TPM has multiple mechanisms of action, including
inhibition of Na
+
and Ca
2 +
currents, blockade of the
AMPA/kainate subtype of glutamate receptor, and facilita-
tion of GABA effects at the GABA
A
receptor. TPM also
inhibits carbonic anhydrase, although, like ZNS, this effect
is not believed to contribute to its antiepileptic action
(Shank et al., 1994).
TPM reduces the number of action potentials within a
burst and decreases burst duration in cultured hippocampal
neurones (DeLorenzo et al., 2000). It also reduces voltage-
dependent Na
+
currents in cultured cerebellar granule cells
(Zona et al., 1996a, 1996b). In this respect, its action is
similar to that of PHT and CBZ, delaying recovery of Na
+
channels from the inactive state. It has also been reported
that TPM blocks whole cell Ca
2 +
currents, possibly by an
action on the high-voltage-activated L-type channel (Zhang
et al., 2000).
In hippocampal neurones, TPM blocks inward currents
evoked by kainate, but not NMDA (Gibbs et al., 2000),
implying a selective effect on the AMPA/kainate subtype of
the glutamate receptor. TPM also interacts with the GABA
A
receptor. It enhances GABA-stimulated Cl

flux into
cerebellar granule neurones and increases Cl

currents
evoked by GABA in mouse cerebral cortical neurones
(White et al., 1997). The GABA
A
receptor effects of TPM
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 29
are not influenced by flumazenil, suggesting that they do not
involve the BZD recognition site on the receptor complex
(White et al., 1995a). Finally, recent evidence suggests that
brain concentrations of GABA, and its metabolites, are
elevated by TPM in patients with refractory epilepsy (Petr-
off et al., 1999). These effects, however, are not reproduced
in experimental animals (Sills et al., 2000).
6. Antiepileptic drugs with unknown mechanisms
of action
6.1. Levetiracetam
Levetiracetam (LEV; Fig. 3) is the S-enantiomer of the
ethyl analogue of piracetam, a widely used nootropic agent
in the elderly (Loscher & Honack, 1993). As the most
recently licensed AED, clinical experience with LEV is
limited (Genton & Van Vleymen, 2000). Results of clinical
trials suggest that it is effective against partial seizures with
or without secondary generalisation (Bialer et al., 1999).
However, more extensive investigation of LEV is required
before its full spectrum of clinical activity is revealed.
LEV appears to have a unique mode of action that, at
this time, remains to be clearly characterised. Exhaustive
preclinical investigations suggest that it does not interact
directly with any of the traditional targets, including Na
+
,
Ca
2 +
, and K
+
channels or the GABA and glutamate
neurotransmitter systems (Noyer et al., 1995). It is believed
to bind to a specific, as yet unidentified, site on the
synaptic plasma membrane. Competitive binding studies
suggest that LEV is not displaced by CBZ, PHT, VPA, PB,
clonazepam, picrotoxin, or bicuculline, but does interact
with ESM, pentylenetetrazol, and bemegride at this site
(Noyer et al., 1995). The implications of these observations
are unclear.
LEV has no effect on whole brain GABA synthesis or
metabolism, or on the concentrations of GABA, glutamate,
or glutamine (Sills et al., 1997). These studies, however, did
not discount the possibility of regionally specific effects on
the GABA system. Accordingly, LEV reduces GABA turn-
over in the striatum of the rat by increasing GABA-T
activity and reducing GAD activity (Loscher et al., 1996).
These modest effects are accompanied by a decrease in
spontaneous neuronal firing of the substantia nigra pars
reticulata, which receives a strong GABAergic input from
the striatum (Loscher et al., 1996).
Other preclinical studies suggest that LEV attenuates
bicuculline-induced increases in neuronal excitability in
the CA3 region of the rat hippocampus (Margineanu &
Wulfert, 1995), possibly by blockade of T-type Ca
2 +
channels (Margineanu & Wulfert, 1997). Suppression of
NMDA-induced bursting has also been reported (Birnstiel
et al., 1997). The relative contribution of any, or all, of the
anecdotal effects described above to the antiepileptic action
of LEV remains to be substantiated.
7. Conclusions
For the purposes of this review, it seemed prudent to
categorise the currently used AEDs according to their
principal mechanisms of action. However, it is becoming
apparent that most, if not all, have multiple cellular effects.
Given that the epilepsies are, by definition, a group of
disorders rather than a single disease entity, it is not
surprising that the currently employed AEDs display such
a diverse range of pharmacological actions.
The immediate challenge is to establish the relative
importance of individual mechanisms to the overall anti-
epileptic effect of each drug. The use of AEDs with known
modes of action has the potential to promote the under-
standing of the underlying pathophysiology of seizure dis-
orders and, in turn, to provide a framework for future
targetted drug development.
Despite familiarity with established AEDs and the intro-
duction of nine new agents in the past decade, up to one-third
of epilepsy patients remain resistant to optimum drug treat-
ment (Kwan & Brodie, 2000). Identifying the mechanisms
that matter has the potential to address the problem of
refractory epilepsy, at least in part. It is equally fundamental
to the development of a rational basis for future pharmaco-
therapy, centred around AED mechanisms of action and
tailored to the individual patient (Brodie et al., 1997).
References
Albus, H., & Williamson, R. (1998). Electrophysiologic analysis of the
actions of valproate on pyramidal neurons in the rat hippocampal slices.
Epilepsia 39, 124139.
Appleton, R. E., Peters, A. C. B., Mumford, J. P., & Shaw, D. E. (1999).
Randomised, placebo-controlled study of vigabatrin as first-line treat-
ment of infantile spasms. Epilepsia 40, 16271633.
Barchi, L. (1998). Ion channel mutations affecting muscle and brain. Curr
Opin Neurol 11, 461468.
Bialer, M., Johannessen, S. I., Kupferberg, H. J., Levy, R. H., Loiseau, P., &
Perucca, E. (1999). Progress report on new antiepileptic drugs: a sum-
mary of the fourth Eilat conference. Epilepsy Res 34, 141.
Birnstiel, S., Wulfert, E., & Beck, S. G. (1997). Levetiracetam (ucb L059)
affects in vitro models of epilepsy in CA3 pyramidal neurons without
altering normal synaptic transmission. Naunyn Schmiedebergs Arch
Pharmacol 356, 611618.
Borden, L. A., Smith, K. E., Hartig, P. R., Branchek, T. A., & Weinshank,
R. L. (1992). Molecular heterogeneity of the g-aminobutyric acid (GA-
BA) transport system. J Biol Chem 267, 2109821104.
Borden, L. A., Dhar, T. G. M., Smith, K. E., Weinshank, R. L., Branchek,
T. A., & Gluchowski, C. (1994). Tiagabine, SKF 89976-A, CI-966,
and NNC-711 are selective for cloned GABA transporter GAT-1. Eur
J Pharmacol 269, 219224.
Bourgeois, B. F. D., Dodson, W. E., & Ferrendelli, J. A. (1983). Primi-
done, phenobarbital and PEMA: I. Seizure protection, neurotoxicity
and therapeutic index of individual compounds in mice. Neurology 33,
283290.
Braestrup, C., Nielsen, G. B., Sonnewald, U., Knutsen, L. J. S., Andersen,
K. E., Jansen, J. A., Frederiksen, K., Andersen, P. H., Mortensen, A.,
& Suzdak, P. D. (1990). (R)-N-[4,4-bis(3-methyl-2-thienyl)but-3-en-1-
yl]nipecotic acid binds with high affinity to the brain g-aminobutyric
acid uptake carrier. J Neurochem 54, 639647.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 30
Brodie, M. J. (1995). Tiagabine pharmacology in profile. Epilepsia 36
(suppl. 6), S7S9.
Brodie, M. J. (1999). Monostars: an aid to choosing an antiepileptic drug as
monotherapy. Epilepsa 40 (suppl. 6), S17S22.
Brodie, M. J., & Dichter, M. A. (1996). Antiepileptic drugs. N Engl J Med
334, 168175.
Brodie, M. J., & Dichter, M. A. (1997). Established antiepileptic drugs.
Seizure 6, 159174.
Brodie, M. J., & French, J. A. (2000). Management of epilepsy in adoles-
cents and adults. Lancet 356, 323329.
Brodie, M. J., & Kwan, P. (2000). The star systems: overview and
use in determining drug choice for patients with epilepsy. CNS Drugs,
in press.
Brodie, M. J., & Yuen, A. W. C. The 105 Study Group (1997). Lamotrigine
substitution studyevidence for synergism with sodium valproate?
Epilepsy Res 26, 423432.
Calabresi, P., De Murtas, M., Stefani, A., Pisani, A., Sancesario, G., Mer-
curi, N. B., & Bernardi, G. (1995). Action of GP 47779, the active
metabolite of oxcarbazepine, on the corticostriatal system I. Modulation
of corticostriatal synaptic transmission. Epilepsia 36, 990996.
Catterall, W. A. (1992). Cellular and molecular biology of voltage-gated
sodium channels. Physiol Rev 72, S15S48.
Catterall, W. A. (1995). Structure and function of voltage-gated ion chan-
nels. Annu Rev Biochem 64, 493531.
Chapman, A. G., Meldrum, B. S., & Mendes, E. (1983). Acute anticonvul-
sant activity of structural analogues of valproic acid and changes in
brain GABA and aspartate content. Life Sci 32, 20232031.
Chapman, A. G., Croucher, M. J., & Meldrum, B. S. (1984). Anticon-
vulsant activity of intracerebroventricularly administered valproate and
valproate analogues. A dose-dependent correlation with changes in
brain aspartate and GABA levels in DBA/2 mice. Biochem Pharma-
col 33, 14591463.
Cheung, H., Kamp, D., & Harris, E. (1992). An in vitro investigation of the
action of lamotrigine on neuronal voltage-activated sodium channels.
Epilepsy Res 13, 107112.
Coffin, V., Cohen-Williams, M., & Barnett, A. (1994). Selective antago-
nism of the anticonvulsant effects of felbamate by glycine. Eur J Phar-
macol 256, R9R10.
Coulter, D. A. (1997). Antiepileptic drug cellular mechanisms of action:
where does lamotrigine fit in? J Child Neurol 12 (suppl. 1), S2S9.
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989a). Calcium currents
in rat thalamocortical relay neurones: kinetic properties of the transient
low-threshold current. J Physiol 414, 587604.
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989b). Specific petit
mal anticonvulsants reduce calcium currents in thalamic neurons. Neu-
rosci Lett 98, 7478.
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989c). Characterization
of ethosuximide reduction of low-threshold calcium currents in thalamic
neurons. Ann Neurol 25, 582593.
Courtney, K. R., & Etter, E. F. (1983). Modulated anticonvulsant block of
sodium channels in nerve and muscle. Eur J Pharmacol 88, 19.
Daikhin, Y., & Yudkoff, M. (2000). Compartmentation of brain glutamate
metabolism in neurons and glia. J Nutr 130, 1026S1031S.
Dailey, J. W., Reith, M. E., Yan, Q. S., Li, M. Y., & Jobe, P. C. (1997a).
Anticonvulsant doses of carbamazepine increase hippocampal extracel-
lular serotonin in genetically epilepsy-prone rats: dose response rela-
tionships. Neurosci Lett 227, 1316.
Dailey, J. W., Reith, M. E., Yan, Q. S., Li, M. Y., & Jobe, P. C. (1997b).
Carbamazepine increases extracellular serotonin concentration: lack
of antagonism by tetrodotoxin or zero Ca
2 +
. Eur J Pharmacol 328,
153162.
Davies, J. A. (1995). Mechanisms of action of antiepileptic drugs. Seizure
4, 267272.
DeLorenzo, R. J., Sombati, S., & Coulter, D. A. (2000). Effects of top-
iramate on sustained repetitive firing and spontaneous recurrent seizure
discharges in cultured hippocampal neurons. Epilepsia 41 (suppl. 1),
S40S44.
Dichter, M. A., & Brodie, M. J. (1996). New antiepileptic drugs. N Engl J
Med 334, 15831590.
Dolphin, A. C. (1995). Voltage-dependent calcium channels and their mod-
ulation by neurotransmitters and G proteins. Exp Physiol 80, 136.
During, M. J., Mattson, R. H., Scheyer, R., Rask, C., Pierce, M.,
McKelvy, J., & Thomas, V. (1992). The effect of tiagabine HCl on
extracellular GABA levels in the human hippocampus. Epilepsia 33
(suppl. 3), S83.
Editorial (1988). Sodium valproate. Lancet ii, 12291231.
Editorial (1989). Oxcarbazepine. Lancet ii, 196198.
Fink-Jensen, A., Suzdak, P. D., Swedberg, M. D., Judge, M. E., Hansen, L.,
& Nielsen, P. G. (1992). The g-aminobutyric acid (GABA) uptake
inhibitor, tiagabine, increases extracellular brain levels of GABA in
awake rats. Eur J Pharmacol 220, 197201.
Gandolfo, G., Romettino, S., Gottesman, C., Van Luijtelaer, G., Coenen, A.,
Bidard, J-N., & Lazdunski, M. (1989). K
+
channel openers decrease
seizures in genetically epileptic rats. Eur J Pharmacol 167, 181183.
Gee, N. S., Brown, J. P., Dissanayake, V. U. K., Offord, J., Thurlow, R.,
& Woodruff, G. N. (1996). The novel anticonvulsant drug, gabapentin
(Neurontin), binds to the a2d subunit of a calcium channel. J Biol
Chem 271, 57685776.
Genton, P., & Van Vleymen, B. (2000). Piracetam and levetiracetam: close
structural similarities but different pharmacological and clinical profiles.
Epileptic Disord 2, 99105.
Gibbs, J. W., Sombati, S., III, DeLorenzo, R. J., & Coulter, D. A. (2000).
Cellular actions of topiramate: blockade of kainate-evoked inward
currents in cultured hippocampal neurons. Epilepsia 41 (suppl. 1),
S10S16.
Goldlust, A., Su, T., Welty, D. F., Taylor, C. P., & Oxender, D. L. (1995).
Effects of the anticonvulsant drug gabapentin on enzymes in the meta-
bolic pathways of glutamate and GABA. Epilepsy Res 22, 111.
Gordon, R., Gels, M., Diamantis, W., & Sofia, R. D. (1991). Interaction of
felbamate and diazepam against maximal electroshock seizures and
chemoconvulsants in mice. Pharmacol Biochem Behav 40, 109113.
Gotz, E., Feuerstein, T. J., & Meyer, D. K. (1993). Effects of gabapentin on
release of g-aminobutyric acid from slices of rat neostriatum. Drug Res
43, 636638.
Granger, P., Biton, B., Faure, C., Vige, X., Depoortere, H., Graham, D.,
Langer, S. Z., Scatton, B., & Avenet, P. (1995). Modulation of the g-
aminobutyric acid type A receptor by the antiepileptic drugs carbama-
zepine and phenytoin. Mol Pharmacol 47, 11891196.
Guastella, J., Nelson, N., Nelson, H., Czyzyk, L., Keynan, S., & Miedel,
M. C. (1990). Cloning and expression of a rat brain GABA trans-
porter. Science 249, 13031306.
Hofmann, F., Biel, M., & Flockerzi, V. (1994). Molecular basis for Ca
2 +
channel diversity. Annu Rev Neurosci 17, 399418.
Hough, C. J., Irwin, R. P., Gao, X. M., Rogawski, M. A., & Chuang, D. M.
(1996). Carbamazepine inhibition of N-methyl-D-aspartate-evoked cal-
cium influx in rat cerebellar granule cells. J Pharmacol Exp Ther 276,
143149.
Johnston, G. A. R. (1996). GABA
A
receptor pharmacology. Pharmacol
Ther 69, 173198.
Jung, M. J., Lippert, B., Metcalf, B., Bohlen, P., & Schechter, P. J. (1977).
g-Vinyl GABA (4-amino-hex-5-enoic acid), a new irreversible inhibitor
of GABA-T: effects on brain GABA metabolism in mice. J Neurochem
29, 797802.
Kalviainen, R., Nousiainen, I., Mantyjarvi, M., Nikoskelainen, E., Parta-
nen, J., Partanen, K., & Riekkinen, P., Sr. (1999). Vigabatrin, a ga-
baergic antiepileptic drug, causes concentric visual field defects.
Neurology 53, 922926.
Kawata, Y., Okada, M., Murakami, T., Mizuno, K., Wada, K., Kondo, T., &
Kaneko, S. (1999). Effects of zonisamide on K
+
and Ca
2 +
evoked
release of monoamine as well as K
+
evoked intracellular Ca
2 +
mobi-
lization in rat hippocampus. Epilepsy Res 35, 173182.
Kelly, K. M., Gross, R. A., & Macdonald, R. L. (1990). Valproic acid
selectively reduces the low-threshold (T) calcium current in rat nodose
neurons. Neurosci Lett 116, 233238.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 31
Krogsgaard-Larsen, P., Falch, E., Larsson, O. M., & Schousboe, A. (1987).
GABA uptake inhibitors: relevance to antiepileptic drug research. Epi-
lepsy Res 1, 7793.
Kuo, C. C., & Lu, L. (1997). Characterization of lamotrigine inhibition of
Na
+
channels in rat hippocampal neurones. Br J Pharmacol 121,
12311238.
Kuo, C. C., Chen, R. S., Lu, L., & Chen, R. C. (1997). Carbamazepine
inhibition of neuronal Na
+
currents: quantitative distinction from phe-
nytoin and possible therapeutic implications. Mol Pharmacol 51,
10771083.
Kwan, P., & Brodie, M. J. (2000). Early identification of refractory epi-
lepsy. N Engl J Med 342, 314319.
Kyllerman, M., & Ben-Menachem, E. (1998). Zonisamide for progressive
myoclonus epilepsy: long-term observations in seven patients. Epilepsy
Res 29, 109114.
Lampl, I., Schwindt, P., & Crill, W. (1998). Reduction of cortical pyramidal
neuron excitability by the action of phenytoin on persistent Na
+
cur-
rent. J Pharmacol Exp Ther 284, 228237.
Lang, D. G., Wang, C. M., & Cooper, B. R. (1993). Lamotrigine, phenytoin
and carbamazepine interactions on the sodium current present in
N4TG1 mouse neuroblastoma cells. J Pharmacol Exp Ther 266,
829835.
Leach, J. P., & Brodie, M. J. (1995). Lamotrigine: clinical use. In R. Levy,
R. H. Mattson, B. S. Meldrum, & J. K. Penry (Eds.), Antiepileptic
Drugs, 4th edn. ( pp. 889895). New York: Raven Press.
Leach, J. P., & Brodie, M. J. (1998). Tiagabine. Lancet 351, 203207.
Leach, J. P., Sills, G. J., Majid, A., Butler, E., Carswell, A., Thompson,
G. G., & Brodie, M. J. (1996). Effects of tiagabine and vigabatrin on
GABA uptake into rat cortical astrocytes in primary culture. Seizure 5,
229234.
Leach, J. P., Sills, G. J., Butler, E., Forrest, G., Thompson, G. G., & Brodie,
M. J. (1997). Neurochemical actions of gabapentin in mouse brain.
Epilepsy Res 27, 175180.
Leach, M. J., Marden, C. M., & Miller, A. A. (1986). Pharmacological
studies on lamotrigine, a novel potential antiepileptic drug: II. Neuro-
chemical studies on the mechanism of action. Epilepsia 27, 490497.
Leppik, I. E. (1999). Zonisamide. Epilepsia 40 (suppl. 5), S23S29.
Leppik, I. E., Willmore, L. J., Homan, R. W., Fromm, G., Oommen, K. J.,
Penry, J. K., Sackellares, J. C., Smith, D. B., Lesser, R. P., Wallace,
J. D., Trudeau, J. L., Lamoreaux, L. K., & Spenser, M. (1993).
Efficacy and safety of zonisamide: results of a multicenter study.
Epilepsy Res 14, 165173.
Lippert, B., Metcalf, B. W., Jung, M. J., & Casara, P. (1977). 4-Amino-hex-
5-eonic acid, a selective catalytic inhibitor of 4-aminobutyric-acid ami-
notransferase in mammalian brain. Eur J Biochem 4, 441445.
Loscher, W. (1998). New visions in the pharmacology of anticonvulsion.
Eur J Pharmacol 342, 113.
Loscher, W. (1999). Valproate: a reappraisal of its pharmacodynamic prop-
erties and mechanisms of action. Prog Neurobiol 58, 3159.
Loscher, W., & Honack, D. (1993). Profile of ucb L059, a novel anticon-
vulsant drug, in models of partial and generalized epilepsy in mice and
rats. Eur J Pharmacol 232, 147158.
Loscher, W., Honack, D., & Bloms-Funke, P. (1996). The novel antiepi-
leptic drug levetiracetam (ucb L059) induces alterations in GABA me-
tabolism and turnover in discrete areas of rat brain and reduces neuronal
activity in substantia nigra pars reticulata. Brain Res 735, 208216.
Macdonald, R. L., & Greenfield, L. J., Jr. (1997). Mechanisms of action of
new antiepileptic drugs. Curr Opin Neurol 10, 121128.
Macdonald, R. L., & Kelly, K. M. (1995). Antiepileptic drug mechanisms
of action. Epilepsia 36 (suppl. 2), S2S12.
Macdonald, R. L., Rogers, C. J., & Twyman, R. E. (1989). Barbiturate
regulation of kinetic properties of the GABA
A
receptor channel of
mouse spinal neurones in culture. J Physiol 417, 483500.
Marangos, P. J., Post, R. M., Patel, J., Zander, A., Parma, K., & Weiss, S.
(1983). Specific and potent interactions of carbamazepine with brain
adenosine receptors. Eur J Pharmacol 93, 175182.
Margineanu, D. G., & Wulfert, E. (1995). ucb L059, a novel anticonvulsant,
reduces bicuculline-induced hyperexcitability in rat hippocampal CA3
in vivo. Eur J Pharmacol 286, 321325.
Margineanu, D. G., & Wulfert, E. (1997). Inhibition by levetiracetam of a
non-GABA
A
receptor-associated epileptiform effect of bicuculline in rat
hippocampus. Br J Pharmacol 122, 11461150.
Mattson, R. H., Cramer, J. A., Collins, J. F., Smith, D. B., Delgado-
Escueta, A. V., Browne, T. R., Williamson, P. D., Treiman, D. M.,
McNamara, J. O., McCutchen, C. B., Homan, R. W., Crill, W. E.,
Lubozynski, M. F., Rosenthal, N. P., & Mayersdorf, A. (1985). Com-
parison of carbamazepine, phenobarbital, phenytoin and primidone in
partial and secondary generalised tonic-clonic seizures. N Engl J Med
313, 145151.
McCabe, R. T., Wasterlain, C. G., Kucharczyk, N., Sofia, R. D., & Vogel,
J. R. (1993). Evidence of anticonvulsant and neuroprotective action of
felbamate mediated by strychnine-insensitive glycine receptors. J Phar-
macol Exp Ther 264, 248252.
McLean, M. J., & Macdonald, R. L. (1983). Multiple actions of phenytoin
on mouse spinal cord neurons in cell culture. J Pharmacol Exp Ther
227, 779789.
McLean, M. J., & Macdonald, R. L. (1986). Sodium valproate, but not
ethosuximide, produces use- and voltage-dependent limitation of high
frequency repetitive firing of action potentials of mouse central neurons
in cell culture. J Pharmacol Exp Ther 237, 10011011.
McLean, M. J., Schmutz, M., Wamil, A. W., Olpe, H. R., Portet, C., &
Feldmann, K. F. (1994). Oxcarbazepine: mechanisms of action. Epilep-
sia 35 (suppl. 3), S5S9.
Meldrum, B. S. (1995). Neurotransmission in epilepsy. Epilepsia 36
(suppl. 1), S30S35.
Meldrum, B. S. (1996). Update on the mechanism of action of antiepileptic
drugs. Epilepsia 37 (suppl. 6), S4S11.
Meldrum, B. S. (2000). Glutamate as a neurotransmitter in the brain: review
of physiology and pathology. J Nutr 130, 1007S1015S.
Meldrum, B. S., & Chapman, A. G. (1999). Basic mechanisms of Gabitril
(tiagabine) and future potential developments. Epilepsia 40 (suppl. 9),
S2S6.
Meldrum, B. S., Akbar, M. T., & Chapman, A. G. (1999). Glutamate
receptors and transporters in genetic and acquired models of epilepsy.
Epilepsy Res 362, 189204.
Merritt, H. H., & Putnam, T. J. (1938). Sodium diphenyl hydantoinate in the
treatment of convulsive disorders. JAMA 111, 10681073.
Meunier, H., Carraz, G., Meunier, Y., Eymard, P., & Aimard, M. (1963).
Proprietes pharmacodynamiques de lacide n-propylacetique. 1
er
mem-
oire: proprietes antiepileptique. Therapie 18, 435438.
Mimaki, T., Suzuki, Y., Tagawa, T., Tanaka, J., Itoh, N., & Yabuuchi, H.
(1988). [
3
H]Zonisamide binding in rat brain. Jpn J Psychiatry Neurol
42, 640642.
Mody, I. (1998). Ion channels in epilepsy. Int Rev Neurobiol 42, 199226.
Nobile, M., & Vercellino, P. (1997). Inhibition of delayed rectifier K
+
channels by phenytoin in rat neuroblastoma cells. Br J Pharmacol
120, 647652.
Noyer, M., Gillard, M., Matagne, A., Henichart, J. P., & Wulfert, E.
(1995). The novel antiepileptic drug levetiracetam (ucb L059) appears
to act via a specific binding site in CNS membranes. Eur J Pharmacol
286, 137146.
Okada, M., Kaneko, S., Hirano, T., Mizuno, K., Kondo, T., Otani, K., &
Fukushima, Y. (1995). Effects of zonisamide on dopaminergic system.
Epilepsy Res 22, 193205.
Olsen, R. W., & Avoli, M. (1997). GABA and epileptogenesis. Epilepsia
38, 399407.
Ophoff, R. A., Terwindt, G. M., Frants, R. R., & Ferrari, M. D. (1998).
P/Q-type Ca
2 +
channel defects in migraine, ataxia and epilepsy.
Trends Pharmacol Sci 19, 121127.
Pellock, J. M., & Brodie, M. J. (1997). Felbamate: an update. Epilepsia 38,
12611264.
Petroff, O. A. C., & Rothman, D. L. (1998). Measuring human brain GABA
in vivo: effects of GABA-transaminase inhibition with vigabatrin. Mol
Neurobiol 16, 97121.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 32
Petroff, O. A. C., Rothman, D. L., Behar, K. L., Lamoureux, D., & Mattson,
R. H. (1996a). The effect of gabapentin on brain g-aminobutyric acid in
patients with epilepsy. Ann Neurol 39, 9599.
Petroff, O. A. C., Rothman, D. L., Behar, K. L., & Mattson, R. H. (1996b).
Human brain GABA levels rise after initiation of vigabatrin therapy but
fail to rise further with increasing dose. Neurology 46, 14591463.
Petroff, O. A. C., Hyder, F., Mattson, R. H., & Rothman, D. L. (1999).
Topiramate increases brain GABA, homocarnosine, and pyrrolidinone
in patients with epilepsy. Neurology 52, 473478.
Pisani, A., Stefani, A., Siniscalchi, A., Mercuri, N. B., Bernardi, G., &
Calabresi, P. (1995). Electrophysiological actions of felbamate on rat
striatal neurones. Br J Pharmacol 16, 20532061.
Pongs, O. (1999). Voltage-gated potassium channels: from hyperexcitability
to excitement. FEBS Lett 452, 3135.
Porter, R. J., & Rogawski, M. A. (1992). New antiepileptic drugs: from
serendipity to rational discovery. Epilepsia 33 (suppl. 1), S1S6.
Pugliese, A. M., & Corradetti, R. (1996). Effects of the antiepileptic drug
felbamate on long-term potentiation in the CA1 region of rat hippo-
campal slices. Neurosci Lett 215, 2124.
Rabow, L. E., Russek, S. J., & Farb, D. H. (1995). From ion currents to
genomic analysis: recent advances in GABA
A
receptor research. Syn-
apse 21, 189274.
Ragsdale, D. S., & Avoli, M. (1998). Sodium channels as molecular targets
for antiepileptic drugs. Brain Res Rev 26, 1628.
Ragsdale, D. S., McPhee, J. C., Scheuer, T., & Catterall, W. A. (1996).
Common molecular determinants of local anesthetic, antiarrhythmic,
and anticonvulsant block of voltage-gated Na
+
channels. Proc Natl
Acad Sci USA 93, 92709275.
Reynolds, E. H., Milner, G., Matthews, D. M., & Chanarin, I. (1966).
Anticonvulsant therapy, megaloblastic haemopoiesis and folic acid me-
tabolism. Q J Med 35, 521537.
Rho, J. M., & Sankar, R. (1999). The pharmacologic basis of antiepileptic
drug action. Epilepsia 40, 14711483.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1994). Mechanism of
action of the anticonvulsant felbamate: opposing effects on N-methyl-
D-aspartate and g-aminobutyric acid
A
receptors. Ann Neurol 35,
229234.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1996). Direct activation of
GABA
A
receptors by barbiturates in cultured rat hippocampal neurons.
J Physiol 497, 509522.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1997). Barbiturate-like
actions of the propanediol dicarbamates felbamate and meprobamate. J
Pharmacol Exp Ther 280, 13831391.
Rock, D. M., Macdonald, R. L., & Taylor, C. P. (1989). Blockade of
sustained repetitive action potentials in cultured spinal cord neurons
by zonisamide (AD 810, CI 912): a novel anticonvulsant. Epilepsy
Res 3, 138143.
Roepstorff, A., & Lambert, J. D. (1992). Comparison of the effect of
GABA uptake blockers, tiagabine and nipecotic acid, on inhibitory
synaptic efficacy in hippocampal CA1 neurones. Neurosci Lett 146,
131134.
Rogawski, M. A. (2000). KCNQ2/KCNQ3 K
+
channels and the molecular
pathogenesis of epilepsy:implications for therapy. Trends Neurosci 23,
393398.
Rogawski, M. A., & Porter, R. J. (1990). Antiepileptic drugs: pharmaco-
logical mechanisms and clinical efficacy with consideration of promis-
ing developmental stage compounds. Pharmacol Rev 42, 223286.
Rostock, A., Tober, C., Rundfeldt, C., Bartsch, R., Engel, J., Polymer-
opoulos, E. E., Kutscher, B., Loscher, W., Honack, D., White, H. S.,
& Wolf, H. H. (1996). D-23129: a new anticonvulsant with a broad
spectrum activity in animal models of epileptic seizures. Epilepsy Res
23, 211223.
Rowley, H. L., Marsden, C. A., & Martin, K. F. (1995). Differential effects
of phenytoin and sodium valproate on seizure-induced changes in g-
aminobutyric acid and glutamate release in vivo. Eur J Pharmacol 294,
541546.
Rundfeldt, C., & Netzer, R. (2000). The novel anticonvulsant retigabine
activates M-currents in Chinese hamster ovary-cells transfected with
human KCNQ2/3 subunits. Neurosci Lett 282, 7376.
Schachter, S. C. (1995). Review of the mechanisms of action of antiepi-
leptic drugs. CNS Drugs 4, 469477.
Schauf, C. L. (1987). Zonisamide enhances slow sodium inactivation in
Myxicola. Brain Res 413, 185188.
Schechter, P. J., Tranier, Y., Jung, M. J., & Bohlen, P. (1977). Audiogenic
seizure protection by elevated brain GABA concentration in mice: ef-
fects of g-acetylenic GABA and g-vinyl GABA, two irreversible GA-
BA-T inhibitors. Eur J Pharmacol 45, 319328.
Schechter, P. J., Tranier, Y., & Grove, J. (1978). Effect of n-dipropylacetate
on amino acid concentrations in mouse brain: correlation with anticon-
vulsant activity. J Neurochem 31, 13251327.
Schumacher, T. B., Beck, H., Steinhauser, C., Schramm, J., & Elger, C. E.
(1998). Effects of phenytoin, carbamazepine, and gabapentin on calci-
um channels in hippocampal granule cells from patients with temporal
lobe epilepsy. Epilepsia 39, 355363.
Schwartz, J. R., & Grigat, G. (1989). Phenytoin and carbamazepine: poten-
tial- and frequency-dependent block of Na
+
currents in mammalian
myelinated nerve fibers. Epilepsia 30, 286294.
Shank, R. P., Gardocki, J. F., Vaught, J. L., Davis, C. B., Schupsky, J. J.,
Raffa, R. B., Dodgson, S. J., Nortey, S. O., & Maryanoff, B. E. (1994).
Topiramate: preclinical evaluation of a structurally novel anticonvul-
sant. Epilepsia 35, 450460.
Sills, G. J., Leach, J. P., Butler, E., Carswell, A., Thompson, G. G., &
Brodie, M. J. (1996). Antiepileptic drug action in primary cultures of
rat cortical astrocytes. Epilepsia 37 (suppl. 4), 116.
Sills, G. J., Leach, J. P., Fraser, C. M., Forrest, G., Patsalos, P. N., & Brodie,
M. J. (1997). Neurochemical studies with the novel anticonvulsant
levetiracetam in mouse brain. Eur J Pharmacol 325, 3540.
Sills, G. J., Butler, E., Thompson, G. G., & Brodie, M. J. (1999). Vigabatrin
and tiagabine are pharmacologically different drugs. A pre-clinical
study. Seizure 8, 404411.
Sills, G. J., Leach, J. P., Kilpatrick, W. S., Fraser, C. M., Thompson, G. G.,
& Brodie, M. J. (2000). Concentrationeffect studies with topiramate
on selected enzymes and intermediates of the GABA shunt. Epilepsia
41 (suppl. 1), S30S34.
Song, J. H., Nagata, K., Huang, C. S., Yeh, J. Z., & Narahashi, T. (1996).
Differential block of two types of sodium channels by anticonvulsants.
Neuroreport 7, 30313036.
Squires, R. F., & Braestrup, C. (1977). Benzodiazepine receptors in rat
brain. Nature 266, 732734.
Srinivasan, J., Richens, A., & Davies, J. A. (1996). Effects of felbamate on
veratridine- and K
+
-stimulated release of glutamate from mouse cortex.
Eur J Pharmacol 315, 285288.
Stefani, A., Pisani, A., De Murtas, M., Mercuri, N. B., Marciani, M. G., &
Calabresi, P. (1995). Action of GP 47779, the active metabolite of
oxcarbazepine, on the corticostriatal system. II. Modulation of high-
voltage-activated calcium currents. Epilepsia 336, 9971002.
Stefani, A., Calabresi, P., Pisani, A., Mercuri, N. B., Siniscalchi, A., &
Bernardi, G. (1996a). Felbamate inhibits dihydropyridine-sensitive
calcium channels in central neurons. J Pharmacol Exp Ther 277,
121127.
Stefani, A., Spadoni, F., Siniscalchi, A., & Bernadi, G. (1996b). Lamotri-
gine inhibits Ca
2 +
currents in cortical neurons: functional implications.
Eur J Pharmacol 307, 113116.
Stefani, A., Spadoni, F., & Bernardi, G. (1997). Voltage-activated cal-
cium channels: targets of antiepileptic drug therapy? Epilepsia 38,
959965.
Steriade, M., & Llinas, R. (1988). The functional states of the thalamus and
the associated neuronal interplay. Physiol Rev 68, 649742.
Study, R. E., & Barker, J. L. (1981). Diazepam and ( )-pentobarbital:
fluctuation analysis reveals different mechanisms for potentiation of
g-aminobutyric acid responses in cultured central neurons. Proc Natl
Acad Sci USA 78, 71807184.
Su, T. Z., Lunney, E., Campbell, G., & Oxender, D. L. (1995). Transport of
gabapentin, a g-amino acid drug, by system l g-amino acid transporters:
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 33
a comparative study in astrocytes, synaptosomes, and CHO cells. J
Neurochem 64, 21252131.
Suzuki, S., Kawakami, K., Nishimura, S., Watanabe, Y., Yagi, K.,
Seino, M., & Miyamoto, K. (1992). Zonisamide blocks T-type calcium
channels in cultured neurons of rat cerebral cortex. Epilepsy Res 12,
2127.
Taglialatela, M., Ongini, E., Brown, A. M., Di Renzo, G., & Annunziato, L.
(1996). Felbamate inhibits cloned voltage-dependent Na
+
channels
from human and rat brain. Eur J Pharmacol 316, 373377.
Taylor, C. P., Rock, D. M., Weinkauf, R. J., & Ganong, A. H. (1988). In
vitro and in vivo electrophysiology effects of the anticonvulsant gaba-
pentin. Soc Neurosci Abstr 14, 866.
Taylor, C. P., Vartanian, M. G., Andruszkiewicz, R., & Silverman, R. B.
(1992). 3-Alkyl GABA and 3-alkylglutamic acid analogues: two new
classes of anticonvulsant agents. Epilepsy Res 11, 103110.
Taylor, C. P., Gee, N. S., Su, T-Z., Kocsis, J. D., Welty, D. F., Brown, J. P.,
Dooley, D. J., Boden, P., & Singh, L. (1998). A summary of mech-
anistic hypotheses of gabapentin pharmacology. Epilepsy Res 29,
233249.
Taylor, L. A., McQuade, R. D., & Tice, M. A. (1995). Felbamate, a novel
antiepileptic drug, reverses N-methyl-D-aspartate/glycine-stimulated in-
creases in intracellular Ca
2 +
concentration. Eur J Pharmacol 289,
229233.
Tecoma, E. S. (1999). Oxcarbazepine. Epilepsia 40 (suppl. 5), S37S46.
Teoh, H., Fowler, L. J., & Bowery, N. G. (1995). Effect of lamotrigine
on the electrically-evoked release of endogenous amino acids from
slices of dorsal horn of the rat spinal cord. Neuropharmacology 34,
12731278.
Ticku, M. K., Kamatchi, G. L., & Sofia, R. D. (1991). Effect of anticon-
vulsant felbamate on GABA receptor system. Epilepsia 32, 389391.
Treiman, D. M., Meyers, P. D., Walton, N. Y., Collins, J. F., Colling, C.,
Rowan, A. J., Handforth, A., Faught, E., Calabrese, V. P., Uthman, B.,
Ramsay, R. E., & Mamdani, B. (1998). A comparison of four treatments
for generalized convulsive status epilepticus. N Engl J Med 339, 792
798.
Trist, D. G. (2000). Excitatory amino acid agonists and antagonists: phar-
macology and therapeutic applications. Pharm Acta Helv 74, 221229.
Tunnicliff, G. (1996). Basis of the antiseizure action of phenytoin. Gen
Pharmacol 27, 10911097.
Twyman, R. E., Rogers, C. J., & Macdonald, R. L. (1989). Differential
regulation of g-aminobutyric acid receptor channels by diazepam and
phenobarbital. Ann Neurol 25, 213220.
Upton, N. (1994). Mechanisms of action of new antiepileptic drugs: rational
design and serendipitous findings. Trends Pharmacol Sci 15, 456463.
van Dongen, A. M. J., van Erp, M. G., & Voskuyl, R. A. (1986). Valproate
reduces excitability by blockage of sodium and potassium conductance.
Epilepsia 27, 177182.
Waldmeier, P. C., Baumann, P. A., Wicki, P., Feldtrauer, J. J., Stierlin, C., &
Schmutz, M. (1995). Similar potency of carbamazepine, oxcarbazepine,
and lamotrigine in inhibiting the release of glutamate and other neuro-
transmitters.. Neurology 45, 19071913.
Wamil, A. W., & McLean, M. J. (1994). Limitation by gabapentin of high
frequency action potential firing by mouse central neurons in culture.
Epilepsy Res 17, 111.
Wang, C. M., Lang, D. G., & Cooper, B. R. (1993). Lamotrigine effects
on ion channels in cultured neuronal cells. Epilepsia 34 (suppl. 6),
117118.
Wang, S. J., Huang, C. C., Hsu, K. S., Tsai, J. J., & Gean, P. W. (1996).
Inhibition of N-type calcium currents by lamotrigine in rat amygdalar
neurones. Neuroreport 7, 30373040.
White, H. S. (1999). Comparative anticonvulsant and mechanistic profile
of established and newer antiepileptic drugs. Epilepsia 40 (suppl. 5),
S2S10.
White, H. S., Brown, D., Skeen, G. A., Wolf, H. H., & Twyman, R. E.
(1995a). The anticonvulsant topiramate displays a unique ability to
potentiate GABA-evoked chloride currents. Epilepsia 36 (suppl. 3),
S39S40.
White, H. S., Harmsworth, W. L., Sofia, R. D., & Wolf, H. H. (1995b).
Felbamate modulates the strychnine-insensitive glycine receptor. Epi-
lepsy Res 20, 4148.
White, H. S., Brown, S. D., Woodhead, J. H., Skeen, G. A., & Wolf, H. H.
(1997). Topiramate enhances GABA-mediated chloride flux and
GABA-evoked chloride currents in murine brain neurons and increases
seizure threshold. Epilepsy Res 28, 167179.
Wilson, E. A., & Brodie, M. J. (1996). New antiepileptic drugs. In: M. J.
Brodie, & D. M. Brodie (Eds.), Ballie`res Clinical Neurology. Modern
Management of Epilepsy ( pp. 723747). London: Ballie`re-Tindall.
Yamaguchi, S., & Rogawski, M. A. (1992). Effects of anticonvulsant drugs
on 4-aminopyridine-induced seizures in mice. Epilepsy Res 11, 916.
Zhang, X., Velumian, A. A., Jones, O. T., & Carlen, P. L. (2000). Modu-
lation of high-voltage-activated calcium channels in dentate granule
cells by topiramate. Epilepsia 41 (suppl. 1), S52S60.
Zona, C., & Avoli, M. (1990). Effects induced by the antiepileptic drug
valproic acid upon the ionic currents recorded in rat neocortical neurons
in cell culture. Exp Brain Res 81, 313317.
Zona, C., & Avoli, M. (1997). Lamotrigine reduces voltage-gated sodium
currents in rat central neurons in culture. Epilepsia 38, 522525.
Zona, C., Barbarosie, M., Kawasaki, H., & Avoli, M. (1996a). Effects
induced by the anticonvulsant drug topiramate on voltage-gated sodium
currents generated by cerebellar granule cells in tissue culture. Epilepsia
37 (suppl. 5), 24.
Zona, C., Ciotti, M. T., & Avoli, M. (1996b). Topiramate attenuates volt-
age-gated sodium currents in rat cerebellar granule cells. Neurosci Lett
231, 123126.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 2134 34

S-ar putea să vă placă și