Vous êtes sur la page 1sur 14

Pharmacology & Therapeutics 90 (2001) 21 – 34

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. Ca2+ 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

* 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).

0163-7258/01/$ – see front matter D 2001 Elsevier Science Inc. All rights reserved.
PII: S 0 1 6 3 - 7 2 5 8 ( 0 1 ) 0 0 1 2 2 - X
22 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

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 1.1. Classification of mechanisms of action

The serendipitous discovery of phenobarbital (PB) in Practical approaches to drug treatment have resorted
1912 marked the beginning of the modern pharmacotherapy primarily to symptom control, i.e., suppression of seiz-
of epilepsy (Brodie & Dichter, 1996). In the ensuing 70 ures. Although the mechanisms of action of the currently
years, phenytoin (PHT), carbamazepine (CBZ), ethosuxi- marketed AEDs are still not completely understood, they
mide (ESM), sodium valproate (VPA), and a range of ultimately involve alteration of the balance between neu-
benzodiazepines (BZDs) became available and can be ronal excitation and inhibition (White, 1999). At the
regarded as ‘‘established’’ antiepileptic drugs (AEDs; Bro- cellular level, three basic mechanisms are recognised:
die & Dichter, 1996). In the past decade, nine new agents modulation of voltage-dependent ion channels (Na + ,
have been licensed as add-on treatment for difficult-to- Ca 2 + , K + ), enhancement of g-aminobutyric acid
control epilepsy (Table 1) and still more are under evalua- (GABA)-mediated inhibitory neurotransmission, and
tion (Dichter & Brodie, 1996). attenuation of excitatory (particularly glutamate-mediated)
The undoubtedly beneficial expansion of the pharmaco- transmission (Meldrum, 1996).
logical armamentarium for the treatment of epilepsy does, In this review, we will attempt to introduce the concepts
however, complicate selection of the most suitable AED (or of ion channel and neurotransmitter modulation and, there-
combination of AEDs) for individual patients. With limited after, group the currently used AEDs according to their
clinical experience of the new agents, the mechanisms of principal mechanisms of action (Table 1). It is, however,
action of individual AEDs may be an important criterion in increasingly recognised that many AEDs possess multiple
this decision-making process (Brodie, 1999; Brodie & primary mechanisms. In addition, many less well-charac-
Kwan, 2000). 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
Table 1 remains to be discovered.
Proposed mechanisms of antiepileptic drug action
# Na + # Ca2 + " K+ " Inhibitory # Excitatory
channels channels channels transmission transmission
Established AEDs 2. Targets for antiepileptic drug action
PHT +++
CBZ +++ 2.1. Ion channels
ESM +++
PB + +++ +
BZDs +++ 2.1.1. Na+ channels
VPA + + ++ + In the nervous system, voltage-gated ion channels control
the flow of cations across surface and internal cell mem-
New AEDs branes (Barchi, 1998). Of these, the Na + channel is argu-
LTG +++ +
OXC +++ + +
ably of principal importance. Voltage-dependent Na +
ZNS ++ ++ channels are responsible for the upstroke of the neuronal
VGB +++ action potential, and ultimately control the intrinsic excit-
TGB +++ ability of the nervous system (Porter & Rogawski, 1992).
GBP + + ++ The neuronal Na + channel has a multi-subunit structure that
FBM ++ ++ ++ ++
TPM ++ ++ ++ ++
forms a Na + -selective, voltage-gated pore through the
LEV + + + plasma membrane. The protein structure undergoes confor-
+++, primary action; ++, probable action; +, possible action. mational alterations in response to changes in membrane
Data from Upton (1994), Schachter (1995), Macdonald & Kelly (1995), potential, regulating conductance through the intrinsic pore
Meldrum (1996), Coulter (1997), and White (1999). (Ragsdale & Avoli, 1998).
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 23

The main structural component of the neuronal Na + subtype-specific manner, an effect that may contribute to
channel is the a-subunit, which forms the ion-conducting their antiepileptic actions (Stefani et al., 1997).
pore and confers voltage dependency (Catterall, 1992). In
the mammalian brain, the a-subunit associates with two 2.1.3. K+ channels
auxiliary subunits designated b1 and b2. The b-subunits are Neuronal K + channels are large protein complexes that
not required for basic Na + channel activity, but they form tetrameric structures, the monomers of which are
modulate the expression and function of individual channels structurally and genetically related to the a- and a1-subunits
(Ragsdale & Avoli, 1998). of the Na + and Ca2 + channel, respectively (Barchi, 1998).
At normal membrane potentials, most Na + channels The association of four subunits (monomers) in the neuronal
exist in a closed, resting state. Upon depolarisation, the membrane is required for the formation of a K + -sensitive
channel activates, facilitating ion flux. Thereafter, the Na + pore and, therefore, channel function (Pongs, 1999). More
channel enters an inactivated state, from which it is not than 40 distinct K + channel subunits have been identified,
readily re-activated. Repolarisation of the neuronal mem- together with several auxiliary subunits. Given heterologous
brane rapidly converts the channel back to a resting state, arrangement, it is possible that countless populations of K +
from which it can respond to subsequent depolarisations channels, with individual functions and distributions, are
(Catterall, 1992; Ragsdale & Avoli, 1998). Neuronal Na + expressed in the mammalian brain (Pongs, 1999).
channels can cycle through these functional states within a At the neuronal level, K + channels are intimately
few milliseconds. This characteristic is essential for sustain- involved in excitability. They are responsible for the action
ing the rapid bursts of action potentials necessary for some potential downstroke or, more specifically, repolarisation of
normal brain functions, and is implicated in the production the plasma membrane in the aftermath of Na + channel
of epileptic discharges. The neuronal Na + channel repre- activation (Pongs, 1999). Direct activation of voltage-
sents one of the most important targets for AED action dependent K + channels hyperpolarises the neuronal mem-
(Upton, 1994; Macdonald & Kelly, 1995; Meldrum, 1996; brane and limits action potential firing (Porter & Rogaw-
White, 1999). ski, 1992). Accordingly, K + channel activators have
anticonvulsant effects in some experimental seizure models
2.1.2. Ca2+ channels (Gandolfo et al., 1989; Rostock et al., 1996), whereas K +
Voltage-dependent Ca2 + channels share key structural channel blockers precipitate seizures (Yamaguchi &
elements and sequence homology with their Na + channel Rogawski, 1992).
counterparts (Barchi, 1998). The a1-subunit of the Ca2 + Potentiation of voltage-sensitive K + channel currents
channel is the homologue of the a-subunit of the Na + may prove to be an important target for future AED
channel. It forms the Ca2 + -sensitive channel pore and development. The novel antiepileptic agent retigabine, cur-
confers voltage dependency (Catterall, 1995). In the mam- rently undergoing Phase II clinical trial, is believed to exert
malian brain, the a1-subunit heterogeneously associates its effects, at least in part, by activation of the KCNQ2/
with other subunits designated b, g, and d. KCNQ3 K + channels (Rundfeldt & Netzer, 2000). Muta-
Voltage-sensitive Ca2 + channels can be broadly classi- tions in the KCNQ2/KCNQ3 channels have been reported
fied into low or high threshold, according to the membrane in benign neonatal familial convulsions, a generalised epi-
potential at which they are activated (Hofmann et al., 1994). lepsy syndrome (Rogawski, 2000).
The low-threshold T-type Ca2 + channel is expressed pre-
dominantly in thalamocortical relay neurones, where it is 2.2. g-Aminobutyric acid-mediated inhibition
believed to be instrumental in the generation of the rhythmic
3-Hz spike-and-wave discharge that is characteristic of GABA is the predominant inhibitory neurotransmitter in
generalised absence seizures (Coulter et al., 1989a). High- the mammalian CNS, where it is released at up to 40% of all
threshold Ca2 + channels are subclassified by their pharma- synapses (Olsen & Avoli, 1997). Impairment of GABA
cological properties into L-, N-, P-, Q-, and R-types (Hof- function is widely recognised to provoke seizures, whereas
mann et al., 1994; Catterall, 1995; Dolphin, 1995). These facilitation has an anticonvulsant effect (Löscher, 1999).
channels are distributed throughout the nervous system on GABA is synthesised from glutamate, exclusively in
dendrites, cell bodies, and nerve terminals. The N-, P-, and GABAergic neurones, by the action of the enzyme glutamic
Q-type channels, in particular, have been implicated in the acid decarboxylase (GAD; Löscher, 1999). Upon synaptic
control of neurotransmitter release at the synapse (Stefani release, GABA acts on its three specific receptors, GABAA,
et al., 1997). GABAB, and the newly characterised GABAC. GABA
Interest in Ca2 + channels has heightened in recent years, receptors are distinguished by their pharmacology and
following the identification of subunit-specific genetic function (Johnston, 1996). The GABAA receptor belongs
mutations that can alter channel structure and/or function to the ligand-gated ion channel superfamily, and responds to
and that have been implicated in several human neurological GABA binding by increasing Cl  conductance, resulting in
diseases (Ophoff et al., 1998). Several AEDs have been neuronal hyperpolarisation (Rabow et al., 1995). GABAB
reported to block voltage-sensitive Ca2 + channels in a receptors are G-protein-linked, activation of which leads to
24 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

an increase in K + conductance (Olsen & Avoli, 1997). It and glial cells by the action of several specific transporters
has recently been proposed that GABAA and GABAB (Meldrum et al., 1999). Glial glutamate uptake is of princi-
receptors may have evolved from the GABAC receptor, pal importance. Glial cells convert glutamate into glutamine
which is comparatively simpler in structure and pharmacol- by the action of the enzyme glutamine synthetase. Gluta-
ogy (Johnston, 1996). mine is subsequently transferred to glutamatergic neurones,
Of the three receptor subtypes, the GABAA receptor, completing the cycle (Daikhin & Yudkoff, 2000).
with its pentameric subunit array and central Cl  ion pore, Like GABA receptors, ionotropic glutamate receptors are
is perhaps the best understood (Rabow et al., 1995). comprised of various combinations of subunits forming
GABAA receptors are composed of various combinations tetrameric and pentameric arrays. They are classified into
of 5 subunits, a(1 – 6), b(1 – 3), g(1 – 3), d, and e. Receptor three specific subtypes, a-amino-3-hydroxy-5-methyl-iso-
physiology, pharmacology, and distribution differs, depend- xazole-4-propionic acid (AMPA), kainate, and N-methyl-D-
ing on subunit composition (Mody, 1998). Theoretically, aspartate (NMDA), which form ligand-gated ion channels,
tens of thousands of different GABAA receptors could permeable to Na + and, depending on subtype and subunit
exist, but naturally only 10 or fewer combinations of composition, Ca2 + ions (Trist, 2000). The NMDA receptor
subtypes are encountered. is further distinguished by having glycine as a co-agonist.
Following receptor activation, GABA is removed from The AMPA and kainate subtypes of the glutamate receptor
the synaptic cleft into localised nerve terminals and glial are implicated in fast excitatory neurotransmission, whereas
cells, by specific membrane-bound transport molecules. the NMDA receptor, quiescent at resting membrane poten-
Currently, four active transport systems, GABA transporter tial, is recruited during periods of prolonged depolarisation
(GAT)-1, GAT-2, GAT-3, and betaine GAT (BGT)-1, have (Meldrum, 2000). The metabotropic family of glutamate
been described (Borden et al., 1992). GABA has a variable receptors, also classified into three distinct subtypes (Groups
affinity for these transporters, and only GAT-1, predomi- I, II, and III), are G-protein linked and predominantly
nantly located in the cerebral cortex and hippocampus, has presynaptic, possibly controlling neurotransmitter release
GABA as its principal substrate (Guastella et al., 1990). (Meldrum, 2000).
After removal from the synapse, GABA is either recycled to Although none of the commonly used AEDs exert their
the readily releasable neurotransmitter pool (GABAergic pharmacological effects solely by an action on the glutamate
nerve terminals only) or metabolised (neurones and glial system, blockade of ionotropic glutamate receptors is
cells) to the inactive molecule succinic acid semialdehyde believed to contribute to the antiepileptic activity of several
by the action of the mitochondrial enzyme GABA-trans- compounds (Upton, 1994; Macdonald & Kelly, 1995; Mel-
aminase (GABA-T; Meldrum, 1995). drum, 1996; White, 1999). In addition, several AEDs have
Several AEDs exert their effects, at least in part, by been reported to reduce glutamate release, although this
actions on the GABAergic system. Increased GABA syn- effect may be more indicative of their actions on neuronal
thesis, increased release, allosteric receptor facilitation, and Ca2 + channels than a direct effect on the glutamate system
reduced inactivation have all been implicated in the mech- (Stefani et al., 1997).
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 3. Modulation of ion channels by antiepileptic drugs
compounds (Löscher, 1998).
3.1. Phenytoin
2.3. Glutamate-mediated excitation
PHT was discovered following a search to identify a
Glutamate is the principal excitatory neurotransmitter in nonsedative analogue of PB (Fig. 1; Merritt & Putnam,
the mammalian brain (Meldrum, 2000). Focal injection of 1938). It has become a first-line treatment for partial and
glutamate induces seizures in animals, and over-activation generalised tonic-clonic seizures (Brodie & Dichter, 1996).
of glutamatergic transmission or abnormal glutamate recep- PHT is believed to exert its anticonvulsant effect pri-
tor properties are observed in certain experimental seizure marily by an action on voltage-dependent Na + channels
models and human epilepsy syndromes (Meldrum, 1995). (Tunnicliff, 1996), binding to the fast inactivated state of
Inhibition of the neuronal release of glutamate and blockade the channel and reducing the frequency of sustained repet-
of its receptors have received considerable attention in the itive firing of action potentials, without affecting their
search for novel AEDs (Meldrum, 2000). amplitude or duration (McLean & Macdonald, 1983).
Glutamate is synthesised from glutamine by the action of PHT inhibits high-frequency repetitive firing in a voltage-
the enzyme glutaminase in glutamatergic neurones (Daikhin dependent manner, with limitation of firing increased after
& Yudkoff, 2000). Following synaptic release, glutamate depolarisation and removed by hyperpolarisation (Schwartz
exerts its pharmacological effects on several receptors, & Grigat, 1989). Na + channel blockade with PHT is also
classified into ionotropic and metabotropic families. Gluta- frequency-dependent, being more effective at higher fre-
mate is removed from the synaptic cleft into nerve terminals quencies of neuronal stimulation. An effect on persistent
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 25

Fig. 1. Molecular structures of the established AEDs.

Na + currents has also been suggested (Lampl et al., 1998). tems have been reported. Whether these additional actions
The precise binding site for PHT on the Na + channel is contribute to the anticonvulsant effects of the drug is unclear.
unclear (Ragsdale et al., 1996), and there may be a
preferential effect on different subtypes of Na + channels 3.3. Lamotrigine
(Song et al., 1996).
PHT has also been reported to block high voltage- Lamotrigine (LTG; Fig. 2) is a new AED that was
activated Ca2 + channels (Schumacher et al., 1998), to developed as a result of a once-presumed link between
attenuate post-ictal glutamate release (Rowley et al., 1995) anticonvulsant and antifolate properties (Reynolds et al.,
and, paradoxically, to reduce K + currents (Nobile & 1966). It has proved to have a broad spectrum of activity,
Vercellino, 1997). There is further unsubstantiated evi- with efficacy for partial, absence, myoclonic, and tonic-
dence to suggest that PHT potentiates the action of GABA clonic seizures (Leach & Brodie, 1995). There is no
at specific molecular subtypes of the GABAA receptor evidence that the anticonvulsant activity of LTG is related
(Granger et al., 1995). to its weak antifolate effect (Macdonald & Kelly, 1995).
Instead, like PHT and CBZ, it inhibits sustained repetitive
3.2. Carbamazepine firing of action potentials (Cheung et al., 1992; Wang et al.,
1993) by blocking Na + channels in a voltage- and use-
CBZ is chemically related to the tricyclic antidepressants dependent manner (Cheung et al., 1992; Lang et al., 1993;
(Fig. 1). First introduced in 1963, it is widely used in the Zona & Avoli, 1997).
treatment of partial and generalised tonic-clonic seizures The broad clinical profile of LTG suggests that its effects
(Brodie & French, 2000). CBZ has been reported to stabilise on the Na + channel may differ from those observed with
the inactive form of the Na + channel in a voltage-, PHT and CBZ. Unlike PHT, LTG acts principally on the
frequency-, and time-dependent fashion (Courtney & Etter, slow inactivated state of the channel (Kuo & Lu, 1997). In
1983). Although similar to PHT in this respect, subtle addition, LTG may exhibit differential sensitivity for the
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
Ca2 + 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 Ca2 + channels or potentiates the actions of
GABA. However, effects on the serotonin (Dailey et al.,
1997a, 1997b) and adenosine (Marangos et al., 1983) sys- Fig. 2. Molecular structures of some new AEDs.
26 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

various a-subunits of the Na + channel (Coulter, 1997). al., 1995, 1997). Interestingly, unlike any other licensed
These subunits have a markedly different regional distribu- AED, OXC may additionally increase K + channel conduc-
tion in the brain (Catterall, 1992). Indeed, it has been tance (McLean et al., 1994).
suggested that LTG may selectively target Na + channels
on neurones that synthesise glutamate and aspartate (Leach 3.5. Ethosuximide
et al., 1986).
In addition to Na + channel effects, LTG reduces whole- ESM has been used in the treatment of generalised
cell Ca2 + currents in rat amygdalar neurones, possibly via absence seizures for over 30 years (Fig. 1; Brodie & Dichter,
the N- and P-type channels that have been implicated in 1997). It has no consistent efficacy for any other seizure
neurotransmitter release (Stefani et al., 1996b, 1997; Wang type. ESM exerts its anti-absence effects by reducing T-type
et al., 1996). This effect might explain the inhibition of Ca2 + currents in thalamocortical relay neurones (Coulter et
electrically stimulated glutamate release from rat spinal al., 1989b, 1989c). As discussed in Section 2.1.2, the low-
dorsal horn slices observed with the drug (Teoh et al., threshold T-type Ca2 + channel predominates in these neuro-
1995). It remains possible, however, that LTG possesses nes, where it is believed to play a fundamental role in the
additional unidentified mechanisms that confer its relatively generation of the characteristic 3-Hz spike-and-wave dis-
broader clinical spectrum when compared with other Na + charge of absence epilepsy (Coulter et al., 1989a). ESM
channel-blocking agents. blocks the T-type Ca2 + channel and prevents synchronised
firing. It has no other known mechanisms of action (Rogaw-
3.4. Oxcarbazepine ski & Porter, 1990).

Oxcarbazepine (OXC; Fig. 3) is a relatively novel AED, 3.6. Zonisamide


with widespread geographical approval for clinical use
(Tecoma, 1999). It is closely related to CBZ in structure. The development of zonisamide (ZNS; Fig. 3) was, until
The keto substitutions at the 10 and 11 positions of the recently, suspended following observations linking the drug
dibenzazepine nucleus do not affect the therapeutic profile to an increased incidence of renal calculi (Leppik et al.,
of the drug when compared with CBZ, but result in altered 1993; Leppik, 1999). However, long-term experience in
biotransformation and better tolerability (White, 1999). The Japan, where the drug is licensed, failed to substantiate
structural modifications circumvent the 10,11-epoxide these concerns. Clinical evidence to date suggests that ZNS
metabolite of CBZ that is believed to be responsible for is effective against partial and generalised seizures, and has
many of its side effects and its ability to induce cytochrome particular efficacy in the progressive myoclonic epilepsies
P450-dependent hepatic metabolism (Tecoma, 1999). OXC that are often resistant to AED treatment (Dichter & Brodie,
is essentially a pro-drug, and is rapidly and completely 1996; Kyllerman & Ben-Menachem, 1998).
reduced in the liver to its active metabolite, the monohy- The principal pharmacological action of ZNS involves
droxy derivative (10,11-dihydro-10-hydroxy CBZ; Edito- modulation of voltage-dependent ion channels. Like LTG,
rial, 1989). ZNS enhances slow Na + channel inactivation (Schauf,
In terms of mechanisms of action, OXC appears to exert 1987) and reduces sustained repetitive firing in spinal cord
its pharmacological effects by blockade of voltage-depend- neurones (Rock et al., 1989). It also blocks low-threshold T-
ent Na + channels in a manner similar to that reported for type Ca2 + currents, which may account for its anti-absence
PHT and CBZ (McLean et al., 1994). It also reduces effects (Suzuki et al., 1992).
presynaptic glutamate release, possibly by blocking high- ZNS also inhibits carbonic anhydrase, although this
threshold Ca2 + currents (Calabresi et al., 1995; Stefani et 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 GABAA
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
Fig. 3. Molecular structures of remaining new AEDs. their sedative, anaesthetic, and anticonvulsant properties.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 27

PB (Fig. 1) is still commonly prescribed worldwide for of tolerance to their pharmacological effects has, however,
epilepsy, although its cognitive and behavioural side effects restricted their use in chronic treatment regimens (Brodie &
have limited its use, particularly in the developed world Dichter, 1996).
(Mattson et al., 1985; Brodie & Dichter, 1997). The use of The BZDs bind to a specific site in the brain (Squires &
primidone (PRM; Fig. 1) has been similarly restricted. It is Braestrup, 1977), now recognised as the a-subunit of the
largely metabolised to PB, although there is evidence to GABAA receptor (Macdonald & Kelly, 1995). Binding
suggest that it possesses an additional active metabolite. results in allosteric activation of the receptor, increasing
Variations in the experimental anticonvulsant profiles of PB the frequency of Cl  channel opening, without affecting
and PRM (Bourgeois et al., 1983), and the relatively greater open duration or channel conductance (Study & Barker,
toxicity of PRM (Mattson et al., 1985), may be conferred by 1981; Twyman et al., 1989). Unlike the barbiturates, the
this metabolite. BZDs are unable to directly activate the GABAA receptor in
PB exerts its pharmacological effects by allosteric acti- the absence of GABA (White, 1999).
vation of the GABAA receptor, increasing the duration of Augmentation of GABAergic inhibition in the thalamus
Cl  channel opening, without affecting the frequency of can result in the de-inactivation of T-type Ca2 + channels,
opening or channel conductance (Macdonald et al., 1989; triggering a strong low-threshold burst and enhancing
Twyman et al., 1989). The barbiturates can also activate the development of the thalamocortical rhythmicity that is
GABAA receptor directly, in the absence of GABA, an characteristic of absence seizures (Coulter, 1997). Some
effect that may underlie their sedative properties (Rho et al., GABAergic AEDs accordingly exacerbate absence seizures;
1996; White, 1999). Additional mechanisms of barbiturate others are without effect, while the BZDs are notable by
action have been reported, including blockade of high- their efficacy (Dichter & Brodie, 1996).
voltage-activated Ca2 + channels (Rogawski & Porter, This phenomenon is explained by selective augmentation
1990) and an inhibitory effect on the AMPA/kainate subtype of GABAA-mediated inhibition within the nucleus reticula-
of glutamate receptor (Davies, 1995). 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
4.2. Benzodiazepines a-subunits of the GABAA receptor between the nucleus
reticularis thalami and other thalamic nuclei may underlie
More than 50 chemically distinct BZDs are marketed this apparent contradiction (Coulter, 1997).
worldwide. Diazepam, lorazepam, clobazam, and clonaze-
pam are those most commonly used as AEDs (Fig. 4; Brodie 4.3. Vigabatrin
& Dichter, 1996; Dichter & Brodie, 1996). The antiepileptic
BZDs have a broad spectrum of clinical activity, with In 1989, vigabatrin (VGB; Fig. 2) became the first of the
efficacy in the partial and idiopathic generalised epilepsies new generation of AEDs to be licensed in the United
(Dichter & Brodie, 1996) and for the acute treatment of Kingdom. It was initially approved as adjunctive therapy
status epilepticus (Treiman et al., 1998). The development 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
(Kälviäinen 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
Fig. 4. Structures of BDZ compounds commonly used in epilepsy. of new enzyme protein over a period of 4 – 5 days
28 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

(Schechter et al., 1977). Similar effects are also observed in VPA has been reported to block voltage-dependent Na +
humans (Petroff et al., 1996b; Petroff & Rothman, 1998). channels. It reduces sustained repetitive firing of mouse
Preliminary evidence suggests that, in addition to an effect neurones in culture (McLean & Macdonald, 1986), inhibits
on GABA-T, VGB may block the uptake of GABA into Na + channels in Xenopus leavis myelinated neurones (van
glial cells (Leach et al., 1996). Dongen et al., 1986), and reduces Na + currents in neo-
cortical neurones (Zona & Avoli, 1990). However, rat
4.4. Tiagabine hippocampal slice studies suggest that, unlike PHT and
CBZ, VPA has no effect on the recovery of Na + channels
Tiagabine (TGB) is a novel AED, recently licensed from the inactivated state (Albus & Williamson, 1998).
widely for the adjunctive treatment of partial seizures with VPA may also block T-type Ca2 + channels in a manner
or without secondary generalisation (Fig. 3; Leach & similar to that reported for ESM. Such an effect would
Brodie, 1998). It is an analogue of nipecotic acid, a proto- explain its efficacy against generalised absence seizures.
typic GABA uptake blocker, which is widely recognised to However, the reduction of T-type Ca2 + currents observed
prevent GABA transport into both neurones and glial cells with VPA in rat primary afferent neurones is modest and
(Krogsgaard-Larsen et al., 1987). Nipecotic acid, however, requires relatively high drug concentrations (Kelly et al.,
fails to cross the blood-brain barrier following systemic 1990). In addition, VPA appears to have no effect on Ca2 +
administration. This problem is overcome by linking it to a channel conductance in rat thalamic neurones (Coulter
lipophilic anchor to form TGB, which is able to cross the et al., 1989c).
blood-brain barrier more readily (Brodie, 1995). There is evidence to suggest that VPA elevates whole
TGB inhibits GABA uptake into synaptosomal mem- brain GABA levels and potentiates GABA responses,
branes, neurones, and glial cells. It has a greater affinity possibly by enhancing GAD activity and inhibiting
(2.5-fold) for glial than for neuronal uptake (Braestrup et al., GABA degradation (Löscher, 1999). Anecdotal reports
1990). TGB has a selective action on the GAT-1 GABA suggest that the drug also augments GABA release
transporter, with little or no activity on GAT-2, GAT-3, or (Rowley et al., 1995) and blocks GABA uptake (Sills et
BGT-1 (Borden et al., 1994). As a result, the pharmaco- al., 1996). The reproducibility of these effects has, how-
logical effect of TGB reflects the regional distribution of ever, been questioned (Rogawski & Porter, 1990). It is
GAT-1, and is mainly restricted to the cerebral cortex and suggested that the GABAergic effects of VPA exhibit a
hippocampus (Meldrum & Chapman, 1999). degree of regional specificity within the brain and that
Unlike VGB, which elevates whole brain GABA con- inconsistent results reflect the resolution of individual
centrations, TGB temporarily prolongs the presence of studies (Löscher, 1999).
GABA in the synaptic cleft (Sills et al., 1999). This Single doses of VPA decrease brain levels of the
important distinction has been confirmed by in vivo micro- excitatory amino acid aspartate, without influencing those
dialysis studies in rats (Fink-Jensen et al., 1992) and in the of glutamate or GABA (Schechter et al., 1978). Decreases
hippocampus of a patient with refractory partial epilepsy in aspartate concentration have been shown to correspond
(During et al., 1992). In rat hippocampal slices, TGB pro- with the period of anticonvulsant activity in animal models
longs the duration, but not the magnitude, of the peak (Chapman et al., 1983). The relative anticonvulsant poten-
inhibitory postsynaptic current, consistent with delayed cies of a series of valproate analogues also correlates more
clearance of GABA from the synapse (Roepstorff & Lam- closely with their ability to reduce brain aspartate levels
bert, 1992). Inhibition of GABA uptake is the only known than with their effects on GABA concentration (Chapman
mechanism of TGB action (Brodie, 1995). et al., 1984). The potential contribution of any of the
above effects to the clinical activity of VPA remains to
be determined.
5. Antiepileptic drugs with multiple mechanisms
of action 5.2. Gabapentin

5.1. Sodium valproate Gabapentin (GBP) is a novel compound, structurally


related to GABA, which is effective in the adjunctive
The antiepileptic properties of VPA were discovered treatment of partial seizures, with or without secondary
serendipitously when valproic acid was employed in animal generalisation (Fig. 2; Dichter & Brodie, 1996). It was
studies as a solvent for drugs under formal investigation originally designed as a GABAmimetic that could freely
(Fig. 1; Meunier et al., 1963). VPA has since proved to be cross the blood-brain barrier. However, subsequent studies
an extremely useful AED, with a broad spectrum of activity have shown that GBP does not directly interact with GABA
and particular efficacy in the generalised epilepsies (Edito- receptors (Taylor et al., 1998) or transporters (Su et al.,
rial, 1988; Brodie & Dichter, 1996). However, the precise 1995; Macdonald & Greenfield, 1997).
mechanisms by which it exerts its antiepileptic effects There is evidence that GBP may increase the synthesis
remain to be conclusively determined. (Taylor et al., 1992) and nonvesicular release of GABA
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 29

(Gotz et al., 1993), and may prevent its metabolism (Leach FBM facilitates the effects of diazepam against exper-
et al., 1997). Using 1H NMR spectroscopy, GBP has been imentally induced seizures, implying an additional action at
shown to elevate GABA concentrations in the occipital the GABAA receptor (Gordon et al., 1991). Early studies
cortex of epileptic patients (Petroff et al., 1996a). Whether failed to demonstrate an effect of FBM on ligand binding to
this observation is the result of enhanced synthesis, the GABA, BZD, or picrotoxin recognition sites on the
increased release, or reduced metabolism of GABA remains receptor complex (Ticku et al., 1991). However, subsequent
to be determined. investigations suggested that, in addition to inhibiting
Early efforts to identify the mechanism of action of GBP NMDA responses, FBM potentiated GABAA receptor cur-
proposed an interaction with the L-amino acid transport rents (Rho et al., 1994). The precise binding site for FBM on
system, resulting in alterations in the cytosolic and extrac- the GABAA receptor remains to be identified, although a
ellular concentrations of several amino acids, including L- weak interaction with the barbiturate recognition site has
leucine, L-valine, and L-phenylalanine (Su et al., 1995). been mooted (Rho et al., 1997).
Inhibition of branched chain amino acid transferase and Additional actions of FBM on voltage-dependent ion
augmentation of glutamate dehydrogenase activities have channels have been reported. It reduces voltage-dependent
also been mooted (Goldlust et al., 1995). Investigations of Na + currents in striatal neurones (Pisani et al., 1995) and
Na + channel function suggest that GBP does not affect stabilises the inactive state of the a-subunit of Na +
sustained repetitive firing in spinal cord neurones (Taylor et channels transiently expressed in Xenopus oocytes (Taglia-
al., 1988), although prolonged exposure reduces high-fre- latela et al., 1996). In cortical and neostriatal neurones,
quency action potential firing in central neurones (Wamil & FBM reduces high-voltage-activated Ca2 + channels, an
McLean, 1994). effect reversed by the L-type channel antagonist nifedipine,
The identification of a specific binding site for GBP in but not by N- or P-type channel blockers (Stefani et al.,
the mammalian brain, and its subsequent unveiling as the 1996a). The effects of FBM on voltage-activated ion
a2d-subunit of the L-type voltage-dependent Ca2 + channel channels may explain the decrease in veratridine-induced
(Gee et al., 1996), suggested another potential pharmaco- release of glutamate observed with the drug (Srinivasan
logical mechanism. The implications of these findings et al., 1996).
remain to be fully investigated, but the lack of effect of
GBP on whole cell Ca2 + currents in human dentate granule 5.4. Topiramate
cells acutely isolated from patients with temporal lobe
epilepsy (Schumacher et al., 1998) questions the pharmaco- The sulfamate derivative topiramate (TPM; Fig. 2) is
logical relevance of this finding. active against partial-onset and generalised seizures in
humans (Wilson & Brodie, 1996; Brodie & French,
5.3. Felbamate 2000). TPM has multiple mechanisms of action, including
inhibition of Na + and Ca2 + currents, blockade of the
Felbamate (FBM; Fig. 2) was licensed in the United AMPA/kainate subtype of glutamate receptor, and facilita-
States in 1993 as monotherapy and add-on treatment for tion of GABA effects at the GABAA receptor. TPM also
partial-onset and primary generalised tonic-clonic seizures inhibits carbonic anhydrase, although, like ZNS, this effect
in adults and for children with Lennox– Gastaut syndrome is not believed to contribute to its antiepileptic action
(Dichter & Brodie, 1996). Its initial clinical success was (Shank et al., 1994).
tempered by the significant incidence of aplastic anaemia TPM reduces the number of action potentials within a
and hepatotoxicity revealed in postmarketing surveillance burst and decreases burst duration in cultured hippocampal
(Pellock & Brodie, 1997). neurones (DeLorenzo et al., 2000). It also reduces voltage-
FBM is believed to be the first effective AED with a dependent Na + currents in cultured cerebellar granule cells
direct action on the NMDA subtype of glutamate receptor. It (Zona et al., 1996a, 1996b). In this respect, its action is
inhibits NMDA/glycine-stimulated increases in intracellular similar to that of PHT and CBZ, delaying recovery of Na +
Ca2 + (Taylor et al., 1995), reduces inward currents evoked channels from the inactive state. It has also been reported
by NMDA application to striatal neurones (Pisani et al., that TPM blocks whole cell Ca2 + currents, possibly by an
1995), and blocks NMDA receptor-mediated excitatory action on the high-voltage-activated L-type channel (Zhang
postsynaptic potentials (Pugliese & Corradetti, 1996). Con- et al., 2000).
siderable evidence suggests that FBM interacts with the In hippocampal neurones, TPM blocks inward currents
strychnine-insensitive glycine recognition site on the evoked by kainate, but not NMDA (Gibbs et al., 2000),
NMDA receptor complex (White et al., 1995b). FBM implying a selective effect on the AMPA/kainate subtype of
inhibits the binding of high-affinity glycine antagonists at the glutamate receptor. TPM also interacts with the GABAA
this site (McCabe et al., 1993), and its anticonvulsant effects receptor. It enhances GABA-stimulated Cl  flux into
in several experimental models are blocked by glycine and cerebellar granule neurones and increases Cl  currents
synthetic glycine site compounds (Coffin et al., 1994; White evoked by GABA in mouse cerebral cortical neurones
et al., 1995b). (White et al., 1997). The GABAA receptor effects of TPM
30 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

are not influenced by flumazenil, suggesting that they do not 7. Conclusions


involve the BZD recognition site on the receptor complex
(White et al., 1995a). Finally, recent evidence suggests that For the purposes of this review, it seemed prudent to
brain concentrations of GABA, and its metabolites, are categorise the currently used AEDs according to their
elevated by TPM in patients with refractory epilepsy (Petr- principal mechanisms of action. However, it is becoming
off et al., 1999). These effects, however, are not reproduced apparent that most, if not all, have multiple cellular effects.
in experimental animals (Sills et al., 2000). 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
6. Antiepileptic drugs with unknown mechanisms a diverse range of pharmacological actions.
of action The immediate challenge is to establish the relative
importance of individual mechanisms to the overall anti-
6.1. Levetiracetam epileptic effect of each drug. The use of AEDs with known
modes of action has the potential to promote the under-
Levetiracetam (LEV; Fig. 3) is the S-enantiomer of the standing of the underlying pathophysiology of seizure dis-
ethyl analogue of piracetam, a widely used nootropic agent orders and, in turn, to provide a framework for future
in the elderly (Löscher & Hönack, 1993). As the most targetted drug development.
recently licensed AED, clinical experience with LEV is Despite familiarity with established AEDs and the intro-
limited (Genton & Van Vleymen, 2000). Results of clinical duction of nine new agents in the past decade, up to one-third
trials suggest that it is effective against partial seizures with of epilepsy patients remain resistant to optimum drug treat-
or without secondary generalisation (Bialer et al., 1999). ment (Kwan & Brodie, 2000). Identifying the ‘‘mechanisms
However, more extensive investigation of LEV is required that matter’’ has the potential to address the problem of
before its full spectrum of clinical activity is revealed. refractory epilepsy, at least in part. It is equally fundamental
LEV appears to have a unique mode of action that, at to the development of a rational basis for future pharmaco-
this time, remains to be clearly characterised. Exhaustive therapy, centred around AED mechanisms of action and
preclinical investigations suggest that it does not interact tailored to the individual patient (Brodie et al., 1997).
directly with any of the traditional targets, including Na + ,
Ca2 + , and K + channels or the GABA and glutamate
neurotransmitter systems (Noyer et al., 1995). It is believed References
to bind to a specific, as yet unidentified, site on the
synaptic plasma membrane. Competitive binding studies Albus, H., & Williamson, R. (1998). Electrophysiologic analysis of the
suggest that LEV is not displaced by CBZ, PHT, VPA, PB, actions of valproate on pyramidal neurons in the rat hippocampal slices.
clonazepam, picrotoxin, or bicuculline, but does interact Epilepsia 39, 124 – 139.
Appleton, R. E., Peters, A. C. B., Mumford, J. P., & Shaw, D. E. (1999).
with ESM, pentylenetetrazol, and bemegride at this site Randomised, placebo-controlled study of vigabatrin as first-line treat-
(Noyer et al., 1995). The implications of these observations ment of infantile spasms. Epilepsia 40, 1627 – 1633.
are unclear. Barchi, L. (1998). Ion channel mutations affecting muscle and brain. Curr
LEV has no effect on whole brain GABA synthesis or Opin Neurol 11, 461 – 468.
metabolism, or on the concentrations of GABA, glutamate, Bialer, M., Johannessen, S. I., Kupferberg, H. J., Levy, R. H., Loiseau, P., &
Perucca, E. (1999). Progress report on new antiepileptic drugs: a sum-
or glutamine (Sills et al., 1997). These studies, however, did mary of the fourth Eilat conference. Epilepsy Res 34, 1 – 41.
not discount the possibility of regionally specific effects on Birnstiel, S., Wulfert, E., & Beck, S. G. (1997). Levetiracetam (ucb L059)
the GABA system. Accordingly, LEV reduces GABA turn- affects in vitro models of epilepsy in CA3 pyramidal neurons without
over in the striatum of the rat by increasing GABA-T altering normal synaptic transmission. Naunyn Schmiedebergs Arch
Pharmacol 356, 611 – 618.
activity and reducing GAD activity (Löscher et al., 1996).
Borden, L. A., Smith, K. E., Hartig, P. R., Branchek, T. A., & Weinshank,
These modest effects are accompanied by a decrease in R. L. (1992). Molecular heterogeneity of the g-aminobutyric acid (GA-
spontaneous neuronal firing of the substantia nigra pars BA) transport system. J Biol Chem 267, 21098 – 21104.
reticulata, which receives a strong GABAergic input from Borden, L. A., Dhar, T. G. M., Smith, K. E., Weinshank, R. L., Branchek,
the striatum (Löscher et al., 1996). T. A., & Gluchowski, C. (1994). Tiagabine, SKF 89976-A, CI-966,
Other preclinical studies suggest that LEV attenuates and NNC-711 are selective for cloned GABA transporter GAT-1. Eur
J Pharmacol 269, 219 – 224.
bicuculline-induced increases in neuronal excitability in Bourgeois, B. F. D., Dodson, W. E., & Ferrendelli, J. A. (1983). Primi-
the CA3 region of the rat hippocampus (Margineanu & done, phenobarbital and PEMA: I. Seizure protection, neurotoxicity
Wulfert, 1995), possibly by blockade of T-type Ca2 + and therapeutic index of individual compounds in mice. Neurology 33,
channels (Margineanu & Wulfert, 1997). Suppression of 283 – 290.
NMDA-induced bursting has also been reported (Birnstiel Braestrup, C., Nielsen, G. B., Sonnewald, U., Knutsen, L. J. S., Andersen,
K. E., Jansen, J. A., Frederiksen, K., Andersen, P. H., Mortensen, A.,
et al., 1997). The relative contribution of any, or all, of the & Suzdak, P. D. (1990). (R)-N-[4,4-bis(3-methyl-2-thienyl)but-3-en-1-
anecdotal effects described above to the antiepileptic action yl]nipecotic acid binds with high affinity to the brain g-aminobutyric
of LEV remains to be substantiated. acid uptake carrier. J Neurochem 54, 639 – 647.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 31

Brodie, M. J. (1995). Tiagabine pharmacology in profile. Epilepsia 36 Dichter, M. A., & Brodie, M. J. (1996). New antiepileptic drugs. N Engl J
(suppl. 6), S7 – S9. Med 334, 1583 – 1590.
Brodie, M. J. (1999). Monostars: an aid to choosing an antiepileptic drug as Dolphin, A. C. (1995). Voltage-dependent calcium channels and their mod-
monotherapy. Epilepsa 40 (suppl. 6), S17 – S22. ulation by neurotransmitters and G proteins. Exp Physiol 80, 1 – 36.
Brodie, M. J., & Dichter, M. A. (1996). Antiepileptic drugs. N Engl J Med During, M. J., Mattson, R. H., Scheyer, R., Rask, C., Pierce, M.,
334, 168 – 175. McKelvy, J., & Thomas, V. (1992). The effect of tiagabine HCl on
Brodie, M. J., & Dichter, M. A. (1997). Established antiepileptic drugs. extracellular GABA levels in the human hippocampus. Epilepsia 33
Seizure 6, 159 – 174. (suppl. 3), S83.
Brodie, M. J., & French, J. A. (2000). Management of epilepsy in adoles- Editorial (1988). Sodium valproate. Lancet ii, 1229 – 1231.
cents and adults. Lancet 356, 323 – 329. Editorial (1989). Oxcarbazepine. Lancet ii, 196 – 198.
Brodie, M. J., & Kwan, P. (2000). The ‘‘star’’ systems: overview and Fink-Jensen, A., Suzdak, P. D., Swedberg, M. D., Judge, M. E., Hansen, L.,
use in determining drug choice for patients with epilepsy. CNS Drugs, & Nielsen, P. G. (1992). The g-aminobutyric acid (GABA) uptake
in press. inhibitor, tiagabine, increases extracellular brain levels of GABA in
Brodie, M. J., & Yuen, A. W. C. The 105 Study Group (1997). Lamotrigine awake rats. Eur J Pharmacol 220, 197 – 201.
substitution study—evidence for synergism with sodium valproate? Gandolfo, G., Romettino, S., Gottesman, C., Van Luijtelaer, G., Coenen, A.,
Epilepsy Res 26, 423 – 432. Bidard, J-N., & Lazdunski, M. (1989). K + channel openers decrease
Calabresi, P., De Murtas, M., Stefani, A., Pisani, A., Sancesario, G., Mer- seizures in genetically epileptic rats. Eur J Pharmacol 167, 181 – 183.
curi, N. B., & Bernardi, G. (1995). Action of GP 47779, the active Gee, N. S., Brown, J. P., Dissanayake, V. U. K., Offord, J., Thurlow, R.,
metabolite of oxcarbazepine, on the corticostriatal system I. Modulation & Woodruff, G. N. (1996). The novel anticonvulsant drug, gabapentin
of corticostriatal synaptic transmission. Epilepsia 36, 990 – 996. (Neurontin), binds to the a2d subunit of a calcium channel. J Biol
Catterall, W. A. (1992). Cellular and molecular biology of voltage-gated Chem 271, 5768 – 5776.
sodium channels. Physiol Rev 72, S15 – S48. Genton, P., & Van Vleymen, B. (2000). Piracetam and levetiracetam: close
Catterall, W. A. (1995). Structure and function of voltage-gated ion chan- structural similarities but different pharmacological and clinical profiles.
nels. Annu Rev Biochem 64, 493 – 531. Epileptic Disord 2, 99 – 105.
Chapman, A. G., Meldrum, B. S., & Mendes, E. (1983). Acute anticonvul- Gibbs, J. W., Sombati, S., III, DeLorenzo, R. J., & Coulter, D. A. (2000).
sant activity of structural analogues of valproic acid and changes in Cellular actions of topiramate: blockade of kainate-evoked inward
brain GABA and aspartate content. Life Sci 32, 2023 – 2031. currents in cultured hippocampal neurons. Epilepsia 41 (suppl. 1),
Chapman, A. G., Croucher, M. J., & Meldrum, B. S. (1984). Anticon- S10 – S16.
vulsant activity of intracerebroventricularly administered valproate and Goldlust, A., Su, T., Welty, D. F., Taylor, C. P., & Oxender, D. L. (1995).
valproate analogues. A dose-dependent correlation with changes in Effects of the anticonvulsant drug gabapentin on enzymes in the meta-
brain aspartate and GABA levels in DBA/2 mice. Biochem Pharma- bolic pathways of glutamate and GABA. Epilepsy Res 22, 1 – 11.
col 33, 1459 – 1463. Gordon, R., Gels, M., Diamantis, W., & Sofia, R. D. (1991). Interaction of
Cheung, H., Kamp, D., & Harris, E. (1992). An in vitro investigation of the felbamate and diazepam against maximal electroshock seizures and
action of lamotrigine on neuronal voltage-activated sodium channels. chemoconvulsants in mice. Pharmacol Biochem Behav 40, 109 – 113.
Epilepsy Res 13, 107 – 112. Gotz, E., Feuerstein, T. J., & Meyer, D. K. (1993). Effects of gabapentin on
Coffin, V., Cohen-Williams, M., & Barnett, A. (1994). Selective antago- release of g-aminobutyric acid from slices of rat neostriatum. Drug Res
nism of the anticonvulsant effects of felbamate by glycine. Eur J Phar- 43, 636 – 638.
macol 256, R9 – R10. Granger, P., Biton, B., Faure, C., Vige, X., Depoortere, H., Graham, D.,
Coulter, D. A. (1997). Antiepileptic drug cellular mechanisms of action: Langer, S. Z., Scatton, B., & Avenet, P. (1995). Modulation of the g-
where does lamotrigine fit in? J Child Neurol 12 (suppl. 1), S2 – S9. aminobutyric acid type A receptor by the antiepileptic drugs carbama-
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989a). Calcium currents zepine and phenytoin. Mol Pharmacol 47, 1189 – 1196.
in rat thalamocortical relay neurones: kinetic properties of the transient Guastella, J., Nelson, N., Nelson, H., Czyzyk, L., Keynan, S., & Miedel,
low-threshold current. J Physiol 414, 587 – 604. M. C. (1990). Cloning and expression of a rat brain GABA trans-
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989b). Specific petit porter. Science 249, 1303 – 1306.
mal anticonvulsants reduce calcium currents in thalamic neurons. Neu- Hofmann, F., Biel, M., & Flockerzi, V. (1994). Molecular basis for Ca2 +
rosci Lett 98, 74 – 78. channel diversity. Annu Rev Neurosci 17, 399 – 418.
Coulter, D. A., Huguenard, J. R., & Prince, D. A. (1989c). Characterization Hough, C. J., Irwin, R. P., Gao, X. M., Rogawski, M. A., & Chuang, D. M.
of ethosuximide reduction of low-threshold calcium currents in thalamic (1996). Carbamazepine inhibition of N-methyl-D-aspartate-evoked cal-
neurons. Ann Neurol 25, 582 – 593. cium influx in rat cerebellar granule cells. J Pharmacol Exp Ther 276,
Courtney, K. R., & Etter, E. F. (1983). Modulated anticonvulsant block of 143 – 149.
sodium channels in nerve and muscle. Eur J Pharmacol 88, 1 – 9. Johnston, G. A. R. (1996). GABAA receptor pharmacology. Pharmacol
Daikhin, Y., & Yudkoff, M. (2000). Compartmentation of brain glutamate Ther 69, 173 – 198.
metabolism in neurons and glia. J Nutr 130, 1026S – 1031S. Jung, M. J., Lippert, B., Metcalf, B., Böhlen, P., & Schechter, P. J. (1977).
Dailey, J. W., Reith, M. E., Yan, Q. S., Li, M. Y., & Jobe, P. C. (1997a). g-Vinyl GABA (4-amino-hex-5-enoic acid), a new irreversible inhibitor
Anticonvulsant doses of carbamazepine increase hippocampal extracel- of GABA-T: effects on brain GABA metabolism in mice. J Neurochem
lular serotonin in genetically epilepsy-prone rats: dose response rela- 29, 797 – 802.
tionships. Neurosci Lett 227, 13 – 16. Kälviäinen, R., Nousiainen, I., Mäntyjärvi, M., Nikoskelainen, E., Parta-
Dailey, J. W., Reith, M. E., Yan, Q. S., Li, M. Y., & Jobe, P. C. (1997b). nen, J., Partanen, K., & Riekkinen, P., Sr. (1999). Vigabatrin, a ga-
Carbamazepine increases extracellular serotonin concentration: lack baergic antiepileptic drug, causes concentric visual field defects.
of antagonism by tetrodotoxin or zero Ca2 + . Eur J Pharmacol 328, Neurology 53, 922 – 926.
153 – 162. Kawata, Y., Okada, M., Murakami, T., Mizuno, K., Wada, K., Kondo, T., &
Davies, J. A. (1995). Mechanisms of action of antiepileptic drugs. Seizure Kaneko, S. (1999). Effects of zonisamide on K + and Ca2 + evoked
4, 267 – 272. release of monoamine as well as K + evoked intracellular Ca2 + mobi-
DeLorenzo, R. J., Sombati, S., & Coulter, D. A. (2000). Effects of top- lization in rat hippocampus. Epilepsy Res 35, 173 – 182.
iramate on sustained repetitive firing and spontaneous recurrent seizure Kelly, K. M., Gross, R. A., & Macdonald, R. L. (1990). Valproic acid
discharges in cultured hippocampal neurons. Epilepsia 41 (suppl. 1), selectively reduces the low-threshold (T) calcium current in rat nodose
S40 – S44. neurons. Neurosci Lett 116, 233 – 238.
32 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

Krogsgaard-Larsen, P., Falch, E., Larsson, O. M., & Schousboe, A. (1987). reduces bicuculline-induced hyperexcitability in rat hippocampal CA3
GABA uptake inhibitors: relevance to antiepileptic drug research. Epi- in vivo. Eur J Pharmacol 286, 321 – 325.
lepsy Res 1, 77 – 93. Margineanu, D. G., & Wulfert, E. (1997). Inhibition by levetiracetam of a
Kuo, C. C., & Lu, L. (1997). Characterization of lamotrigine inhibition of non-GABAA receptor-associated epileptiform effect of bicuculline in rat
Na + channels in rat hippocampal neurones. Br J Pharmacol 121, hippocampus. Br J Pharmacol 122, 1146 – 1150.
1231 – 1238. Mattson, R. H., Cramer, J. A., Collins, J. F., Smith, D. B., Delgado-
Kuo, C. C., Chen, R. S., Lu, L., & Chen, R. C. (1997). Carbamazepine Escueta, A. V., Browne, T. R., Williamson, P. D., Treiman, D. M.,
inhibition of neuronal Na + currents: quantitative distinction from phe- McNamara, J. O., McCutchen, C. B., Homan, R. W., Crill, W. E.,
nytoin and possible therapeutic implications. Mol Pharmacol 51, Lubozynski, M. F., Rosenthal, N. P., & Mayersdorf, A. (1985). Com-
1077 – 1083. parison of carbamazepine, phenobarbital, phenytoin and primidone in
Kwan, P., & Brodie, M. J. (2000). Early identification of refractory epi- partial and secondary generalised tonic-clonic seizures. N Engl J Med
lepsy. N Engl J Med 342, 314 – 319. 313, 145 – 151.
Kyllerman, M., & Ben-Menachem, E. (1998). Zonisamide for progressive McCabe, R. T., Wasterlain, C. G., Kucharczyk, N., Sofia, R. D., & Vogel,
myoclonus epilepsy: long-term observations in seven patients. Epilepsy J. R. (1993). Evidence of anticonvulsant and neuroprotective action of
Res 29, 109 – 114. felbamate mediated by strychnine-insensitive glycine receptors. J Phar-
Lampl, I., Schwindt, P., & Crill, W. (1998). Reduction of cortical pyramidal macol Exp Ther 264, 248 – 252.
neuron excitability by the action of phenytoin on persistent Na + cur- McLean, M. J., & Macdonald, R. L. (1983). Multiple actions of phenytoin
rent. J Pharmacol Exp Ther 284, 228 – 237. on mouse spinal cord neurons in cell culture. J Pharmacol Exp Ther
Lang, D. G., Wang, C. M., & Cooper, B. R. (1993). Lamotrigine, phenytoin 227, 779 – 789.
and carbamazepine interactions on the sodium current present in McLean, M. J., & Macdonald, R. L. (1986). Sodium valproate, but not
N4TG1 mouse neuroblastoma cells. J Pharmacol Exp Ther 266, ethosuximide, produces use- and voltage-dependent limitation of high
829 – 835. frequency repetitive firing of action potentials of mouse central neurons
Leach, J. P., & Brodie, M. J. (1995). Lamotrigine: clinical use. In R. Levy, in cell culture. J Pharmacol Exp Ther 237, 1001 – 1011.
R. H. Mattson, B. S. Meldrum, & J. K. Penry (Eds.), Antiepileptic McLean, M. J., Schmutz, M., Wamil, A. W., Olpe, H. R., Portet, C., &
Drugs, 4th edn. ( pp. 889 – 895). New York: Raven Press. Feldmann, K. F. (1994). Oxcarbazepine: mechanisms of action. Epilep-
Leach, J. P., & Brodie, M. J. (1998). Tiagabine. Lancet 351, 203 – 207. sia 35 (suppl. 3), S5 – S9.
Leach, J. P., Sills, G. J., Majid, A., Butler, E., Carswell, A., Thompson, Meldrum, B. S. (1995). Neurotransmission in epilepsy. Epilepsia 36
G. G., & Brodie, M. J. (1996). Effects of tiagabine and vigabatrin on (suppl. 1), S30 – S35.
GABA uptake into rat cortical astrocytes in primary culture. Seizure 5, Meldrum, B. S. (1996). Update on the mechanism of action of antiepileptic
229 – 234. drugs. Epilepsia 37 (suppl. 6), S4 – S11.
Leach, J. P., Sills, G. J., Butler, E., Forrest, G., Thompson, G. G., & Brodie, Meldrum, B. S. (2000). Glutamate as a neurotransmitter in the brain: review
M. J. (1997). Neurochemical actions of gabapentin in mouse brain. of physiology and pathology. J Nutr 130, 1007S – 1015S.
Epilepsy Res 27, 175 – 180. Meldrum, B. S., & Chapman, A. G. (1999). Basic mechanisms of Gabitril
Leach, M. J., Marden, C. M., & Miller, A. A. (1986). Pharmacological (tiagabine) and future potential developments. Epilepsia 40 (suppl. 9),
studies on lamotrigine, a novel potential antiepileptic drug: II. Neuro- S2 – S6.
chemical studies on the mechanism of action. Epilepsia 27, 490 – 497. Meldrum, B. S., Akbar, M. T., & Chapman, A. G. (1999). Glutamate
Leppik, I. E. (1999). Zonisamide. Epilepsia 40 (suppl. 5), S23 – S29. receptors and transporters in genetic and acquired models of epilepsy.
Leppik, I. E., Willmore, L. J., Homan, R. W., Fromm, G., Oommen, K. J., Epilepsy Res 362, 189 – 204.
Penry, J. K., Sackellares, J. C., Smith, D. B., Lesser, R. P., Wallace, Merritt, H. H., & Putnam, T. J. (1938). Sodium diphenyl hydantoinate in the
J. D., Trudeau, J. L., Lamoreaux, L. K., & Spenser, M. (1993). treatment of convulsive disorders. JAMA 111, 1068 – 1073.
Efficacy and safety of zonisamide: results of a multicenter study. Meunier, H., Carraz, G., Meunier, Y., Eymard, P., & Aimard, M. (1963).
Epilepsy Res 14, 165 – 173. Proprietes pharmacodynamiques de l’acide n-propylacetique. 1er mem-
Lippert, B., Metcalf, B. W., Jung, M. J., & Casara, P. (1977). 4-Amino-hex- oire: proprietes antiepileptique. Therapie 18, 435 – 438.
5-eonic acid, a selective catalytic inhibitor of 4-aminobutyric-acid ami- Mimaki, T., Suzuki, Y., Tagawa, T., Tanaka, J., Itoh, N., & Yabuuchi, H.
notransferase in mammalian brain. Eur J Biochem 4, 441 – 445. (1988). [3H]Zonisamide binding in rat brain. Jpn J Psychiatry Neurol
Löscher, W. (1998). New visions in the pharmacology of anticonvulsion. 42, 640 – 642.
Eur J Pharmacol 342, 1 – 13. Mody, I. (1998). Ion channels in epilepsy. Int Rev Neurobiol 42, 199 – 226.
Löscher, W. (1999). Valproate: a reappraisal of its pharmacodynamic prop- Nobile, M., & Vercellino, P. (1997). Inhibition of delayed rectifier K +
erties and mechanisms of action. Prog Neurobiol 58, 31 – 59. channels by phenytoin in rat neuroblastoma cells. Br J Pharmacol
Löscher, W., & Hönack, D. (1993). Profile of ucb L059, a novel anticon- 120, 647 – 652.
vulsant drug, in models of partial and generalized epilepsy in mice and Noyer, M., Gillard, M., Matagne, A., Henichart, J. P., & Wulfert, E.
rats. Eur J Pharmacol 232, 147 – 158. (1995). The novel antiepileptic drug levetiracetam (ucb L059) appears
Löscher, W., Hönack, D., & Bloms-Funke, P. (1996). The novel antiepi- to act via a specific binding site in CNS membranes. Eur J Pharmacol
leptic drug levetiracetam (ucb L059) induces alterations in GABA me- 286, 137 – 146.
tabolism and turnover in discrete areas of rat brain and reduces neuronal Okada, M., Kaneko, S., Hirano, T., Mizuno, K., Kondo, T., Otani, K., &
activity in substantia nigra pars reticulata. Brain Res 735, 208 – 216. Fukushima, Y. (1995). Effects of zonisamide on dopaminergic system.
Macdonald, R. L., & Greenfield, L. J., Jr. (1997). Mechanisms of action of Epilepsy Res 22, 193 – 205.
new antiepileptic drugs. Curr Opin Neurol 10, 121 – 128. Olsen, R. W., & Avoli, M. (1997). GABA and epileptogenesis. Epilepsia
Macdonald, R. L., & Kelly, K. M. (1995). Antiepileptic drug mechanisms 38, 399 – 407.
of action. Epilepsia 36 (suppl. 2), S2 – S12. Ophoff, R. A., Terwindt, G. M., Frants, R. R., & Ferrari, M. D. (1998).
Macdonald, R. L., Rogers, C. J., & Twyman, R. E. (1989). Barbiturate P/Q-type Ca2 + channel defects in migraine, ataxia and epilepsy.
regulation of kinetic properties of the GABAA receptor channel of Trends Pharmacol Sci 19, 121 – 127.
mouse spinal neurones in culture. J Physiol 417, 483 – 500. Pellock, J. M., & Brodie, M. J. (1997). Felbamate: an update. Epilepsia 38,
Marangos, P. J., Post, R. M., Patel, J., Zander, A., Parma, K., & Weiss, S. 1261 – 1264.
(1983). Specific and potent interactions of carbamazepine with brain Petroff, O. A. C., & Rothman, D. L. (1998). Measuring human brain GABA
adenosine receptors. Eur J Pharmacol 93, 175 – 182. in vivo: effects of GABA-transaminase inhibition with vigabatrin. Mol
Margineanu, D. G., & Wulfert, E. (1995). ucb L059, a novel anticonvulsant, Neurobiol 16, 97 – 121.
P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34 33

Petroff, O. A. C., Rothman, D. L., Behar, K. L., Lamoureux, D., & Mattson, activates M-currents in Chinese hamster ovary-cells transfected with
R. H. (1996a). The effect of gabapentin on brain g-aminobutyric acid in human KCNQ2/3 subunits. Neurosci Lett 282, 73 – 76.
patients with epilepsy. Ann Neurol 39, 95 – 99. Schachter, S. C. (1995). Review of the mechanisms of action of antiepi-
Petroff, O. A. C., Rothman, D. L., Behar, K. L., & Mattson, R. H. (1996b). leptic drugs. CNS Drugs 4, 469 – 477.
Human brain GABA levels rise after initiation of vigabatrin therapy but Schauf, C. L. (1987). Zonisamide enhances slow sodium inactivation in
fail to rise further with increasing dose. Neurology 46, 1459 – 1463. Myxicola. Brain Res 413, 185 – 188.
Petroff, O. A. C., Hyder, F., Mattson, R. H., & Rothman, D. L. (1999). Schechter, P. J., Tranier, Y., Jung, M. J., & Bohlen, P. (1977). Audiogenic
Topiramate increases brain GABA, homocarnosine, and pyrrolidinone seizure protection by elevated brain GABA concentration in mice: ef-
in patients with epilepsy. Neurology 52, 473 – 478. fects of g-acetylenic GABA and g-vinyl GABA, two irreversible GA-
Pisani, A., Stefani, A., Siniscalchi, A., Mercuri, N. B., Bernardi, G., & BA-T inhibitors. Eur J Pharmacol 45, 319 – 328.
Calabresi, P. (1995). Electrophysiological actions of felbamate on rat Schechter, P. J., Tranier, Y., & Grove, J. (1978). Effect of n-dipropylacetate
striatal neurones. Br J Pharmacol 16, 2053 – 2061. on amino acid concentrations in mouse brain: correlation with anticon-
Pongs, O. (1999). Voltage-gated potassium channels: from hyperexcitability vulsant activity. J Neurochem 31, 1325 – 1327.
to excitement. FEBS Lett 452, 31 – 35. Schumacher, T. B., Beck, H., Steinhauser, C., Schramm, J., & Elger, C. E.
Porter, R. J., & Rogawski, M. A. (1992). New antiepileptic drugs: from (1998). Effects of phenytoin, carbamazepine, and gabapentin on calci-
serendipity to rational discovery. Epilepsia 33 (suppl. 1), S1 – S6. um channels in hippocampal granule cells from patients with temporal
Pugliese, A. M., & Corradetti, R. (1996). Effects of the antiepileptic drug lobe epilepsy. Epilepsia 39, 355 – 363.
felbamate on long-term potentiation in the CA1 region of rat hippo- Schwartz, J. R., & Grigat, G. (1989). Phenytoin and carbamazepine: poten-
campal slices. Neurosci Lett 215, 21 – 24. tial- and frequency-dependent block of Na + currents in mammalian
Rabow, L. E., Russek, S. J., & Farb, D. H. (1995). From ion currents to myelinated nerve fibers. Epilepsia 30, 286 – 294.
genomic analysis: recent advances in GABAA receptor research. Syn- Shank, R. P., Gardocki, J. F., Vaught, J. L., Davis, C. B., Schupsky, J. J.,
apse 21, 189 – 274. Raffa, R. B., Dodgson, S. J., Nortey, S. O., & Maryanoff, B. E. (1994).
Ragsdale, D. S., & Avoli, M. (1998). Sodium channels as molecular targets Topiramate: preclinical evaluation of a structurally novel anticonvul-
for antiepileptic drugs. Brain Res Rev 26, 16 – 28. sant. Epilepsia 35, 450 – 460.
Ragsdale, D. S., McPhee, J. C., Scheuer, T., & Catterall, W. A. (1996). Sills, G. J., Leach, J. P., Butler, E., Carswell, A., Thompson, G. G., &
Common molecular determinants of local anesthetic, antiarrhythmic, Brodie, M. J. (1996). Antiepileptic drug action in primary cultures of
and anticonvulsant block of voltage-gated Na + channels. Proc Natl rat cortical astrocytes. Epilepsia 37 (suppl. 4), 116.
Acad Sci USA 93, 9270 – 9275. Sills, G. J., Leach, J. P., Fraser, C. M., Forrest, G., Patsalos, P. N., & Brodie,
Reynolds, E. H., Milner, G., Matthews, D. M., & Chanarin, I. (1966). M. J. (1997). Neurochemical studies with the novel anticonvulsant
Anticonvulsant therapy, megaloblastic haemopoiesis and folic acid me- levetiracetam in mouse brain. Eur J Pharmacol 325, 35 – 40.
tabolism. Q J Med 35, 521 – 537. Sills, G. J., Butler, E., Thompson, G. G., & Brodie, M. J. (1999). Vigabatrin
Rho, J. M., & Sankar, R. (1999). The pharmacologic basis of antiepileptic and tiagabine are pharmacologically different drugs. A pre-clinical
drug action. Epilepsia 40, 1471 – 1483. study. Seizure 8, 404 – 411.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1994). Mechanism of Sills, G. J., Leach, J. P., Kilpatrick, W. S., Fraser, C. M., Thompson, G. G.,
action of the anticonvulsant felbamate: opposing effects on N-methyl- & Brodie, M. J. (2000). Concentration – effect studies with topiramate
D-aspartate and g-aminobutyric acidA receptors. Ann Neurol 35, on selected enzymes and intermediates of the GABA shunt. Epilepsia
229 – 234. 41 (suppl. 1), S30 – S34.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1996). Direct activation of Song, J. H., Nagata, K., Huang, C. S., Yeh, J. Z., & Narahashi, T. (1996).
GABAA receptors by barbiturates in cultured rat hippocampal neurons. Differential block of two types of sodium channels by anticonvulsants.
J Physiol 497, 509 – 522. Neuroreport 7, 3031 – 3036.
Rho, J. M., Donevan, S. D., & Rogawski, M. A. (1997). Barbiturate-like Squires, R. F., & Braestrup, C. (1977). Benzodiazepine receptors in rat
actions of the propanediol dicarbamates felbamate and meprobamate. J brain. Nature 266, 732 – 734.
Pharmacol Exp Ther 280, 1383 – 1391. Srinivasan, J., Richens, A., & Davies, J. A. (1996). Effects of felbamate on
Rock, D. M., Macdonald, R. L., & Taylor, C. P. (1989). Blockade of veratridine- and K + -stimulated release of glutamate from mouse cortex.
sustained repetitive action potentials in cultured spinal cord neurons Eur J Pharmacol 315, 285 – 288.
by zonisamide (AD 810, CI 912): a novel anticonvulsant. Epilepsy Stefani, A., Pisani, A., De Murtas, M., Mercuri, N. B., Marciani, M. G., &
Res 3, 138 – 143. Calabresi, P. (1995). Action of GP 47779, the active metabolite of
Roepstorff, A., & Lambert, J. D. (1992). Comparison of the effect of oxcarbazepine, on the corticostriatal system. II. Modulation of high-
GABA uptake blockers, tiagabine and nipecotic acid, on inhibitory voltage-activated calcium currents. Epilepsia 336, 997 – 1002.
synaptic efficacy in hippocampal CA1 neurones. Neurosci Lett 146, Stefani, A., Calabresi, P., Pisani, A., Mercuri, N. B., Siniscalchi, A., &
131 – 134. Bernardi, G. (1996a). Felbamate inhibits dihydropyridine-sensitive
Rogawski, M. A. (2000). KCNQ2/KCNQ3 K + channels and the molecular calcium channels in central neurons. J Pharmacol Exp Ther 277,
pathogenesis of epilepsy:implications for therapy. Trends Neurosci 23, 121 – 127.
393 – 398. Stefani, A., Spadoni, F., Siniscalchi, A., & Bernadi, G. (1996b). Lamotri-
Rogawski, M. A., & Porter, R. J. (1990). Antiepileptic drugs: pharmaco- gine inhibits Ca2 + currents in cortical neurons: functional implications.
logical mechanisms and clinical efficacy with consideration of promis- Eur J Pharmacol 307, 113 – 116.
ing developmental stage compounds. Pharmacol Rev 42, 223 – 286. Stefani, A., Spadoni, F., & Bernardi, G. (1997). Voltage-activated cal-
Rostock, A., Tober, C., Rundfeldt, C., Bartsch, R., Engel, J., Polymer- cium channels: targets of antiepileptic drug therapy? Epilepsia 38,
opoulos, E. E., Kutscher, B., Löscher, W., Honack, D., White, H. S., 959 – 965.
& Wolf, H. H. (1996). D-23129: a new anticonvulsant with a broad Steriade, M., & Llinas, R. (1988). The functional states of the thalamus and
spectrum activity in animal models of epileptic seizures. Epilepsy Res the associated neuronal interplay. Physiol Rev 68, 649 – 742.
23, 211 – 223. Study, R. E., & Barker, J. L. (1981). Diazepam and (  )-pentobarbital:
Rowley, H. L., Marsden, C. A., & Martin, K. F. (1995). Differential effects fluctuation analysis reveals different mechanisms for potentiation of
of phenytoin and sodium valproate on seizure-induced changes in g- g-aminobutyric acid responses in cultured central neurons. Proc Natl
aminobutyric acid and glutamate release in vivo. Eur J Pharmacol 294, Acad Sci USA 78, 7180 – 7184.
541 – 546. Su, T. Z., Lunney, E., Campbell, G., & Oxender, D. L. (1995). Transport of
Rundfeldt, C., & Netzer, R. (2000). The novel anticonvulsant retigabine gabapentin, a g-amino acid drug, by system l g-amino acid transporters:
34 P. Kwan et al. / Pharmacology & Therapeutics 90 (2001) 21–34

a comparative study in astrocytes, synaptosomes, and CHO cells. J Waldmeier, P. C., Baumann, P. A., Wicki, P., Feldtrauer, J. J., Stierlin, C., &
Neurochem 64, 2125 – 2131. Schmutz, M. (1995). Similar potency of carbamazepine, oxcarbazepine,
Suzuki, S., Kawakami, K., Nishimura, S., Watanabe, Y., Yagi, K., and lamotrigine in inhibiting the release of glutamate and other neuro-
Seino, M., & Miyamoto, K. (1992). Zonisamide blocks T-type calcium transmitters.. Neurology 45, 1907 – 1913.
channels in cultured neurons of rat cerebral cortex. Epilepsy Res 12, Wamil, A. W., & McLean, M. J. (1994). Limitation by gabapentin of high
21 – 27. frequency action potential firing by mouse central neurons in culture.
Taglialatela, M., Ongini, E., Brown, A. M., Di Renzo, G., & Annunziato, L. Epilepsy Res 17, 1 – 11.
(1996). Felbamate inhibits cloned voltage-dependent Na + channels Wang, C. M., Lang, D. G., & Cooper, B. R. (1993). Lamotrigine effects
from human and rat brain. Eur J Pharmacol 316, 373 – 377. on ion channels in cultured neuronal cells. Epilepsia 34 (suppl. 6),
Taylor, C. P., Rock, D. M., Weinkauf, R. J., & Ganong, A. H. (1988). In 117 – 118.
vitro and in vivo electrophysiology effects of the anticonvulsant gaba- Wang, S. J., Huang, C. C., Hsu, K. S., Tsai, J. J., & Gean, P. W. (1996).
pentin. Soc Neurosci Abstr 14, 866. Inhibition of N-type calcium currents by lamotrigine in rat amygdalar
Taylor, C. P., Vartanian, M. G., Andruszkiewicz, R., & Silverman, R. B. neurones. Neuroreport 7, 3037 – 3040.
(1992). 3-Alkyl GABA and 3-alkylglutamic acid analogues: two new White, H. S. (1999). Comparative anticonvulsant and mechanistic profile
classes of anticonvulsant agents. Epilepsy Res 11, 103 – 110. of established and newer antiepileptic drugs. Epilepsia 40 (suppl. 5),
Taylor, C. P., Gee, N. S., Su, T-Z., Kocsis, J. D., Welty, D. F., Brown, J. P., S2 – S10.
Dooley, D. J., Boden, P., & Singh, L. (1998). A summary of mech- White, H. S., Brown, D., Skeen, G. A., Wolf, H. H., & Twyman, R. E.
anistic hypotheses of gabapentin pharmacology. Epilepsy Res 29, (1995a). The anticonvulsant topiramate displays a unique ability to
233 – 249. potentiate GABA-evoked chloride currents. Epilepsia 36 (suppl. 3),
Taylor, L. A., McQuade, R. D., & Tice, M. A. (1995). Felbamate, a novel S39 – S40.
antiepileptic drug, reverses N-methyl-D-aspartate/glycine-stimulated in- White, H. S., Harmsworth, W. L., Sofia, R. D., & Wolf, H. H. (1995b).
creases in intracellular Ca2 + concentration. Eur J Pharmacol 289, Felbamate modulates the strychnine-insensitive glycine receptor. Epi-
229 – 233. lepsy Res 20, 41 – 48.
Tecoma, E. S. (1999). Oxcarbazepine. Epilepsia 40 (suppl. 5), S37 – S46. White, H. S., Brown, S. D., Woodhead, J. H., Skeen, G. A., & Wolf, H. H.
Teoh, H., Fowler, L. J., & Bowery, N. G. (1995). Effect of lamotrigine (1997). Topiramate enhances GABA-mediated chloride flux and
on the electrically-evoked release of endogenous amino acids from GABA-evoked chloride currents in murine brain neurons and increases
slices of dorsal horn of the rat spinal cord. Neuropharmacology 34, seizure threshold. Epilepsy Res 28, 167 – 179.
1273 – 1278. Wilson, E. A., & Brodie, M. J. (1996). New antiepileptic drugs. In: M. J.
Ticku, M. K., Kamatchi, G. L., & Sofia, R. D. (1991). Effect of anticon- Brodie, & D. M. Brodie (Eds.), Ballière’s Clinical Neurology. Modern
vulsant felbamate on GABA receptor system. Epilepsia 32, 389 – 391. Management of Epilepsy ( pp. 723 – 747). London: Ballière-Tindall.
Treiman, D. M., Meyers, P. D., Walton, N. Y., Collins, J. F., Colling, C., Yamaguchi, S., & Rogawski, M. A. (1992). Effects of anticonvulsant drugs
Rowan, A. J., Handforth, A., Faught, E., Calabrese, V. P., Uthman, B., on 4-aminopyridine-induced seizures in mice. Epilepsy Res 11, 9 – 16.
Ramsay, R. E., & Mamdani, B. (1998). A comparison of four treatments Zhang, X., Velumian, A. A., Jones, O. T., & Carlen, P. L. (2000). Modu-
for generalized convulsive status epilepticus. N Engl J Med 339, 792 – lation of high-voltage-activated calcium channels in dentate granule
798. cells by topiramate. Epilepsia 41 (suppl. 1), S52 – S60.
Trist, D. G. (2000). Excitatory amino acid agonists and antagonists: phar- Zona, C., & Avoli, M. (1990). Effects induced by the antiepileptic drug
macology and therapeutic applications. Pharm Acta Helv 74, 221 – 229. valproic acid upon the ionic currents recorded in rat neocortical neurons
Tunnicliff, G. (1996). Basis of the antiseizure action of phenytoin. Gen in cell culture. Exp Brain Res 81, 313 – 317.
Pharmacol 27, 1091 – 1097. Zona, C., & Avoli, M. (1997). Lamotrigine reduces voltage-gated sodium
Twyman, R. E., Rogers, C. J., & Macdonald, R. L. (1989). Differential currents in rat central neurons in culture. Epilepsia 38, 522 – 525.
regulation of g-aminobutyric acid receptor channels by diazepam and Zona, C., Barbarosie, M., Kawasaki, H., & Avoli, M. (1996a). Effects
phenobarbital. Ann Neurol 25, 213 – 220. induced by the anticonvulsant drug topiramate on voltage-gated sodium
Upton, N. (1994). Mechanisms of action of new antiepileptic drugs: rational currents generated by cerebellar granule cells in tissue culture. Epilepsia
design and serendipitous findings. Trends Pharmacol Sci 15, 456 – 463. 37 (suppl. 5), 24.
van Dongen, A. M. J., van Erp, M. G., & Voskuyl, R. A. (1986). Valproate Zona, C., Ciotti, M. T., & Avoli, M. (1996b). Topiramate attenuates volt-
reduces excitability by blockage of sodium and potassium conductance. age-gated sodium currents in rat cerebellar granule cells. Neurosci Lett
Epilepsia 27, 177 – 182. 231, 123 – 126.

Vous aimerez peut-être aussi