1 s2.0 S0014299919303115 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

European Journal of Pharmacology 855 (2019) 175–182

Contents lists available at ScienceDirect

European Journal of Pharmacology


journal homepage: www.elsevier.com/locate/ejphar

Full length article

Effects of cenobamate (YKP3089), a newly developed anti-epileptic drug, on T


voltage-gated sodium channels in rat hippocampal CA3 neurons
Michiko Nakamuraa,b, Jin-Hwa Choa, Hyewon Shinc, Il-Sung Janga,b,∗
a
Department of Pharmacology, School of Dentistry, Kyungpook National University, 2177 Dalgubeol-daero, Jung-gu, Daegu, 700-412, Republic of Korea
b
Brain Science & Engineering Institute, Kyungpook National University, 2177 Dalgubeol-daero, Jung-gu, Daegu, 700-412, Republic of Korea
c
Department of Pharmacology, SK Biopharmaceuticals, Co., Ltd., 221 Pangyoyeok-ro, Seongnam, Gyeonggi, 305-712, Republic of Korea

ARTICLE INFO ABSTRACT

Keywords: New, more effective pharmacologic treatments for epilepsy are needed, as a substantial portion of patients
Epilepsy (> 30%) are refractory to currently available anti-epileptic drugs. Cenobamate (YKP3089) is an investigational
Anti-epileptic drugs anti-epileptic drug in clinical development. Two completed adequate and well-controlled studies demonstrated a
Voltage-gated Na+ channels significant reduction in focal seizures with cenobamate in patients with epilepsy. In this study, we characterized
Hippocampal neurons
the effects of cenobamate on voltage-gated Na+ channels in acutely isolated rat hippocampal CA3 neurons using
Patch clamp
a whole-cell patch-clamp technique. While cenobamate had little effect on the peak component of transient Na+
current (INaT) induced by brief depolarizing step pulses, it potently inhibited the non-inactivating persistent
component of INa (INaP). In addition, cenobamate potently inhibited the current by slow voltage-ramp stimuli.
Cenobamate significantly shifted the steady-state fast inactivation relationship toward a hyperpolarizing range,
indicating that cenobamate binds to voltage-gated Na+ channels at the inactivated state with a higher affinity.
Cenobamate also accelerated the development of inactivation and retarded recovery from inactivation of vol-
tage-gated Na+ channels. In current clamp experiments, cenobamate hyperpolarized membrane potentials in a
concentration-dependent manner, and these effects were mediated by inhibiting the INaP. Cenobamate also in-
creased the threshold for generation of action potentials, and decreased the number of action potentials elicited
by depolarizing current injection. Given that the INaP plays a pivotal role in the repetitive and/or burst gen-
eration of action potentials, the cenobamate-mediated preferential blockade of INaP might contribute to anti-
epileptic activity.

1. Introduction existence of slowly inactivating or non-inactivating Na+ currents


mediated by voltage-gated Na+ channels, called persistent Na+ cur-
Epilepsy is a common neurological disorder characterized by re- rents (INaP), has been identified in central as well as peripheral neurons
current and unprovoked seizures (Brodie et al., 2012; Chen et al., 2018; (Crill, 1996; Wu et al., 2005; Xie et al., 2011). In addition to its role in
Kwan and Brodie, 2000; Tian et al., 2018). Excessive excitability in the excitability and firing pattern in neurons of many brain regions
neural tissues is believed to contribute to epilepsy (Das et al., 2012). (Bennett et al., 2000; Jackson et al., 2004; Taddese and Bean, 2002; Yue
Anti-epileptic drugs (AEDs) acting on voltage-gated Na+ channels have et al., 2005), the INaP plays a crucial role in several pathological con-
long been utilized for the pharmacologic treatment of epilepsy because ditions, such as epilepsy and pain (Hains and Waxman, 2007; Lossin
voltage-gated Na+ channels play a pivotal role in the generation and et al., 2002; Stafstrom, 2007). In fact, several point mutations in vol-
conduction of action potentials (Catterall, 2014). However, despite the tage-gated Na+ channels that exhibit increased INaP have been identi-
use of many currently available AEDs, seizures remain uncontrolled in fied in patients with epilepsy (Lossin et al., 2002; Stafstrom, 2007).
over 30% of patients with epilepsy (Brodie et al., 2012; Chen et al., Moreover, in subicular burst firing neurons resected from patients with
2018; Kwan and Brodie, 2000; Tian et al., 2018). temporal lobe epilepsy, a large increase of INaP has been recorded, with
Transient Na+ currents (INaT) are generally quickly activated and an amplitude up to half of the total sodium current (Vreugdenhil et al.,
then inactivated upon membrane depolarization. However, the 2004). Several AEDs targeting voltage-gated Na+ channels, such as


Corresponding author. Professor Department of Pharmacology, School of Dentistry Kyungpook National University, 2177 Dalgubeol-daero, Jung-gu, Daegu, 700-
412, Republic of Korea.
E-mail addresses: michiko21a@hotmail.com (M. Nakamura), cjinhwa@knu.ac.kr (J.-H. Cho), hyewon.shin@sk.com (H. Shin), jis7619@knu.ac.kr (I.-S. Jang).

https://doi.org/10.1016/j.ejphar.2019.05.007
Received 6 December 2018; Received in revised form 30 April 2019; Accepted 3 May 2019
Available online 04 May 2019
0014-2999/ © 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/BY/4.0/).
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

phenytoin, valproic acid, and lamotrigine, are known to inhibit this (P-97; Sutter Instrument Co., Novato, CA, USA). The resistance of the
increased INaP to some extent (Spadoni et al., 2002; Stafstrom, 2007). A recording pipettes filled with the internal solution (in mM; 140 CsF, 10
more recent study has also shown the anti-epileptic activity of two CsCl, 2 EGTA, 2 ATP-Na2, and 10 HEPES, with the pH adjusted to 7.2
compounds that preferentially block INaP in animal models (Anderson with Tris-base) was 0.8–1.5 MΩ. In current-clamp experiments, 140 CsF
et al., 2014), indicating that the INaP may be a promising target for the and 10 CsCl were replaced with equimolar 140 KF and 10 KCl, re-
development of new AEDs. spectively. Membrane potentials were corrected for the liquid junction
Cenobamate (YKP3089) is an investigational AED that has shown a potential. Neurons were viewed under phase contrast on an inverted
broad-spectrum of anticonvulsant activity in rodent epilepsy models microscope (TE-2000; Nikon, Tokyo, Japan). Membrane currents were
(Bialer et al., 2013). In mice and rats, cenobamate displayed an antic- filtered at 2–5 kHz, digitized at 10–20 kHz, and stored on a computer
onvulsant activity in the maximal electroshock test and prevented sei- equipped with pCLAMP 10.3 (Molecular Devices). Capacitative and
zures induced by chemical convulsants such as pentylenetetrazol and leakage currents were subtracted by the P/4 protocol using pCLAMP
picrotoxin. In addition, cenobamate was reported to be effective in two program. During recordings, 10 mV hyperpolarizing step pulses (30 ms
models of focal seizure, the hippocampal kindled rat and the mouse in duration) were periodically applied to monitor the access resistance,
6 Hz psychomotor seizure models. Two completed adequate and well- and recordings were discontinued if access resistance changed by more
controlled clinical studies demonstrated a significant reduction in focal than 10%. All experiments were performed at room temperature
seizures with cenobamate in patients with epilepsy (clinicaltrials.gov (22°C–25 °C). To record voltage-gated Na+ currents (INa), the K+-free
NCT01397968 and NCT01866111), and a long-term open-label phase 3 external solution routinely contained 100 μM Cd2+ to block voltage-
safety clinical trial is currently ongoing (clinicaltrials.gov gated Ca2+ channels. Depolarizing step pulses to evoke the INa were
NCT02535091). In this study, we characterized the mode of action of applied with an interval of 10 s (except where indicated), which was
cenobamate in acutely isolated rat hippocampal CA3 pyramidal neu- sufficiently long enough to recover from the inactivation of voltage-
rons using a whole-cell patch-clamp technique. gated Na+ channels. To record membrane currents of ligand-gated ion
channels, CsF within the pipette solution was replaced with equimolar
2. Materials and methods Cs-methanesulfonate.

2.1. Preparation 2.3. Data analysis

All experiments complied with the guiding principles for the care The amplitude of INa was measured by subtracting the baseline from
and use of animals approved by the Kyungpook National University and the peak amplitude of INa by using pCLAMP 10.3 (Molecular Devices).
Council of the Physiological Society of Korea, and every effort was The continuous curves for the concentration-response relationship were
made to minimize both the number of animals used and their suffering. fitted using a least-squares fit method to the following equation, I = 1 –
Sprague Dawley rats (12–17 day-old, either sex) were decapitated [Cn/(Cn + IC50n)], where I is the relative amplitude of INa changed by
under ketamine anesthesia (100 mg/kg, i.p.). The brain was dissected cenobamate, C is the cenobamate concentration, IC50 is the cenobamate
and transversely sliced at a thickness of 400 μm using a microslicer concentration for the half-maximal response, and n is the Hill coefficient.
(VT1000S; Leica, Nussloch, Germany). Slices containing the hippo- The amplitude of INa was transformed into conductance (G) using the
campus were kept in an incubation medium (in mM; 124 NaCl, 3 KCl, following equation; G = I/(V – ENa), where ENa is the Na+ equilibrium
1.5 KH2PO4, 24 NaHCO3, 2 CaCl2, 1.3 MgSO4 and 10 glucose) saturated potential calculated by the Nernst equation. The voltage-activation and
with 95% O2 and 5% CO2 at room temperature (22-24 °C) for at least voltage-inactivation relationships of Na+ channels were fitted to the
1 h before the mechanical dissociation. For dissociation, slices were Boltzmann equations, respectively; G/Gmax = 1/{1 + exp[(V50,act – V)/
transferred into a 35 mm culture dish (Primaria 3801; Becton k]} and I/Imax = 1–1/{1 + exp[(V50,inact – V)/k]}, where Gmax and Imax
Dickinson, Rutherford, NJ, USA), which contained a standard external are the maximum conductance and current amplitude, respectively,
solution (in mM; 150 NaCl, 3 KCl, 2 CaCl2, 1 MgCl2, 10 glucose, and 10 V50,act and V50,inact are half-maximal potentials for activation and fast
HEPES, adjusted to a pH of 7.4 with Tris-base), and the hippocampal inactivation, respectively, and k is the slope factor. The fast (τfast) and
CA3 region was identified under a binocular microscope (SMZ-1; Nikon, slow (τslow) time constants of the decay of single INa and the kinetic data
Tokyo, Japan). Details of the mechanical dissociation have been de- for the recovery from inactivation were fitted to the following equation; I
scribed previously (Akaike and Moorhouse, 2003; Rhee et al., 1999). (t) = A0 + Afast x [1 – exp(–t/τfast)] + Aslow x [1 – exp(–t/τslow)], and the
Briefly, mechanical dissociation was accomplished using a custom-built kinetic data for the development of inactivation were fitted to the fol-
vibration device and a fire-polished glass pipette oscillating at about lowing equation; I(t) = A0 + Afast x [exp(–t/τfast)] + Aslow x [exp(–t/
50–60 Hz (0.3–0.5 mm) on the surface of the CA3 pyramidal layer. τslow)], where I(t) is the amplitude of INa at time t, and Afast and Aslow are
Slices were removed and mechanically dissociated neurons allowed to the amplitude fraction of τfast, and τslow, respectively. The weighted time
settle and to adhere to the bottom of the dish for 15 min. These dis- constant (τWD or τweighted) was calculated by using the following equa-
sociated neurons lost the most distal processes, but a short portion tion: τWD or τweighted = [(τfast x Afast) + (τslow x Aslow)]/(Afast + Aslow).
(∼100 μm in a length) of their thick proximal dendrites was retained, The binding affinity of cenobamate for resting or inactivated Na+
as previously shown (Jang et al., 2006) (Fig. 1A). channels was fitted to the following equation (Bean, 1984): exp
(ΔV50,inact/k) = [1 + (C/KI)]/[1 + (C/KR)], where ΔV50,inact, C, KI, and
2.2. Electrophysiology KR represent the voltage shift, cenobamate concentration, dissociation
constants for inactivated and resting states, respectively. Numerical va-
All electrical measurements were performed using conventional lues are provided as the mean ± S.E.M. using values normalized to the
whole-cell patch recordings and a patch-clamp amplifier (Axopatch control. Significant differences in the mean amplitude were tested using
200B; Molecular Devices, Union City, CA, USA) with a K+-free external Student's paired two-tailed t-test, except where indicated, with absolute
solution (in mM; 130 NaCl, 20 tetraethylammonium-Cl, 3 CsCl, 2 CaCl2, values rather than normalized ones. Values of P < 0.05 were considered
1 MgCl2, 10 glucose, and 10 HEPES, adjusted to a pH of 7.4 with Tris- significantly different.
base). In a subset of experiments using a low Na+ external solution,
100 mM NaCl was replaced with equimolar N-methyl-D-glucamine-Cl. 2.4. Drugs
Neurons were voltage-clamped at a holding potential (VH) of −80 mV,
except where indicated. Patch pipettes were made from borosilicate The drugs and chemicals used in the present study were purchased
capillary glass (G-1.5; Narishige, Tokyo, Japan) using a pipette puller from Sigma-Aldrich (St. Louis, MO, USA). Cenobamate (Batch PD187-

176
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

Fig. 1. Effects of cenobamate on voltage-


gated Na+ currents in acutely isolated hip-
pocampal CA3 neurons. A: A typical phase
contrast image of an acutely isolated CA3
pyramidal neuron. B: Typical traces of INaT
and INaP in the absence and presence of
100 μM cenobamate. C: Onset time (a) and
decay time constant (b) of INaT in the ab-
sence (open column) and presence (closed
column) of 100 μM cenobamate. Each
column represents the mean ± S.E.M. from
7 experiments (7 animals). n.s., not sig-
nificant. D: Concentration-response re-
lationships of cenobamate against the peak
(open circles) and steady state components
(closed circles) of INaT. Each point re-
presents the mean ± S.E.M. from 7 ex-
periments (7 animals).

14-002N) was provided by SK Biopharmaceuticals, Co., Ltd (Gyeonggi, to inhibit persistent currents but not transient Na+ currents at central
Republic of Korea). Cenobamate (YKP3089) was dissolved in dimethyl neurons (Urbani and Belluzzi, 2000), or 300 nM tetrodotoxin, a specific
sulfoxide to give a stock solution of 100–300 mM, and the final con- voltage-gated Na+ channel blocker (Fig. 2Ab), suggesting that the slow
centration of dimethyl sulfoxide applied to the external solution voltage-ramp-induced membrane currents share properties with INaP
was < 0.1% v/v. All solutions containing drugs were applied using the mediated by voltage-gated Na+ channels as shown in previous studies
“Y-tube system” for rapid solution exchange (Murase et al., 1989). (Bellingham, 2011; Park et al., 2013; Parri and Crunelli, 1998). Next,
the effects of cenobamate on the currents by slow voltage-ramps in
3. Results hippocampal CA3 neurons were examined. Cenobamate decreased the
amplitude of the currents in a concentration-dependent manner
3.1. Effects of cenobamate on voltage-gated Na+ channels in acutely (IC50 = 53.1 ± 4.8 μM, n = 6). At 100 μM cenobamate, the current
isolated hippocampal CA3 neurons amplitude decreased to 31.5 ± 1.4% of control (n = 6, P < 0.01,
Fig. 2B and D). The conductance of currents recorded at each voltage of
INa were induced by depolarizing step pulses (−80 mV to −20 mV, ramp command was calculated and normalized to the maximal con-
every 5 s) in acutely isolated hippocampal CA3 neurons. The INaT was ductance in the control condition, and the data were fitted to the
quickly decayed with a time constant of 0.40 ± 0.01 ms (n = 7, Fig. 1B Boltzmann function (Fig. 2C). Cenobamate (100 μM) did not shift V50,act
and C), but the INa was not completely returned to baseline even after (−47.8 ± 3.5 for control and −51.1 ± 4.0 mV for cenobamate,
200 ms of sustained depolarizing stimuli (Fig. 1B). The transient com- n = 6, P = 0.45, Fig. 2C). In addition, cenobamate did not change the
ponent of INa was 10.6 ± 1.6 nA (n = 22), and the persistent compo- slope factor of voltage dependence (Fig. 2Cb).
nent of INa was 135.1 ± 21.8 pA (n = 22) (Fig. 1B). This persistent
component during sustained depolarizing stimuli was similar to the 3.3. Effects of cenobamate on the voltage dependence of Na+ channels
non-desensitizing or INaP shown in previous studies (Kohling, 2002;
Urbani and Belluzzi, 2000). We next examined the effects of cen- The effect of cenobamate on the voltage-activation relationship of
obamate on INa in acutely isolated rat hippocampal CA3 neurons. Na+ channels was examined. In these experiments, extracellular Na+
Cenobamate (100 μM) hardly affected the INaT amplitude, as shown by a concentration ([Na+]o) was decreased to 30 mM to improve the quality
slight decrease to 94.6 ± 2.5% of vehicle control (n = 7, P < 0.01, of the voltage-clamp. In these conditions, the INa was induced by 50 ms
Fig. 1B and D). In contrast, cenobamate (100 μM) potently decreased depolarizing test pulses from −80 to +30 mV in 10 mV increments
the INaP to 25.6 ± 4.4% of control (n = 7, P < 0.01, Fig. 1B and D). before and during application of cenobamate (Fig. 3A). However,
The cenobamate-mediated inhibition of the transient and persistent cenobamate at a concentration of 100 μM concentration had little effect
components was concentration-dependent (IC50 > 500 μM for the INaT, on the current-voltage relationship (Fig. 3B). When the conductance at
and IC50 = 53.1 ± 4.2 μM for the INaP, n = 7, respectively, Fig. 1D). each voltage was normalized to the maximal conductance in the control
Cenobamate (100 μM) had no effect on the time to peak or the weighted condition, and the pooled data were fitted to the Boltzmann function,
decay time constant (τWD) of transient INa (time to peak; 0.40 ± 0.01 cenobamate concentrations at ≤100 μM did not change the half-max-
for control and 0.41 ± 0.02 for cenobamate, n = 7, P = 0.88, τWD; imal voltage for activation (V50,act) (−38.2 ± 4.1 mV and
1.1 ± 0.04 for control and 1.0 ± 0.03 for cenobamate, n = 7, −37.9 ± 4.3 mV in the absence and presence of 100 μM cenobamate,
P = 0.36, Fig. 1C). Cenobamate produced a more potent and pre- 0.3 ± 0.2 mV shift, n = 8, P = 0.41, Fig. 3C and D). In addition, cen-
ferential inhibition of INaP current compared with carbamazepine and obamate did not change the slope factor of voltage dependence (data
lamotrigine (Supplementary Fig. S1). not shown).
Next, the effect of cenobamate on the voltage dependence for the
3.2. Effect of cenobamate on Na+ currents by slow voltage-ramps steady-state fast inactivation of Na+ channels was examined. The INa
was induced by 50 ms depolarization test pulses to −20 mV after
The membrane currents induced by the slow voltage-ramp com- 500 ms prepulses from −120 to −30 mV in 10 mV increments
mand (−80 mV to +10 mV, 6 s, in every 10 s) were recorded from (Fig. 3E). The peak amplitudes of INa recorded at each prepulse voltages
hippocampal CA3 neurons. These slowly inactivating currents were were normalized to the maximal currents in the control condition, and
completely blocked either by 10 μM riluzole (Fig. 2Aa), which is known the data were fitted to the Boltzmann function (Fig. 3E and F).

177
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

Fig. 2. Effects of cenobamate on persistent Na+


currents in hippocampal CA3 neurons. A: Typical
traces of currents in the absence and presence of
10 μM riluzole (a) and 300 nM tetrodotoxin (TTX, b)
induced by slow voltage-ramp commands (15 mV/s).
The slow voltage-ramp current was obtained by
subtraction of raw traces (a–b). B: Typical traces of
slow voltage-ramp current in the absence and pre-
sence of 100 μM cenobamate. Ca: Conductance-vol-
tage relationships of voltage-gated Na+ channels in
the absence (open circles) and presence (closed cir-
cles) of 100 μM cenobamate. Each point represents
the mean ± S.E.M. from 6 experiments. The con-
ductance was calculated by using the following
equation: G = I/(V – Vrev), where I is the amplitude
of the INaP and Vrev is the reversal potential of Na+.
Cb: Effects of cenobamate (100 μM) on the midpoint
voltage for activation (V50,act, left) and slope factor.
Each column represents the mean ± S.E.M. from 6
experiments (6 animals). n.s., not significant. D:
Concentration-response relationships of cenobamate
against the INaP. Each point represents the
mean ± S.E.M. from 6 experiments (6 animals).

Cenobamate shifted the midpoint voltage for inactivation (V50,inact) control and 6.1 ± 0.4 for cenobamate, n = 8, P = 0.19). It should be
toward a hyperpolarizing range in a concentration-dependent manner; noted that cenobamate did not reduce the INa at hyperpolarized mem-
100 μM cenobamate shifted the V50,inact from −59.1 ± 3.1 to brane potentials (−120 to −100 mV) (Fig. 3F). When the extent of
−65.0 ± 2.6 mV (−5.9 ± 2.6 mV shift, n = 8, P < 0.01) (Fig. 3F cenobamate-induced hyperpolarizing shift was plotted and fitted (see
and G). However, cenobamate did not change the slope factor of vol- Methods), the corresponding KI and KR values of cenobamate were
tage dependence for steady-state fast inactivation (5.5 ± 0.2 for 48.2 ± 6.8 μM and 797.7 ± 155.4 μM, respectively (Fig. 3H).

Fig. 3. Effects of cenobamate on the current-voltage relationship of voltage-gated Na+ channels in hippocampal CA3 neurons. Aa: A schematic illustration of voltage
step pulses. The INa was induced by 50 ms depolarization pulses from −80 to +30 mV in 10 mV increments at a VH of
−80 mV. Ab: Typical traces of the INa induced by voltage step pulses in the absence (left) and presence (right) of 100 μM cenobamate. Each point represents the
mean ± S.E.M. from 8 experiments (8 animals). B: Current-voltage relationships of voltage-gated Na+ channels in the absence (open circles) and presence (closed
circles) of 100 μM cenobamate. Each point represents mean ± S.E.M. from 8 experiments (8 animals). C: Conductance-voltage relationships of voltage-gated Na+
channels in the absence (open circles) and presence (closed circles) of 100 μM cenobamate. Each point represents the mean ± S.E.M. from 8 experiments. The
conductance was calculated by using the following equation: G = I/(V – Vrev), where I is the peak amplitude of the INa and Vrev is the reversal potential of Na+. D:
Cenobamate-induced changes in the midpoint voltage for activation (V50,act) of voltage-gated Na+ channels. Each column represents the mean ± S.E.M. from 8
experiments (8 animals). *P < 0.05; n.s., not significant. Note that cenobamate at ≤100 μM concentrations did not change the midpoint voltage for activation. Ea: A
schematic illustration of voltage step pulses. The INa was induced by 50 ms depolarization pulses to −20 mV after 500 ms prepulses from −120 to −30 mV in 10 mV
increments. Eb: Typical traces of the INa induced by voltage step pulses in the absence (left) and presence (right) of 100 μM cenobamate. Each point represents the
mean ± S.E.M. from 8 experiments (8 animals). F: Current-voltage relationships of voltage-gated Na+ channels in the absence (open circles) and presence (closed
circles) of 100 μM cenobamate. Each point represents the mean ± S.E.M. from 8 experiments (8 animals). G: Cenobamate-induced changes in the midpoint voltage
for inactivation (V50,inact) of voltage-gated Na+ channels. Each column represents the mean ± S.E.M. from 8 experiments (8 animals). **P < 0.01; n.s., not
significant. H: The exp(ΔV50,inact/k) values were plotted against the cenobamate concentration. Each point represents the mean ± S.E.M. from 6 to 8 experiments (8
animals). The KR and KI values were 797.7 μM and 48.2 μM, respectively.

178
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

Fig. 4. Effect of cenobamate on the onset of


inactivation of voltage-gated Na+ channels
in hippocampal CA3 neurons. A: A sche-
matic illustration of voltage step pulses. The
INa was induced by the two-pulse protocol,
where the first conditioning pulses (P1; −20
mV depolarization, 2–8000 ms duration)
were followed by the second test pulses (P2;
−20 mV depolarization, 50 ms duration).
The second INa was recovered with an in-
terpulse interval of 20 ms at a potential of
−80 mV. B: Typical first (P1) and second
(P2) traces of INa induced by the two-pulse
protocol in the absence (left) and presence
(right) of 100 μM cenobamate. C: Kinetics
for the development of inactivation of vol-
tage-gated Na+ channels in the absence
(open circles) and presence (closed circles)
of 100 μM cenobamate. Each point re-
presents the mean ± S.E.M. from 9 ex-
periments (9 animals). D: Effects of cen-
obamate on the kinetic parameters (τfast,
τslow, Afast, Aslow, and τweighted) for the development of inactivation of voltage-gated Na+ channels. Each column represents the mean ± S.E.M. from 9 experiments
(9 animals). *P < 0.05, **P < 0.01.

3.4. Effect of cenobamate on inactivation kinetics of voltage-gated Na+ 62.4 ± 6.1 ms for cenobamate, n = 5, P < 0.01, Fig. 5D), and induced
channels a statistically significant decrease in Afast but not in Aslow (Afast;
0.66 ± 0.01 for control and 0.55 ± 0.01 ms for cenobamate, n = 5,
The effects of cenobamate on the development of inactivation of P < 0.01, Aslow; 0.36 ± 0.01 for control and 0.38 ± 0.12 ms for
voltage-gated Na+ channels were examined by use of the two-pulse cenobamate, n = 5, P = 0.26, respectively, Fig. 5D).
protocol, where the first conditioning pulse (P1) varying from 2 to
8000 ms in duration was followed by a second test pulse (P2) with an 3.5. Effect of cenobamate on the excitability of CA3 neurons
interpulse interval of 20 ms (Fig. 4A and B). The fraction of voltage-
gated Na+ channels available in the second test pulse was determined Persistent Na+ currents are closely related to the basal neuronal
as the P2/P1 ratio. Then, the P2/P1 ratio was plotted against the dura- excitability, including resting membrane potentials and firing patterns
tion of first conditioning pulse. The extent of the development of in- (Stafstrom, 2007). Therefore, we next examined the effect of cen-
activation was well-fitted to the double exponential function with the obamate on the excitability of CA3 pyramidal neurons in a current-
fast (τfast) and slow time constants (τslow) (Fig. 4C). Cenobamate clamp mode. In current-clamp conditions, resting membrane potentials
(100 μM) significantly decreased the τfast and τslow to 47.4 ± 3.2% and were −60.3 ± 2.1 mV (n = 9). Application of cenobamate hyperpo-
66.6 ± 7.8% of control, respectively (τfast; 445.2 ± 76.1 ms for con- larized membrane potentials in a concentration-dependent manner with
trol and 211.2 ± 12.8 ms for cenobamate, τslow; 6097.8 ± 631.3 ms an EC50 value of 89.8 ± 12.3 μM (n = 7, Fig. 6Aa and Ba). At a10 μM
for control and 4059.2 ± 454.5 ms for cenobamate, n = 9, P < 0.05, concentration, cenobamate elicited a hyperpolarization of
respectively, Fig. 4D). Cenobamate (100 μM) also decreased the 1.5 ± 0.3 mV (n = 7, P < 0.01). The cenobamate-induced hyperpo-
weighted time constant (τweighted) to 39.1 ± 7.9% of control larization completely disappeared in the presence of either 10 μM ri-
(5584.7 ± 551.3 ms for control and 2126.4 ± 257.0 ms for cen- luzole (−5.6 ± 0.5 mV and −0.2 ± 0.1 mV in the absence and pre-
obamate, n = 9, P < 0.01, Fig. 4D), and induced statistically sig- sence of riluzole, n = 6, P < 0.01) or 300 nM tetrodotoxin
nificant decreases in Afast and Aslow (Afast; 0.67 ± 0.01 for control and (−5.3 ± 0.4 mV and −0.1 ± 0.1 mV in the absence and presence of
0.61 ± 0.01 for cenobamate, n = 9, P < 0.05, Aslow; 0.34 ± 0.02 for tetrodotoxin, n = 6, P < 0.01) (Fig. 6Ab and Bb). In addition, riluzole
control and 0.23 ± 0.02 for cenobamate, n = 9, P < 0.01, respec- (10 μM) and tetrodotoxin (300 nM) by themselves hyperpolarized
tively, Fig. 4D). membrane potentials to −7.6 ± 1.1 mV (n = 6) and −11.4 ± 0.8 mV
The effects of cenobamate on the recovery from inactivation of (n = 6), respectively (Fig. 6Ab). The addition of either riluzole (10 μM)
voltage-gated Na+ channels were examined by use of the two-pulse or tetrodotoxin (300 nM) immediately after the application of cen-
protocol, where Na+ channels were fully inactivated by the first pulse obamate (100 μM) further hyperpolarized membrane potentials
(P1, 500 ms duration) and allowed to recover during interpulse inter- (Fig. 6Ac). These results suggest the riluzole-sensitive INaP is involved in
vals varying from 1 to 5000 ms before applying the second test pulse the excitability of CA3 pyramidal neurons, and that cenobamate can
(P2, 50 ms duration) (Fig. 5A and B). The fraction of voltage-gated Na+ reduce neuronal excitability by inhibiting the INaP. On the other hand,
channels available in the second test pulse was determined as the P2/P1 cenobamate had no effect on voltage-gated Ca+ or K+ channels
ratio. And then, the P2/P1 ratio was plotted against the duration of (Supplementary Fig. S2) or excitatory receptors, such as α-amino-3-
interpulse intervals. The extent of the recovery from inactivation was hydroxy-5-methylisoxazole-4-propionate (AMPA), kainic acid (KA),
well-fitted to the double exponential function (Fig. 5C). Cenobamate and N-methyl-D-aspartate (NMDA) receptors (Supplementary Fig. S3).
(100 μM) significantly increased the τfast and τslow to 248.9 ± 27.7% The effect of cenobamate was further examined on the firing pat-
and 523.7 ± 66.2% of the control, respectively (τfast, 4.7 ± 0.3 ms for terns in response to a depolarizing current injection (Fig. 6C and D).
control and 11.7 ± 1.1 ms for cenobamate; τslow, 25.7 ± 3.1 ms for While cenobamate (100 μM) significantly increased the rheobase cur-
control and 134.6 ± 11.8 ms for cenobamate, n = 5, P < 0.01, re- rents (65.3 ± 15.1 pA for control and 130.6 ± 19.9 pA for cen-
spectively, Fig. 5D). Cenobamate (100 μM) also increased the τweighted obamate, n = 9, P < 0.01, Fig. 6Da), the input resistance was not
to 520.2 ± 34.7% of control (12.0 ± 1.0 ms for control and changed by 100 μM cenobamate (251.1 ± 49.9 MΩ for control and

179
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

Fig. 5. Effect of cenobamate on the re-


covery from inactivation of voltage-gated
Na+ channels in hippocampal CA3 neurons.
A: A schematic illustration of voltage step
pulses. The INa was induced by the two-
pulse protocol, where the first conditioning
pulses (P1; −20 mV depolarization, 500 ms
duration) were followed by the second test
pulses (P2; −20 mV depolarization, 50 ms
duration). The second INa was recovered
with various interpulse intervals of
1–900 ms at a potential of −80 mV. B:
Typical first (P1) and second (P2) traces of
INa induced by the two-pulse protocol in the
absence (left) and presence (right) of
100 μM cenobamate. C: Kinetics for the re-
covery from inactivation of voltage-gated
Na+ channels in the absence (open circles)
and presence (closed circles) of 100 μM
cenobamate. Each point is the P2/P1 ratio of
INa induced by two depolarizing step pulses
and represents the mean ± S.E.M. from 5 experiments (5 animals). D: Effects of cenobamate on the kinetic parameters (τfast, τslow, Afast, Aslow, and τweighted) for the
recovery from inactivation of voltage-gated Na+ channels. Each column represents the mean ± S.E.M. from 5 experiments (5 animals). **P < 0.01; n.s., not
significant.

295.7 ± 49.2 MΩ for cenobamate, n = 9, P = 0.08, Fig. 6Db). Cen- of INaT, indicating that cenobamate had little effect on the transient
obamate (100 μM) also decreased the number of action potentials component of INa. However, cenobamate did potently inhibit the per-
generated by depolarizing current stimuli (Fig. 6Dc). sistent component of INa. We also found that cenobamate potently in-
hibited slow voltage-ramp-induced current in a concentration-depen-
4. Discussion dent manner, with an IC50 value of 53.1 μM. This IC50 value was similar
to that for the inhibition of INaP, suggesting that INaP might be com-
In this study, the investigational AED cenobamate did not sig- parable to the slow voltage-ramp-induced current. The much higher
nificantly alter the peak amplitude, onset time, or decay time constant IC50 value (> 500 μM) associated with inhibition of the INaT would

Fig. 6. Effects of cenobamate on the excitability of acutely isolated CA3 pyramidal neurons. Aa: Typical traces of voltage responses during the application of various
concentrations of cenobamate (3 μM–300 μM). Ab: Typical traces of voltage responses during the application of 100 μM cenobamate in the absence (left) and
presence of 10 μM riluzole (middle) or 300 nM TTX (right). TTX, tetrodotoxin. Ac: Typical traces of voltage responses during the consecutive application of 100 μM
cenobamate (C) and 10 μM riluzole (R) or 300 nM TTX (T). Ba: Concentration-response relationships of cenobamate (open circles) against the membrane hy-
perpolarization. The voltage of maximal hyperpolarization for cenobamate was set to 11.4 mV, the same value at which TTX induced membrane hyperpolarization.
Each point represents the mean ± S.E.M. from 4 to 8 experiments (8 animals). Note that either 10 μM riluzole (closed circle) or 300 nM TTX (closed rectangle)
greatly induced membrane hyperpolarization, as also shown in Ab. TTX, tetrodotoxin; Rilu, riluzole. Bb: Cenobamate (100 μM)-induced membrane hyperpolarization
in the absence and presence of 10 μM riluzole or 300 nM TTX. Each column represents the mean ± S.E.M. from 6 experiments (6 animals). **P < 0.01, n.s.; not
significant; TTX, tetrodotoxin. Note that cenobamate did not induce any membrane hyperpolarization after the blockade of INaP by riluzole or TTX. C: Typical traces
of voltage responses during depolarizing current injection in the absence (upper) and presence (lower) of 100 μM cenobamate. D: Effects of cenobamate (100 μM) on
the rheobase currents (a) and input resistance (Rinput, b). Each column represents the mean ± S.E.M. from 9 experiments (9 animals). **P < 0.01; CNB, cen-
obamate; Cont, control; n.s., not significant. Dc: Effect of cenobamate (100 μM) on the number of action potentials elicited by each depolarizing current injection
(1T–4T). T indicates the amplitude of minimal current to elicit action potentials. Each point represents the mean ± S.E.M. from 9 experiments.

180
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

suggest that cenobamate may preferentially inhibit the INaP. resulting in the facilitation of repetitive firing (Stafstrom, 2007). In
Furthermore, cenobamate showed a higher preference for the INaP in- pathological conditions, evidence suggests that the INaP is closely re-
hibition compared to either carbamazepine or lamotrigine; this may lated to epilepsy. For example, point mutations in the sodium channel,
contribute to the relative differences in efficacy and tolerability among voltage-gated alpha subunit NaV1.1, such as T875M, W1204R, and
these known agents. In addition, under the concentrations examined in R1648H, show impaired inactivation of Na+ channels, resulting in
these analyses, cenobamate would be expected to significantly inhibit generalized epilepsy with febrile seizures plus (GEFS+) (Lossin et al.,
the persistent component of INa at clinically therapeutic doses 2002). In addition, mutations in NaV1.2 (Scn2aQ54) also lead to im-
(equivalent to doses ranging from 100 to 400 mg/day). paired inactivation of Na+ channels and are associated with severe
As shown in this study, slow voltage-ramp-induced membrane epileptic behaviors (Bergren et al., 2005; Kearney et al., 2001). These
currents were recorded from the majority of acutely isolated CA3 pyr- mutant Na+ channels also show increased non-inactivating INa and/or
amidal neurons from rats. The major portion of these currents appeared INaP, suggesting possible roles of INaP in the pathophysiology of epilepsy
to be INaP-mediated by voltage-gated Na+ channels, as they were al- (Stafstrom, 2007; Sugawara et al., 2001). Furthermore, increased INaP
most completely blocked either by tetrodotoxin, a specific voltage- in the pilocarpine-induced status epilepticus rat model of temporal lobe
gated Na+ channel blocker, or riluzole, which is known to inhibit INaP epilepsy contributed to establishing chronic epilepsy in this rodent
(Urbani and Belluzzi, 2000). On the other hand, either riluzole or te- model (Chen et al., 2011). A recent study has shown that preferential
trodotoxin by itself induced membrane hyperpolarization in current- INaP blockade has anti-epileptic effects in Scn2aQ54 animals (Anderson
clamp experiments, suggesting that tonic INaP might participate in the et al., 2014). In fact, Na+ channel blockers that are used to treat epi-
regulation of resting membrane potentials of acutely isolated CA3 lepsy, such as phenytoin, valproic acid, carbamazepine, and lamo-
pyramidal neurons. Although the tetrodotoxin-induced membrane hy- trigine, are also known to block the INaP to some degree (Kohling, 2002;
perpolarization in the same neuron was found in our previous study Niespodziany et al., 2004; Spadoni et al., 2002). Given that the INaP
(Jang et al., 2006), further examination would be needed to elucidate blockade underlies the anti-epileptic effect mediated by cenobamate,
the detailed mechanism underlying the involvement of tonic INaP in the further investigation is needed to determine whether cenobamate in-
regulation of resting membrane potentials. hibits the INaP from mutated Na+ channels. In addition, it would be
As many commercially available AEDs affect the voltage depen- worth evaluating the clinical effect of cenobamate on epileptic patients
dence of Na+ channels (Catterall, 2014; Macdonald and Kelly, 1995; having mutated Na+ channels.
Spadoni et al., 2002), we also examined the effects of cenobamate on In current-clamp experiments, cenobamate hyperpolarized
the voltage-dependent activation and inactivation of Na+ channels. We membrane potentials in a concentration-dependent manner,
found that changes in V50,act values induced by cenobamate were not as reaching statistical significance at ≥10 μM cenobamate. The cen-
prominent as the dose-dependent changes in V50,inact values, suggesting obamate-induced hyperpolarization was completely blocked either
that the effect of cenobamate on Na+ channels is mainly mediated by by tetrodotoxin or riluzole, suggesting the involvement of INaP in
modulating the voltage-dependent fast inactivation rather than acti- such hyperpolarization. A hyperpolarization of membrane potentials
vation. Since cenobamate had no inhibitory effect on the INa at hy- can be mediated by the inhibition of excitatory receptors such as
perpolarized membrane potentials, a negative shift of steady-state fast AMPA, KA and NMDA receptors. However, these receptors were not
inactivation relationship might be responsible for the cenobamate-in- likely to be involved in the cenobamate-induced hyperpolarization,
duced decrease in transient INa at −80 mV. In addition, the cenobamate because cenobamate at a 100 μM concentration did not affect other
inhibition of peak INa of voltage-gated Na+ channels was more potent at voltage-gated ion channels and/or ionotropic glutamate receptors
depolarized membrane potentials, indicating that cenobamate has a that are able to induce membrane hyperpolarization (see
much higher affinity for the inactivated than for resting Na+ channels. Supplementary Figs. S2 and S3). We also found that cenobamate
This was further supported by the KI and KR values (48.2 μM and increased the rheobase currents rather than input resistance in CA3
797.8 μM, respectively). As a number of other Na+ channel blocking pyramidal neurons. Considering the minimal effect on the transient
AEDs are state-dependent and have a higher affinity for inactivated INa, the cenobamate-induced membrane hyperpolarization might be
Na+ channels, a higher affinity of cenobamate for the inactivated responsible for the increase in threshold for the generation of action
channels than for resting Na+ channels would be expected to contribute potentials. More importantly, cenobamate reduced the number of
to its anti-epileptic effects (Macdonald and Kelly, 1995). The kinetics action potentials triggered by depolarizing current stimuli. Given
regarding the inactivation property of voltage-gated Na+ channels, that the INaP is involved in the repetitive and burst generation of
such as the development of inactivation and recovery from fast in- action potentials in central neurons (Wu et al., 2005; Yue et al.,
activation, are essential for excessive neuronal excitability, such as re- 2005), the preferential blockade of INaP by cenobamate may con-
petitive and bursting generation of action potentials. In the present tribute to its anti-epileptic activity against bursting neurons.
study, cenobamate significantly accelerated the overall inactivation In addition to the INaP, hyperpolarization-gated and cyclic nu-
kinetics, and the extent of this acceleration was higher in the fast (τfast) cleotide-activated cation (HCN) currents mediated by HCN channels
than in the slow component (τslow). Assuming that the two time con- are also involved in the repetitive and rhythmic generation of action
stants reflect the extent of fast and slow inactivation, respectively, potentials in various central neurons (Biel et al., 2009; Khaliq and
cenobamate might preferentially accelerate the fast inactivation during Bean, 2010; Puopolo et al., 2007). In the present study, we could not
sustained membrane depolarization. In addition, we found that cen- record any HCN channel-mediated voltage responses in acutely iso-
obamate significantly decreased the overall recovery kinetics, and the lated CA3 pyramidal neurons. Furthermore, cenobamate did not af-
extent of the decrease was greater in the τslow than in the τfast, in- fect the voltage changes in response to hyperpolarizing current sti-
dicating that cenobamate would further retard the recovery from slow muli, suggesting that the INaP rather than HCN channel-mediated
inactivation. In this stage, the mechanisms underlying the cenobamate- currents is a pharmacologic target for cenobamate. In fact, HCN
induced changes in inactivation kinetics are unknown. However, given channel inhibitors do not show effects on the firing pattern in several
that the INaP is a non-inactivating component of INa during sustained neurons that display repetitive firing activity (Khaliq and Bean,
membrane depolarization, and that cenobamate preferentially inhibits 2010; Taddese and Bean, 2002).
the INaP than the INaT, the suppression of INaP by cenobamate might In conclusion, we have examined the effects of cenobamate, a newly
result in the change of inactivation and/or recovery kinetics of Na+ developed AED, on neuronal voltage-gated Na+ channels. In addition to
channels. its modulation of several properties of voltage-gated Na+ channels,
Despite the small amplitude, the INaP, in particular at subthreshold cenobamate acts as a preferential INaP inhibitor in neuronal voltage-
membrane potentials, plays pivotal roles in neuronal excitability, gated Na+ channels to exert its anti-epileptic efficacy.

181
M. Nakamura, et al. European Journal of Pharmacology 855 (2019) 175–182

Conflicts of interest Conference (EILAT XI). Epilepsy Res. 103 (1), 2–30.
Biel, M., Wahl-Schott, C., Michalakis, S., Zong, X., 2009. Hyperpolarization-activated
cation channels: from genes to function. Physiol. Rev. 89 (3), 847–885.
MN: Declarations of interest: none. Brodie, M.J., Barry, S.J., Bamagous, G.A., Norrie, J.D., Kwan, P., 2012. Patterns of
JHC: Declarations of interest: none. treatment response in newly diagnosed epilepsy. Neurology 78 (20), 1548–1554.
Catterall, W.A., 2014. Sodium channels, inherited epilepsy, and antiepileptic drugs. Annu.
HS: Employee of SK Biopharmaceuticals. Rev. Pharmacol. Toxicol. 54, 317–338.
ISJ: Declarations of interest: none. Chen, S., Su, H., Yue, C., Remy, S., Royeck, M., Sochivko, D., Opitz, T., Beck, H., Yaari, Y.,
2011. An increase in persistent sodium current contributes to intrinsic neuronal
bursting after status epilepticus. J. Neurophysiol. 105 (1), 117–129.
Authors’ contributions Chen, Z., Brodie, M.J., Liew, D., Kwan, P., 2018. Treatment outcomes in patients with
newly diagnosed epilepsy treated with established and new antiepileptic drugs: a 30-
year longitudinal cohort study. JAMA Neurol 75 (3), 279–286.
MN carried out the electrophysiological studies and drafted/criti-
Crill, W.E., 1996. Persistent sodium current in mammalian central neurons. Annu. Rev.
cally reviewed the manuscript. Physiol. 58, 349–362.
JHC carried out the electrophysiological studies. Das, N., Dhanawat, M., Shrivastava, S.K., 2012. An overview on antiepileptic drugs. Drug
Discov Ther 6 (4), 178–193.
HS participated in designing the study and critically reviewed the Hains, B.C., Waxman, S.G., 2007. Sodium channel expression and the molecular patho-
manuscript. physiology of pain after SCI. Prog. Brain Res. 161, 195–203.
ISJ participated in designing the study, carried out the electro- Jackson, A.C., Yao, G.L., Bean, B.P., 2004. Mechanism of spontaneous firing in dor-
somedial suprachiasmatic nucleus neurons. J. Neurosci. 24 (37), 7985–7998.
physiological studies, and drafted/critically reviewed the manuscript. Jang, I.S., Nakamura, M., Ito, Y., Akaike, N., 2006. Presynaptic GABAA receptors facilitate
All authors read and approved the final manuscript. spontaneous glutamate release from presynaptic terminals on mechanically dis-
sociated rat CA3 pyramidal neurons. Neuroscience 138 (1), 25–35.
Kearney, J.A., Plummer, N.W., Smith, M.R., Kapur, J., Cummins, T.R., Waxman, S.G.,
Prior presentation/publication Goldin, A.L., Meisler, M.H., 2001. A gain-of-function mutation in the sodium channel
gene Scn2a results in seizures and behavioral abnormalities. Neuroscience 102 (2),
307–317.
Nakamura M et al. Mechanism of action of cenobamate: preferential Khaliq, Z.M., Bean, B.P., 2010. Pacemaking in dopaminergic ventral tegmental area
inhibition of the persistent sodium current. Presented at the 2018 neurons: depolarizing drive from background and voltage-dependent sodium con-
Annual Meeting of the American Academy of Neurology, April 21–27, ductances. J. Neurosci. 30 (21), 7401–7413.
Kohling, R., 2002. Voltage-gated sodium channels in epilepsy. Epilepsia 43 (11),
Los Angeles, CA. Poster P5.278. 1278–1295.
Kwan, P., Brodie, M.J., 2000. Early identification of refractory epilepsy. N. Engl. J. Med.
342 (5), 314–319.
Funding & role of sponsor Lossin, C., Wang, D.W., Rhodes, T.H., Vanoye, C.G., George Jr., A.L., 2002. Molecular
basis of an inherited epilepsy. Neuron 34 (6), 877–884.
This study was supported by SK Biopharmaceuticals Co., Ltd. Macdonald, R.L., Kelly, K.M., 1995. Antiepileptic drug mechanisms of action. Epilepsia 36
(Suppl. 2), S2–S12.
Medical writing support was funded by SK Life Science, Inc. As an Murase, K., Ryu, P.D., Randic, M., 1989. Excitatory and inhibitory amino acids and
employee of the study sponsor, HS participated in designing the study, peptide-induced responses in acutely isolated rat spinal dorsal horn neurons.
Neurosci. Lett. 103 (1), 56–63.
critically reviewed the manuscript, and read and approved the final
Niespodziany, I., Klitgaard, H., Margineanu, D.G., 2004. Is the persistent sodium current a
manuscript. specific target of anti-absence drugs? Neuroreport 15 (6), 1049–1052.
Park, Y.Y., Johnston, D., Gray, R., 2013. Slowly inactivating component of Na+ current in
peri-somatic region of hippocampal CA1 pyramidal neurons. J. Neurophysiol. 109
Acknowledgments (5), 1378–1390.
Parri, H.R., Crunelli, V., 1998. Sodium current in rat and cat thalamocortical neurons:
The authors thank Sarah Mizne, PharmD, and Miriam Gitler, PhD, of role of a non-inactivating component in tonic and burst firing. J. Neurosci. 18 (3),
854–867.
MedVal Scientific Information Services, LLC, for providing medical Puopolo, M., Raviola, E., Bean, B.P., 2007. Roles of subthreshold calcium current and
writing and editorial support, which were funded by SK Life Science, sodium current in spontaneous firing of mouse midbrain dopamine neurons. J.
Neurosci. 27 (3), 645–656.
Inc. This manuscript was prepared according to the International Rhee, J.S., Ishibashi, H., Akaike, N., 1999. Calcium channels in the GABAergic pre-
Society for Medical Publication Professionals' “Good Publication synaptic nerve terminals projecting to meynert neurons of the rat. J. Neurochem. 72
Practice for Communicating Company-Sponsored Medical Research: (2), 800–807.
Spadoni, F., Hainsworth, A.H., Mercuri, N.B., Caputi, L., Martella, G., Lavaroni, F.,
GPP3.” Bernardi, G., Stefani, A., 2002. Lamotrigine derivatives and riluzole inhibit INa,P in
cortical neurons. Neuroreport 13 (9), 1167–1170.
Appendix A. Supplementary data Stafstrom, C.E., 2007. Persistent sodium current and its role in epilepsy. Epilepsy Curr. 7
(1), 15–22.
Sugawara, T., Tsurubuchi, Y., Agarwala, K.L., Ito, M., Fukuma, G., Mazaki-Miyazaki, E.,
Supplementary data to this article can be found online at https:// Nagafuji, H., Noda, M., Imoto, K., Wada, K., Mitsudome, A., Kaneko, S., Montal, M.,
Nagata, K., Hirose, S., Yamakawa, K., 2001. A missense mutation of the Na+ channel
doi.org/10.1016/j.ejphar.2019.05.007.
aII subunit gene Nav1.2 in a patient with febrile and afebrile seizures causes channel
dysfunction. Proc. Natl. Acad. Sci. U. S. A. 98 (11), 6384–6389.
References Taddese, A., Bean, B.P., 2002. Subthreshold sodium current from rapidly inactivating
sodium channels drives spontaneous firing of tuberomammillary neurons. Neuron 33
(4), 587–600.
Akaike, N., Moorhouse, A.J., 2003. Techniques: applications of the nerve-bouton pre- Tian, N., Boring, M., Kobau, R., Zack, M.M., Croft, J.B., 2018. Active epilepsy and seizure
paration in neuropharmacology. Trends Pharmacol. Sci. 24 (1), 44–47. control in adults - United States, 2013 and 2015. MMWR Morb. Mortal. Wkly. Rep. 67
Anderson, L.L., Thompson, C.H., Hawkins, N.A., Nath, R.D., Petersohn, A.A., Rajamani, (15), 437–442.
S., Bush, W.S., Frankel, W.N., Vanoye, C.G., Kearney, J.A., George Jr., A.L., 2014. Urbani, A., Belluzzi, O., 2000. Riluzole inhibits the persistent sodium current in mam-
Antiepileptic activity of preferential inhibitors of persistent sodium current. Epilepsia malian CNS neurons. Eur. J. Neurosci. 12 (10), 3567–3574.
55 (8), 1274–1283. Vreugdenhil, M., Hoogland, G., van Veelen, C.W., Wadman, W.J., 2004. Persistent sodium
Bean, B.P., 1984. Nitrendipine block of cardiac calcium channels: high-affinity binding to current in subicular neurons isolated from patients with temporal lobe epilepsy. Eur.
the inactivated state. Proc. Natl. Acad. Sci. U. S. A. 81 (20), 6388–6392. J. Neurosci. 19 (10), 2769–2778.
Bellingham, M.C., 2011. A review of the neural mechanisms of action and clinical effi- Wu, N., Enomoto, A., Tanaka, S., Hsiao, C.F., Nykamp, D.Q., Izhikevich, E., Chandler,
ciency of riluzole in treating amyotrophic lateral sclerosis: what have we learned in S.H., 2005. Persistent sodium currents in mesencephalic v neurons participate in
the last decade? CNS Neurosci. Ther. 17 (1), 4–31. burst generation and control of membrane excitability. J. Neurophysiol. 93 (5),
Bennett, B.D., Callaway, J.C., Wilson, C.J., 2000. Intrinsic membrane properties under- 2710–2722.
lying spontaneous tonic firing in neostriatal cholinergic interneurons. J. Neurosci. 20 Xie, R.G., Zheng, D.W., Xing, J.L., Zhang, X.J., Song, Y., Xie, Y.B., Kuang, F., Dong, H.,
(22), 8493–8503. You, S.W., Xu, H., Hu, S.J., 2011. Blockade of persistent sodium currents contributes
Bergren, S.K., Chen, S., Galecki, A., Kearney, J.A., 2005. Genetic modifiers affecting se- to the riluzole-induced inhibition of spontaneous activity and oscillations in injured
verity of epilepsy caused by mutation of sodium channel Scn2a. Mamm. Genome 16 DRG neurons. PLoS One 6 (4), e18681.
(9), 683–690. Yue, C., Remy, S., Su, H., Beck, H., Yaari, Y., 2005. Proximal persistent Na+ channels
Bialer, M., Johannessen, S.I., Levy, R.H., Perucca, E., Tomson, T., White, H.S., 2013. drive spike afterdepolarizations and associated bursting in adult CA1 pyramidal cells.
Progress report on new antiepileptic drugs: a summary of the Eleventh Eilat J. Neurosci. 25 (42), 9704–9720.

182

You might also like