Next Article in Journal
Microbiota Dysbiosis and Gut Barrier Dysfunction Associated with Non-Alcoholic Fatty Liver Disease Are Modulated by a Specific Metabolic Cofactors’ Combination
Next Article in Special Issue
Acute Effects of Ocrelizumab Infusion in Multiple Sclerosis Patients
Previous Article in Journal
circSMARCA5 Is an Upstream Regulator of the Expression of miR-126-3p, miR-515-5p, and Their mRNA Targets, Insulin-like Growth Factor Binding Protein 2 (IGFBP2) and NRAS Proto-Oncogene, GTPase (NRAS) in Glioblastoma
Previous Article in Special Issue
Ella versus Simoa Serum Neurofilament Assessment to Monitor Treatment Response in Highly Active Multiple Sclerosis Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rufinamide, a Triazole-Derived Antiepileptic Drug, Stimulates Ca2+-Activated K+ Currents While Inhibiting Voltage-Gated Na+ Currents

1
Department of Pediatrics, Chi-Mei Medical Center, Tainan 71004, Taiwan
2
Department of Physiology, College of Medicine, National Cheng Kung University, Tainan 70101, Taiwan
3
Institute of Basic Medical Sciences, College of Medicine, National Cheng Kung University, Tainan 70101, Taiwan
4
Department of Neurology, National Cheng Kung University Hospital, College of Medicine, National Cheng Kung University, Tainan 70101, Taiwan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13677; https://doi.org/10.3390/ijms232213677
Submission received: 29 September 2022 / Revised: 27 October 2022 / Accepted: 3 November 2022 / Published: 8 November 2022
(This article belongs to the Special Issue New Mechanisms and Therapeutics in Neurological Diseases 2.0)

Abstract

:
Rufinamide (RFM) is a clinically utilized antiepileptic drug that, as a triazole derivative, has a unique structure. The extent to which this drug affects membrane ionic currents remains incompletely understood. With the aid of patch clamp technology, we investigated the effects of RFM on the amplitude, gating, and hysteresis of ionic currents from pituitary GH3 lactotrophs. RFM increased the amplitude of Ca2+-activated K+ currents (IK(Ca)) in pituitary GH3 lactotrophs, and the increase was attenuated by the further addition of iberiotoxin or paxilline. The addition of RFM to the cytosolic surface of the detached patch of membrane resulted in the enhanced activity of large-conductance Ca2+-activated K+ channels (BKCa channels), and paxilline reversed this activity. RFM increased the strength of the hysteresis exhibited by the BKCa channels and induced by an inverted isosceles-triangular ramp pulse. The peak and late voltage-gated Na+ current (INa) evoked by rapid step depolarizations were differentially suppressed by RFM. The molecular docking approach suggested that RFM bound to the intracellular domain of KCa1.1 channels with amino acid residues, thereby functionally affecting BKCa channels’ activity. This study is the first to present evidence that, in addition to inhibiting the INa, RFM effectively modifies the IK(Ca), which suggests that it has an impact on neuronal function and excitability.

1. Introduction

Rufinamide (RFM, Banzel®, Inovelon®, 1-((2,6-difluorophenyl)methyl)-1H-1,2,3-triazole-4carboxamide), is recognized as a unique anticonvulsant drug because, as a triazole derivative, its structure is dissimilar to other currently marketed antiepileptic drugs [1,2,3,4]. It is increasingly being used in combination with other medications and therapies to treat Lennox-Gastaut syndrome, severe epileptic encephalopathy, and other seizure disorders [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26]. RFM has been shown to decrease the incidence and severity of seizures associated with Lennon-Gastaut syndrome with a favorable cognitive side-effect profile [6,21,22,27]. However, the answers that have been provided so far to the questions of whether and how RFM causes any integrated effects of transmembrane ionic currents are rather speculative.
Although its mechanism of action as an antiepileptic drug is still unclear, RFM modulates the activity of NaV channels by prolonging the inactive state of these channels [28,29,30,31,32,33,34] and the specific stabilization of the intermediate inactivated NaV channels [35]. It has previously been shown to improve functional and behavioral deficits through a mechanism linked to its blockade of tetrodotoxin-resistant Na+ currents [32]. However, whether this drug modifies the amplitude and gating of other types of ionic currents is unclear.
RFM has been shown to exert a neuroprotective effect against kainic acid-induced excitotoxic neuronal death in the hippocampus [36,37]. It has also shown promise by improving cognition and ischemia-reperfusion injury and by increasing neurogenesis in the gerbil hippocampus [38]. It has also proven beneficial in reducing myotonia in skeletal muscles [39]. However, there is evidence that atonic seizures may be aggravated by RFM [40].
Large-conductance Ca2+-activated K+ (BKCa or BK) channels (KCa1.1, KCNMA1, Slo1) belong to a family of voltage-gated K+ channels, and they are activated by an increase in either the cytosolic concentration of Ca2+ or the membrane potential, or both. The activation of BKCa channels can conduct large amounts of K+ ions across the cell membrane. Owing to its high-conductance state and a single-channel conductance of about 150–250 pS, the BKCa channel is also thought to be a maxi- or large-K+ channel. This family of K+ channels is functionally expressed in an array of excitable and non-excitable cells, and its activity can play a role in numerous physiological or pathological events, including membrane excitability, neurotransmitter release, stimulus-secretion coupling, muscle relaxation, and motor coordination [41,42,43,44,45,46,47,48]. Small natural and synthetic molecules have been shown to be important regulators of BKCa-channel activity [45,49,50,51,52]. It has also been noted that the BKCa channel shows promise in approaches to treating epilepsy [41,42,43,53,54].
In view of the considerations stated above, the purpose of this work is to establish whether RFM causes any modifications in membrane ionic currents in electrically excitable cells (e.g., pituitary GH3 lactotrophs). The types of membrane ionic currents that we studied include the Ca2+-activated K+ current (IK(Ca)) and the voltage-gated Na+ current (INa). The present study is the first to present evidence that, in addition to inhibiting the INa, RFM is capable of effectively stimulating activity in BKCa channels.

2. Results

2.1. Effect of RFM on CA2+-Activated K+ Current (IK(CA)) Identified in GH3 Cells

In the first stage of experiments, the modification of the IK(Ca) caused by exposing GH3 cells to RFM was evaluated. We bathed the cells in normal Tyrode’s solution, which contained 1.8 mM CaCl2, and the recording pipette was backfilled with a K+-containing isotonic solution. The membrane patch under the electrode tip was broken through by gentle suction in order to conduct the whole-cell current recordings. We applied the voltage clamp technique to each tested cell at 0 mV, and we subjected it to a 300-ms depolarizing voltage command pulse to +50 mV. Under this experimental protocol, a large and noisy outward current in response to the depolarizing step was robustly evoked, and the magnitude of the current evidently rose with increasing membrane depolarizations. Such ionic currents have been characterized as IK(Ca) [52,53,54,55]. Of note is the finding that, after the cells had been exposed for 1 min to RFM at a concentration of 3 μM or 10 μM, a concentration-dependent increase in the amplitude of the IK(Ca) was observed (see Figure 1A). More specifically, compared to a control value of 153 ± 19 pA (n = 8), the addition of 3 μM RFM produced an increase in the current amplitude at the end of the step depolarization to 243 ± 32 pA (n = 8, p < 0.05), whereas the addition of 10 μM RFM produced an increase to 298 ± 39 pA (n = 8, p < 0.05). After washing out the RFM, the current amplitude returned to 159 ± 21 pA (n = 8).
The varying stimulating effects of different concentrations of RFM on the amplitude of the IK(Ca) elicited by the 300 ms depolarizing step were further examined, with the effects on GH3 cells shown in Figure 1B. According to a Hill equation detailed in the section titled Materials and Methods, the EC50 value of RFM required for evoking the IK(Ca) was estimated to be 3.9 μM, with a Hill coefficient of 1.2.

2.2. Effect of RFM on the Current-Voltage (I–V) Relationship the IK(Ca) in GH3 Cells

We next investigated the effect of RFM on the IK(Ca) for various membrane potentials. In these experiments, the IK(Ca) was robustly evoked as each tested cell was depolarized from 0 mV by a series of depolarizing voltage steps to a range of potentials between 0 mV and +90 mV in 10-mV increments (see Figure 2A). The mean I–V relationships of the IK(Ca) with and without the addition of RFM are shown in Figure 2B. The whole-cell conductance of the IK(Ca) was measured at membrane potentials between +40 mV and +90 mV, and the addition of 10 μM RFM caused the IK(Ca) conductance to increase from 7.2 ± 0.6 nS to 15.1 ± 0.9 nS (n = 8, p < 0.05). After washing out the RFM, the conductance returned to 7.3 ± 0.7 nS (n = 8).

2.3. Comparisons among the Effects of RFM Only, RFM Plus Apamin, RFM Plus Glibenclamide, RFM Plus Iberiotoxin, and RFM Plus Paxilline on the Amplitude of the IK(Ca)

We next examined whether the subsequent addition of apamin, glibenclamide, iberiotoxin, and paxilline, while retaining the RFM, would alter the stimulatory effect of RFM on the IK(Ca) in GH3 cells. Evidence has been found that apamin blocks the activity of small-conductance Ca2+-activated K+ channels, glibenclamide is an inhibitor of ATP-sensitive K+ (KATP) channels, and iberiotoxin and paxilline are effective at suppressing BKCa-channel activity [51,56,57]. As demonstrated in the scatter graph presented in Figure 3, in the continued presence of 10 μM RFM, the addition of neither apamin nor glibenclamide modified the RFM-induced increase in the amplitude of the IK(Ca). However, the subsequent addition of iberiotoxin and paxilline reversed the stimulation of the IK(Ca) produced by the presence of RFM. It, thus, seems likely that the RFM-mediated stimulation of the IK(Ca) is caused predominantly by the current flow through the BKCa channels present in GH3 cells.

2.4. Stimulatory Effect of RFM on Large-Conductance CA2+-Activated K+ (BKCa) Channels in GH3 Cells

In terms of its biophysical characteristics, the IK(Ca) is a large, noisy, voltage-dependent, Ca2+-sensitive K+ current, and its strength largely comes from the opening of the BKCa channels inherently in GH3 cells [45]. Therefore, we continued to explore the possible effects that RFM has on the activity of BKCa channels, and to this end, the inside-out patch clamp technique was performed. The recordings of the currents in cells that were bathed in a symmetrical K+ concentration (145 mM) and bath medium contained 0.1 μM Ca2+. While taking measurements, we kept the examined cell in a voltage clamp at a holding potential of +60 mV. As illustrated in Figure 4A, after the introduction of 10 μM RFM into the part of the ion channel near the cytosolic membrane leaflet, a drastic elevation in the open-state probability of the channel was detected. Recorded from the detached patches of GH3 cells, these probabilities for the BKCa channels significantly and consistently increased from 0.093 ± 0.012 to 0.142 ± 0.019 (n = 7, p < 0.05) during exposure to 3 μM RFM and to 0223 ± 0.025 (n = 7, p < 0.05) during exposure to 10 μM RFM. After washing out the compound, the channel activity was reduced to 0.101 ± 0.014 (n = 7, p < 0.05).
As shown in Figure 4B, in the continued presence of 10 μM RFM, the subsequent addition of 1 μM paxilline effectively attenuated the RFM-stimulated activity of the channels, which was not the case when 1 μM TRAM-34 was introduced. Paxilline is known to be an inhibitor of the BKCa channels, while TRAM-34 has been shown to suppress the activity of intermediate-conductance Ca2+-activated K+ (IKCa) channels [57,58]. Moreover, the mean open times of the BKCa channels in the absence and presence of 10 μM RFM did not differ significantly (2.32 ± 0.11 for the control versus 2.33 ± 0.12 in the presence of RFM, n = 7, p > 0.05). However, the slow component of the mean closed time of the channels in the presence of 10 μM RFM decreased significantly to 6.17 ± 0.57 (n = 7, p < 0.05) from a control value of 12.22 ± 0.89 (n = 7).

2.5. Effect of RFM on the Voltage-Dependent Hysteresis of BKCa-Channel Activity Evoked in Response to a Long-Lasting Inverted Isosceles-Triangular Ramp Pulse

The voltage-dependent hysteresis of ionic currents (i.e., a lag in current strength as the linear voltage command is changed in the opposite direction) has recently been shown to have a significant impact reminiscent of electrical activity, as when an action potential fires in a neuron (i.e., initial depolarization and late repolarization) [59,60,61,62]. In view of such findings, we further explored whether voltage-dependent hysteresis occurred in the BKCa-channel activity recorded in GH3 cells. In this series of inside-out patch clamp experiments, we used a long (1.6 s in duration) inverted isosceles-triangular ramp pulse with a ramp speed of ±450 mV/s to measure the characteristics of the hysteretic behavior (Figure 5). The trajectory of the channel activity evoked by the downsloping ramp pulse (i.e., a change in voltage from +180 to −180 mV) and the upsloping ramp pulse (i.e., a change from −180 to +180 mV) as a function of time was clearly distinguishable. In other words, in the control phase (i.e., when the RFM was not present), the relative open probability of the channels evoked by the descending (forward) end of the inverted triangular voltage ramp was higher than that in response to the upsloping (backward) end (Figure 5C).
With the goal of quantifying our observations, we evaluated the degree of voltage-dependent hysteresis based on the voltage separation between the downsloping and upsloping branches at 50% of the relative probability of the BKCa channels being open. In a manner dependent on the concentration used, the presence of RFM increased the overall strength of the hysteresis exhibited by the BKCa channels in the GH3 cells (Figure 5D). More specifically, as the single-channel recordings were taken, the addition of 3 μM RFM led to an increase in the strength of the hysteresis to 22.1 ± 1.3 mV (n = 7, p < 0.05), whereas the addition of 10 μM RFM increased it to 25.1 ± 1.6 mV (n = 7, p < 0.05), from a control value of 14.9 ± 1.1 mV (n = 7).

2.6. Effect of RFM on Voltage-Gated NA+ Currents (INa) Recorded in GH3 Cells

Earlier studies have demonstrated the effectiveness of RFM in altering the magnitude of the INa in different cell types [28,29,30,31,32,33,34,35]. We further examined whether the addition of RFM would exert any perturbations on the magnitude of the INa. In these experiments, to preclude the contamination of Ca2+ and K+ currents, we bathed the cells in Ca2+-free Tyrode’s solution containing 10 mM TEA and 0.5 mM CdCl2, and the recording electrodes were backfilled with Cs+-containing solution [63,64]. The cells were exposed to 3 μM RFM, and the peak and late components of the INa activated by a 30 ms depolarizing step from −80 mV to −10 mV were robustly decreased. Concomitant with these results, the time course of INa inactivation was shortened in the presence of 3 μM RFM (Figure 6A,B). The application of 3 μM RFM decreased the peak amplitude of the INa to 1201 ± 96 pA (n = 8, p < 0.05) from a control value of 1287 ± 103 pA (n = 8) and the late amplitude of the INa to 21 ± 3 pA (n = 8, p < 0.05) from a control value of 48 ± 4 pA (n = 8). After washing out the RFM, the peak and late INa returned to 1209 ± 99 pA (n = 8) and 46 ± 4 pA (n = 8), respectively. Furthermore, the exposure of the GH3 cells to 3 μM RFM resulted in an escalated time course of INa inactivation, as shown by the reduction in τinact(S) (the slow component of the inactivation time constant of the INa) from 11.2 ± 0.8 ms to 4.1 ± 0.2 ms (n = 8, p < 0.05).
The effect that the addition of different concentrations of RFM had on suppressing the peak and late INa evoked by the 30 ms depolarizing step was further examined. This inhibitory effect on the amplitude of the INa in the GH3 cells is shown in Figure 6C. According to a Hill equation presented in the section titled Materials and Methods, the IC50 values of the RFM required for inhibiting the peak and late INa activated by rapid step depolarizations were 22.7 μM and 3.1 μM, respectively.

2.7. Effect of RFM on Mean I–V Relationship of the Peak INa Identified in GH3 Cells

We further examined the I–V relationships of the peak INa obtained with and without the application of RFM. These voltage-clamp experiments were conducted in cells held at −80 mV, and a series of voltage pulses ranging between −80 mV and +40 mV was then applied to the tested cells. As shown in Figure 7, the presence of 3 μM RFM did not alter the overall I–V relationship of the peak INa, although it did increase the peak amplitude of the INa, particularly at the level of −10 mV. The values of the reversal potential and the threshold potential of the peak INa recorded both in the absence and in the presence of RFM were indistinguishable. The I–V curves obtained during the control period and during cell exposure to 3 μM RFM were optimally fitted with a Boltzmann function, as indicated in the section titled Materials and Methods (In control: (i.e., the absence of RFM), G = 3.8 ± 0.3 nS, Vh = −20.8 ± 1.9 mV, k = 3.3 ± 0.3 (n = 8), while in the presence of 3 μM RFM: G = 2.9 ± 0.3 nS, Vh = −20.4 ± 1.8 mV, k = 3.4 ± 0.3 (n = 8)). It is thus likely that the steady-state activation curve of the INa was not changed during cell exposure to 3 μM RFM, although this concentration of RFM decreased the whole-cell conductance of the peak INa.

2.8. RFM-Mediated Attenuation of the Stimulation of the INa Produced by Tefluthrin (Tef)

Tefluthrin (Tef), a type-I pyrethroid insecticide, has been shown to activate the INa [63,64]. We further investigated whether the subsequent addition of RFM would modify the Tef-activated INa in GH3 cells. As shown in Figure 8, when the cells were exposed to 10 mM Tef, the peak amplitude and the τinact(S) of the INa activated by the application of abrupt step depolarizations from −80 mV to −10 mV were 669 ± 31 pA and 26 ± 9 ms (n = 8), respectively. In the continued presence of 10 mM Tef, the further addition of 10 μM RFM reduced the current amplitude and the tinact(S) to 432 ± 27 pA and 21 ± 7 ms (n = 8, p < 0.05), respectively, whereas adding 30 μM RFM resulted in the current amplitude and the tinact(S) being decreased to 258 ± 19 pA and 16 ± 5 ms (n = 8, p < 0.05), respectively. It is clear, therefore, that while the cells were exposed to Tef, the addition of RFM was effective at modifying the Tef-activated INa in the GH3 cells.

3. Discussion

This study has shown that, as cells were continuously exposed to RFM, depending on its concentration, the amplitude of IK(Ca) increased with an EC50 value of 3.9 μM. The BKCa channels were stimulated by the presence of RFM, with their activity detected using the inside-out configuration of the patch-clamp technique, although no discernible change was detected in the single-channel conductance. Therefore, the BKCa channel is expected to be a relevant target for RFM treatment.
In our study, the RFM-mediated increase in BKCa-channel activity in GH3 cells was not associated with a change in single-channel amplitude, as was confirmed by the absence of a noticeable difference in the single-channel conductance of the channels measured with and without the addition of RFM. However, the RFM-stimulated activity of the BKCa channels was attenuated by the further addition of paxilline, which was not the case when TRAM-34 was added. Additionally, the presence of RFM was shown to shorten the slow component of the mean closed time of the channel. This may be a major factor in the RFM-mediated activation of the BKCa channels, although no change in the mean open time of the channel was found.
It needs to be emphasized that the amplitude of IK(Ca) shown in Figure 1A in control conditions at the level of +50 mV was around 150 pA. Since the single-channel conductance of the BKCa channels was estimated to be around 200 pS, we might expect a single-channel current of 10 pA at +50 mV (Figure 4). According to this, the whole-cell current demonstrated herein is very low, indicating a very small open probability of the BKCa channel. The results are also consistent with the BKCa-channel activity at +50 mV shown in Figure 5. This remark applied to the RFM-stimulated current, which was just twice the control. Therefore, it is worthy of being further investigated, whether RFM-stimulated IK(Ca) magnitude could be greatly enhanced as the channel number in different cell types is raised.
The IC50 values for the RFM-mediated inhibition of the peak and late INa in this study were estimated to be 22.7 μM and 3.1 μM, respectively. Our findings showed the effectiveness of RFM in causing the differential inhibition of the peak and late components of the INa in GH3 cells and in hippocampal mHippoE-124 neurons (see the Supplementary Information). In the continued presence of Tef, the addition of RFM led to an attenuation of the Tef-activated INa. It also needs to be emphasized that BKCa-channel activity has been identified in human cardiac fibroblasts [65]. It is thus possible that the added RFM interacts with BKCa channels to raise the amplitude of IK(Ca) in those cells which might be electrically coupled to heart cells.
Another important finding in this study is the occurrence of the voltage-dependent hysteresis of the activity of single BKCa channels activated in response to an inverted isosceles-triangular ramp pulse, in situations where the intracellular surfaces of detached membrane patches from GH3 cells were exposed to varying concentrations of RFM. With the increase in the RFM concentration, the strength of the hysteresis exhibited by the channels (i.e., the difference in voltage between the forward and backward limbs at 50% channel open probability) was noticeably enhanced. The results suggest that as the RFM concentration was increased, the voltage-dependent change in the channel activity (i.e., the voltage sensor domain) was intrinsically modified so that the difference between the V1/2 values measured at the downsloping and upsloping ends of the inverted triangular ramp widened. However, no change in the single-channel conductance of the BKCa channels activated by this ramp pulse was observed with either the absence or presence of various concentrations of RFM. Therefore, it is likely that the observed change in the voltage-dependent hysteresis in the presence of RFM does not occur in the pore region of the channel, although the channel activity exhibited intrinsic hysteresis.
The maximum plasma concentrations of RFM at dosages of 10 mg/kg/day and 30 mg/kg/day have previously been reported to be 4.01 μg/mL (16.8 μM) and 8.68 μg/mL (36.4 μM), respectively [8]. Similar results have also been obtained in other studies [7,8,66]. As such, both IC50 values found in its inhibition of INa and EC50 in its stimulation of IK(Ca) were noted to fall within clinically achieved concentrations. Hence, the explanations for the RFM-mediated perturbation of ionic currents presented in the current study could have therapeutical and pharmacological relevance.
In this study, we also conducted a docking study using PyRx software to investigate how RFM and the protein of the KCa1.1 might fit together. The predicted binding sites of the RFM are shown in Figure 9. Given that RFM forms hydrogen bonds with the amino acid residues Asn427, Asn808, and Ile810 and that it forms hydrophobic interactions with the amino acid residues His350, Tyr429, and Asn809, we conclude that RFM can bind to the intracellular domain of KCa1.1 channels and that the RFM-induced binding site is not located in the pore regions of the channels.

4. Materials and Methods

4.1. Chemicals, Drugs, and Solutions Used in This Work

Rufinamide (RFM, Banzel®, Inovelon®, R8404, 1-((2,6-difluorophenyl)methyl)-1H-1,2,3-triazole-4carboxamide, CAS number: 106308-44-5, C10H8F2N4O), E-4031, glibenclamide, tefluthrin (Tef), tetraethylammonium chloride (TEA), and tetrodotoxin (TTX) were acquired from Sigma-Aldrich (Merck, Taipei, Taiwan). Apamin, iberiotoxin, and paxilline were supplied by Alomone (Asia Bioscience, Taipei, Taiwan), and TRAM-34 by Togenesis Technologies (Taipei, Taiwan). For cell preparations, we obtained culture media, fetal bovine/calf serum, horse serum, L-glutamine, and trypsin/EDTA from HyCloneTM (Thermo Fisher; Level Biotech, Tainan, Taiwan), and other chemicals, such as CdCl2, CsCl, and CsOH, were of analytical research grade.
The ionic composition of the normal Tyrode’s solution used as the extracellular bath solution was (in mM): NaCl 136.5, KCl 5.4, CaCl2 1.8, MgCl2 0.53, glucose 5.5, and HEPES-NaOH buffer 5.5 (pH 7.4). The composition of the intracellular pipette (internal) solution for measuring the ions flowing through the whole-cell K+ currents was (in mM): K-aspartate 130, KCl 20, MgCl2 1, Na2ATP 3, Na2GTP 0.1, EGTA 0.1, and HEPES-KOH buffer 5 (pH 7.2). To measure the INa, we substituted K+ ions in the internal solution for equimolar Cs+ ions, the pH value was adjusted to 7.2 by adding CsOH, and the cells were bathed in Ca2+-free Tyrode’s solution containing 10 mM TEA. For single BKCa-channel recordings, the high K+ bathing solution was (in mM): KCl 145, MgCl20.53, and HEPES-KOH buffer 5 (pH 7.4), and the pipette solution was (in mM): KCl 145, MgCl2 2, and HEPES-KOH buffer 5 (pH 7.2). A dissociation constant of 0.1 μM for EGTA and Ca2+ (at pH 7.2) was the basis for estimating the cytosolic-free Ca2+ concentration. The bath and pipette solutions were filtered on the day of use with a syringe filter equipped with a 0.22-μm Supor® nylon membrane (#4612; Pall Corp; Genechain Biotechnology, Kaohsiung, Taiwan).

4.2. Cell Preparations

The GH3 pituitary cell line was supplied by the Bioresources Collection and Research Center (BCRC-60015; Hsinchu, Taiwan). The GH3 cell line was maintained by growing cells in 50 mL plastic culture flasks in 5 mL of Ham’s F-12 medium, supplemented with 2.5% fetal calf serum (v/v), 15% horse serum (v/v), and 2 mM L-glutamine. The growth medium was replaced twice a week, and the cells were split into subcultures once a week. Electrophysiological recordings were conducted five or six days after the cells were cultured up to 60–80% confluence.

4.3. Electrophysiological Recordings

Shortly before each measurement, GH3 cells were gently dispersed, after which, a few drops of cell suspension were quickly placed into a custom-built recording chamber and allowed to settle at the bottom of the chamber. The recording chamber was positioned on the stage of an inverted phase-contrast microscope (Diaphot-200; Nikon; Lin Trading Co., Taipei, Taiwan) which was equipped with a video camera system with a magnification capability of up to 1500× magnification to monitor the cell size during the experiments. The cells were kept in a bath at room temperature (20–25 °C) in normal Tyrode’s solution containing 1.8 mM CaCl2. The patch electrodes were prepared from Kimax®-51 capillaries with a 1.5–1.8 mm outer diameter (#34500; Kimble; Dogger, New Taipei City, Taiwan) using a two-state PP-830 puller (Narishige; Taiwan Instrument, Tainan, Taiwan), and the tips were then fire-polished with an MF-83 microforge (Narishige). When filled with pipette solution, the resistances ranged between 3 MΩ and 5 MΩ. We performed standard patch-clamp recordings in the cell-attached, inside-out, or whole-cell configurations using an RK-400 amplifier (Bio-Logic, Claix, France) [67]. Junction potentials, which develop at the electrode tip when the composition of the intracellular solution differs from that of the bath, were nulled before the start of the formation of each gigaseal, and the junction potential corrections were then applied to the whole-cell data. During measurement, the recorded signals were stored online at 10 kHz or more with an ASUSPRO-BN401 LG laptop computer (ASUS, Tainan, Taiwan) equipped with a Digidata-1440A device (Molecular Devices; Advanced Biotech, New Taipei City, Taiwan) and controlled with the pCLAMP 10.6 software (Molecular Devices) [68].

4.4. Whole-Cell Current Analyses

To determine the percentage increase in the IK(Ca) evoked by the presence of RFM, the amplitude of the current at the concentration of 300 μM RFM was taken as 100%, and the current amplitudes during cell exposure to different concentrations of RFM (1–300 μM) were analyzed and compared. To measure the IK(Ca), we kept cells bathed in normal Tyrode’s solution containing 1.8 mM CaCl2, and a depolarizing voltage command pulse from 0 mV to +50 mV was applied. The amplitudes of the IK(Ca) measured during the addition of RFM were compared with those measured after the subsequent addition of paxilline (1 μM). The concentration-response data for the activation of the IK(Ca) were then fitted to the modified Hill equation (i.e., a multi-parameter logistic equation) using the least-squares method, as follows:
P e r c e n t a g e   i n c r e a s e   % = C n H × E m a x C n H + E C 50 n H
where [C] is the RFM concentration added; Emax is the maximal stimulation of the IK(Ca) (i.e., the paxilline-sensitive current) caused by the addition of RFM; and nH and EC50 are the Hill coefficient and the RFM concentration required to achieve 50% stimulation, respectively.
To determine the concentration-dependent inhibition of the peak and late components of the INa caused by RFM, the GH3 cells were kept bathed in Ca2+-free Tyrode’s solution, to which 1 μM TTX and 0.5 mM CdCl2 were added. The tested cell was voltage-clamped at –80 mV and a 30 ms step depolarization from –80 mV to −10 mV was delivered to it. The amplitudes of the currents (recorded at the beginning and end of the depolarization) evoked by the depolarizing voltage command pulse to –10 mV were measured during the control period (i.e., in the absence of RFM) and during cell exposure to varying concentrations of RFM (0.1–300 μM). The concentration required to suppress 50% (i.e., IC50) of the peak and late INa was estimated on the basis of the goodness of fit test with another modified form of the Hill equation, as follows:
R e l a t i v e   a m p l i t u d e = RFM n H × 1 a I C 50 n H + RFM n H + a
where [RFM] represents the different concentrations of RFM used; and nH and IC50 are the Hill coefficient inherent to the concentration-response relationship and the concentration at which the inhibition of 50% of the peak or late INa is observed, respectively. At this point, the maximal inhibition (1−a) of the peak and late INa was also estimated.
The I–V relationship of the peak INa obtained with and without the addition of RFM was derived and fitted to a Boltzmann function given as:
I I m a x = G 1 + e x p V V h / k × V E r e v
where V is the voltage in mV; Erev is the reversal potential of the INa (fixed at +45 mV); G is the Na+ conductance expressed in nS; I is the current expressed in pA; and Vh and k are the gating parameters.

4.5. Single-Channel Analyses

The amplitudes of single BKCa channels identified in the GH3 cells were determined by fitting Gaussian distributions to the amplitude histograms of the closed (resting) and open states. The open-state probability of a channel in a patch was expressed as N·PO and was estimated using the following equation:
N · P O = A 1 + 2 A 2 + + n A n A 0 + A 1 + A 2 + + A n
where A0 indicates the area under the curve of an all-points histogram that corresponds to the closed (resting) state; A1An are the histogram areas indicating the levels of a distinct open state for 1 to n channels in the patch; and N represents the number of active channels in the patch. To perform the analysis of the open and closed time in the channel, only one single channel in the patch was used.
The relationship between the membrane potential and the relative probability of the BKCa channels being in the open state in response to a triangular ramp pulse with and without the addition of RFM was established and well fitted to the Boltzmann equation (or the Fermi-Dirac distribution) as follows:
relative   N · P O = n 1 + e x p ( V V 1 2 ) q F R T
where N is the number of channels in the patch; n is the maximum relative N·PO; V is the membrane potential expressed in mV; V1/2 and q are the potential for half-maximal activation and the apparent gating charge, respectively; and F, R, and T are the Faraday constant, the universal gas constant, and the absolute temperature, respectively.
To evaluate the effect of RFM on the strength of the hysteresis exhibited by the BKCa channels, a 1.6 s inverted triangular ramp pulse from +180 to −180 mV with a ramp speed of ±450 mV/s at a rate of 0.05 Hz was applied to the detached patch of membrane by converting the electrical signal from digital to analog. To establish the relative probability of the channels opening in response to the exposure of the cell to different concentrations of RFM, the amplitudes of the current in each single channel evoked by 20-voltage ramps were averaged, and each point of the averaged current was divided by the single-channel amplitude measured for each potential after a correction was performed for a leak component. The number of active channels in the patch, N, was counted at the end of the experiments by introducing a solution with 100 μM Ca2+, which was then used to normalize the open-state probability. To acquire values for the gating charge and the half-maximal activation of the current, the curve obtained at the descending (forward), or ascending (backward) limb of the inverted triangular ramp pulse was approximately fitted with a Boltzmann function as described above.

4.6. Curve-Fitting Procedures and Statistical Analyses

The curves were fitted to the experimental data by performing a linear or nonlinear regression (i.e., exponential or sigmoid function) using the pCLAMP 10.7 software (Molecular Devices), the 64-bit OriginPro 2021 software (OriginLab®; Scientific Formosa, Kaohsiung, Taiwan), or the Microsoft Excel 2013 software. The results of the analyses of the data, which consisted of different types of ionic currents, are presented as the mean ± standard error of the mean (SEM). The sample size (n) indicates the number of cells examined. SEM error bars were plotted. The differences are considered statistically significant when p < 0.05, as indicated in the legends for the figures.

5. Conclusions

The important findings of this study are: (a) the presence of RFM caused an increase in the amplitude of the IK(Ca), the size of which depended on the RMF concentration; (b) the RFM-activated IK(Ca) was reversed with the addition of iberiotoxin and paxilline but not apamin or glibenclamide; (c) RFM added to the cytosolic surface of the detached membrane patch increased the open-state probability of BKCa channels, but no evident change in single-channel conductance was detected in its presence; (d) the RFM-induced increase in the BKCa-channel activity was suppressed by paxilline, but not by TRAM-34; (e) the presence of RFM enhanced the strength of the hysteresis exhibited by the BKCa channels activated by an inverted isosceles-triangular ramp pulse; and (f) the presence of RFM differentially suppressed the peak and late components of the INa activated by the rapid depolarization of the cell membrane. Taken together, these findings suggest that the RFM-mediated stimulation of BKCa channels demonstrated herein brings to light an as-yet unidentified but important ionic mechanism underlying the actions produced by RFM, through which it affects neuronal activities, including neuronal excitability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232213677/s1.

Author Contributions

Conceptualization by M.-C.L., S.-N.W. and C.-W.H.; methodology by S.-N.W. and C.-W.H.; validation by M.-C.L., S.-N.W. and C.-W.H.; formal analysis by S.-N.W.; investigation by M.-C.L., S.-N.W. and C.-W.H.; data curation by M.-C.L., S.-N.W. and C.-W.H.; writing—original draft preparation by M.-C.L., S.-N.W. and C.-W.H.; writing—review and editing by M.-C.L., S.-N.W. and C.-W.H.; and funding acquisition by S.-N.W. and C.-W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the Ministry of Science and Technology, Taiwan (MOST-109-2314-B-006-034-MY3, MOST-111-2314-B-006-103-MY2).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request from the corresponding authors.

Acknowledgments

The authors would like to express their appreciation for the comments and the analysis of the voltage-dependent hysteresis of ionic currents in this work. We gratefully recognize Shin-Yen Chao and Tzu-Hsien Chuang for their technical assistance.

Conflicts of Interest

The authors declare that there are no conflicts of interest.

References

  1. Arroyo, S. Rufinamide. Neurotherapeutics 2007, 4, 155–162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. McCormack, P.L. Rufinamide: A pharmacoeconomic profile of its use as adjunctive therapy in Lennox-Gastaut syndrome. Pharmacoeconomics 2012, 30, 247–256. [Google Scholar] [CrossRef] [PubMed]
  3. Farrokh, S.; Bon, J.; Erdman, M.; Tesoro, E. Use of newer anticonvulsants for the treatment of status epilepticus. Pharmacotherapy 2019, 39, 297–316. [Google Scholar] [CrossRef]
  4. Panebianco, M.; Prabhakar, H.; Marson, A.G. Rufinamide add-on therapy for refractory epilepsy. Cochrane Database Sys. Rev. 2018, 4, CD011772. [Google Scholar] [CrossRef] [PubMed]
  5. Albini, M.; Morano, A.; Fanelia, M.; Lapenta, L.; Casciato, S.; Fattouch, J.; Manfredi, M.; Giallonardo, A.T.; Di Bonaventura, C. Effectiveness of rufinamide in the treatment of idiopathic generalized epilepsy with atypical evolution: Case report and review of the literature. Clin. EEG Neurosci. 2016, 47, 162–166. [Google Scholar] [CrossRef]
  6. Heaney, D.; Walker, M.C. Rufinamide. Drugs Today 2007, 43, 455–460. [Google Scholar] [CrossRef]
  7. Kim, S.H.; Kang, H.C.; Lee, J.S.; Kim, H.D. Rufinamide efficacy and safety in children aged 1–4 years with Lennox-Gastaut syndrome. Brain Dev. 2018, 40, 897–903. [Google Scholar] [CrossRef]
  8. Wier, H.A.; Cerna, A.; So, T.Y. Rufinamide for pediatric patients with Lennox-Gastaut syndrome. Pediatric. Drugs 2011, 13, 97–106. [Google Scholar] [CrossRef]
  9. Kessler, S.K.; McCarthy, A.; Cnaan, A.; Dlugos, D.J. Retention rates of rufinamide in pediatric epilepsy patients with and without Lennox-Gastaut syndrome. Epilespy Res. 2015, 112, 18–26. [Google Scholar] [CrossRef] [Green Version]
  10. McMurray, R.; Striano, P. Treatment of adults with Lennox-Gastaut syndrome: Further analysis of efficacy and safety/tolerability of rufinamide. Neurol. Ther. 2016, 5, 35–43. [Google Scholar] [CrossRef]
  11. Ohtsuka, Y.; Yoshinaga, H.; Shirasaka, Y.; Takayama, R.; Takano, H.; Iyoda, K. Long-term safety and seizure outcome in Japanese patients with Lennox-Gastaut syndrome receiving adjunctive rufinamide therapy: An open-label study following a randomized clinical trial. Epilepsy Res. 2016, 121, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Thompson, A.G.B.; Cock, H.R. Successful treatment of super-refractory tonic status epilepticus with rufinamide: First clinical report. Seizure 2016, 39, 1–4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Xu, Z.; Zhao, H.; Chen, Z. The efficacy and safety of rufinamide in drug-resistant epilepsy: A meta-analysis of double-blind, randomized, placebo controlled trials. Epilepsy Res. 2016, 120, 104–110. [Google Scholar] [CrossRef] [PubMed]
  14. Cross, J.H.; Auvin, S.; Falip, M.; Striano, P.; Arzimanoglou, A. Expert opinion on the management of Lennox-Gastaut syndrome: Treatment algorithms and practical considerations. Front. Neurol. 2017, 8, 505. [Google Scholar] [CrossRef]
  15. Jaraba, S.; Santamarina, E.; Miró, J.; Toledo, M.; Molins, A.; Burcet, J.; Becerra, J.L.; Raspall, M.; Pico, G.; Miravet, E.; et al. Rufinamide in children and adults in routine clinical practice. Acta Neurol. Scand. 2017, 135, 122–128. [Google Scholar] [CrossRef]
  16. Kothare, S.; Kluger, G.; Sachdeo, R.; Williams, B.; Olhaye, O.; Perdomo, C.; Bibbiani, F. Dosing considerations for rufinamide in patients with Lennox-Gastaut syndrome: Phase III trial results and real-world clinical data. Seizure 2017, 47, 25–33. [Google Scholar] [CrossRef] [Green Version]
  17. Nikanorova, M.; Brandt, C.; Auvin, S.; McMurray, R. Real-world data on rufinamide treatment in patients with Lennox-Gastaut syndrome: Results from a European noninterventional registry study. Epilepsy Behav. 2017, 76, 63–70. [Google Scholar] [CrossRef]
  18. Ostendorf, A.P.; Ng, Y.T. Treatment-resistant Lennox-Gastaut syndrome: Therapeutic trends challenges and future directions. Neuropsychiatr. Dis. Treat. 2017, 13, 1131–1140. [Google Scholar] [CrossRef] [Green Version]
  19. Zhao, T.; Feng, X.; Liu, J.; Gao, J.; Zhou, C. Evaluate the efficacy and safety of anti-epileptic medications for partial seizures of epilepsy: A network meta-analysis. J. Cell. Biochem. 2017, 118, 2850–2864. [Google Scholar] [CrossRef]
  20. Arzimanoglou, A.; Ferreira, J.; Satlin, A.; Olhaye, O.; Kumar, D.; Dhadda, S.; Bibbiani, F. Evaluation of long-term safety, tolerability, and behavioral outcomes with adjunctive rufinamide in pediatric patients (≥1 to <4 years old) with Lennox-Gastaut syndrome: Final results from randomized study 303. Eur. J. Paediatr. Neurol. 2019, 23, 126–135. [Google Scholar] [CrossRef]
  21. Asadi-Pooya, A.A. Lennox-Gastaut syndrome: A comprehensive review. Neurol. Sci. 2018, 39, 403–414. [Google Scholar] [CrossRef] [PubMed]
  22. Striano, P.; McMurray, R.; Santamarina, E.; Falip, M. Rufinamide for the treatment of Lennox-Gastaut syndrome: Evidence for clinical trials and clinical practice. Epileptic Disord. 2018, 20, 13–29. [Google Scholar] [CrossRef] [PubMed]
  23. Yıldız, E.P.; Hızlı, Z.; Bektaş, G.; Ulak-Özkan, M.; Tatlı, B.; Aydınlı, N.; Çaslıskan, M.; Özmen, M. Efficacy of rufinamide in childhood refractory epilepsy. Turk. J. Pediatr. 2018, 60, 238–243. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, Y.C.; Lai, Y.C.; Chou, P.; Hsueh, S.W.; Lin, T.H.; Huang, C.S.; Wang, R.W.; Yang, Y.C.; Kuo, C.C. How Can an Na+ Channel Inhibitor Ameliorate Seizures in Lennox-Gastaut Syndrome? Ann. Neurol. 2021, 89, 1099–1113. [Google Scholar] [CrossRef]
  25. Heger, K.; Skipsfjord, J.; Kiselev, Y.; Burns, M.L.; Aaberg, K.M.; Johannessen, S.I.; Skurtveit, S.; Landmark, C.J. Changes in the use of antiseizure medications in children and adolescents in Norway, 2009–2018. Epilepsy Res. 2022, 181, 106872. [Google Scholar] [CrossRef]
  26. Yamamoto, Y.; Inoue, Y.; Usui, N.; Imai, K.; Kagawa, Y.; Takahashi, Y. Therapeutic drug monitoring for rufinamide in Japanese patients with epilepsy: Focus on drug interactions, tolerability, and clinical effectiveness. Ther. Drug Monit. 2022, 44, 585–591. [Google Scholar] [CrossRef]
  27. Balagura, G.; Riva, A.; Marchese, F.; Verrotti, A.; Striano, P. Adjunctive rufinamide in children with Lennox-Gastaut syndrome: A literature review. Neuropsychiatr. Dis. Treat. 2020, 16, 369–379. [Google Scholar] [CrossRef] [Green Version]
  28. Vohora, D.; Saraogi, P.; Yazdani, M.A.; Bhowmik, M.; Khanam, R.; Pillai, K.K. Recent advances in adjunctive therapy for epilepsy: Focus on sodium channel blockers as third-generation antiepileptic drugs. Drugs Today 2010, 46, 265–277. [Google Scholar] [CrossRef]
  29. Stephen, L.J.; Brodie, M.J. Pharmacotherapy of epilepsy: Newly approved and developmental agents. CNS Drugs 2011, 25, 89–107. [Google Scholar] [CrossRef]
  30. Suter, M.R.; Kirschmann, G.; Laedermann, C.J.; Abriel, H.; Decosterd, I. Rufinamide attenuates mechanical allodynia in a model of neuropathic pain in the mouse and stabilizes voltage-gated sodium channel inactivated state. Anesthesiology 2013, 118, 160–172. [Google Scholar] [CrossRef]
  31. Gilchrist, J.; Dutton, S.; Diaz-Bustamante, M.; McPherson, A.; Olivares, N.; Kalia, J.; Escayg, A.; Bosmans, F. Nav1.1 modulation by a novel triazole compound attenuates epileptic seizure in rodents. ACS Chem. Biol. 2014, 9, 1204–1212. [Google Scholar] [CrossRef] [PubMed]
  32. Kharatmal, S.B.; Singh, J.N.; Sharma, S.S. Rufinamide improves functional and behavioral deficits via blockade of tetrodotoxin-resistance sodium channels in diabetic neuropathy. Curr. Neurovasc. Res. 2015, 12, 262–268. [Google Scholar] [CrossRef] [PubMed]
  33. Brodie, M.J. Sodium channel blockers in the treatment of epilepsy. CNS Drugs 2017, 31, 527–534. [Google Scholar] [CrossRef] [PubMed]
  34. Gáll, Z.; Vancea, S.; Szilagyi, T.; Gáll, O.; Kolcsár, M. Dose-dependent pharmacokinetics and brain penetration of rufinamide following intravenous and oral administration to rats. Eur. J. Pharm. Sci. 2015, 68, 106–113. [Google Scholar] [CrossRef] [PubMed]
  35. Lin, Y.C.; Lai, Y.C.; Lin, T.H.; Yang, Y.C.; Kuo, C.C. Selective stabilization of the intermediate inactivated Na+ channel by the new-generation anticonvulsant rufinamide. Biochem. Pharmacol. 2022, 197, 114928. [Google Scholar] [CrossRef] [PubMed]
  36. Das, A.; McDowell, M.; O’Dell, C.M.; Busch, M.E.; Smith, J.A.; Ray, S.K.; Banik, N.L. Post-treatment with voltage-gated Na+ channel blocker attenuates kainic acid-induced apoptosis in rat primary hippocampal neurons. Neurochem. Res. 2010, 35, 2175–2183. [Google Scholar] [CrossRef] [Green Version]
  37. Park, J.A.; Lee, C.H. Effect of rufinamide on the kainic acid-induced excitotoxic neuronal death in the mouse hippocampus. Arch. Pharm. Res. 2018, 41, 776–783. [Google Scholar] [CrossRef]
  38. Chen, P.C.; Ruan, J.S.; Wu, S.N. Evidence of decreased activity in intermediate-conductance calcium-activated potassium channels during retinoic acid-induced differentiation in motor neuron-like NSC-34 cells. Cell. Physiol. Biochem. 2018, 48, 2374–2388. [Google Scholar] [CrossRef]
  39. Skov, M.; de Paoli, F.V.; Nielsen, O.B.; Pedersen, T.H. The anti-convulsants lacosamide, lamotrigine, and rufinamide reduce myotonia in isolated human and rat skeletal muscle. Muscle Nerve 2017, 56, 136–142. [Google Scholar] [CrossRef]
  40. Bektaş, G.; Çalışkan, M.; Aydın, A.; Pembegül Yıldız, E.; Tatlı, B.; Aydınlı, N.; Özmen, M. Aggravation of atonic seizures by rufinamide: A case report. Brain Dev. 2016, 38, 654–657. [Google Scholar] [CrossRef]
  41. Lee, U.S.; Cui, J. BK channel activation: Structural and functional insights. Trends Neurosci. 2010, 33, 415–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Liu, J.; Ye, J.; Zou, X.; Xu, Z.; Feng, Y.; Zou, X.; Chen, Z.; Li, Y.; Cang, Y. CRL4A(CRBN) E3 ubiquitin ligase restricts BK channel activity and prevents epileptogenesis. Nat. Commun. 2014, 5, 3924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Contet, C.; Goulding, S.P.; Kuljis, D.A.; Barth, A.L. BK channels in the central nervous system. Int. Rev. Neurobiol. 2016, 128, 281–342. [Google Scholar]
  44. Grigouli, M.; Sgritta, M.; Cherubini, E. Presynaptic BK channels control transmitter release: Physiological relevance and potential therapeutic implications. J. Physiol. 2016, 594, 3489–3500. [Google Scholar] [CrossRef] [Green Version]
  45. Wu, S.N.; Chern, J.H.; Shen, S.; Chen, H.H.; Hsu, T.Y.; Lee, C.C.; Chan, M.H.; Lai, M.C.; Shie, F.S. Stimulatory actions of a novel thiourea derivative on large-conductance, calcium-activated potassium channels. J. Cell. Physiol. 2017, 232, 3409–3421. [Google Scholar] [CrossRef]
  46. Niday, Z.; Bean, B.P. BK channel regulation of afterpotentials and burst firing in cerebellar Purkinje neurons. J. Neurosci. 2021, 41, 2854–2869. [Google Scholar] [CrossRef] [PubMed]
  47. Shah, K.R.; Guan, X.; Yan, J. Structural and functional coupling of calcium-activated BK channels and calcium-permeable channels within nanodomain signaling complexes. Front. Physiol. 2022, 12, 796540. [Google Scholar] [CrossRef]
  48. Watanave, M.; Takahashi, N.; Hosoi, N.; Konno, A.; Yamamoto, H.; Yasui, H.; Kawachi, M.; Horii, T.; Matsuzaki, Y.; Hatada, J.; et al. Protein kinase Cγ in cerebellar Purkinje cells regulates Ca2+-activated large-conductance K+ channels and motor coordination. Proc. Natl. Acad. Sci. USA 2022, 119, e2113336119. [Google Scholar] [CrossRef]
  49. Wu, S.N.; Peng, H.; Chen, B.S.; Wang, Y.J.; Wu, P.Y.; Lin, M.W. Potent activation of large-conductance Ca2+-activated K+ channels by the diphenylurea 1,3-bis-[2-hydroxy-5-(trifluoromethyl)phenyl]urea (NS1643) in pituitary tumor (GH3) cells. Mol. Pharmacol. 2008, 74, 1696–1704. [Google Scholar] [CrossRef] [Green Version]
  50. Wu, S.N.; Wang, Y.J.; Lin, M.W. Potent stimulation of large-conductance Ca2+-activated K+ channels by rottlerin, an inhibitor of protein kinase C-Δ, in pituitary tumor (GH3) cells and in cortical neuronal (HCN-1A) cells. J. Cell. Physiol. 2007, 210, 655–666. [Google Scholar] [CrossRef]
  51. Chang, W.T.; Wu, S.N. Effective activation of BKCa channels by QO-40 (5-(chloromethyl)-3-(naphthalen-1yl)-2-(trifluoromethyl)pyrazolo[1,5-a]pyrimidin-7(4H)-one), known to be an opener of KCNQ2/Q3 channels. Pharmaceuticals 2021, 14, 388. [Google Scholar] [CrossRef] [PubMed]
  52. Zuccolini, P.; Ferrera, L.; Remigante, A.; Picco, C.; Barbieri, R.; Bertelli, S.; Moran, O.; Gavazzo, P.; Pusch, M. The VRAC blocker DCPIB directly gates the BK channels and increases intracellular Ca2+ in melanoma and pancreatic duct adenocardinoma cell lines. Br. J. Pharmacol. 2022, 179, 3452–3469. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, J.; Krishnamoorthy, G.; Saxena, A.; Zhang, G.; Shi, J.; Yang, H.; Delaloye, K.; Sept, D.; Cui, J. An epilepsy/dyskinesia-associated mutation enhances BK channel activation by potentiating Ca2+ sensing. Neuron 2010, 66, 871–883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Leo, A.; Citraro, R.; Constanti, A.; De Sarro, G.; Russo, E. Are big potassium Ca2+-activated potassium channels in viable target for the treatment of epilepsy? Expert Opin. Ther. Targets 2015, 19, 911–926. [Google Scholar] [CrossRef]
  55. Chen, B.H.; Ahn, J.H.; Park, J.H.; Song, M.; Kim, H.; Lee, T.K.; Lee, J.C.; Kim, Y.M.; Hwang, I.K.; Kim, D.W.; et al. Rufinamide, an antiepileptic drug, improves cognition and increases neurogenesis in the aged gerbil hippocampal dentate gyrus via increasing expressions of IGF-1, IGF-1R and pCREB. Chem. Biol. Interact. 2018, 286, 71–77. [Google Scholar] [CrossRef] [PubMed]
  56. Knaus, H.G.; McManus, O.B.; Lee, S.H.; Schmalhofer, W.A.; Garcia-Calvo, M.; Helms, L.M.; Sanchez, M.; Giangiacomo, K.; Reuben, J.P.; Smith, A.B. 3rd.; et al. Tremorgenic indole alkaloids potently inhibit smooth muscle high-conductance calcium-activated potassium channels. Biochemistry 1994, 33, 5819–5828. [Google Scholar] [CrossRef]
  57. Bennekou, P.; Barksmann, T.L.; Jensen, L.R.; Kristensen, B.I.; Christophersen, P. Voltage activation and hysteresis of the non-selective voltage-dependent channel in the intact human red cell. Bioelectrochemistry 2004, 62, 181–185. [Google Scholar] [CrossRef]
  58. Rappaport, S.M.; Teijido, O.; Hoogerheide, D.P.; Rostovtseva, T.K.; Berezhkovskii, A.M.; Bezrukov, S.M. Conductance hysteresis in the voltage-dependent anion channel. Eur. Biophys. J. 2015, 44, 465–472. [Google Scholar] [CrossRef] [Green Version]
  59. Villalba-Galea, C.A.; Chiem, A.T. Hysteretic behavior in voltage-gated channels. Front. Pharmacol. 2020, 11, 579596. [Google Scholar] [CrossRef]
  60. Mahmood, A.; Echtenkamp, W.; Street, M.; Wang, J.L.; Cao, S.; Komesu, T.; Dowben, P.A.; Buragohain, P.; Lu, H.; Gruverman, A.; et al. Voltage controlled Néel vector rotation in zero magnetic field. Nat. Commun. 2021, 12, 1674. [Google Scholar] [CrossRef]
  61. Chang, W.T.; Wu, S.N. Inhibitory effectiveness of gomisin A, a dibenzocyclooctadiene lignan isolated from Schizandra chinensis, on the amplitude and gating of voltage-gated Na+ current. Int. J. Mol. Sci. 2020, 21, 8816. [Google Scholar] [CrossRef] [PubMed]
  62. Chuang, T.H.; Cho, H.Y.; Wu, S.N. Effective accentuation of voltage-gated sodium current caused by apocynin (4′-hydroxy-3′-methoxyacetophenone), a known NADPH-oxidase inhibitor. Biomedicines 2021, 9, 1146. [Google Scholar] [CrossRef] [PubMed]
  63. So, E.C.; Wu, S.N.; Lo, Y.C.; Su, K. Differential regulation of tefluthrin and telmisartan on the gating charges of INa activation and inactivation as well as on resurgent and persistent INa in a pituitary cell line (GH3). Toxicol. Lett. 2018, 285, 104–112. [Google Scholar] [CrossRef]
  64. Wu, G.; Li, Q.; Liu, X.; Li-Byarlay, H.; He, B. Differential state-dependent effects of deltamethrin and tefluthrin on sodium channels in central neurons of Helicoverpa armigera. Pestic. Biochem. Physiol. 2021, 175, 104836. [Google Scholar] [CrossRef]
  65. Wang, Y.J.; Sung, R.J.; Lin, M.W.; Wu, S.N. Contribution of BKCa-channel activity in human cardiac fibroblasts to electrical coupling of cardiomyocytes-fibroblasts. J. Membr. Biol. 2006, 213, 175–185. [Google Scholar] [CrossRef] [PubMed]
  66. Perucca, E.; Cloyd, J.; Critchley, D.; Fuseau, E. Rufinamide: Clinical pharmacokinetics and concentration-response relationships in patients with epilepsy. Epilepsia 2008, 49, 1123–1141. [Google Scholar] [CrossRef]
  67. Lai, M.C.; Wu, S.N.; Huang, C.W. Zingerone modulates neuronal voltage-gated Na+ and L-Type Ca2+ currents. Int. J. Mol. Sci. 2022, 23, 3123. [Google Scholar] [CrossRef]
  68. Hung, T.Y.; Wu, S.N.; Huang, C.W. The integrated effects of brivaracetam, a selective analog of levetiracetam, on ionic currents and neuronal excitability. Biomedicines. 2021, 9, 369. [Google Scholar] [CrossRef]
Figure 1. Effect of RFM on Ca2+-activated K+ current (IK(Ca)) measured from pituitary GH3 cells. Cells were bathed in normal Tyrode’s solution containing 1.8 mM CaCl2, the ionic composition of which is detailed in Materials and Methods. (A) Representative IK(Ca) traces obtained in the control (i.e., absence of RFM, a), during the exposure to 3 μM RFM (b) or 10 μM RFM (c), and after washout of RFM (d). The voltage-clamp protocol is illustrated in the upper part. (B) Concentration-dependent stimulation of RFM on IK(Ca) in response to membrane depolarization (mean ± SEM; n = 8 for each point). Current amplitude was measured at the end of the 300 ms depolarizing step to +50 mV from a holding potential of 0 mV. The IK(Ca) amplitude during cell exposure to 300 μM RFM was taken as 100%, and those in the presence of different RFM concentrations were thereafter compared. The Hill equation indicated in Materials and Methods was well-fitted to the experimental data (solid red line).
Figure 1. Effect of RFM on Ca2+-activated K+ current (IK(Ca)) measured from pituitary GH3 cells. Cells were bathed in normal Tyrode’s solution containing 1.8 mM CaCl2, the ionic composition of which is detailed in Materials and Methods. (A) Representative IK(Ca) traces obtained in the control (i.e., absence of RFM, a), during the exposure to 3 μM RFM (b) or 10 μM RFM (c), and after washout of RFM (d). The voltage-clamp protocol is illustrated in the upper part. (B) Concentration-dependent stimulation of RFM on IK(Ca) in response to membrane depolarization (mean ± SEM; n = 8 for each point). Current amplitude was measured at the end of the 300 ms depolarizing step to +50 mV from a holding potential of 0 mV. The IK(Ca) amplitude during cell exposure to 300 μM RFM was taken as 100%, and those in the presence of different RFM concentrations were thereafter compared. The Hill equation indicated in Materials and Methods was well-fitted to the experimental data (solid red line).
Ijms 23 13677 g001
Figure 2. Effect of RFM on mean current-voltage (I–V) relationships of IK(Ca) identified in GH3 cells. (A) Representative current traces obtained in the absence (upper) and presence (lower) of 10 μM RFM. The uppermost part shows the voltage-clamp protocol applied. The potential traces labeled in different colors correspond with current ones acquired without or with the RFM application. The duration in each depolarizing step is different for better illustrations, and the blue solid arrow indicates the outwardly rectifying properties of IK(Ca) with increasing positive voltage. (B) Mean I–V relationships of IK(Ca) amplitude acquired in the control (blue squares), during exposure to 10 μM RFM (red squares) and washout of RFM (black squares) (mean ± SEM; n = 8 for each point). Current amplitude was measured at the end of each depolarizing step.
Figure 2. Effect of RFM on mean current-voltage (I–V) relationships of IK(Ca) identified in GH3 cells. (A) Representative current traces obtained in the absence (upper) and presence (lower) of 10 μM RFM. The uppermost part shows the voltage-clamp protocol applied. The potential traces labeled in different colors correspond with current ones acquired without or with the RFM application. The duration in each depolarizing step is different for better illustrations, and the blue solid arrow indicates the outwardly rectifying properties of IK(Ca) with increasing positive voltage. (B) Mean I–V relationships of IK(Ca) amplitude acquired in the control (blue squares), during exposure to 10 μM RFM (red squares) and washout of RFM (black squares) (mean ± SEM; n = 8 for each point). Current amplitude was measured at the end of each depolarizing step.
Ijms 23 13677 g002
Figure 3. Summary scatter graph showing comparisons on IK(Ca) amplitude produced in the presence of RFM, RFM plus apamin, RFM plus glibenclamide (Glib), RFM plus iberiotoxin, and RFM plus paxilline in GH3 cells. In these experiments, IK(Ca) was evoked by a 300 ms depolarizing pulse from 0 to +50 mV, and the current amplitude was measured at the end of the command pulse. Each point represents the mean ± SEM (n = 8). * Significantly different from control (p < 0.05) and ** significantly different from RFM (10 μM) alone group (p < 0.05).
Figure 3. Summary scatter graph showing comparisons on IK(Ca) amplitude produced in the presence of RFM, RFM plus apamin, RFM plus glibenclamide (Glib), RFM plus iberiotoxin, and RFM plus paxilline in GH3 cells. In these experiments, IK(Ca) was evoked by a 300 ms depolarizing pulse from 0 to +50 mV, and the current amplitude was measured at the end of the command pulse. Each point represents the mean ± SEM (n = 8). * Significantly different from control (p < 0.05) and ** significantly different from RFM (10 μM) alone group (p < 0.05).
Ijms 23 13677 g003
Figure 4. Effect of RFM on the activity of large-conductance Ca2+-activated K+ (BKCa) channels recorded from GH3 cells. The single-channel experiments were conducted as cells were exposed to high-K+ concentration (145 mM) containing 0.1 μM Ca2+, the pipette solution contained K+-enriched solution, and an inside-out configuration with a holding potential of +60 mV was created. (A) Representative tracings of BKCa channels in the absence (left, blue color) and presence (right, red color) of 10 μM RFM. RFM was applied to the bath medium. The lower five traces were expanded records from the uppermost current ones (15 s in duration) in the absence (left) and presence (right) of 10 μM RFM. The upper deflection indicates the opening event of the channel recorded at the holding potential of +60 mV. Of note, the BKCa-channel activity was conceivably enhanced as RFM was applied to the bath (i.e., in the cytosolic leaflet of the detached membrane patch). (B) Summary scatter graph demonstrating the effect of RFM, RFM plus paxilline, and RFM plus TRAM-34 on the probability of BKCa-channel openings (mean ± SEM; n = 7 for each point). Inside-out current recordings were performed in these experiments, the potential was set at +60 mV, and each tested compound was applied to bath medium. * Significantly different from control (p < 0.05) and ** significantly different from RFM (10 μM)-alone group (p < 0.05). (C) Mean I–V relationships of single BKCa-channel currents acquired in the absence (blue filled circles) and presence (orange open circles) of 10 μM RFM (mean ± SEM; n = 7 for each point). The dashed linear line was pointed toward 0 mV (i.e., reversal potential). Of note, these two linear I–V relationships of BKCa channels between the absence and presence of RFM are superimposed.
Figure 4. Effect of RFM on the activity of large-conductance Ca2+-activated K+ (BKCa) channels recorded from GH3 cells. The single-channel experiments were conducted as cells were exposed to high-K+ concentration (145 mM) containing 0.1 μM Ca2+, the pipette solution contained K+-enriched solution, and an inside-out configuration with a holding potential of +60 mV was created. (A) Representative tracings of BKCa channels in the absence (left, blue color) and presence (right, red color) of 10 μM RFM. RFM was applied to the bath medium. The lower five traces were expanded records from the uppermost current ones (15 s in duration) in the absence (left) and presence (right) of 10 μM RFM. The upper deflection indicates the opening event of the channel recorded at the holding potential of +60 mV. Of note, the BKCa-channel activity was conceivably enhanced as RFM was applied to the bath (i.e., in the cytosolic leaflet of the detached membrane patch). (B) Summary scatter graph demonstrating the effect of RFM, RFM plus paxilline, and RFM plus TRAM-34 on the probability of BKCa-channel openings (mean ± SEM; n = 7 for each point). Inside-out current recordings were performed in these experiments, the potential was set at +60 mV, and each tested compound was applied to bath medium. * Significantly different from control (p < 0.05) and ** significantly different from RFM (10 μM)-alone group (p < 0.05). (C) Mean I–V relationships of single BKCa-channel currents acquired in the absence (blue filled circles) and presence (orange open circles) of 10 μM RFM (mean ± SEM; n = 7 for each point). The dashed linear line was pointed toward 0 mV (i.e., reversal potential). Of note, these two linear I–V relationships of BKCa channels between the absence and presence of RFM are superimposed.
Ijms 23 13677 g004
Figure 5. BKCa-channel activity activated by inverted isosceles-triangular ramp pulse obtained with or without the RFM application. Inside-out current recordings were taken, bath medium contained 0.1 μM Ca2+, and an inverted triangular ramp pulse between +180 and −180 mV with a duration of 1.6 s (i.e., ramp pulse = ±450 mV/s) was applied to the patch. (A) Representative current traces (pink colors) obtained in the control period. The upper part shows the voltage-clamp protocol applied, and the solid arrow indicates the direction in which time passes. (B) Instantaneous relationship of BKCa-channel current versus membrane potential evoked by descending (blue open symbols in data points) or ascending limb (organ open symbols) activated during the inverted isosceles-triangular ramp pulse. (C) Hysteretic strength of BKCa-channel opening activated by 1.6 s long-lasting isosceles-triangular ramp pulse. Note that the relationship of relative open probability versus membrane potential is overly distinguishable between the descending (forward) and ascending (backward) limbs of the triangular ramp pulse. (D) Summary scatter graph showing the effect of RFM (3 or 10 μM) on voltage-dependent hysteresis of BKCa channels evoked by inverted triangular ramp pulse (mean ± SEM; n = 7 for each point). The hysteretic strength was measured at the voltage separation between the descending (forward) and ascending (backward) limb at 50% of the relative open probability. An Inside-out configuration was created, and an inverted isosceles-triangular ramp pulse with a ramp speed of ±450 mV/s was applied to the detached patch. * Significantly different from control (p < 0.05) and ** significantly different from RFM (3 μM)-alone group (p < 0.05).
Figure 5. BKCa-channel activity activated by inverted isosceles-triangular ramp pulse obtained with or without the RFM application. Inside-out current recordings were taken, bath medium contained 0.1 μM Ca2+, and an inverted triangular ramp pulse between +180 and −180 mV with a duration of 1.6 s (i.e., ramp pulse = ±450 mV/s) was applied to the patch. (A) Representative current traces (pink colors) obtained in the control period. The upper part shows the voltage-clamp protocol applied, and the solid arrow indicates the direction in which time passes. (B) Instantaneous relationship of BKCa-channel current versus membrane potential evoked by descending (blue open symbols in data points) or ascending limb (organ open symbols) activated during the inverted isosceles-triangular ramp pulse. (C) Hysteretic strength of BKCa-channel opening activated by 1.6 s long-lasting isosceles-triangular ramp pulse. Note that the relationship of relative open probability versus membrane potential is overly distinguishable between the descending (forward) and ascending (backward) limbs of the triangular ramp pulse. (D) Summary scatter graph showing the effect of RFM (3 or 10 μM) on voltage-dependent hysteresis of BKCa channels evoked by inverted triangular ramp pulse (mean ± SEM; n = 7 for each point). The hysteretic strength was measured at the voltage separation between the descending (forward) and ascending (backward) limb at 50% of the relative open probability. An Inside-out configuration was created, and an inverted isosceles-triangular ramp pulse with a ramp speed of ±450 mV/s was applied to the detached patch. * Significantly different from control (p < 0.05) and ** significantly different from RFM (3 μM)-alone group (p < 0.05).
Ijms 23 13677 g005aIjms 23 13677 g005b
Figure 6. Inhibitory effects of RFM on voltage-gated Na+ current (INa) identified in GH3 cells. The experiments were conducted when cells were bathed in Ca2+-free, Tyrode’s solution containing 10 mM TEA and 0.5 mM CdCl2, and the electrode that we used was filled with a Cs+-enriched solution. (A) Representative potential (upper) and current traces (lower) obtained in the control period (a) and during exposure to 3 μM RFM (b). (B) Representative expanded traces (indicated on a faster time scale) shown from the dashed box in (A). Note that cell exposure to RFM differentially suppresses the peak and late components of INa in response to rapid membrane depolarization. (C) Concentration-dependent inhibition of RFM on the peak (blue open squares) and late (red open circles) components of INa (mean ± SEM; n = 8 for each point). Current amplitude was taken at the beginning and end of the 30 ms depolarizing step applied from −80 to −10 mV. The modified Hill equation indicated in Materials and Methods was reasonably fitted to the experimental data (blue and red smooth lines).
Figure 6. Inhibitory effects of RFM on voltage-gated Na+ current (INa) identified in GH3 cells. The experiments were conducted when cells were bathed in Ca2+-free, Tyrode’s solution containing 10 mM TEA and 0.5 mM CdCl2, and the electrode that we used was filled with a Cs+-enriched solution. (A) Representative potential (upper) and current traces (lower) obtained in the control period (a) and during exposure to 3 μM RFM (b). (B) Representative expanded traces (indicated on a faster time scale) shown from the dashed box in (A). Note that cell exposure to RFM differentially suppresses the peak and late components of INa in response to rapid membrane depolarization. (C) Concentration-dependent inhibition of RFM on the peak (blue open squares) and late (red open circles) components of INa (mean ± SEM; n = 8 for each point). Current amplitude was taken at the beginning and end of the 30 ms depolarizing step applied from −80 to −10 mV. The modified Hill equation indicated in Materials and Methods was reasonably fitted to the experimental data (blue and red smooth lines).
Ijms 23 13677 g006
Figure 7. Effect of RFM on mean I–V relationship of peak INa in GH3 cells. The examined cell was held at −80 mV and a series of voltage steps between −80 and +40 mV in 10-mV increments was thereafter applied to it. Current amplitude measured at the start of the abrupt depolarizing step was taken. The data points indicated in blue-filled squares are controls (i.e., absence of RFM), and those in red open squares were taken in the presence of 3 μM RFM. Each data point represents the mean ± SEM (n = 7). The smooth continuous lines were least-squares fitted to a modified Boltzmann function detailed in Materials and Methods.
Figure 7. Effect of RFM on mean I–V relationship of peak INa in GH3 cells. The examined cell was held at −80 mV and a series of voltage steps between −80 and +40 mV in 10-mV increments was thereafter applied to it. Current amplitude measured at the start of the abrupt depolarizing step was taken. The data points indicated in blue-filled squares are controls (i.e., absence of RFM), and those in red open squares were taken in the presence of 3 μM RFM. Each data point represents the mean ± SEM (n = 7). The smooth continuous lines were least-squares fitted to a modified Boltzmann function detailed in Materials and Methods.
Ijms 23 13677 g007
Figure 8. Effect of tefluthrin (Tef) and Tef plus RFM (10 or 30 μM) on INa evoked by the rapid depolarizing step in GH3 cells. (A) Representative current traces obtained in the presence of 10 μM Tef (a) and during the exposure to 10 μM Tef plus 10 μM RFM (b) or 10 μM Tef plus 30 μM RFM (c). The voltage−clamp protocol applied is displayed in the upper part. (B) Current traces taken from the dashed box in (A). Of notice, in the continued presence of Tef, further application of RFM remains effective in reversing Tef-stimulated INa in these cells.
Figure 8. Effect of tefluthrin (Tef) and Tef plus RFM (10 or 30 μM) on INa evoked by the rapid depolarizing step in GH3 cells. (A) Representative current traces obtained in the presence of 10 μM Tef (a) and during the exposure to 10 μM Tef plus 10 μM RFM (b) or 10 μM Tef plus 30 μM RFM (c). The voltage−clamp protocol applied is displayed in the upper part. (B) Current traces taken from the dashed box in (A). Of notice, in the continued presence of Tef, further application of RFM remains effective in reversing Tef-stimulated INa in these cells.
Ijms 23 13677 g008
Figure 9. Docking results of KCa1.1 channel and rufinamide (RFM). The protein structure of the KCa1.1 channel was obtained from PDB (PDB ID: 6V3G), while the structure of RFM was acquired from PubChem (Compound CID: 129228). The structure of the KCa1.1 channel was docked with the RFM molecule through PyRx (https://pyrx.sourceforge.io/, URL accessed on 2 November 2022). The diagram of the interaction between the KCa1.1 channel and the RFM molecule was generated by LigPlot+ (https://www.ebi.ac.uk/thornton-srv/software/LIGPLOT/, accessed on 2 November 2022). Red arcs with spokes radiating toward the ligand denote the hydrophobic contact, while the green dotted line depicts the hydrogen bond.
Figure 9. Docking results of KCa1.1 channel and rufinamide (RFM). The protein structure of the KCa1.1 channel was obtained from PDB (PDB ID: 6V3G), while the structure of RFM was acquired from PubChem (Compound CID: 129228). The structure of the KCa1.1 channel was docked with the RFM molecule through PyRx (https://pyrx.sourceforge.io/, URL accessed on 2 November 2022). The diagram of the interaction between the KCa1.1 channel and the RFM molecule was generated by LigPlot+ (https://www.ebi.ac.uk/thornton-srv/software/LIGPLOT/, accessed on 2 November 2022). Red arcs with spokes radiating toward the ligand denote the hydrophobic contact, while the green dotted line depicts the hydrogen bond.
Ijms 23 13677 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lai, M.-C.; Wu, S.-N.; Huang, C.-W. Rufinamide, a Triazole-Derived Antiepileptic Drug, Stimulates Ca2+-Activated K+ Currents While Inhibiting Voltage-Gated Na+ Currents. Int. J. Mol. Sci. 2022, 23, 13677. https://doi.org/10.3390/ijms232213677

AMA Style

Lai M-C, Wu S-N, Huang C-W. Rufinamide, a Triazole-Derived Antiepileptic Drug, Stimulates Ca2+-Activated K+ Currents While Inhibiting Voltage-Gated Na+ Currents. International Journal of Molecular Sciences. 2022; 23(22):13677. https://doi.org/10.3390/ijms232213677

Chicago/Turabian Style

Lai, Ming-Chi, Sheng-Nan Wu, and Chin-Wei Huang. 2022. "Rufinamide, a Triazole-Derived Antiepileptic Drug, Stimulates Ca2+-Activated K+ Currents While Inhibiting Voltage-Gated Na+ Currents" International Journal of Molecular Sciences 23, no. 22: 13677. https://doi.org/10.3390/ijms232213677

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop