Skip to main content
Neuroscience Bulletin logoLink to Neuroscience Bulletin
. 2017 May 9;33(4):455–477. doi: 10.1007/s12264-017-0134-1

Ion Channel Genes and Epilepsy: Functional Alteration, Pathogenic Potential, and Mechanism of Epilepsy

Feng Wei 1,2,#, Li-Min Yan 1,#, Tao Su 1, Na He 1, Zhi-Jian Lin 1, Jie Wang 1, Yi-Wu Shi 1, Yong-Hong Yi 1, Wei-Ping Liao 1,
PMCID: PMC5567559  PMID: 28488083

Abstract

Ion channels are crucial in the generation and modulation of excitability in the nervous system and have been implicated in human epilepsy. Forty-one epilepsy-associated ion channel genes and their mutations are systematically reviewed. In this paper, we analyzed the genotypes, functional alterations (funotypes), and phenotypes of these mutations. Eleven genes featured loss-of-function mutations and six had gain-of-function mutations. Nine genes displayed diversified funotypes, among which a distinct funotype-phenotype correlation was found in SCN1A. These data suggest that the funotype is an essential consideration in evaluating the pathogenicity of mutations and a distinct funotype or funotype-phenotype correlation helps to define the pathogenic potential of a gene.

Keywords: Epilepsy, Ion channel gene, Epilepsy gene, Genetics, Gene function, Pathogenic mechanism

Introduction

Epilepsy is a group of chronic brain disorders characterized by recurrent seizures due to abnormal excessive electrical discharges of cerebral neurons [1]. It is generally believed that genetic factors play an important role in the etiopathogenesis of epilepsy. Recent studies have demonstrated that 977 genes are associated with epilepsy, among which genes encoding ion channels predominate [2].

Ion channels are pore-forming membrane proteins. Their functions include establishing action potentials and maintaining homeostasis by gating the ionic flow traversing the cell membrane, managing the ionic flow across cells, and regulating cell volume. Since these functions are essential to the excitability of neuron, ion channels potentially play a critical role in epileptogenesis [3]. The association between ion channel genes and epilepsy may provide insights into the mechanisms underlying epilepsy.

CHRNA4, which encodes the α4 subunit of the ligand-gated ion channel nAChR (nicotinic acetylcholine receptor), was the first epilepsy gene identified in patients with autosomal-dominant nocturnal frontal lobe epilepsy (ADNFLE) in 1995 [4]. Since then, many ion channel genes have been reported to be potentially associated with epilepsy [2]. However, the associations differ. Genes like CHRNA4 have been confirmed to be epilepsy genes by familial co-segregation, multi-source validations, and functional alteration [47], whereas some genes warrant further investigation. Functional studies are used to determine the impairments caused by gene mutations and provide insights into the underlying mechanism of epilepsy. Evidence from functional studies is also helpful in evaluating the pathogenicity of a gene and its mutations, especially when considered together with clinical features.

In this review, we summarize the epilepsy-associated ion channel genes, the mutations, the functional changes in mutants, and the corresponding phenotypes and inheritance, aiming to provide clues for evaluating the association between ion channel genes and epilepsy and understanding the mechanisms of epilepsy.

Gene Searching and Analysis Strategies

Based on several databases (OMIM (Online Mendelian Inheritance in Man), HGMD (Human Gene Mutation Database), and EpilepsyGene) and recent publications in PubMed, we previously retrieved 977 epilepsy-associated genes [2], among which 60 are ion channel genes, including 28 epilepsy genes and 32 potential epilepsy genes (not yet in the OMIM database). We systematically searched all the publications for mutational and functional studies of these genes. For each gene, we searched the PubMed database using the terms gene symbol, gene full name, and the corresponding gene-encoded product, like “CHRNA4”, “cholinergic receptor nicotinic alpha 4 subunit”, and “nAChRα4” or “α4 subunit of nAChR”. Additional searches were performed according to the reference list of the publications. We included all epilepsy genes and potential epilepsy-associated genes with functional studies performed on their mutants. For genes with updated reviews, such as SCN1A, we summarized the data from recent publications [8].

Functional alterations (functional types, or funotypes) are generally classified into gain-of-function (GOF), loss-of-function (LOF; that refers to the complete loss of function), and partial loss-of-function (pLOF; that denotes mutants with residual current (function)) as in our previous report [9]. To facilitate understanding of the functional consequences, gene mutations are classified into destructive and missense mutations. Destructive mutations are those causing gross protein malformations, including truncating mutations (nonsense and frameshifting mutations), splice-site mutations, and mutations with genomic rearrangement, which mainly lead to functional deficiency and haploinsufficiency. The functional consequences of missense mutations need to be determined in biophysiological studies.

When referring to animal models of these genes, data from mutation-specific and phenotype-related models are described and cited in this work. More information on genetic knock-out or knock-in mouse models can be retrieved from the Mouse Genome Informatics database (http://www.informatics.jax.org/).

Gene Mutations, Functional Alterations, and Pathogenic Mechanisms

Forty-one ion channel genes were included in the analysis, covering 28 epilepsy genes, one epilepsy-related gene (GRIN1), and 12 potential epilepsy genes. Based on their biophysical and physiological characteristics, these genes were classified into eight main groups (Table 1). More than 1,600 mutations have been identified in these genes, most of which are associated with more than one epileptic phenotype. In the following we analyze the corresponding mutations and their functional changes in each phenotype.

Table 1.

Summary of human ion channels implicated in epilepsies.

Gating categories Main functions Gene (Protein)
Sodium Channels Responsible for generation and propagation of action potentials SCN1A (NaV1.1), SCN1B (NaVβ1), SCN2A (NaV1.2), SCN3A (NaV1.3), SCN8A (NaV1.6), SCN9A†(NaV1.7)
Potassium Channels
 Voltage-gated Regulation of outward K+ currents and action potentials, modulation of neurotransmitter release KCNA2 (KV1.2), KCNB1 (KV2.1), KCNC1 (KV3.1), KCND2 (KV4.2), KCND3 (KV4.3), KCNH2 (KV11.1), KCNH5 (KV10.2), KCNQ2 (KV7.2), KCNQ3 (KV7.3), KCNV2 (KV8.2)
 (Ca2+-activated) Regulation of neuronal firing properties and circuit excitability KCNMA1 (KCa1.1)
 (Na+-activated) Regulation of delayed outward IKNa currents and contribution to adaptation of firing rate KCNT1 (KCa4.1)
 Calcium channels React to membrane potential depolarization by opening and provide an elevation of Ca2+ ions to drive many processes CACNA1A (CaV2.1), CACNA1H (CaV3.2), CACNA2D2#(CaVα2δ-2), CACNB4 (CaVβ4),
 Chloride channels Maintenance of resting membrane potential and regulation of cell volume CLCN2 (CLC-2), CLCN4 (CLC-4)
 γ-Aminobutyric acid type A receptor Mediation of major inhibitory functions in CNS GABRA1 (GABAAα1), GABRA6 (GABAAα6), GABRB1 (GABAAβ1), GABRB2 (GABAAβ2), GABRB3 (GABAAβ3), GABRD (GABAAδ), GABRG2 (GABAAγ2)
 Ionotropic glutamate receptors Excitatory synaptic transmission, plasticity, and excitotoxicity of the CNS GRIN1 (GluN1), GRIN2A (GluN2A), GRIN2B (GluN2B), GRIN2D (GluN2D)
 Nicotinic acetylcholine receptors Permeation of Na+ and K+ and modulation of neurotransmitter release CHRNA2 (nAChRα2), CHRNA4 (nAChRα4), CHRNA7 (nAChRα7), CHRNB2 (nAChRβ2)
 Hyperpolarization-activated cyclic nucleotide-gated channels Permeation of Na+ and K+ HCN1 (HCN1), HCN2 (HCN2)

Underlined: potential epilepsy-associated genes with functional alterations examined.

SCN9A may be one of the digenic causes of Dravet Syndrome with SCN1A.

# CACNA2D2 may be one of the digenic causes of epilepsy with CELSR3.

Sodium Channel Genes

Voltage-gated Na+ channels in the brain are composed of one large pore-forming α subunit and two smaller β subunits [10]. They are critical for neuronal excitability, including action potential initiation and conduction [11]. The α-subunit is capable of conducting currents on its own, and is expressed in a tissue-specific manner. SCN1A (encoding NaV1.1), SCN2A (encoding NaV1.2), SCN3A (encoding NaV1.3), SCN8A (encoding NaV1.6), and SCN9A (encoding NaV1.7) have been associated with epilepsy. The β-subunits modulate multiple aspects of NaV channel behavior and are essential for the control of neuronal excitability [12]. SCN1B (encoding NaVβ1) is an epilepsy gene (Table 2).

Table 2.

Mutations in epilepsy-associated Na+ channel genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
SCN1A SME de novo (mostly), inherited 42.3% cases have missense mutations Missense mutations in pore region (54.1%) lead to LOF or pLOF, in other regions can cause pLOF, G-LOF, and LOF [8]
57.7% cases have destructive mutations
PE and/or FS+ de novo, inherited 74.5% cases have missense mutations Missense mutations in pore region (42.1%) lead to LOF or pLOF, in other regions can cause pLOF, G-LOF, and LOF
25.5% cases have destructive mutations
GE and/or FS+ Inherited (mostly), de novo 87.0% cases have missense mutations Milder functional alterations, such as increased excitability, decreased excitability, pLOF, or pure GOF, but no LOF and G-LOF
13.0% cases have destructive mutations
SCN1B EFS+ AD R85C, R85H, C121W pLOF [11, 19, 20]
R125L, K208I Not available
c.208-2A>C Destructive
IE Unknown T28I Not available
PS Unknown D25N Not available
SME AR R125C (homozygous) LOF [22]
I106F (homozygous) Not available
SCN2A BFNS AD R1319Q, L1330F pLOF [25]
L1563V, M252V GOF [25, 26]
>12 missense mutations Not available
1 gross insertion and 2 gross deletions Destructive
de novo V261M GOF [26]
FS and GEFS+ AD R188W pLOF [188]
EE de novo A263V GOF [189]
R102X Destructive, LOF [27]
>25 missense mutations Not available
2 gross insertions Destructive
SCN3A CPS Paternal† K354Q GOF [33]
Unknown E1111K GOF [34]
R357Q, D766N Unchanged [34]
PEFS+ Unknown M1323V GOF [34]
GEFS+ de novo N302S pLOF [35]
SCN8A EE de novo T767I, N984K, N1768D, R1617Q, R1872W, R1872L, R1872Q GOF [37, 40]
R223G pLOF, thermolabile mutant [39]
G1451S LOF at 37°C [40]
>30 missense mutations Not available
2 deletions Destructive
SCN9A SME Maternal† I228M# GOF [44]
I684M#, L1123F#, E519K, E1160Q Not available
Paternal† K655R#, I739V#, C699Y Not available
Unknown I1267V#, K655R Not available
FS and FS+ AD N641Y GOF (in mouse model) [45]
Paternal† I739V Not available
Unknown I62V, P149Q, K655R, S490N Not available

AD, autosomal dominant; AR, autosomal recessive; BFNS, benign familial neonatal seizures; CPS, complex partial seizure; EE, epileptic encephalopathy; FS, febrile seizure; GE, generalized epilepsy; GEFS+, generalized epilepsy with febrile seizures plus; GOF, gain-of-function; IE, idiopathic epilepsy; LOF, loss-of-function; pLOF, partial loss-of-function; PE, partial epilepsy; PEFS+, partial epilepsy with febrile seizures plus; PS, partial seizure; SME, severe myoclonic epilepsy.

† Incomplete penetrance; transmitter not affected.

# Combined with SCN1A mutation.

SCN1A is expressed at a high level in the central nervous system (CNS), and NaV1.1 is found predominantly in the somata and dendrites of neurons [13]. SCN1A is one of the most important causative genes in epilepsy. To date, >1,257 epilepsy-related mutations have been reported [8], mainly in patients having epilepsy with antecedent febrile seizures (FS). Severe myoclonic epilepsy (SME) in infancy is the most severe phenotype and is frequently associated with destructive or missense mutations located in the pore region which cause LOF of NaV1.1. In contrast, mild generalized epilepsy with febrile seizures plus (GEFS+) or FS has the highest frequency of missense mutations that are usually located outside the pore region and cause mild functional changes. Partial epilepsy with FS+ (PEFS+) is an intermediate phenotype in terms of both clinical severity and mutation impairment. These data suggest that LOF of NaV1.1 is the primary basis for epilepsies with FS+, and the clinical severity is correlated with the functional impairment in a quantity-dependent manner. Experiments in Scn1a knock-out mice have demonstrated that the Na+ current density is reduced in inhibitory interneurons, but not in excitatory pyramidal neurons, explaining how the LOF of Nav1.1 would impair inhibitory functions in the brain and lead to hyperexcitability and epilepsy. Inhibitory interneurons are generally distributed locally with heterogeneity in different brain areas [14], explaining the common partial seizures in SME and PEFS+ [8, 1517].

SCN1B encodes NaVβ1 that can influence many cardinal conformational changes of NaV channels during the action potential process [18]. SCN1B mutations were initially identified in families with epilepsy and FS. Functional studies on mutants (R85C, R85H, C121W, and R125C) revealed LOF of β1 and subsequently impaired function of Na+ channels [11, 19, 20]. Two homozygous missense mutations (I106F and R125C) have been identified in more severe cases (SME patients) [21, 22], indicating a quantity-dependent feature. The SCN1B phenotype shows clinical features similar to those of SCN1A, suggesting that the mechanism underlying the pathogenicity of SCN1B mutations potentially involves impaired function of NaV1.1.

The temporal expression pattern of SCN2A in the brain is similar to SCN1A, but NaV1.2 is specifically localized in axons and terminals [13]. SCN2A mutations were initially identified in families with benign familial neonatal-infantile seizures [23, 24]. Functional studies showed pLOF with decreased channel availability in two mutants and GOF in another two mutants [2328]. It is hard to explain the pathogenicity of heterozygous mutations with pure LOF or pLOF, since heterozygous knock-out of Scn2a in mice does not result in seizure activity [29]. There is no mutation-specific knock-in model to show whether a mutation with GOF would be pathogenic. SCN2A is transcribed in different splice forms during neonatal and adult stages. The neonatal splice isoform is less excitable than that of adults, and mutants would change the channels to a more excitable status than the neonatal isoform but at a level similar to adult channels [30]. This may be one of the explanations for the pathogenicity of SCN2A mutations in neonates. Multiple de novo mutations have been identified in patients with epileptic encephalopathies (EEs) through next-generation sequencing. However, their roles in the pathogenicity of EEs are currently uncertain due to a lack of evidence.

Scn3a in rodents is expressed at the highest level in the embryonic and early postnatal brains and gradually disappears thereafter [31]. In contrast, SCN3A is expressed in small amounts in the adult human brain, and NaV1.3 shows a somatodendritic localization [32]. SCN3A has been potentially associated with epilepsy in several publications [3335]. Functional analyses have shown GOF in three [33, 34], pLOF in one [35], and no significant changes in two of the mutants [34]. The functional changes are generally slight in these mutants. The relationship between SCN3A and epilepsy remains to be clarified.

SCN8A is highly expressed in cerebellar granule cells and in pyramidal and granule cells of the hippocampus [13]. SCN8A has been associated with EEs in recent years. More than 40 de novo SCN8A mutations have been identified in cases with various EEs. All the mutations are missense, except two that are destructive. Nine of the missense mutations have been characterized in functional studies. A majority of the mutations, including T767I, N984K, N1768D, R1617Q, R1872W, R1872L, and R1872Q, have demonstrated GOF; whereas G1451S and R223G display LOF or pLOF with thermosensitivity [3640]. No distinct genotype-phenotype or funotype-genotype association has been found. Considering that Scn8a-null heterozygote mice are seizure-resistant [41, 42], mutants with LOF are unlikely to be pathogenic. Further studies are required to determine the role of SCN8A mutations in EEs and the underlying mechanism.

SCN9A is expressed predominantly in the peripheral nervous system and slightly in the CNS. The first suspicion of an association between SCN9A and epilepsy came from linkage analysis that located an FS-related locus in the genomic region containing SCN1A, SCN2A, and SCN3A[43], and SCN9A is located nearby. Several SCN9A mutations have been identified in patients with FS-related epilepsies [44, 45]. In an SME cohort, six of nine patients with SCN9A missense variants also harbored SCN1A mutations [45], suggesting that SCN9A may be one of the digenic causes of SME. A mouse model with knock-in of N641Y presents susceptibility to epileptic seizures, suggesting that SCN9A may be a modifier or susceptibility gene of epilepsy [45].

Although voltage-gated Na+ channels have molecular and physiological characteristics in common, their associations with epilepsy differ in many aspects, including phenotype, pathogenic funotype, and the underlying pathogenic mechanism.

Potassium Channel Genes

K+ channels control the resting membrane potential and enable rapid repolarization of the action potential by producing outward K+ currents, thus limiting neuronal excitability [46]. K+ channels are composed of four pore-forming α subunits and modulatory β subunits. Voltage-gated K+ (KV) channels are the largest ion channel group that are expressed substantially in the CNS. KV channels, including Ca2+-activated and Na+-activated K+ channels, have been associated with epilepsies (Table 3).

Table 3.

Mutations in the epilepsy-associated K+ channel genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
KCNA2 EE de novo I263T, P405L LOF with dominant-negative effect [47]
R297Q, L298F GOF
L290R Not available
KCNB1 EE de novo R306C, S347R†, T374I†, V378A†, G379R†, G401R LOF [5153]
R312H, G381R, F416L Not available
KCNC1 PME de novo R320H LOF with dominant-negative effect [54]
KCND2 EE and autism de novo V404M GOF [57]
TLE Paternal# N587fsX1 Destructive, pLOF [58]
KCND3 GE de novo R293_F295dup pLOF [59]
KCNH2 Epilepsy with LQT2 AD A429P LOF [61]
Y493F pLOF
c. 234-241del Destructive, LOF
de novo R863X Destructive [62]
Unknown I82T LOF [63]
KCNH5 EE de novo R327H GOF [64, 65]
KCNQ2 BFNS AD S247W, G271V, W344R, R353G LOF [70, 71, 79, 190, 191]
E119G, S122L, A196V, L197P, R207W, R207Q, M208V, R214W, N258S, Y284C, A294G, A306T, R333Q, L351F, T359K, R553Q pLOF [6778]
L619R GOF [79]
L351V, Y362C Unchanged [71]
>15 Missense mutations Not available
S247X, S399X, K537X, c.761_770del10insA Destructive, LOF [68, 70, 91]
Q323X, R448X, V589X, P410fsX12, 867ins5bp Destructive, pLOF [70, 72, 75]
R581X,W269X, 2 small insertions, 8 small deletions, 14 splicing Destructive
de novo D212G, R213W pLOF [89, 192]
R333W Not available
F304del Destructive, LOF [101]
1 small deletion, 1 splicing site Destructive
EE de novo A265P, T274M, G290D, A294V LOF [80, 81]
S122L, A196V, I205V, M532W, R560W pLOF [8184]
R144Q, R201C, R201H GOF [86]
>35 Missense mutations Not available
W157X, 2 small indels, 1 small insertion,2 splicing Destructive
Paternal# R213Q pLOF [73, 82]
KCNQ3 BFNS AD I317T,W309R LOF [70, 90]
E299K, D305G, G310V, N821S pLOF [75, 78, 91, 92]
N468S Unchanged [75]
R330C,R330H, G340V, R780C Not available
de novo R364H Not available
BECTS AD A381V, P574S Unchanged [92]
IE Unknown P574T Not available
EE AD R330L pLOF [85]
de novo R230C GOF [86]
KCNV2 PS Maternal# R7K GOF [94]
EE Maternal# M285R GOF
KCNMA1 GS and PD AD D434G GOF [96, 193]
KCNT1 ADNFLE AD G288S, R398Q,Y796H, R928C GOF [97, 100]
de novo M896I GOF [97, 105]
EE de novo G288S,R398Q, R428Q, R474H, M516V, K629N, I760M, Y796H, P924L, A934T GOF [99, 101105]
H257D, R262Q, Q270E, V271F, P409S, R428Q, R429C, R429H, R474C, A477T, K629E, M896K, R933G, R950Q Not available
Paternal‡ A966T (homozygote) GOF [194]

AD, autosomal dominant; ADNFLE, autosomal dominant nocturnal frontal lobe epilepsy; AR, autosomal recessive; BECTS, benign epilepsy of childhood with centrotemporal spikes; BFNS, benign familial neonatal seizures; EE, epileptic encephalopathy; GE, generalized epilepsy; GS, general seizure; IE, idiopathic epilepsy; LQT2, long-QT syndrome type 2; PD, paroxysmal dyskinesia; PME, progressive myoclonic epilepsy; PS, partial seizure; TLE, temporal lobe epilepsy.

† Mutated channel lost K+ selectivity and increased permeability to other positive and negative ions.

# Incomplete penetrance; transmitter not affected.

‡ Patient had paternal isodisomy for chromosome 9; father not affected.

KCNA2 encodes KV1.2 that is expressed in axons and synaptic terminals; it enables efficient repolarization following an action potential [47]. Five missense mutations within KCNA2 have been identified in patients with EEs [47, 48]. Functional studies of the mutations I263T and P405L have shown LOF with a dominant-negative effect [47], predicting hyperexcitable neuronal membranes and repetitive firing due to impaired repolarization. Kcna2-knock-out mice display increased seizure susceptibility and premature death, supporting the role of LOF mutants in epilepsy [49]. Another two mutations (R297Q and L298F) demonstrate GOF, predicting permanently open channels at physiological membrane potentials and electrical silencing by membrane hyperpolarization [47]. Further studies are required to elucidate the mechanism of action of KV1.2 GOF mutants in epileptogenesis.

KCNB1 encodes KV2.1, which is the main contributor to the delayed rectifier K+ current in pyramidal neurons of the hippocampus and cortex [50]. This current is vital for membrane repolarization and for suppressing high-frequency firing. Nine KCNB1 mutations have been reported in EE patients, and most of the mutations are located in the pore region [5153]. Six mutations show LOF, and four of them (S347R, T374I, V378A, and G379R) also cause a loss of K+ selectivity with a dominant-negative effect [51]. Considering the function of suppressing high-frequency firing, LOF of KV2.1 predicts hyperactivity of neuronal networks and an increase in the risk of seizures.

KCNC1 encodes KV3.1, a member of the KV3 subfamily that shows more positively shifted voltage-dependent activation and faster activation and deactivation rates than other KV channels. A de novo mutation (R320H) in KCNC1 has been identified in a patient with progressive myoclonic epilepsy, and shown to display LOF in a functional study [54]. KV3.1 is preferentially expressed in fast-spiking inhibitory GABAergic interneurons and enables them to fire at high frequencies [55]. Lacking KV3.1 function may impair the firing of fast-spiking GABAergic interneurons and subsequently result in hyperexcitability of the brain.

Both KV4.2 (encoded by KCND2) and KV4.3 (encoded by KCND3) are members of the KV4 subfamily, which regulate the rate of low-frequency firing and control the backpropagation of action potentials into the dendritic tree [56]. A de novo mutation within KCND2 (V404M) has been identified in a pair of twins with comorbidity of autism and epilepsy, showing GOF and profound impairment of closed-state inactivation [57]. A paternally-inherited truncated KCND2 mutation (N587fsX1) has been found in a patient with temporal lobe epilepsy (incomplete penetrance, the father was not affected), which showed pLOF and a reduction of the inhibitory current contributing to aberrant neuronal excitability [58]. A de novo duplicated KCND3 mutation (R293_F295dup) has been reported in a patient with generalized epilepsy and shows pLOF with a great depolarizing shift in the voltage-dependence of both KV4.3 activation and inactivation [59]. Due to the limited data and a lack of a genotype (or funotype)-phenotype correlation, it is hard to define the association between KV4 and epilepsy.

KCNH2 (also known as hERG) encodes KV11.1 that is widely expressed in the human brain and heart. In the brain, KV11.1 regulates neuronal firing and modulates the excitability of GABAergic and dopaminergic neurons [60]. KCNH2 mutants were initially reported to be associated with long-QT syndrome type 2 (LQT2). To date, five mutations have been identified in patients with LQT2 and variable seizures. Functional analyses have shown LOF in all mutations [6163], suggesting that LOF of KV11.1 may increase the risk of epilepsy.

KCNH5 encodes KV10.2 that is selectively expressed in interneurons localized to layer IV of the cerebral cortex in multiple areas, especially in numerous excitatory interneurons [64]. A patient with EE and multiple neurodevelopmental deficits has been reported to carry a de novo R327H mutation that confers a GOF change in the KV10.2 channel [64, 65]. Since layer IV contains both excitatory and inhibitory interneurons [65], it is hard to estimate the effect of this KV10.2 mutant on epilepsy.

KCNQ2 encodes KV7.2 and KCNQ3 encodes KV7.3. KV7 channels mediate low-threshold, slowly-activating, non-inactivating muscarinic currents [66]. Opening of homogeneous KV7.2 or heterogeneous KV7.2/KV7.3 complexes inhibits initiation of the action potential and thus suppresses neuronal excitability [66]. Mutations in KCNQ2 were initially identified in patients with benign familial neonatal seizures (BFNS). Functional studies have illustrated LOF or pLOF in a majority of mutants [6778], GOF in one mutation (L619R) [79], and no significant change in two mutations (L351V and Y362C) [71]. Mutations in KCNQ2 have also been identified in patients with EEs. Nine mutations demonstrate LOF or pLOF [8085], while three (R144Q, R201H, and R201C) demonstrate GOF by stabilizing the activated state of the channels [86]. Mutations A196V and S122L have been identified in both benign BFNS and intractable EE [68, 69, 83, 84]. Mice expressing LOF mutant KV7.2 channels display spontaneous seizures, behavioral hyperactivity, and increased hippocampal neuronal excitability and cell death [87]. A 25% reduction in the muscarinic current amplitude is sufficient to cause electrical hyperexcitability and leads to neonatal/infantile epilepsy in humans [78, 88]. Therefore, LOF of KV7.2 leads to neuronal hyperexcitability and induces epileptogenesis. The EE-related R213Q mutation causes significantly more evident kinetic alterations than the BFNS-related R213W mutation [89], suggesting a potential genotype-phenotype correlation. The role of KV7 GOF mutants in epileptogenesis is still under debate [79, 86]. Similarly, mutations in KCNQ3 have been identified in patients with BFNS and mainly show LOF or pLOF [70, 75, 78, 9092]. Another two de novo KCNQ3 mutations have been identified in patients with EEs and each displays pLOF or GOF [85, 86].

KCNV2 encodes KV8.2, which is electrophysiologically silent when expressed as a homotetramer. However, when assembled with KV2 subunits, KV8.2 significantly reduces the membrane expression of heterotetrameric channels and suppresses KV2 currents [93]. The KV8.2 and KV2.1 subunits show a remarkable regional overlap in their CNS expression patterns [60]. Two mutations in KCNV2, R7K and M285R, have been identified in patients with partial seizures and EE, respectively [94]. They show GOF and enhanced Kv8.2-mediated suppression of KV2.1 currents, subsequently reducing KV2.1 currents and leading to epilepsy. The M285R mutant, which was identified in a patient with EE, also causes defects of KV2.1 activation kinetics [94], potentially explaining the more severe phenotype.

KCNMA1 encodes the α-subunit of large-conductance Ca2+-activated KCa1.1 channels. KCa1.1 is predominantly expressed in the axons and presynaptic terminals of neurons and promotes high-frequency firing [95]. A GOF mutation in KCNMA1 has been detected in a large family with generalized epilepsy and paroxysmal dyskinesia [96]. The enhanced Ca2+-activated K+ current (BK current) increases the firing rate and spontaneous non-convulsive seizures in mice [96]. Thus it is possible that GOF of KCa1.1 increases the BK current and enables faster re-priming (removal of inactivation) of Na+ channels, leading to hyperexcitability.

KCNT1 encodes the α-subunit of the Na+-activated channel KCa4.1 (also known as Slack, KCNT1, or Slo2.2), which is highly expressed in many regions of the brain, and significantly found in neurons of the frontal cortex [97]. The precise function of homotetrameric KCa4.1 channels is unclear. Functional heterotetrameric channels consisting of KCa4.1 and KCa4.2 (encoded by KCNT2) subunits contribute to the delayed outward current I KNa, which helps to modulate neuronal excitability and adaptability in response to high-frequency stimulation [98]. Mutations in KCNT1 have been found in ADNFLE and EEs (especially epilepsy of infancy with migrating focal seizures). Two mutations (G288S and R398Q) have been identified in both ADNFLE and EE patients. All known functional consequences of KCNT1 mutations show a strong GOF effect [97, 99105]. Although the actual mechanisms by which GOF mutations lead to neuronal hyperexcitability are uncertain, KCNT1 could be confirmed as an epilepsy gene when clinical evidence is taken into account.

Calcium Channel Genes

Voltage-gated Ca2+ (CaV) channels conduct an inward Ca2+ current after depolarization, mediate action potential firing and membrane oscillations, and thus have widespread effects on neuronal excitability [106]. Each CaV channel consists of one principal α1 subunit, which forms the pore and defines the channel type, and modulates β, α2δ, and possibly γ subunits. CACNA1A, CACNA1H, CACNA2D2, and CACNB4 are associated with epilepsies (Table 4).

Table 4.

Mutations in epilepsy-associated Ca2+-channel genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
CACNA1A EE de novo E101Q, A712T, A713T, S1373L, A1511S Not available
IGE Unknown R477H, R1967Q Not available
Q1397X Destructive
CACNA1H CAE Inherited F161L, E282K, C456S, G499S, P648L, G773D, R788C, V831M, A876T, T1606M, R1892H GOF [109112]
T1733A† pLOF
R744Q, A748V, G784S,G848S, Q1264H, D1463N Unchanged [109, 111]
P314S, P492S, H515Y, Q1264H Not available
MAE Inherited G983S LOF [112]
IGE Inherited R788C, T1606M, A1705T# GOF [112]
A1059T pLOF
E1170K Unchanged
Unknown A480T, P618L, G755D, E1170K Not available
V621fsX33 Destructive
CACNA2D2 EE AR L1040P (homozygous) pLOF [113]
N432fsX (homozygous) Destructive
CACNB4 JME AD R482X Destructive, LOF [115]
IGE AD C104F LOF [115]

AD, autosomal dominant; AR, autosomal recessive; CAE, childhood absence epilepsy; EE, epileptic encephalopathy; IGE, idiopathic generalized epilepsy; JME, juvenile myoclonic epilepsy; MAE, myoclonic-astatic epilepsy.

†Not segregated with epilepsy in the two affected siblings of a CAE family. #A1705T co-segregates with R788C in all carriers.

CACNA1A encodes the α1 subunit of CaV2.1, forming a P/Q-type voltage-gated Ca2+ channel. Mutations in CACNA1A have been identified in an IGE cohort [107]. One recent study demonstrated de novo CACNA1A mutations in patients with EEs [108]. Functional studies on these mutants have not been performed.

CACNA1H encodes the α1 subunit of CaV3.2, a member of the CaV3 subfamily. CaV3 channels are highly expressed in thalamic neurons, conduct low-voltage activated T-type (transient) Ca2+ currents, and play roles in circadian rhythms. Twenty-two mutations in CACNA1H have been identified in patients with childhood absence epilepsy (CAE), and most of them alter the channel kinetics. Based on functional studies and computer simulation, 11 mutations have been shown or predicted to display GOF [109112], while six have been shown or predicted to cause no alteration in channel function [109, 111]. A GOF mutation (R1584P) in Cacna1h has been identified in the “Genetic Absence Epilepsy Rats of Strasbourg” model, and the T-type currents increase with age, mirroring the temporal profile of epilepsy development [106]. In addition, mutations in CACNA1H have been identified in patients with other types of idiopathic generalized epilepsy (IGE), and the changes in channel function are similar to CAE-related mutations [112]. These results suggest that GOF of CACNA1H in humans may increase neuronal firing by decreasing the threshold for rebound burst firing and thus lead to hyperexcitability. LOF has occasionally been identified in IGE- and CAE-related mutations, but the clinical and experimental data are insufficient to ascertain the pathogenicity of these mutants.

CACNA2D2 encodes the α2δ-2 subunit, which co-assembles with the α1 subunit of high-voltage P/Q-type Ca2+ channels (CaV2.1) in the cerebellum and hippocampus. α2δ-2 increases the whole-cell Ca2+ current amplitude and accelerates inactivation. Two homozygous mutations (L1040P and N432fsX) in CACNA2D2 have been identified in patients with EE. Functional analysis of L1040P showed pLOF [113]. Entla mice carrying a nonfunctional α2δ-2 subunit show absence seizures [114]. Deficient α2δ-2 function in humans is expected to slow the inactivation of CaV2.1, thus increasing the action of CaV and leading to epileptogenesis.

CACNB4 encodes the β4 subunit, an auxiliary subunit of CaV2.1 [115]. The β4 subunit may enhance trafficking and expression of the α1 subunit, shift the channel activation to more hyperpolarized potentials, and increase the channel-opening probability [106]. One truncated mutation (R482X) has been identified in a family with juvenile myoclonic epilepsy (JME), and one missense mutation (C104F) has been identified in two families with IGE. A functional study has revealed that C104F exerts an effect similar to the destructive mutation R482X and increases Ca2+ currents [115], probably due to the impaired ability to shift channel activation toward hyperpolarized potentials. Cacnb4 knock-out mice exhibit a “lethargic” phenotype of nonconvulsive seizures, ataxia, and dyskinesias [116]. Specific β4 subunit isoforms have been observed to accumulate in the nucleus, but are suspected to be involved in the pathogenesis of phenotypes other than epilepsy [117]. The involvement of β4 mutants in epileptogenesis is still unclear.

To date, clinical and experimental evidence suggests that Ca2+ channels are implicated in epilepsy. However, the distinct roles of Ca2+ channels in epilepsy phenotypes warrant further clarification.

Chloride Channel Genes

Cl channels (CLCs) are ubiquitously distributed and fulfill diverse functions. The CLC family encompasses nine human proteins, which are divided into two functional subgroups: Cl channels (CLC channels) and chloride-proton (Cl/H+) exchangers (CLC exchangers) [118]. The CLC channels are located in the membranes of excitable and epithelial cells and regulate membrane excitability and the transport of electrolytes, water, and nutrients; the CLC exchangers are mainly expressed intracellularly and may play housekeeping roles [118]. CLCN2 and CLCN4 are reported to be associated with epilepsy (Table 5).

Table 5.

Mutations of epilepsy-associated Cl channel genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
CLCN2 JME Paternal† R235Q pLOF [119]
GTCS Paternal† R644C# Unchanged
Unknown R577Q pLOF
IGE Unknown S719L Not available
IE Paternal† G715E pLOF [120]‡
Unknown G44R, R73H, F82L, S758N, A760V Not available
W570X Destructive
CLCN4 EE de novo L221P, V275M, S534L, G544R§ LOF [64, 122]
A555V, R718W pLOF [122]
D15N Unchanged [122]
Inherited V212G, G731R LOF [122]
G78S, L221V, V536M pLOF
D15fsX18, I626fsX135, intron9+5G>A, 1 intragenic copy number deletion Destructive

EE, epileptic encephalopathy; GTCS, generalized tonic-clonic seizure; IE, idiopathic epilepsy; IGE, idiopathic generalized epilepsy; JME, juvenile myoclonic epilepsy.

† Incomplete penetrance; transmitter not affected.

# Also found in five Indian controls (5/89, 2.8%), but not in Caucasian (386) and North African (263) controls.

‡ Two mutations with false family data were not included. §Two unrelated carriers had different nucleotide substitutions (c.1630G > A and c.1630G > C).

CLCN2 encodes CLC-2, which is an inwardly rectifying channel that opens very slowly upon hyperpolarization. Besides the voltage changes, CLC-2 can be activated by cell swelling. Eleven CLCN2 mutations have been reported to be related to idiopathic epilepsy. Among four mutations with functional studies, three showed pLOF [119, 120]. Clcn2 knock-out mice develop leukodystrophy with vacuoles slowly appearing in the myelin sheaths of central axons [121], but the precise function of CLC-2 in human neurons remains poorly understood.

CLCN4 encodes CLC-4, which is a strongly voltage-dependent 2Cl/H+ exchanger and is expressed widely. The functions of CLC-4 include endosomal acidification and trafficking. CLCN4 mutations have been identified in patients with EE and X-linked intellectual disability. Functional analyses have mainly shown LOF or pLOF [64, 122, 123]. Clcn4 depletion in cultured rodent neurons causes less-branched dendrites and axons [123, 124].

γ-Aminobutyric Acid Type A Receptor (GABAA Receptor) Genes

The GABAA receptors are a group of ligand-gated Cl channels. In the human brain, most GABAA receptors are heteropentamers consisting of two α(1-6), two β(1–3), and one γ(1–3) or δ subunits [125], of which the α1β2γ2 receptor is the most common [126]. Heterodimers (formed by an α and a β subunit) and homopentamers (formed by five β3 subunits) exist in small amounts under physiological conditions. By allowing Cl influx through its pore, the GABAA receptor mediates phasic (synaptic) or tonic (perisynaptic) inhibitory transmission in the brain, leading to hyperpolarization [127]. Epilepsy-associated GABAA receptor genes are listed in Table 6.

Table 6.

Mutations of epilepsy-associated GABAA receptor genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
GABRA1 CAE de novo S326fsX328 Destructive, LOF [129]
JME AD F104C, A322D pLOF [128, 130]
Unknown c.-248+1 G>T Destructive
GEFS+ Paternal† V74I Not available
IGE AD D219N pLOF [134]
K353delins18X Destructive, LOF [134]
Maternal† c.256-8 T>G Destructive
Unknown T20I, L215V, D219N Not available
MAE de novo K306T pLOF [130]
SME de novo S76R, G251S, K306T pLOF [130, 132]
R112Q, L146M, R214H, T292I Not available
EE de novo S76R, R214H pLOF [130]
R112Q, N115D, G251D, P260L, M263I, M263T, V287L, T289P Not available
K401fsX25 Destructive
Unknown T289A Not available
GABRA6 CAE Paternal† R46W pLOF [136, 137]
GABRB1 IS de novo F246S pLOF [139]
EE de novo T287I Not available
GABRB2 GTCS de novo M79T Not available
EE de novo T287P LOF [140]
GABRB3 CAE AD P11S, S15F, G32R pLOF [141]
IS de novo N110D pLOF [131, 139]
LGS de novo D120N, E180G, Y302C LOF
EE de novo L170R, Y182F, Q249K, L256Q, T287I, A305V, A305T Not available
GABRG2 FS AD R177G pLOF [142]
R136X Destructive, pLOF [143, 146]
V462fsX33 Destructive
BECTS de novo R323Q pLOF [144]
AD c.549-3T>G Destructive
CAE with FS AD c.769+2T>G Destructive, pLOF [145]
GEFS+ AD P83S, K328M LOF [147, 150]
R82Q pLOF [150]
M199V Not available
Unknown R304K, R363Q Not available
AD Q390X Destructive, LOF [146, 148]
W429X, Y444fsX51 Destructive, pLOF [146, 151]
E402fsX3 Destructive
GTCS Unknown N79S Unchanged [150]
IGE AD G257R Unchanged# [144]
P59fsX12 Destructive
EE de novo A106T, I107T, P282S, R323W, R323Q, F343L pLOF [83, 126]
Paternal† Q40X Destructive, pLOF [149]
GABRD GEFS+ AD E177A pLOF [156]
R220C Unchanged
JME AD R220H pLOF [156]

BECTS, benign epilepsy of childhood with centrotemporal spikes; CAE, childhood absence epilepsy; EE, epileptic encephalopathy; FS, febrile seizure; GEFS+, generalized epilepsy with febrile seizures plus; GTCS, generalized tonic-clonic seizure; IGE, idiopathic generalized epilepsy; IS, infantile spasms; JME, juvenile myoclonic epilepsy; LGS, Lennox-Gastaut syndrome; MAE, myoclonic-astatic epilepsy; SME, severe myoclonic epilepsy.

† Incomplete penetrance; transmitter not affected.

# Reduced surface expression.

GABRA1 encodes an α1 subunit that is essential for the initiation of GABA-evoked potentials. Mutations in GABRA1 were initially identified in a large family with JME [128], and the phenotypic spectrum was later expanded to other IGEs including CAE [129] and GEFS+ [130], as well as EEs [130133]. Functional studies have demonstrated that all the examined mutants displayed LOF or pLOF [128130, 132, 134] with trafficking impairment causing retention in the endoplasmic reticulum [127]. Heterozygous Gabra1-knock-out mice display spike-wave discharges and absence-like seizures [135]. GABRA6 encodes an α6 subunit. A pLOF mutation in GABRA6 has also been identified in a patient with CAE and disruption of the α6 subunit is associated with δ subunit dysfunction [136, 137].

GABRB1, GABRB2, and GABRB3 encode β1, β2, and β3 subunits, respectively. The β subunits are expressed predominantly in human brain with temporal specificity [138]. The expression of β1 is the most abundant after birth, and then gradually decreases and maintains a lower level in mature neurons. In contrast, the expression of β2 changes more dynamically with development, with the highest expression during childhood and adolescence; the highest β3 level also occurs in early development but remains relatively constant. Two de novo mutations in each GABRB1 and GABRB2 have been identified in patients with EEs. Their functional analyses showed LOF or pLOF consequences. Mutations in GABRB3 have been identified in patients with CAE and EEs and present LOF or pLOF [139141] on current density. Besides decreased current, the CAE-related mutations in GABRB3 also result in hyperglycosylation [141], which might further disturb channel function.

GABRG2 encodes the γ2 subunit that is critical for receptor trafficking, clustering, synaptic maintenance, and current kinetic properties [126]. Twenty-six mutations in GABRG2 have been reported in a broad spectrum of epilepsies. Functional studies have illustrated LOF or pLOF [83, 126, 142151] of these mutations, accompanied commonly by loss or reduction of γ2 subunit protein surface expression [152]. Similarly, the loss of γ2 in heterozygous Gabrg2-knock-out DBA/2J mice results in absence seizures [153], while heterozygous Gabrg2 Q390X-knock-in C57/BL/6J mice display more severe phenotypes, with spontaneous generalized tonic-clonic seizures and probable sudden unexpected death in epilepsy [154]. Recent studies suggest that cellular homeostasis is also disturbed by γ2 mutations [152].

GABRD encodes the δ subunit. The δ-containing GABAA receptors exhibit preferential sensitivity to extracellular GABA concentrations [155], mediating tonic inhibition. Three mutations have been identified in patients with GEFS+ or JME. Functional analysis has shown pLOF in two mutations (E177A and R220H) and no changes in one (R220C) [156].

The mutations in the GABAA receptor genes identified in human epilepsies illustrate important relationships between GABAA receptor function and epileptogenesis. It seems that the LOF or pLOF effects of GABAA receptor genes are common mechanisms underlying the epilepsies caused by GABAA gene mutations. Hints of such a connection also come from knock-out studies of GABAA genes, in which the loss of GABAA function in animals leads to epilepsy-related activities. It is conceivable that impaired function of GABAA receptors would decrease inhibitory effects and lead to impaired coupling of neuronal excitation and inhibition; however, the precise pathogenic mechanism of GABAA receptor gene mutations remains to be clarified.

N-Methyl-D-Aspartate Receptor (NMDAR) Genes

NMDARs are cation channels that are activated by the excitatory neurotransmitter glutamate. NMDARs play roles in excitatory synaptic transmission, plasticity, and excitotoxicity of the CNS [157]. An NMDAR is commonly a bi-heterotetrameric or tri-heterotetrameric channel, consisting of two obligatory GluN1 subunits and two auxiliary GluN2(A-D) or GluN3(A,B) subunits. Mutations of NMDAR subunits are associated with epilepsy and other neurodevelopmental phenotypes (Table 7).

Table 7.

Mutations in epilepsy-associated NMDAR genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
GRIN1 FS de novo R645S Unchanged [158]
S549R Not available
EE de novo G827R LOF [158]
Y647S, G815R pLOF [158]
R844C Unchanged [158]
R417S, D552E†. M641I, N650K, G815V Not available
S560dup Destructive, LOF [159]
Fatal EE Inherited Q556X (homozygous) Destructive, LOF [158]
GRIN2A IFE Unknown A243V GOF [162]
A290V, G295S, R370W, K669N, P699S, E714K, A727T, K772E, N976S, A1276G Not available
W198fsX, Ser547del Destructive
Paternal# P183I, V734L, I814T Not available
L779SfsX5 Destructive
Maternal# I184S, C231Y, G483R, R504W, M705V, D933N Not available
P31SfsX107 Destructive
AD R518H GOF [163]
M1T, P79R, D731N, A716T, I904F Not available
V529TfsX22, R681fsX, Y1387fsX Destructive
de novo F652V GOF [164]
C436R, A548T, I694T, CNV del Not available
FE and/or GE AD T531M GOF [165]
D1251N Not available
Y943fsX, L334X, Q218fsX, c.1007+1G>A, c.1123-2A>G, c.2007+1G>A, 3 chromosomal abnormalities Destructive
de novo M817V Not available
2 chromosomal abnormalities Destructive
Unknown V506A Not available
c.1007+1G>T, 1 chromosomal abnormality Destructive
EE de novo N615K, L812M GOF [166, 167]
Severe unclassified de novo P552R Not available
Unknown L649V Not available
GRIN2B TLE Paternal# E47Q GOF [168]
Unknown E370K Not available
FE de novo R540H GOF [169]
EE de novo N615I, V618Q GOF [169]
C461K Not available
Paternal# c.2011-5_2011-4delTC Destructive
Unclassified epilepsy and ID de novo 1 chromosomal deletion Destructive
Unknown 1 chromosomal deletion Destructive
GRIN2D EE de novo V667I GOF [170]

BECTS, benign epilepsy of childhood with centrotemporal spikes; EE, epileptic encephalopathy; FE, focal epilepsy; GE, generalized epilepsy; ID, intellectual disability; IFE, idiopathic focal epilepsy; IGE, idiopathic generalized epilepsy; TLE, temporal lobe epilepsy.

† Two unrelated carriers had different nucleotide substitutions (c.1656C > A and c.1656 C > G).

# Incomplete penetrance; transmitter not affected.

GRIN1 encodes the ubiquitous GluN1 subunit that binds glycine during activation of NMDARs. GRIN1 mutations have been identified in patients with profound developmental delay and severe intellectual disability [158]. A total of 13 mutations have been associated with epilepsy. Functional analyses of seven mutations suggest prevalent pLOF or LOF in five (Q556X, S560dup, Y647S, G815R, and G827R), although the other two (R645S and R844C) did not show any functional alterations [158, 159]. Homozygous mutation (Q556X) carriers present more severe clinical phenotypes (fatal EE), while homozygous targeted knock-out mice display abnormal glutamate-mediated receptor currents and result in perinatal lethality. These findings indicate that the GluN1 subunit plays an essential role in neurodevelopment. It is therefore possible that dysfunction of the GluN1 subunit may lead to abnormal neurodevelopment as well as epileptogenesis.

GRIN2A, GRIN2B, and GRIN2D, which encode the GluN2A, GluN2B, and GluN2D subunits, respectively, have been associated with epilepsy. The GluN2 subunits have a common binding site with L-glutamate for activation of NMDARs, but show differential spatial and temporal expression patterns throughout the CNS. GRIN2A is mainly expressed in the hippocampus and cerebral cortex at infant and adult stages [157]. In contrast, GRIN2B is expressed in the whole brain during the embryonic period and in the forebrain after adulthood [160]. The expression of GRIN2D is mostly in the limbic system and interneurons in cortico-limbic regions during embryonic stages and is reduced after birth [161].

GRIN2A mutations have been mainly identified in patients with focal epilepsy (FE) and speech disorder, typically in those with Rolandic spikes. Recently, missense mutations have been identified in patients with other phenotypes like EEs or severe unclassified epilepsy. From the published data, there is no distinct relationship between genotype and the severity of epilepsy. Missense mutations of GRIN2A, four (A243V, R518H, T531M, and F652V) from FE patients and two (N615K and L812M) from EE patients, have presented a consequence of GOF [162167]. These GOF mutants display increased activation at low concentrations of agonists and extended durations of channel open and closed states, thus leading to an excessive excitatory drive and epileptogenesis. However, destructive GRIN2A mutations have also been identified in patients within the spectrum of epilepsies, including FE and EE. It remains to be clarified how the destructive mutations impact the function of NMDARs and lead to epilepsy. Considering that GluN2A is not a ubiquitous subunit, it is possible that the destructive GluN2A subunit is substituted by other functionally different subunits such as other GluN2 or GluN3, and thus entails functional changes of NMDARs.

Nine GRIN2B mutations have been identified in patients with epilepsies, including idiopathic focal epilepsy, temporal lobe epilepsy, and EEs. Functional analyses on four missense mutations (E47Q, R540H, N615I, and V618Q) have shown that they all lead to GOF [168, 169]. GRIN2B mutants have a spectrum of genotypes and phenotypes similar to GRIN2A, suggesting a similar mechanism in pathogenicity.

One de novo mutation in GRIN2D (V667I) has been identified in two unrelated patients with EE [170]. Functional analysis has shown a GOF effect with increased current. Transfection of cultured neurons with the V667I mutant causes dendritic swelling and neuronal death, suggestive of excitotoxicity mediated by NMDAR over-activation.

Neuronal Nicotinic Receptor (nAChR) Genes

The nAChRs are a family of pentameric cation channels that are activated by acetylcholine, producing post-synaptic excitation and neurotransmitter release. Sixteen genes encoding nAChRs have been identified in humans. Four nAChR genes, CHRNA2, CHRNA4, CHRNA7, and CHRNB2, have been associated with epilepsy (Table 8).

Table 8.

Mutations in epilepsy-associated nAChR genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
CHRNA2 ADNFLE AD I279N GOF [175]
I297F pLOF [176]
BFIS AD R376W Not available
IGE Unknown T47M Not available
CHRNA4 ADNFLE AD S280F, S284L, T293I GOF [7, 171, 172]
I275F Not available
c.870_872dupGCT Destructive, GOF [171]
NFLE Maternal† R308H Not available
CHRNA7 IGE Unknown 4 gross deletions and 1 gross insertion Destructive
CHRNB2 ADNFLE AD V287L, V287M, L301V GOF [7, 171, 179]
V308A, I312M Not available
IGE AD H35Q, F478L Not available

ADNFLE, autosomal dominant nocturnal frontal lobe epilepsy; BFIS, benign familial infantile seizures; IGE, idiopathic generalized epilepsy; NFLE, sporadic nocturnal frontal lobe epilepsy.

† Incomplete penetrance; transmitter not affected.

CHRNA4 was the first identified epilepsy gene; it encodes the α4 subunit of nAChRs. The α4 subunit is a component of the high-affinity and slowly desensitizing heteropentamer α4β2*, which is one of the two most common nAChRs in the human brain. To date, six CHRNA4 mutations have been identified in nocturnal frontal lobe epilepsy, five in familial cases and one in a sporadic case. The mutations in functional studies commonly display GOF [7, 171, 172]. Several additional variations with undefined pathogenicity have been reported in cases of ADNFLE [173] and other epilepsy phenotypes [174].

CHRNA2 encodes the α2 subunit that composes a heteromeric nAChR with both β2 and β4 subunits. Two CHRNA2 mutations (I279N and I297F) have been reported in two unrelated ADNFLE families. Functional studies have shown GOF of the I279N mutation [175] and pLOF of the I297F mutation [176]. Recently, one mutation (R376W) was identified in a family with benign familial infantile seizures [177].

CHRNB2 encodes the β2 subunit that participates in forming the heteropentamers α4β2* and α2β2β4. The precise function of β2 subunit is unclear. Genetic deletion of the β2 subunit in mice leads to a reduction of dendritic spine density in pyramidal neurons in pre-limbic and infra-limbic areas [178]. Five mutations have been identified in patients with ADNFLE and another two in an IGE cohort. GOF was found in three ADNFLE-related mutants (V287L, V287M, and L301V) [7, 171, 179].

CHRNA7 encodes the α7 subunit that composes a low-affinity and quickly-desensitizing homopentamer. This homopentamer (i.e. (α7)5) is also a common type of nAChR in human thalamus and isocortex. Four chromosome deletions and one chromosome triplication (all including entire CHRNA7) have been identified in patients with IGE. Since CHRNA7 deletion and duplication can be found in affected probands as well as in asymptomatic parents and healthy controls [180], their pathogenicities are uncertain.

Heteromeric nAChRs regulate both excitatory and inhibitory transmission in the frontal cortex, and the delicate balance of excitation and inhibition is crucial for normal neuronal activity. For instance, GOF of nAChRs (by introducing Chrna4-S252F and Chrna4-L264ins in mice) produces abnormally strong GABA release from GABAergic cells and causes synchronization of pyramidal cells [181]. On the other hand, LOF of nAChRs (using dihydro-β-erythroidine to block heteromeric nAChRs in mice) also decreases feedback inhibition of pyramidal cells in the same GABAergic cells and causes hyperexcitability [182]. Due to the complexity of the functional interactions, each nAChR gene or mutant may have a distinct effect on epileptogenesis. Hence, the exact underlying mechanisms of epileptogenesis for nAChRs warrant further clarification.

Hyperpolarization-Activated Cyclic Nucleotide-Gated (HCN) Channel Genes

The HCN channels are a group of cation channels that are dually activated by voltage hyperpolarization and intracellular cAMP, conducting a mixed Na+-K+ inward current[183]. In neurons, the HCN channels are mainly activated by hyperpolarization and contribute to cellular excitability and plasticity. HCN1 and HCN2 are considered to be associated with epilepsy (Table 9).

Table 9.

Mutations in epilepsy-associated HCN channel genes and their functional effects.

Gene Phenotype Inheritance Mutations Functional alteration Ref.
HCN1 EE de novo S272P, R297T LOF [184]
S100F, H279Y, D401H GOF
Unknown G47V Not available
HCN2 FS Maternal† S126L GOF, faster kinetics at higher temperature. [186]
IGE Maternal† R527Q# Unchanged [195]
AR E515K (homozygous) LOF [187]

EE, epileptic encephalopathy; FS, febrile seizure; IGE, idiopathic generalized epilepsy.

† Incomplete penetrance; transmitter not affected.

# The R527Q substitution was absent in the other affected sibling in the same family.

HCN1 encodes the HCN1 channel that is predominantly expressed in dendrites in the neocortex and hippocampus. Six missense mutations have been identified in patients with EEs. Two mutations (S272P and R297T) showed LOF, while another three (S100F, H279Y, and D401H) showed GOF [184]. There is no significant funotype-phenotype relationship. Hcn1-knockout mice display increased excitability and sensitivity to convulsants [185], suggesting that LOF of HCN1 may lead to epileptogenesis.

HCN2 encodes the HCN2 channel that is expressed evenly in the brain and relatively abundant in the thalamus. Three missense mutations have been identified in patients with FS and IGE. Functional studies have shown divergent effects. An FS-related mutation (S126L) shows GOF with faster kinetics at higher temperatures, indicating neuronal hyperexcitability during hyperthermia [186]. An IGE-related mutation (E515K) shows LOF with a reduced threshold of action potential firing and strongly increased excitability and firing frequency in rat primary cortical neurons [187]. The etiology of HCN2 mutants remains unclear.

Discussion

We systematically reviewed 41 ion channel genes that have been associated with epilepsies and analyzed their genotypes, funotypes, and phenotypes. These data are expected to provide clues in evaluating the pathogenic potential of these genes and understanding the underlying mechanism of epilepsy.

We have summarized the funotypes of mutations (Table 10). Genes with mutations featuring LOF include SCN1B, KCNB1, KCNH2, KCNQ2, KCNQ3, CLCN2, CLCN4, GRIN1, GABRA1, GABRB3, and GABRG2, in which both destructive and missense mutations are potentially pathogenic. Genes with mutations featuring GOF include SCN8A, KCNT1, GRIN2A, GRIN2B, CHRNA4, and CHRNB2. Missense mutations are therefore potentially pathogenic in general. Genes like GRIN2A and GRIN2B seem to be exceptional, in that epilepsy-related mutations feature GOF but destructive mutations have also been identified. One possible explanation is that the destructive mutants could be replaced by other subunits of similar function and lead to GOF, since the subunits encoded by GRIN2A and GRIN2B are not ubiquitous. Several genes display diversified funotypes, among which a distinct funotype-phenotype correlation was found in SCN1A. These data suggest that the funotype should be an essential consideration in evaluating the pathogenicity of mutations. A distinct funotype or funotype-phenotype correlation helps in defining the pathogenic potential of a gene.

Table 10.

Summary of pathogenic funotypes of epilepsy-associated ion channel genes.

Main funotype Confirmed by multiple sources To be confirmed
LOF SCN1B
KCNB1, KCNH2, KCNQ2†, KCNQ3
CLCN2†, CLCN4
GRIN1†, GABRA1, GABRB3, GABRG2
KCNC1, KCND3
CACNA1A, CACNA2D2, CACNB2
GABRA6, GABRB1, GABRB2, GABRD
CHRNA7
GOF SCN8A
KCNT1
GRIN2A#, GRIN2B#
CHRNA4, CHRNB2
SCN9A
KCNH5, KCNV2, KCNMA1
GRIN2D
Both LOF and GOF SCN1A (with distinct funotype-phenotype correlation) SCN2A, SCN3A
KCNA2, KCND2
CACNA1H
CHRNA2
HCN1, HCN2

† With a few exceptions.

# Have destructive mutations without examination of whole channel functions.

Most epilepsy-related mutations are heterozygous. Since the human genome is diploid, a heterozygous mutation is generally considered to be potentially pathogenic in a phenotype of dominant inheritance. Therefore, familial co-segregation information is crucial in evaluating their pathogenicities. However, such information is not always available, e.g., the de novo mutations identified in epilepsy phenotypes in recent years. To evaluate the role of these mutations and the related genes in epilepsy, other aspects such as the correlations between genetic impairment and phenotypic severity, genetic knock-out/knock-in outcomes, and pathogenic mechanisms should be taken into consideration. For heterozygous mutations featuring LOF in functional studies, the correlations between impairment and phenotypic severity and the presentations of genetic knock-out animals would help to define whether the variations are pathogenic. The gene expression patterns and underlying pathogenic mechanisms are also essential considerations, especially for genes with mutations featuring GOF and those with limited data or divergent functional changes.

Generally, epilepsies are caused by abnormal synchronized electrical discharges within the brain [60], and ion channels are believed to regulate brain excitability and play critical roles in epilepsies. However, each ion channel plays a unique role in the generation and modulation of neuronal excitability, and thus may be associated with a distinct epilepsy phenotype or phenotype spectrum with a distinct mechanism. Typically, SCN1A has been confirmed to be associated with epilepsy by the distinct phenotype spectrum of FS-related epilepsies; its unequal expression on excitatory pyramidal neurons and inhibitory interneurons is critical in epileptogenesis; and the heterogeneous and relatively local distribution of inhibitory interneurons explains the common partial seizures. Similarly, mutations of KCNQ2 are associated with epilepsy commonly featuring an early onset; it has been confirmed that a 25% reduction in KV7-mediated muscarinic current amplitude is sufficient to result in electrical hyperexcitability. KCNT1 mutations cause a spectrum of focal epilepsies with intellectual disability. Although their precise mechanisms are unclear, all functionally examined mutations in KCNT1 display a definite and strong GOF effect on channel properties. Functional alterations of ion channel genes may directly result in excessive excitability (e.g. GOF in NMDAR mutants) or insufficient inhibitory activity (e.g. LOF in GABAA receptor mutants) in the CNS. In contrast, the pathogenic mechanisms of some genes underlying epileptogenesis, such as SCN1A, KCNC1, and CACNA2D2, are much more complex.

Epilepsy is a complex entity with various phenotypes and phenotype-specific etiologies. Individuals affected by epilepsy may respond differently to anti-epileptic treatments. The updated discovery of epilepsy-associated ion channel genes provides insights into the underlying mechanisms of epileptogenesis and helps to identify novel therapeutic targets for the development of anti-epileptic drugs and individualized treatment.

Conclusions

We systemically reviewed the mutations, funotypes, and phenotype information of 41 epilepsy-associated ion channel genes. These genes are characterized by distinct phenotypes and pathogenic funotypes. The distinct funotypes or funotype-phenotype correlations suggest that funotype should be an essential consideration in evaluating the pathogenicity of mutations. We highlight the complexity of the underlying epileptogenic mechanisms of each ion channel gene. Besides direct contributions to excessive excitability or insufficient inhibition, the phenotypic explanations should be extended to the molecular and ontological mechanisms of the genes.

Acknowledgements

Research work from our laboratory cited in this review was supported by the National Natural Science Foundation of China (81571273, 81571274, 81501124, 81271434, and 81301107), Omics-based precision medicine of epilepsy being entrusted by Key Research Project of the Ministry of Science and Technology of China (2016YFC0904400), the Natural Science Foundation of Guangdong Province, China (2014A030313489), Science and Technology Planning Projects of Guangdong Province, China (2012B031800404 and 2013B051000084), the Department of Education of Guangdong Province, China (2013CXZDA022, 2013KJCX0156, and 2012KJCX009), the Foundation for High-level Talents in Higher Education of Guangdong Province, China (2013-167), Yangcheng Scholar Research Projects of Guangzhou Municipal College (12A016S and 12A017G), and Science and Technology Projects of Guangzhou, Guangdong Province, China (2014J4100069, 201508020011, 201604020161, and 201607010002).

Footnotes

Feng Wei and Li-Min Yan have contributed equally to this work.

References

  • 1.Chang BS, Lowenstein DH. Epilepsy. N Engl J Med. 2003;349:1257–1266. doi: 10.1056/NEJMra022308. [DOI] [PubMed] [Google Scholar]
  • 2.Wang J, Lin ZJ, Liu L, Xu HQ, Shi YW, Yi YH, et al. Epilepsy-associated genes. Seizure. 2017;44:11–20. doi: 10.1016/j.seizure.2016.11.030. [DOI] [PubMed] [Google Scholar]
  • 3.Kohling R, Wolfart J. Potassium channels in epilepsy. Cold Spring Harb Perspect Med 2016, 6, Pii: a022871. [DOI] [PMC free article] [PubMed]
  • 4.Steinlein OK, Mulley JC, Propping P, Wallace RH, Phillips HA, Sutherland GR, et al. A missense mutation in the neuronal nicotinic acetylcholine receptor alpha 4 subunit is associated with autosomal dominant nocturnal frontal lobe epilepsy. Nat Genet. 1995;11:201–203. doi: 10.1038/ng1095-201. [DOI] [PubMed] [Google Scholar]
  • 5.Weiland S, Witzemann V, Villarroel A, Propping P, Steinlein O. An amino acid exchange in the second transmembrane segment of a neuronal nicotinic receptor causes partial epilepsy by altering its desensitization kinetics. FEBS Lett. 1996;398:91–96. doi: 10.1016/S0014-5793(96)01215-X. [DOI] [PubMed] [Google Scholar]
  • 6.Hirose S, Iwata H, Akiyoshi H, Kobayashi K, Ito M, Wada K, et al. A novel mutation of CHRNA4 responsible for autosomal dominant nocturnal frontal lobe epilepsy. Neurology. 1999;53:1749–1753. doi: 10.1212/WNL.53.8.1749. [DOI] [PubMed] [Google Scholar]
  • 7.Steinlein OK, Hoda JC, Bertrand S, Bertrand D. Mutations in familial nocturnal frontal lobe epilepsy might be associated with distinct neurological phenotypes. Seizure. 2012;21:118–123. doi: 10.1016/j.seizure.2011.10.003. [DOI] [PubMed] [Google Scholar]
  • 8.Meng H, Xu HQ, Yu L, Lin GW, He N, Su T, et al. The SCN1A mutation database: updating information and analysis of the relationships among genotype, functional alteration, and phenotype. Hum Mutat. 2015;36:573–580. doi: 10.1002/humu.22782. [DOI] [PubMed] [Google Scholar]
  • 9.Liao WP, Shi YW, Long YS, Zeng Y, Li T, Yu MJ, et al. Partial epilepsy with antecedent febrile seizures and seizure aggravation by antiepileptic drugs: associated with loss of function of Na(v) 1.1. Epilepsia. 2010;51:1669–1678. doi: 10.1111/j.1528-1167.2010.02645.x. [DOI] [PubMed] [Google Scholar]
  • 10.Hartshorne RP, Catterall WA. Purification of the saxitoxin receptor of the sodium channel from rat brain. Proc Natl Acad Sci U S A. 1981;78:4620–4624. doi: 10.1073/pnas.78.7.4620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kruger LC, O’Malley HA, Hull JM, Kleeman A, Patino GA, Isom LL. beta1-C121W is down but not out: epilepsy-associated Scn1b-C121W results in a deleterious gain-of-function. J Neurosci. 2016;36:6213–6224. doi: 10.1523/JNEUROSCI.0405-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Namadurai S, Yereddi NR, Cusdin FS, Huang CL, Chirgadze DY, Jackson AP. A new look at sodium channel beta subunits. Open Biol. 2015;5:140192. doi: 10.1098/rsob.140192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Trimmer JS, Rhodes KJ. Localization of voltage-gated ion channels in mammalian brain. Annu Rev Physiol. 2004;66:477–519. doi: 10.1146/annurev.physiol.66.032102.113328. [DOI] [PubMed] [Google Scholar]
  • 14.Spruston N. Assembling cell ensembles. Cell. 2014;157:1502–1504. doi: 10.1016/j.cell.2014.05.032. [DOI] [PubMed] [Google Scholar]
  • 15.Zuberi SM, Brunklaus A, Birch R, Reavey E, Duncan J, Forbes GH. Genotype-phenotype associations in SCN1A-related epilepsies. Neurology. 2011;76:594–600. doi: 10.1212/WNL.0b013e31820c309b. [DOI] [PubMed] [Google Scholar]
  • 16.Li N, Zhang J, Guo JF, Yan XX, Xia K, Tang BS. Novel mutation of SCN1A in familial generalized epilepsy with febrile seizures plus. Neurosci Lett. 2010;480:211–214. doi: 10.1016/j.neulet.2010.06.040. [DOI] [PubMed] [Google Scholar]
  • 17.Volkers L, Kahlig KM, Verbeek NE, Das JH, van Kempen MJ, Stroink H, et al. Nav 1.1 dysfunction in genetic epilepsy with febrile seizures-plus or Dravet syndrome. Eur J Neurosci. 2011;34:1268–1275. doi: 10.1111/j.1460-9568.2011.07826.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Brackenbury WJ, Isom LL. Na channel beta subunits: overachievers of the ion channel family. Front Pharmacol. 2011;2:53. doi: 10.3389/fphar.2011.00053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Thomas EA, Xu R, Petrou S. Computational analysis of the R85C and R85H epilepsy mutations in Na+ channel beta1 subunits. Neuroscience. 2007;147:1034–1046. doi: 10.1016/j.neuroscience.2007.05.010. [DOI] [PubMed] [Google Scholar]
  • 20.Wallace RH, Wang DW, Singh R, Scheffer IE, George AL, Jr, Phillips HA, et al. Febrile seizures and generalized epilepsy associated with a mutation in the Na+-channel beta1 subunit gene SCN1B. Nat Genet. 1998;19:366–370. doi: 10.1038/448. [DOI] [PubMed] [Google Scholar]
  • 21.Ogiwara I, Nakayama T, Yamagata T, Ohtani H, Mazaki E, Tsuchiya S, et al. A homozygous mutation of voltage-gated sodium channel beta(I) gene SCN1B in a patient with Dravet syndrome. Epilepsia. 2012;53:e200–e203. doi: 10.1111/epi.12040. [DOI] [PubMed] [Google Scholar]
  • 22.Patino GA, Claes LR, Lopez-Santiago LF, Slat EA, Dondeti RS, Chen C, et al. A functional null mutation of SCN1B in a patient with Dravet syndrome. J Neurosci. 2009;29:10764–10778. doi: 10.1523/JNEUROSCI.2475-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Shi X, Yasumoto S, Kurahashi H, Nakagawa E, Fukasawa T, Uchiya S, et al. Clinical spectrum of SCN2A mutations. Brain Dev. 2012;34:541–545. doi: 10.1016/j.braindev.2011.09.016. [DOI] [PubMed] [Google Scholar]
  • 24.Meisler MH, O’Brien JE, Sharkey LM. Sodium channel gene family: epilepsy mutations, gene interactions and modifier effects. J Physiol. 2010;588:1841–1848. doi: 10.1113/jphysiol.2010.188482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Misra SN, Kahlig KM, George AL., Jr Impaired NaV1.2 function and reduced cell surface expression in benign familial neonatal-infantile seizures. Epilepsia. 2008;49:1535–1545. doi: 10.1111/j.1528-1167.2008.01619.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Liao Y, Deprez L, Maljevic S, Pitsch J, Claes L, Hristova D, et al. Molecular correlates of age-dependent seizures in an inherited neonatal-infantile epilepsy. Brain. 2010;133:1403–1414. doi: 10.1093/brain/awq057. [DOI] [PubMed] [Google Scholar]
  • 27.Kamiya K, Kaneda M, Sugawara T, Mazaki E, Okamura N, Montal M, et al. A nonsense mutation of the sodium channel gene SCN2A in a patient with intractable epilepsy and mental decline. J Neurosci. 2004;24:2690–2698. doi: 10.1523/JNEUROSCI.3089-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Ogiwara I, Ito K, Sawaishi Y, Osaka H, Mazaki E, Inoue I, et al. De novo mutations of voltage-gated sodium channel alphaII gene SCN2A in intractable epilepsies. Neurology. 2009;73:1046–1053. doi: 10.1212/WNL.0b013e3181b9cebc. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Planells-Cases R, Caprini M, Zhang J, Rockenstein EM, Rivera RR, Murre C, et al. Neuronal death and perinatal lethality in voltage-gated sodium channel alpha(II)-deficient mice. Biophys J. 2000;78:2878–2891. doi: 10.1016/S0006-3495(00)76829-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Xu R, Thomas EA, Jenkins M, Gazina EV, Chiu C, Heron SE, et al. A childhood epilepsy mutation reveals a role for developmentally regulated splicing of a sodium channel. Mol Cell Neurosci. 2007;35:292–301. doi: 10.1016/j.mcn.2007.03.003. [DOI] [PubMed] [Google Scholar]
  • 31.Beckh S, Noda M, Lubbert H, Numa S. Differential regulation of three sodium channel messenger RNAs in the rat central nervous system during development. EMBO J. 1989;8:3611–3616. doi: 10.1002/j.1460-2075.1989.tb08534.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Escayg A, Goldin AL. Sodium channel SCN1A and epilepsy: mutations and mechanisms. Epilepsia. 2010;51:1650–1658. doi: 10.1111/j.1528-1167.2010.02640.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Holland KD, Kearney JA, Glauser TA, Buck G, Keddache M, Blankston JR, et al. Mutation of sodium channel SCN3A in a patient with cryptogenic pediatric partial epilepsy. Neurosci Lett. 2008;433:65–70. doi: 10.1016/j.neulet.2007.12.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Vanoye CG, Gurnett CA, Holland KD, George AL, Jr, Kearney JA. Novel SCN3A variants associated with focal epilepsy in children. Neurobiol Dis. 2014;62:313–322. doi: 10.1016/j.nbd.2013.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Chen YJ, Shi YW, Xu HQ, Chen ML, Gao MM, Sun WW, et al. Electrophysiological differences between the same pore region mutation in SCN1A and SCN3A. Mol Neurobiol. 2015;51:1263–1270. doi: 10.1007/s12035-014-8802-x. [DOI] [PubMed] [Google Scholar]
  • 36.Wagnon JL, Barker BS, Hounshell JA, Haaxma CA, Shealy A, Moss T, et al. Pathogenic mechanism of recurrent mutations of SCN8A in epileptic encephalopathy. Ann Clin Transl Neurol. 2016;3:114–123. doi: 10.1002/acn3.276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Veeramah KR, O’Brien JE, Meisler MH, Cheng X, Dib-Hajj SD, Waxman SG, et al. De novo pathogenic SCN8A mutation identified by whole-genome sequencing of a family quartet affected by infantile epileptic encephalopathy and SUDEP. Am J Hum Genet. 2012;90:502–510. doi: 10.1016/j.ajhg.2012.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Estacion M, O’Brien JE, Conravey A, Hammer MF, Waxman SG, Dib-Hajj SD, et al. A novel de novo mutation of SCN8A (Nav1.6) with enhanced channel activation in a child with epileptic encephalopathy. Neurobiol Dis. 2014;69:117–123. doi: 10.1016/j.nbd.2014.05.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.de Kovel CG, Meisler MH, Brilstra EH, van Berkestijn FM, van ‘t Slot R, van Lieshout S, et al. Characterization of a de novo SCN8A mutation in a patient with epileptic encephalopathy. Epilepsy Res. 2014;108:1511–1518. doi: 10.1016/j.eplepsyres.2014.08.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Blanchard MG, Willemsen MH, Walker JB, Dib-Hajj SD, Waxman SG, Jongmans MC, et al. De novo gain-of-function and loss-of-function mutations of SCN8A in patients with intellectual disabilities and epilepsy. J Med Genet. 2015;52:330–337. doi: 10.1136/jmedgenet-2014-102813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Martin MS, Tang B, Papale LA, Yu FH, Catterall WA, Escayg A. The voltage-gated sodium channel Scn8a is a genetic modifier of severe myoclonic epilepsy of infancy. Hum Mol Genet. 2007;16:2892–2899. doi: 10.1093/hmg/ddm248. [DOI] [PubMed] [Google Scholar]
  • 42.Blumenfeld H, Lampert A, Klein JP, Mission J, Chen MC, Rivera M, et al. Role of hippocampal sodium channel Nav1.6 in kindling epileptogenesis. Epilepsia. 2009;50:44–55. doi: 10.1111/j.1528-1167.2008.01710.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Peiffer A, Thompson J, Charlier C, Otterud B, Varvil T, Pappas C, et al. A locus for febrile seizures (FEB3) maps to chromosome 2q23-24. Ann Neurol. 1999;46:671–678. doi: 10.1002/1531-8249(199910)46:4<671::AID-ANA20>3.0.CO;2-5. [DOI] [PubMed] [Google Scholar]
  • 44.Estacion M, Han C, Choi JS, Hoeijmakers JG, Lauria G, Drenth JP, et al. Intra- and interfamily phenotypic diversity in pain syndromes associated with a gain-of-function variant of NaV1.7. Mol Pain. 2011;7:92. doi: 10.1186/1744-8069-7-92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Singh NA, Pappas C, Dahle EJ, Claes LR, Pruess TH, De Jonghe P, et al. A role of SCN9A in human epilepsies, as a cause of febrile seizures and as a potential modifier of Dravet syndrome. PLoS Genet. 2009;5:e1000649. doi: 10.1371/journal.pgen.1000649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Anderson PA, Greenberg RM. Phylogeny of ion channels: clues to structure and function. Comp Biochem Physiol B Biochem Mol Biol. 2001;129:17–28. doi: 10.1016/S1096-4959(01)00376-1. [DOI] [PubMed] [Google Scholar]
  • 47.Syrbe S, Hedrich UB, Riesch E, Djemie T, Muller S, Moller RS, et al. De novo loss- or gain-of-function mutations in KCNA2 cause epileptic encephalopathy. Nat Genet. 2015;47:393–399. doi: 10.1038/ng.3239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Pena SD, Coimbra RL. Ataxia and myoclonic epilepsy due to a heterozygous new mutation in KCNA2: proposal for a new channelopathy. Clin Genet. 2015;87:e1–e3. doi: 10.1111/cge.12542. [DOI] [PubMed] [Google Scholar]
  • 49.Brew HM, Gittelman JX, Silverstein RS, Hanks TD, Demas VP, Robinson LC, et al. Seizures and reduced life span in mice lacking the potassium channel subunit Kv1.2, but hypoexcitability and enlarged Kv1 currents in auditory neurons. J Neurophysiol. 2007;98:1501–1525. doi: 10.1152/jn.00640.2006. [DOI] [PubMed] [Google Scholar]
  • 50.Liu PW, Bean BP. Kv2 channel regulation of action potential repolarization and firing patterns in superior cervical ganglion neurons and hippocampal CA1 pyramidal neurons. J Neurosci. 2014;34:4991–5002. doi: 10.1523/JNEUROSCI.1925-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Torkamani A, Bersell K, Jorge BS, Bjork RL, Friedman JR, Bloss CS, et al. De novo KCNB1 mutations in epileptic encephalopathy. Annals of Neurology. 2014;76:529–540. doi: 10.1002/ana.24263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Saitsu H, Akita T, Tohyama J, Goldberg-Stern H, Kobayashi Y, Cohen R, et al. De novo KCNB1 mutations in infantile epilepsy inhibit repetitive neuronal firing. Sci Rep. 2015;5:15199. doi: 10.1038/srep15199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Thiffault I, Speca DJ, Austin DC, Cobb MM, Eum KS, Safina NP, et al. A novel epileptic encephalopathy mutation in KCNB1 disrupts Kv2.1 ion selectivity, expression, and localization. J Gen Physiol. 2015;146:399–410. doi: 10.1085/jgp.201511444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Muona M, Berkovic SF, Dibbens LM, Oliver KL, Maljevic S, Bayly MA, et al. A recurrent de novo mutation in KCNC1 causes progressive myoclonus epilepsy. Nat Genet. 2015;47:39–46. doi: 10.1038/ng.3144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Rudy B, McBain CJ. Kv3 channels: voltage-gated K+ channels designed for high-frequency repetitive firing. Trends Neurosci. 2001;24:517–526. doi: 10.1016/S0166-2236(00)01892-0. [DOI] [PubMed] [Google Scholar]
  • 56.Jerng HH, Pfaffinger PJ, Covarrubias M. Molecular physiology and modulation of somatodendritic A-type potassium channels. Mol Cell Neurosci. 2004;27:343–369. doi: 10.1016/j.mcn.2004.06.011. [DOI] [PubMed] [Google Scholar]
  • 57.Lee H, Lin MC, Kornblum HI, Papazian DM, Nelson SF. Exome sequencing identifies de novo gain of function missense mutation in KCND2 in identical twins with autism and seizures that slows potassium channel inactivation. Hum Mol Genet. 2014;23:3481–3489. doi: 10.1093/hmg/ddu056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Singh B, Ogiwara I, Kaneda M, Tokonami N, Mazaki E, Baba K, et al. A Kv4.2 truncation mutation in a patient with temporal lobe epilepsy. Neurobiol Dis. 2006;24:245–253. doi: 10.1016/j.nbd.2006.07.001. [DOI] [PubMed] [Google Scholar]
  • 59.Smets K, Duarri A, Deconinck T, Ceulemans B, van de Warrenburg BP, Zuchner S, et al. First de novo KCND3 mutation causes severe Kv4.3 channel dysfunction leading to early onset cerebellar ataxia, intellectual disability, oral apraxia and epilepsy. BMC Med Genet. 2015;16:51. doi: 10.1186/s12881-015-0200-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Villa C, Combi R. Potassium channels and human epileptic phenotypes: An updated overview. Front Cell Neurosci. 2016;10:81. doi: 10.3389/fncel.2016.00081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Keller DI, Grenier J, Christe G, Dubouloz F, Osswald S, Brink M, et al. Characterization of novel KCNH2 mutations in type 2 long QT syndrome manifesting as seizures. Can J Cardiol. 2009;25:455–462. doi: 10.1016/S0828-282X(09)70117-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Zamorano-Leon JJ, Yanez R, Jaime G, Rodriguez-Sierra P, Calatrava-Ledrado L, Alvarez-Granada RR, et al. KCNH2 gene mutation: a potential link between epilepsy and long QT-2 syndrome. J Neurogenet. 2012;26:382–386. doi: 10.3109/01677063.2012.674993. [DOI] [PubMed] [Google Scholar]
  • 63.Partemi S, Cestele S, Pezzella M, Campuzano O, Paravidino R, Pascali VL, et al. Loss-of-function KCNH2 mutation in a family with long QT syndrome, epilepsy, and sudden death. Epilepsia. 2013;54:e112–e116. doi: 10.1111/epi.12259. [DOI] [PubMed] [Google Scholar]
  • 64.Veeramah KR, Johnstone L, Karafet TM, Wolf D, Sprissler R, Salogiannis J, et al. Exome sequencing reveals new causal mutations in children with epileptic encephalopathies. Epilepsia. 2013;54:1270–1281. doi: 10.1111/epi.12201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Yang Y, Vasylyev DV, Dib-Hajj F, Veeramah KR, Hammer MF, Dib-Hajj SD, et al. Multistate structural modeling and voltage-clamp analysis of epilepsy/autism mutation Kv10.2-R327H demonstrate the role of this residue in stabilizing the channel closed state. J Neurosci. 2013;33:16586–16593. doi: 10.1523/JNEUROSCI.2307-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Brown DA, Passmore GM. Neural KCNQ (Kv7) channels. Br J Pharmacol. 2009;156:1185–1195. doi: 10.1111/j.1476-5381.2009.00111.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wuttke TV, Penzien J, Fauler M, Seebohm G, Lehmann-Horn F, Lerche H, et al. Neutralization of a negative charge in the S1-S2 region of the KV7.2 (KCNQ2) channel affects voltage-dependent activation in neonatal epilepsy. J Physiol. 2008;586:545–555. doi: 10.1113/jphysiol.2007.143826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Hunter J, Maljevic S, Shankar A, Siegel A, Weissman B, Holt P, et al. Subthreshold changes of voltage-dependent activation of the K(V)7.2 channel in neonatal epilepsy. Neurobiol Dis. 2006;24:194–201. doi: 10.1016/j.nbd.2006.06.011. [DOI] [PubMed] [Google Scholar]
  • 69.Soldovieri MV, Cilio MR, Miceli F, Bellini G, Miraglia del Giudice E, Castaldo P, et al. Atypical gating of M-type potassium channels conferred by mutations in uncharged residues in the S4 region of KCNQ2 causing benign familial neonatal convulsions. J Neurosci. 2007;27:4919–4928. doi: 10.1523/JNEUROSCI.0580-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Soldovieri MV, Boutry-Kryza N, Milh M, Doummar D, Heron B, Bourel E, et al. Novel KCNQ2 and KCNQ3 mutations in a large cohort of families with benign neonatal epilepsy: first evidence for an altered channel regulation by syntaxin-1A. Hum Mutat. 2014;35:356–367. doi: 10.1002/humu.22500. [DOI] [PubMed] [Google Scholar]
  • 71.Ambrosino P, Alaimo A, Bartollino S, Manocchio L, De Maria M, Mosca I, et al. Epilepsy-causing mutations in Kv7.2 C-terminus affect binding and functional modulation by calmodulin. Biochim Biophys Acta. 2015;1852:1856–1866. doi: 10.1016/j.bbadis.2015.06.012. [DOI] [PubMed] [Google Scholar]
  • 72.Volkers L, Rook MB, Das JH, Verbeek NE, Groenewegen WA, van Kempen MJ, et al. Functional analysis of novel KCNQ2 mutations found in patients with Benign Familial Neonatal Convulsions. Neurosci Lett. 2009;462:24–29. doi: 10.1016/j.neulet.2009.06.064. [DOI] [PubMed] [Google Scholar]
  • 73.Miceli F, Vargas E, Bezanilla F, Taglialatela M. Gating currents from Kv7 channels carrying neuronal hyperexcitability mutations in the voltage-sensing domain. Biophys J. 2012;102:1372–1382. doi: 10.1016/j.bpj.2012.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Maljevic S, Naros G, Yalcin O, Blazevic D, Loeffler H, Caglayan H, et al. Temperature and pharmacological rescue of a folding-defective, dominant-negative KV 7.2 mutation associated with neonatal seizures. Hum Mutat. 2011;32:E2283–E2293. doi: 10.1002/humu.21554. [DOI] [PubMed] [Google Scholar]
  • 75.Singh NA, Westenskow P, Charlier C, Pappas C, Leslie J, Dillon J, et al. KCNQ2 and KCNQ3 potassium channel genes in benign familial neonatal convulsions: expansion of the functional and mutation spectrum. Brain. 2003;126:2726–2737. doi: 10.1093/brain/awg286. [DOI] [PubMed] [Google Scholar]
  • 76.Castaldo P, del Giudice EM, Coppola G, Pascotto A, Annunziato L, Taglialatela M. Benign familial neonatal convulsions caused by altered gating of KCNQ2/KCNQ3 potassium channels. J Neurosci 2002, 22: Rc199. [DOI] [PMC free article] [PubMed]
  • 77.Maljevic S, Wuttke TV, Lerche H. Nervous system KV7 disorders: breakdown of a subthreshold brake. J Physiol. 2008;586:1791–1801. doi: 10.1113/jphysiol.2008.150656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Schroeder BC, Kubisch C, Stein V, Jentsch TJ. Moderate loss of function of cyclic-AMP-modulated KCNQ2/KCNQ3 K+ channels causes epilepsy. Nature. 1998;396:687–690. doi: 10.1038/25367. [DOI] [PubMed] [Google Scholar]
  • 79.Richards MC, Heron SE, Spendlove HE, Scheffer IE, Grinton B, Berkovic SF, et al. Novel mutations in the KCNQ2 gene link epilepsy to a dysfunction of the KCNQ2-calmodulin interaction. J Med Genet. 2004;41:e35. doi: 10.1136/jmg.2003.013938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Abidi A, Devaux JJ, Molinari F, Alcaraz G, Michon FX, Sutera-Sardo J, et al. A recurrent KCNQ2 pore mutation causing early onset epileptic encephalopathy has a moderate effect on M current but alters subcellular localization of Kv7 channels. Neurobiol Dis. 2015;80:80–92. doi: 10.1016/j.nbd.2015.04.017. [DOI] [PubMed] [Google Scholar]
  • 81.Orhan G, Bock M, Schepers D, Ilina EI, Reichel SN, Loffler H, et al. Dominant-negative effects of KCNQ2 mutations are associated with epileptic encephalopathy. Ann Neurol. 2014;75:382–394. doi: 10.1002/ana.24080. [DOI] [PubMed] [Google Scholar]
  • 82.Weckhuysen S, Mandelstam S, Suls A, Audenaert D, Deconinck T, Claes LR, et al. KCNQ2 encephalopathy: emerging phenotype of a neonatal epileptic encephalopathy. Ann Neurol. 2012;71:15–25. doi: 10.1002/ana.22644. [DOI] [PubMed] [Google Scholar]
  • 83.Carvill GL, Heavin SB, Yendle SC, McMahon JM, O’Roak BJ, Cook J, et al. Targeted resequencing in epileptic encephalopathies identifies de novo mutations in CHD2 and SYNGAP1. Nat Genet. 2013;45:825–830. doi: 10.1038/ng.2646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Zhang Y, Kong W, Gao Y, Liu X, Gao K, Xie H, et al. Gene mutation analysis in 253 chinese children with unexplained epilepsy and intellectual/developmental disabilities. PLoS One. 2015;10:e0141782. doi: 10.1371/journal.pone.0141782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Miceli F, Striano P, Soldovieri MV, Fontana A, Nardello R, Robbiano A, et al. A novel KCNQ3 mutation in familial epilepsy with focal seizures and intellectual disability. Epilepsia. 2015;56:e15–e20. doi: 10.1111/epi.12887. [DOI] [PubMed] [Google Scholar]
  • 86.Miceli F, Soldovieri MV, Ambrosino P, De Maria M, Migliore M, Migliore R, et al. Early-onset epileptic encephalopathy caused by gain-of-function mutations in the voltage sensor of Kv7.2 and Kv7.3 potassium channel subunits. J Neurosci. 2015;35:3782–3793. doi: 10.1523/JNEUROSCI.4423-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Peters HC, Hu H, Pongs O, Storm JF, Isbrandt D. Conditional transgenic suppression of M channels in mouse brain reveals functions in neuronal excitability, resonance and behavior. Nat Neurosci. 2005;8:51–60. doi: 10.1038/nn1375. [DOI] [PubMed] [Google Scholar]
  • 88.Jentsch TJ. Neuronal KCNQ potassium channels: physiology and role in disease. Nat Rev Neurosci. 2000;1:21–30. doi: 10.1038/35036198. [DOI] [PubMed] [Google Scholar]
  • 89.Miceli F, Soldovieri MV, Ambrosino P, Barrese V, Migliore M, Cilio MR, et al. Genotype-phenotype correlations in neonatal epilepsies caused by mutations in the voltage sensor of K(v)7.2 potassium channel subunits. Proc Natl Acad Sci U S A. 2013;110:4386–4391. doi: 10.1073/pnas.1216867110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Sugiura Y, Nakatsu F, Hiroyasu K, Ishii A, Hirose S, Okada M, et al. Lack of potassium current in W309R mutant KCNQ3 channel causing benign familial neonatal convulsions (BFNC) Epilepsy Res. 2009;84:82–85. doi: 10.1016/j.eplepsyres.2008.12.003. [DOI] [PubMed] [Google Scholar]
  • 91.Bassi MT, Balottin U, Panzeri C, Piccinelli P, Castaldo P, Barrese V, et al. Functional analysis of novel KCNQ2 and KCNQ3 gene variants found in a large pedigree with benign familial neonatal convulsions (BFNC) Neurogenetics. 2005;6:185–193. doi: 10.1007/s10048-005-0012-2. [DOI] [PubMed] [Google Scholar]
  • 92.Neubauer BA, Waldegger S, Heinzinger J, Hahn A, Kurlemann G, Fiedler B, et al. KCNQ2 and KCNQ3 mutations contribute to different idiopathic epilepsy syndromes. Neurology. 2008;71:177–183. doi: 10.1212/01.wnl.0000317090.92185.ec. [DOI] [PubMed] [Google Scholar]
  • 93.Czirjak G, Toth ZE, Enyedi P. Characterization of the heteromeric potassium channel formed by kv2.1 and the retinal subunit kv8.2 in Xenopus oocytes. J Neurophysiol. 2007;98:1213–1222. doi: 10.1152/jn.00493.2007. [DOI] [PubMed] [Google Scholar]
  • 94.Jorge BS, Campbell CM, Miller AR, Rutter ED, Gurnett CA, Vanoye CG, et al. Voltage-gated potassium channel KCNV2 (Kv8.2) contributes to epilepsy susceptibility. Proc Natl Acad Sci U S A. 2011;108:5443–5448. doi: 10.1073/pnas.1017539108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Gu N, Vervaeke K, Storm JF. BK potassium channels facilitate high-frequency firing and cause early spike frequency adaptation in rat CA1 hippocampal pyramidal cells. J Physiol. 2007;580:859–882. doi: 10.1113/jphysiol.2006.126367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Du W, Bautista JF, Yang H, Diez-Sampedro A, You SA, Wang L, et al. Calcium-sensitive potassium channelopathy in human epilepsy and paroxysmal movement disorder. Nat Genet. 2005;37:733–738. doi: 10.1038/ng1585. [DOI] [PubMed] [Google Scholar]
  • 97.Heron SE, Smith KR, Bahlo M, Nobili L, Kahana E, Licchetta L, et al. Missense mutations in the sodium-gated potassium channel gene KCNT1 cause severe autosomal dominant nocturnal frontal lobe epilepsy. Nat Genet. 2012;44:1188–1190. doi: 10.1038/ng.2440. [DOI] [PubMed] [Google Scholar]
  • 98.Budelli G, Hage TA, Wei A, Rojas P, Jong YJ, O’Malley K, et al. Na+-activated K+ channels express a large delayed outward current in neurons during normal physiology. Nat Neurosci. 2009;12:745–750. doi: 10.1038/nn.2313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Barcia G, Fleming MR, Deligniere A, Gazula VR, Brown MR, Langouet M, et al. De novo gain-of-function KCNT1 channel mutations cause malignant migrating partial seizures of infancy. Nat Genet. 2012;44:1255–1259. doi: 10.1038/ng.2441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Kim GE, Kronengold J, Barcia G, Quraishi IH, Martin HC, Blair E, et al. Human slack potassium channel mutations increase positive cooperativity between individual channels. Cell Rep. 2014;9:1661–1672. doi: 10.1016/j.celrep.2014.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Ishii A, Shioda M, Okumura A, Kidokoro H, Sakauchi M, Shimada S, et al. A recurrent KCNT1 mutation in two sporadic cases with malignant migrating partial seizures in infancy. Gene. 2013;531:467–471. doi: 10.1016/j.gene.2013.08.096. [DOI] [PubMed] [Google Scholar]
  • 102.Allen NM, Conroy J, Shahwan A, Lynch B, Correa RG, Pena SD, et al. Unexplained early onset epileptic encephalopathy: Exome screening and phenotype expansion. Epilepsia. 2016;57:e12–e17. doi: 10.1111/epi.13250. [DOI] [PubMed] [Google Scholar]
  • 103.Rizzo F, Ambrosino P, Guacci A, Chetta M, Marchese G, Rocco T, et al. Characterization of two de novoKCNT1 mutations in children with malignant migrating partial seizures in infancy. Mol Cell Neurosci. 2016;72:54–63. doi: 10.1016/j.mcn.2016.01.004. [DOI] [PubMed] [Google Scholar]
  • 104.Mikati MA, Jiang YH, Carboni M, Shashi V, Petrovski S, Spillmann R, et al. Quinidine in the treatment of KCNT1-positive epilepsies. Ann Neurol. 2015;78:995–999. doi: 10.1002/ana.24520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Milligan CJ, Li M, Gazina EV, Heron SE, Nair U, Trager C, et al. KCNT1 gain of function in 2 epilepsy phenotypes is reversed by quinidine. Ann Neurol. 2014;75:581–590. doi: 10.1002/ana.24128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Rajakulendran S, Hanna MG. The role of calcium channels in epilepsy. Cold Spring Harb Perspect Med. 2016;6:a022723. doi: 10.1101/cshperspect.a022723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Klassen T, Davis C, Goldman A, Burgess D, Chen T, Wheeler D, et al. Exome sequencing of ion channel genes reveals complex profiles confounding personal risk assessment in epilepsy. Cell. 2011;145:1036–1048. doi: 10.1016/j.cell.2011.05.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Epi KCEaekce, Epi KC. De Novo Mutations in SLC1A2 and CACNA1A are important causes of epileptic encephalopathies. Am J Hum Genet 2016, 99: 287–298. [DOI] [PMC free article] [PubMed]
  • 109.Khosravani H, Altier C, Simms B, Hamming KS, Snutch TP, Mezeyova J, et al. Gating effects of mutations in the Cav3.2 T-type calcium channel associated with childhood absence epilepsy. Journal of Biological Chemistry. 2004;279:9681–9684. doi: 10.1074/jbc.C400006200. [DOI] [PubMed] [Google Scholar]
  • 110.Chen Y, Lu J, Pan H, Zhang Y, Wu H, Xu K, et al. Association between genetic variation of CACNA1H and childhood absence epilepsy. Ann Neurol. 2003;54:239–243. doi: 10.1002/ana.10607. [DOI] [PubMed] [Google Scholar]
  • 111.Vitko I, Chen Y, Arias JM, Shen Y, Wu XR, Perez-Reyes E. Functional characterization and neuronal modeling of the effects of childhood absence epilepsy variants of CACNA1H, a T-type calcium channel. J Neurosci. 2005;25:4844–4855. doi: 10.1523/JNEUROSCI.0847-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Heron SE, Khosravani H, Varela D, Bladen C, Williams TC, Newman MR, et al. Extended spectrum of idiopathic generalized epilepsies associated with CACNA1H functional variants. Ann Neurol. 2007;62:560–568. doi: 10.1002/ana.21169. [DOI] [PubMed] [Google Scholar]
  • 113.Edvardson S, Oz S, Abulhijaa FA, Taher FB, Shaag A, Zenvirt S, et al. Early infantile epileptic encephalopathy associated with a high voltage gated calcium channelopathy. J Med Genet. 2013;50:118–123. doi: 10.1136/jmedgenet-2012-101223. [DOI] [PubMed] [Google Scholar]
  • 114.Brill J, Klocke R, Paul D, Boison D, Gouder N, Klugbauer N, et al. entla, a novel epileptic and ataxic Cacna2d2 mutant of the mouse. J Biol Chem. 2004;279:7322–7330. doi: 10.1074/jbc.M308778200. [DOI] [PubMed] [Google Scholar]
  • 115.Escayg A, De Waard M, Lee DD, Bichet D, Wolf P, Mayer T, et al. Coding and noncoding variation of the human calcium-channel beta4-subunit gene CACNB4 in patients with idiopathic generalized epilepsy and episodic ataxia. Am J Hum Genet. 2000;66:1531–1539. doi: 10.1086/302909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Burgess DL, Jones JM, Meisler MH, Noebels JL. Mutation of the Ca2+ channel beta subunit gene Cchb4 is associated with ataxia and seizures in the lethargic (lh) mouse. Cell. 1997;88:385–392. doi: 10.1016/S0092-8674(00)81877-2. [DOI] [PubMed] [Google Scholar]
  • 117.Ronjat M, Kiyonaka S, Barbado M, De Waard M, Mori Y. Nuclear life of the voltage-gated Cacnb4 subunit and its role in gene transcription regulation. Channels (Austin) 2013;7:119–125. doi: 10.4161/chan.23895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Stolting G, Fischer M, Fahlke C. CLC channel function and dysfunction in health and disease. Front Physiol. 2014;5:378. doi: 10.3389/fphys.2014.00378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Saint-Martin C, Gauvain G, Teodorescu G, Gourfinkel-An I, Fedirko E, Weber YG, et al. Two novel CLCN2 mutations accelerating chloride channel deactivation are associated with idiopathic generalized epilepsy. Human Mutation. 2009;30:397–405. doi: 10.1002/humu.20876. [DOI] [PubMed] [Google Scholar]
  • 120.Haug K, Warnstedt M, Alekov AK, Sander T, Ramirez A, Poser B, et al. Retraction: Mutations in CLCN2 encoding a voltage-gated chloride channel are associated with idiopathic generalized epilepsies. Nat Genet. 1043;2009:41. doi: 10.1038/ng0909-1043. [DOI] [PubMed] [Google Scholar]
  • 121.Blanz J, Schweizer M, Auberson M, Maier H, Muenscher A, Hubner CA, et al. Leukoencephalopathy upon disruption of the chloride channel ClC-2. J Neurosci. 2007;27:6581–6589. doi: 10.1523/JNEUROSCI.0338-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Palmer EE, Stuhlmann T, Weinert S, Haan E, Van Esch H, Holvoet M, et al. De novo and inherited mutations in the X-linked gene CLCN4 are associated with syndromic intellectual disability and behavior and seizure disorders in males and females. Mol Psychiatry. 2016 doi: 10.1038/mp.2016.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Hu H, Haas SA, Chelly J, Van Esch H, Raynaud M, de Brouwer AP, et al. X-exome sequencing of 405 unresolved families identifies seven novel intellectual disability genes. Molecular Psychiatry. 2016;21:133–148. doi: 10.1038/mp.2014.193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Hur J, Jeong HJ, Park J, Jeon S. Chloride channel 4 is required for nerve growth factor-induced TrkA signaling and neurite outgrowth in PC12 cells and cortical neurons. Neuroscience. 2013;253:389–397. doi: 10.1016/j.neuroscience.2013.09.003. [DOI] [PubMed] [Google Scholar]
  • 125.Baumann SW, Baur R, Sigel E. Forced subunit assembly in alpha1beta2gamma2 GABAA receptors. Insight into the absolute arrangement. J Biol Chem. 2002;277:46020–46025. doi: 10.1074/jbc.M207663200. [DOI] [PubMed] [Google Scholar]
  • 126.Shen D, Hernandez CC, Shen W, Hu N, Poduri A, Shiedley B, et al. De novo GABRG2 mutations associated with epileptic encephalopathies. Brain. 2016 doi: 10.1093/brain/aww272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Hirose S. Mutant GABA(A) receptor subunits in genetic (idiopathic) epilepsy. Prog Brain Res. 2014;213:55–85. doi: 10.1016/B978-0-444-63326-2.00003-X. [DOI] [PubMed] [Google Scholar]
  • 128.Cossette P, Liu L, Brisebois K, Dong H, Lortie A, Vanasse M, et al. Mutation of GABRA1 in an autosomal dominant form of juvenile myoclonic epilepsy. Nat Genet. 2002;31:184–189. doi: 10.1038/ng885. [DOI] [PubMed] [Google Scholar]
  • 129.Maljevic S, Krampfl K, Cobilanschi J, Tilgen N, Beyer S, Weber YG, et al. A mutation in the GABA(A) receptor alpha(1)-subunit is associated with absence epilepsy. Ann Neurol. 2006;59:983–987. doi: 10.1002/ana.20874. [DOI] [PubMed] [Google Scholar]
  • 130.Johannesen K, Marini C, Pfeffer S, Moller RS, Dorn T, Niturad C, et al. Phenotypic spectrum of GABRA1: From generalized epilepsies to severe epileptic encephalopathies. Neurology. 2016;87:1140–1151. doi: 10.1212/WNL.0000000000003087. [DOI] [PubMed] [Google Scholar]
  • 131.Allen AS, Berkovic SF, Cossette P, Delanty N, Dlugos D, Eichler EE, et al. De novo mutations in epileptic encephalopathies. Nature. 2013;501:217–221. doi: 10.1038/nature12439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Carvill GL, Weckhuysen S, McMahon JM, Hartmann C, Moller RS, Hjalgrim H, et al. GABRA1 and STXBP1: novel genetic causes of Dravet syndrome. Neurology. 2014;82:1245–1253. doi: 10.1212/WNL.0000000000000291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Kodera H, Ohba C, Kato M, Maeda T, Araki K, Tajima D, et al. De novo GABRA1 mutations in Ohtahara and West syndromes. Epilepsia. 2016;57:566–573. doi: 10.1111/epi.13344. [DOI] [PubMed] [Google Scholar]
  • 134.Lachance-Touchette P, Brown P, Meloche C, Kinirons P, Lapointe L, Lacasse H, et al. Novel alpha1 and gamma2 GABAA receptor subunit mutations in families with idiopathic generalized epilepsy. Eur J Neurosci. 2011;34:237–249. doi: 10.1111/j.1460-9568.2011.07767.x. [DOI] [PubMed] [Google Scholar]
  • 135.Arain FM, Boyd KL, Gallagher MJ. Decreased viability and absence-like epilepsy in mice lacking or deficient in the GABAA receptor alpha1 subunit. Epilepsia. 2012;53:e161–e165. doi: 10.1111/j.1528-1167.2012.03596.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Dibbens LM, Harkin LA, Richards M, Hodgson BL, Clarke AL, Petrou S, et al. The role of neuronal GABA(A) receptor subunit mutations in idiopathic generalized epilepsies. Neurosci Lett. 2009;453:162–165. doi: 10.1016/j.neulet.2009.02.038. [DOI] [PubMed] [Google Scholar]
  • 137.Hernandez CC, Gurba KN, Hu N, Macdonald RL. The GABRA6 mutation, R46W, associated with childhood absence epilepsy, alters 6beta22 and 6beta2 GABA(A) receptor channel gating and expression. J Physiol. 2011;589:5857–5878. doi: 10.1113/jphysiol.2011.218883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Fillman SG, Duncan CE, Webster MJ, Elashoff M, Weickert CS. Developmental co-regulation of the beta and gamma GABAA receptor subunits with distinct alpha subunits in the human dorsolateral prefrontal cortex. Int J Dev Neurosci. 2010;28:513–519. doi: 10.1016/j.ijdevneu.2010.05.004. [DOI] [PubMed] [Google Scholar]
  • 139.Janve VS, Hernandez CC, Verdier KM, Hu N, Macdonald RL. Epileptic encephalopathy de novo GABRB mutations impair GABAA receptor function. Ann Neurol. 2016 doi: 10.1002/ana.24631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Ishii A, Kang JQ, Schornak CC, Hernandez CC, Shen W, Watkins JC, et al. A de novo missense mutation of GABRB2 causes early myoclonic encephalopathy. J Med Genet. 2016 doi: 10.1136/jmedgenet-2016-104083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Tanaka M, Olsen RW, Medina MT, Schwartz E, Alonso ME, Duron RM, et al. Hyperglycosylation and reduced GABA currents of mutated GABRB3 polypeptide in remitting childhood absence epilepsy. Am J Hum Genet. 2008;82:1249–1261. doi: 10.1016/j.ajhg.2008.04.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Audenaert D, Schwartz E, Claeys KG, Claes L, Deprez L, Suls A, et al. A novel GABRG2 mutation associated with febrile seizures. Neurology. 2006;67:687–690. doi: 10.1212/01.wnl.0000230145.73496.a2. [DOI] [PubMed] [Google Scholar]
  • 143.Johnston AJ, Kang JQ, Shen W, Pickrell WO, Cushion TD, Davies JS, et al. A novel GABRG2 mutation, p. R136*, in a family with GEFS+ and extended phenotypes. Neurobiol Dis. 2014;64:131–141. doi: 10.1016/j.nbd.2013.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Reinthaler EM, Dejanovic B, Lal D, Semtner M, Merkler Y, Reinhold A, et al. Rare variants in gamma-aminobutyric acid type A receptor genes in rolandic epilepsy and related syndromes. Ann Neurol. 2015;77:972–986. doi: 10.1002/ana.24395. [DOI] [PubMed] [Google Scholar]
  • 145.Tian M, Macdonald RL. The intronic GABRG2 mutation, IVS6+2T->G, associated with childhood absence epilepsy altered subunit mRNA intron splicing, activated nonsense-mediated decay, and produced a stable truncated gamma2 subunit. J Neurosci. 2012;32:5937–5952. doi: 10.1523/JNEUROSCI.5332-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Wang J, Shen D, Xia G, Shen W, Macdonald RL, Xu D, et al. Differential protein structural disturbances and suppression of assembly partners produced by nonsense GABRG2 epilepsy mutations: implications for disease phenotypic heterogeneity. Sci Rep. 2016;6:35294. doi: 10.1038/srep35294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Baulac S, Huberfeld G, Gourfinkel-An I, Mitropoulou G, Beranger A, Prud’homme JF, et al. First genetic evidence of GABA(A) receptor dysfunction in epilepsy: a mutation in the gamma2-subunit gene. Nat Genet. 2001;28:46–48. doi: 10.1038/ng0501-46. [DOI] [PubMed] [Google Scholar]
  • 148.Harkin LA, Bowser DN, Dibbens LM, Singh R, Phillips F, Wallace RH, et al. Truncation of the GABA(A)-Receptor γ2 Subunit in a Family with Generalized Epilepsy with Febrile Seizures Plus. Am J Hum Genet. 2002;70:530–536. doi: 10.1086/338710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Hirose S. A new paradigm of channelopathy in epilepsy syndromes: intracellular trafficking abnormality of channel molecules. Epilepsy Res. 2006;70(Suppl 1):S206–S217. doi: 10.1016/j.eplepsyres.2005.12.007. [DOI] [PubMed] [Google Scholar]
  • 150.Huang X, Hernandez CC, Hu N, Macdonald RL. Three epilepsy-associated GABRG2 missense mutations at the gamma+/beta- interface disrupt GABAA receptor assembly and trafficking by similar mechanisms but to different extents. Neurobiol Dis. 2014;68:167–179. doi: 10.1016/j.nbd.2014.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Sun H, Zhang Y, Liang J, Liu X, Ma X, Wu H, et al. SCN1A, SCN1B, and GABRG2 gene mutation analysis in Chinese families with generalized epilepsy with febrile seizures plus. J Hum Genet. 2008;53:769–774. doi: 10.1007/s10038-008-0306-y. [DOI] [PubMed] [Google Scholar]
  • 152.Kang JQ, Macdonald RL. Molecular pathogenic basis for GABRG2 mutations associated with a spectrum of epilepsy syndromes, from generalized absence epilepsy to dravet syndrome. JAMA Neurol. 2016;73:1009–1016. doi: 10.1001/jamaneurol.2016.0449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Reid CA, Kim T, Phillips AM, Low J, Berkovic SF, Luscher B, et al. Multiple molecular mechanisms for a single GABAA mutation in epilepsy. Neurology. 2013;80:1003–1008. doi: 10.1212/WNL.0b013e3182872867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kang JQ, Shen W, Zhou C, Xu D, Macdonald RL. The human epilepsy mutation GABRG2(Q390X) causes chronic subunit accumulation and neurodegeneration. Nat Neurosci. 2015;18:988–996. doi: 10.1038/nn.4024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Carver CM, Reddy DS. Neurosteroid structure-activity relationships for functional activation of extrasynaptic deltaGABA(A) receptors. J Pharmacol Exp Ther. 2016;357:188–204. doi: 10.1124/jpet.115.229302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Dibbens LM, Feng HJ, Richards MC, Harkin LA, Hodgson BL, Scott D, et al. GABRD encoding a protein for extra- or peri-synaptic GABAA receptors is a susceptibility locus for generalized epilepsies. Hum Mol Genet. 2004;13:1315–1319. doi: 10.1093/hmg/ddh146. [DOI] [PubMed] [Google Scholar]
  • 157.Cull-Candy S, Brickley S, Farrant M. NMDA receptor subunits: diversity, development and disease. Curr Opin Neurobiol. 2001;11:327–335. doi: 10.1016/S0959-4388(00)00215-4. [DOI] [PubMed] [Google Scholar]
  • 158.Lemke JR, Geider K, Helbig KL, Heyne HO, Schutz H, Hentschel J, et al. Delineating the GRIN1 phenotypic spectrum: A distinct genetic NMDA receptor encephalopathy. Neurology. 2016;86:2171–2178. doi: 10.1212/WNL.0000000000002740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Hamdan FF, Gauthier J, Araki Y, Lin DT, Yoshizawa Y, Higashi K, et al. Excess of de novo deleterious mutations in genes associated with glutamatergic systems in nonsyndromic intellectual disability. Am J Hum Genet. 2011;88:306–316. doi: 10.1016/j.ajhg.2011.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Monyer H, Burnashev N, Laurie DJ, Sakmann B, Seeburg PH. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron. 1994;12:529–540. doi: 10.1016/0896-6273(94)90210-0. [DOI] [PubMed] [Google Scholar]
  • 161.Suryavanshi PS, Ugale RR, Yilmazer-Hanke D, Stairs DJ, Dravid SM. GluN2C/GluN2D subunit-selective NMDA receptor potentiator CIQ reverses MK-801-induced impairment in prepulse inhibition and working memory in Y-maze test in mice. Br J Pharmacol. 2014;171:799–809. doi: 10.1111/bph.12518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Lemke JR, Lal D, Reinthaler EM, Steiner I, Nothnagel M, Alber M, et al. Mutations in GRIN2A cause idiopathic focal epilepsy with rolandic spikes. Nat Genet. 2013;45:1067–1072. doi: 10.1038/ng.2728. [DOI] [PubMed] [Google Scholar]
  • 163.Conroy J, McGettigan PA, McCreary D, Shah N, Collins K, Parry-Fielder B, et al. Towards the identification of a genetic basis for Landau-Kleffner syndrome. Epilepsia. 2014;55:858–865. doi: 10.1111/epi.12645. [DOI] [PubMed] [Google Scholar]
  • 164.Lesca G, Rudolf G, Bruneau N, Lozovaya N, Labalme A, Boutry-Kryza N, et al. GRIN2A mutations in acquired epileptic aphasia and related childhood focal epilepsies and encephalopathies with speech and language dysfunction. Nat Genet. 2013;45:1061–1066. doi: 10.1038/ng.2726. [DOI] [PubMed] [Google Scholar]
  • 165.Carvill GL, Regan BM, Yendle SC, O’Roak BJ, Lozovaya N, Bruneau N, et al. GRIN2A mutations cause epilepsy-aphasia spectrum disorders. Nat Genet. 2013;45:1073–1076. doi: 10.1038/ng.2727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Endele S, Rosenberger G, Geider K, Popp B, Tamer C, Stefanova I, et al. Mutations in GRIN2A and GRIN2B encoding regulatory subunits of NMDA receptors cause variable neurodevelopmental phenotypes. Nat Genet. 2010;42:1021–1026. doi: 10.1038/ng.677. [DOI] [PubMed] [Google Scholar]
  • 167.Yuan H, Hansen KB, Zhang J, Pierson TM, Markello TC, Fajardo KV, et al. Functional analysis of a de novo GRIN2A missense mutation associated with early-onset epileptic encephalopathy. Nat Commun. 2014;5:3251. doi: 10.1038/ncomms4251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Hildebrand MS, Myers CT, Carvill GL, Regan BM, Damiano JA, Mullen SA, et al. A targeted resequencing gene panel for focal epilepsy. Neurology. 2016;86:1605–1612. doi: 10.1212/WNL.0000000000002608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Lemke JR, Hendrickx R, Geider K, Laube B, Schwake M, Harvey RJ, et al. GRIN2B mutations in West syndrome and intellectual disability with focal epilepsy. Ann Neurol. 2014;75:147–154. doi: 10.1002/ana.24073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Li D, Yuan H, Ortiz-Gonzalez XR, Marsh ED, Tian L, McCormick EM, et al. GRIN2D recurrent de novo dominant mutation causes a severe epileptic encephalopathy treatable with NMDA receptor channel blockers. Am J Hum Genet. 2016;99:802–816. doi: 10.1016/j.ajhg.2016.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Son CD, Moss FJ, Cohen BN, Lester HA. Nicotine normalizes intracellular subunit stoichiometry of nicotinic receptors carrying mutations linked to autosomal dominant nocturnal frontal lobe epilepsy. Mol Pharmacol. 2009;75:1137–1148. doi: 10.1124/mol.108.054494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Leniger T, Kananura C, Hufnagel A, Bertrand S, Bertrand D, Steinlein OK. A new Chrna4 mutation with low penetrance in nocturnal frontal lobe epilepsy. Epilepsia. 2003;44:981–985. doi: 10.1046/j.1528-1157.2003.61102.x. [DOI] [PubMed] [Google Scholar]
  • 173.Chen Z, Wang L, Wang C, Chen Q, Zhai Q, Guo Y, et al. Mutational analysis of CHRNB2, CHRNA2 and CHRNA4 genes in Chinese population with autosomal dominant nocturnal frontal lobe epilepsy. Int J Clin Exp Med. 2015;8:9063–9070. [PMC free article] [PubMed] [Google Scholar]
  • 174.Rozycka A, Steinborn B, Trzeciak WH. The 1674+11C>T polymorphism of CHRNA4 is associated with juvenile myoclonic epilepsy. Seizure. 2009;18:601–603. doi: 10.1016/j.seizure.2009.06.007. [DOI] [PubMed] [Google Scholar]
  • 175.Aridon P, Marini C, Di Resta C, Brilli E, De Fusco M, Politi F, et al. Increased sensitivity of the neuronal nicotinic receptor alpha 2 subunit causes familial epilepsy with nocturnal wandering and ictal fear. Am J Hum Genet. 2006;79:342–350. doi: 10.1086/506459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Conti V, Aracri P, Chiti L, Brusco S, Mari F, Marini C, et al. Nocturnal frontal lobe epilepsy with paroxysmal arousals due to CHRNA2 loss of function. Neurology. 2015;84:1520–1528. doi: 10.1212/WNL.0000000000001471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Trivisano M, Terracciano A, Milano T, Cappelletti S, Pietrafusa N, Bertini ES, et al. Mutation of CHRNA2 in a family with benign familial infantile seizures: Potential role of nicotinic acetylcholine receptor in various phenotypes of epilepsy. Epilepsia. 2015;56:e53–e57. doi: 10.1111/epi.12967. [DOI] [PubMed] [Google Scholar]
  • 178.Ballesteros-Yanez I, Benavides-Piccione R, Bourgeois JP, Changeux JP, DeFelipe J. Alterations of cortical pyramidal neurons in mice lacking high-affinity nicotinic receptors. Proc Natl Acad Sci U S A. 2010;107:11567–11572. doi: 10.1073/pnas.1006269107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Gullo F, Manfredi I, Lecchi M, Casari G, Wanke E, Becchetti A. Multi-electrode array study of neuronal cultures expressing nicotinic beta2-V287L subunits, linked to autosomal dominant nocturnal frontal lobe epilepsy. An in vitro model of spontaneous epilepsy. Front Neural Circuits 2014, 8: 87. [DOI] [PMC free article] [PubMed]
  • 180.Gillentine MA, Berry LN, Goin-Kochel RP, Ali MA, Ge J, Guffey D, et al. The cognitive and behavioral phenotypes of individuals with CHRNA7 duplications. J Autism Dev Disord. 2016 doi: 10.1007/s10803-016-2961-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Klaassen A, Glykys J, Maguire J, Labarca C, Mody I, Boulter J. Seizures and enhanced cortical GABAergic inhibition in two mouse models of human autosomal dominant nocturnal frontal lobe epilepsy. Proc Natl Acad Sci U S A. 2006;103:19152–19157. doi: 10.1073/pnas.0608215103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Aracri P, Consonni S, Morini R, Perrella M, Rodighiero S, Amadeo A, et al. Tonic modulation of GABA release by nicotinic acetylcholine receptors in layer V of the murine prefrontal cortex. Cereb Cortex. 2010;20:1539–1555. doi: 10.1093/cercor/bhp214. [DOI] [PubMed] [Google Scholar]
  • 183.DiFrancesco JC, DiFrancesco D. Dysfunctional HCN ion channels in neurological diseases. Front Cell Neurosci. 2015;6:174. doi: 10.3389/fncel.2015.00071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Nava C, Dalle C, Rastetter A, Striano P, de Kovel CG, Nabbout R, et al. De novo mutations in HCN1 cause early infantile epileptic encephalopathy. Nat Genet. 2014;46:640–645. doi: 10.1038/ng.2952. [DOI] [PubMed] [Google Scholar]
  • 185.Huang Z, Walker MC, Shah MM. Loss of dendritic HCN1 subunits enhances cortical excitability and epileptogenesis. J Neurosci. 2009;29:10979–10988. doi: 10.1523/JNEUROSCI.1531-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Nakamura Y, Shi X, Numata T, Mori Y, Inoue R, Lossin C, et al. Novel HCN2 mutation contributes to febrile seizures by shifting the channel’s kinetics in a temperature-dependent manner. PLoS One. 2013;8:e80376. doi: 10.1371/journal.pone.0080376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.DiFrancesco JC, Barbuti A, Milanesi R, Coco S, Bucchi A, Bottelli G, et al. Recessive loss-of-function mutation in the pacemaker HCN2 channel causing increased neuronal excitability in a patient with idiopathic generalized epilepsy. J Neurosci. 2011;31:17327–17337. doi: 10.1523/JNEUROSCI.3727-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Sugawara T, Tsurubuchi Y, Agarwala KL, Ito M, Fukuma G, Mazaki-Miyazaki E, et al. A missense mutation of the Na+ channel alpha II subunit gene Na(v)1.2 in a patient with febrile and afebrile seizures causes channel dysfunction. Proc Natl Acad Sci U S A. 2001;98:6384–6389. doi: 10.1073/pnas.111065098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Liao Y, Anttonen AK, Liukkonen E, Gaily E, Maljevic S, Schubert S, et al. SCN2A mutation associated with neonatal epilepsy, late-onset episodic ataxia, myoclonus, and pain. Neurology. 2010;75:1454–1458. doi: 10.1212/WNL.0b013e3181f8812e. [DOI] [PubMed] [Google Scholar]
  • 190.Wang J, Li Y, Hui Z, Cao M, Shi R, Zhang W, et al. Functional analysis of potassium channels in Kv7.2 G271V mutant causing early onset familial epilepsy. Brain Res. 2015;1616:112–122. doi: 10.1016/j.brainres.2015.04.060. [DOI] [PubMed] [Google Scholar]
  • 191.Dedek K, Fusco L, Teloy N, Steinlein OK. Neonatal convulsions and epileptic encephalopathy in an Italian family with a missense mutation in the fifth transmembrane region of KCNQ2. Epilepsy Res. 2003;54:21–27. doi: 10.1016/S0920-1211(03)00037-8. [DOI] [PubMed] [Google Scholar]
  • 192.Miceli F, Soldovieri MV, Lugli L, Bellini G, Ambrosino P, Migliore M, et al. Neutralization of a unique, negatively-charged residue in the voltage sensor of K V 7.2 subunits in a sporadic case of benign familial neonatal seizures. Neurobiol Dis. 2009;34:501–510. doi: 10.1016/j.nbd.2009.03.009. [DOI] [PubMed] [Google Scholar]
  • 193.Lee US, Cui J. Beta subunit-specific modulations of BK channel function by a mutation associated with epilepsy and dyskinesia. J Physiol. 2009;587:1481–1498. doi: 10.1113/jphysiol.2009.169243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Martin HC, Kim GE, Pagnamenta AT, Murakami Y, Carvill GL, Meyer E, et al. Clinical whole-genome sequencing in severe early-onset epilepsy reveals new genes and improves molecular diagnosis. Hum Mol Genet. 2014;23:3200–3211. doi: 10.1093/hmg/ddu030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Tang B, Sander T, Craven KB, Hempelmann A, Escayg A. Mutation analysis of the hyperpolarization-activated cyclic nucleotide-gated channels HCN1 and HCN2 in idiopathic generalized epilepsy. Neurobiol Dis. 2008;29:59–70. doi: 10.1016/j.nbd.2007.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Neuroscience Bulletin are provided here courtesy of Springer

RESOURCES