U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Noebels JL, Avoli M, Rogawski MA, et al., editors. Jasper's Basic Mechanisms of the Epilepsies. 5th edition. New York: Oxford University Press; 2024. doi: 10.1093/med/9780197549469.003.0046

Cover of Jasper's Basic Mechanisms of the Epilepsies

Jasper's Basic Mechanisms of the Epilepsies. 5th edition.

Show details

Chapter 46High-Voltage-Activated Calcium Channels in Epilepsy

Lessons from Humans and Rodents

.

Abstract

Voltage-gated calcium channels support an array of fundamental biological processes in various systems. In the brain, they determine neuronal responses to stimulation, dendritic properties, neuronal firing mode, and synaptic release. They also regulate multiple critical developmental processes, including neuronal proliferation, migration, connectivity, and structural activity-dependent plasticity. For these reasons, mutations in voltage-gated calcium channels result in a variety of neurological disorders displaying epilepsy as a core feature. This chapter summarizes the molecular, cellular, and network mechanisms by which mutations in high-voltage-activated calcium channels result in epilepsy, with a particular focus on CaV1.2 (CACNA1C), CaV1.3 (CACNA1D), CaV2.1 (CACNA1A), and CaV2.3 (CACNA1E)-associated disorders.

Introduction

Calcium ions are key regulators of a range of fundamental biological processes, including cellular proliferation, migration, survival, signaling, and secretion, and system-specific function such as muscle contraction, cardiac rhythmicity, and gut motility. Calcium ions also act as secondary messengers during signal transduction, triggering activity-regulated gene transcription while supporting a range of enzymatic reactions in response to specific stimuli. In the brain, calcium ions are essential for neuronal proliferation, survival, migration, axonal guidance, synapse formation and pruning, and synaptic transmission and plasticity, as detailed in this review. Disruption of calcium homeostasis thus results in a range of neurological disorders, often with epilepsy as a core symptom, together with systemic and autonomic manifestations. This chapter provides an overview of genetic epilepsies associated with mutations in high-voltage-activated (HV) voltage-gated calcium channels (VGCC) genes, while discussing some of the underlying disease mechanisms based on findings in relevant cellular and rodent models.

Voltage-Gated Calcium Channels: An Overview

VGCCs are multisubunit complexes composed of an alpha (α) pore-forming subunit and a combination of auxiliary subunits (Fig. 46–1A). Each α subunit is encoded by a different gene and determines the fundamental properties of the encoded channel (Fig. 46–1B). VGCCs are divided between high-voltage-activated (HVA) and low-voltage-activated (LVA) channels, reflecting their different behavior in response to changes in membrane potential (Fig. 46–1B; see reviews: Perez-Reyes, 2003; Catterall et al., 2005, 2020). Specific subtypes of VGCCs also differ in their biophysical properties, gating kinetics, and in their sensitivity to different pharmacological agents and toxins. HVA channels include CaV1 channels (CaV1.1 to CaV1.4) that mediate L-type currents, activate at depolarized potentials (≈–40 to –30 mV), inactivate slowly, and thus produce prolonged Ca2+ transients, blocked by 1,4-dihydropyridine (DHP). HVA channels also include CaV2 channels that mediate P/Q-type currents (CaV2.1), N-type currents (CaV2.2), and R-type currents (CaV2.3), also activated by depolarization but displaying more rapid inactivation than L-type currents. These channels are respectively inhibited by ω-Agatoxin-IVA, ω-Conotoxin-GVIA, and nickel (NiCl2), or the tarantula toxin SNX 482 (Mintz et al., 1992; Williams et al., 1992, 1994; Soong et al., 1993; McDonough et al., 1997; Newcomb et al., 1998). By contrast, CaV3 channels are LVA. They activate following depolarizations at more hyperpolarized potentials, display low-amplitude brief currents, are insensitive to DHP but highly sensitive to nickel (Tsien et al., 1988, Soong et al., 1993; see review: Perez-Reyes, 2003).

Figure 46–1.. General overview of voltage-gated calcium channels.

Figure 46–1.

General overview of voltage-gated calcium channels. A. Voltage-gated calcium channels (VGCC) are composed of a α1 pore-forming subunit, together with various auxiliary subunits. (Inspired from Catterall and Few, 2008). B. Phylogenetic tree illustrating (more...)

Important structure-function principles are shared between VGCCs and inform our understanding of the functional impact of various disease-causing mutations (see reviews: Catterall, 2000, Catterall et al., 2005, 2017, 2020; Campiglio and Flucher, 2015). The α subunit contains four homologous domains (labeled I to IV), each composed of six trans-membranous helical segments (S1–S6; Fig. 46–1C). The S5 and S6 segments of each domain form the ionic pore, which is lined by charged amino acids that determine the channels’ affinity for Ca2+ ions. The S5 to S6 loops of each domain interact to form the pore entry and the selectivity filter. The S4 segments are voltage sensors. Their outward movement upon depolarization drives conformational changes that open the channel pore. Mutations involving this voltage-sensitive motif tend to impact voltage activation threshold. Fast voltage-dependent channel inactivation occurs through movement of the inter-domain DII-DIII linker loop toward the channel pore, contacting the cytoplasmic ends of the S4–S6 segments. Disease-associated variants in these regions often impact channel inactivation and result in a functional gain-of-function (GoF) effect. A slower inactivation process has also been described, involving secondary messengers and modulation by G protein–coupled receptors (Pietrobon, 2010). Accessory subunits interact with the α subunit (Fig. 46–1A) and regulate channel assembly, trafficking, membrane targeting, and gating properties (Lacinova, 2005; Campiglio and Flucher, 2015).

Different VGCCs usually coexist within a single neuron to control various important biological processes in the dendrites, soma, and nerve terminals (Fig. 46–1D). For instance, different HVA VGCCs mediate synaptic release from a single neuron, with coexpression of different VGCCs often occurring within single release sites (Mintz et al., 1995; Wu et al., 1999; Eltes et al., 2017), although not in all cell types (Yamamoto and Kobayashi, 2018). This suggests that, in many neuronal populations, the loss of one VGCC may be partly compensated by the upregulation of other VGCCs. However, different VGCCs show different efficacy in controlling neurotransmitter release, reflecting their relative position with regard to the active zone: CaV2.1 channels are tightly bound to presynaptic structural elements in the active zone, while CaV2.3 channels tend to be localized further away from the release sites (Sheng et al., 1994; Mintz et al., 1995; Rettig et al., 1996; Wu et al., 1999; Rozov et al., 2001). Their different gating kinetics, binding to intracellular regulators (calmodulin, G protein–coupled receptors) and post-translational processes result in different plasticity rules, such that synaptic transmission that relies on CaV2.1 channels is characterized by short-term facilitation following repeated stimulations, while its loss (and its functional compensation by other HVA channels) results in short-term depression (see review: Catterall et al., 2013). Interestingly, individual release sites from a single neuron, which connect to different cellular targets, typically rely almost exclusively on one type of VGCC in a target-specific manner (Markram et al., 1998; Zaitsev et al., 2007; Rossignol et al., 2013; Eltes et al., 2017) and thus display different short-term synaptic plasticity rules (Reyes et al., 1998; Koester and Johnston, 2005; Eltes et al., 2017; Yamamoto and Kobayashi, 2018), which provide target-specific VGCC modulation at individual synapses (Eltes et al., 2017). Thus, even if the loss of function of a particular α subunit induces a functional compensation by other VGCCs, the different biophysical properties of specific α subunits render them not entirely interchangeable, resulting in functional deficits that are often cell type or network specific, as detailed below.

The functional diversity of VGCCs is further expanded by alternative splicing, which results in the production of multiple different α subunit isoforms, with specific cell type, tissue type, and age-dependent expression patterns (Rettig et al., 1996; Sakurai et al., 1996; Khan et al., 1998; Pereverzev et al., 1998, 2022; Grabsch et al., 1999, Schramm et al., 1999, Mitchell et al., 2002; Striessnig et al., 2006; Ernst and Noebels, 2009; Sinnegger-Brauns et al., 2009; Lipscombe and Andrade, 2015). Thus, some disease-associated mutations that affect selected isoforms result in cell-type-specific and tissue-specific pathological outcomes, in an age-dependent fashion.

The current chapter focuses on the implication of HVA VGCCs in epilepsy, with a particular focus on L-type (CaV1.2 and CaV1.3), P/Q-type (CaV2.1), and R-type (CaV2.3) VGCCs, given the wealth of recent data implicating these channels in various human epilepsy syndromes. Readers are referred to other extensive reviews concerning the important roles of LVA channels and various auxiliary subunits in epilepsy (Perez-Reyes, 2003; Cain and Snutch, 2012; Campiglio and Flucher, 2015; Lory et al., 2020; Weiss and Zamponi, 2020).

L-Type VGCCs in Epilepsy

CaV1.1 channels are expressed most prominently in skeletal muscles (Curtis and Catterall, 1984), with truncated isoforms found in cardiomyocytes and neurons (Diebold et al., 1992; Khan et al., 1998; Wielowieyski et al., 2001). CACNA1S mutations thus give rise to various muscle disorders, including hypokalemic periodic paralysis, malignant hyperthermia, and congenital myopathies (Weber and Lehmann-Horn, 1993; Ptacek et al., 1994; Monnier et al., 1997; Davies et al., 2001; Stewart et al., 2001; Miller et al., 2004; Kim et al., 2007; Carpenter et al., 2009; Matthews et al., 2009; Toppin et al., 2010; Schartner et al., 2017; Flucher, 2020; Brugnoni et al., 2021), with no overt central nervous system (CNS) phenotype described to date. Similarly, CaV1.4 channels are expressed solely in the retina, and their dysfunction results in visual disorders (congenital stationary night blindness with poor visual acuity and nystagmus), without epilepsy or other CNS deficits (Haeseleer et al., 2004; Dembla et al., 2020; Waldner et al., 2020; Koschak et al., 2021).

By contrast, the CACNA1C and CACNA1D genes, respectively encoding the L-type VGCCs CaV1.2 and CaV1.3, have both been associated with epilepsy and various neurodevelopmental disorders, through a range of physiological processes involved in ictogenesis and epileptogenesis, as detailed below. Notably, classical studies on the function of L-type VGCCs in neurons largely relied on pharmacological tools, including the use of classical VGCC blockers: 1,4-dihydropyridines (DHP) (nifedipine, amlodipine), phenylalkylamines (verapamil), and benzothiazepine (diltiazem) (Catterall et al., 2005, 2020); they could not efficiently differentiate the role of different L-type VGCCs in specific biological processes. However, the recent advent of genetically engineered mouse models affecting isolated L-type VGCCs is providing emerging insights on the selective roles of CaV1.2 and CaV1.3 channels in various physiological and pathological processes, as summarized below.

L-Type CaV1.2 Channels: CACNA1C-Associated Disorders in Humans

CaV1.2 channels are expressed broadly in the brain, including in the cortex, hippocampus, amygdala, thalamus, and cerebellum (Splawski et al., 2004). They are also expressed in multiple other organs, including the heart, blood vessels, eyes, developing digits, smooth muscles, lymph nodes, and pancreas (Splawski et al., 2004). Therefore, de novo missense mutations in the CACNA1C gene have been associated with a multisystemic disorder called Timothy syndrome (MIM 601005), which includes life-threatening cardiac arrhythmias (long QT syndrome, second degree AV bloc and ventricular fibrillation), cardiac structural malformations (patent ductus arteriosus, ventricular septal defects, Tetralogy of Fallot), dysmorphic traits (syndactyly, low-set ears, flat nasal bridge, receding upper jaw, small displaced teeth), intermittent hypoglycemia, immune deficiency, and a range of neurocognitive deficits (intellectual deficiency and/or autism spectrum disorder) (Reichenbach et al., 1992; Splawski et al., 2004; see reviews: Napolitano et al., 1993; Bidaud and Lory, 2011; Franklin and Laubham, 2021). Notably, close to a quarter of children with Timothy syndrome develop epilepsy (Splawski et al., 2004). Timothy-syndrome-associated mutations typically induce a GoF effect, mostly through a deficit of channel voltage-dependent fast inactivation (Splawski et al., 2004; Barrett and Tsien, 2008), often with a leftward shift of activation toward more hyperpolarized potentials (Dick et al., 2016), resulting in persistent calcium “window” currents, calcium overload, and cellular toxicity (Splawski et al., 2004).

More recently, missense CACNA1C mutations have been reported in patients with isolated neurological symptoms, without the classical cardiac or systemic manifestations of Timothy syndrome (Rodan et al., 2021; Fig. 46–2A). Nonsense truncating variants or intragenic deletions typically present with autism, intellectual deficiency, or isolated language delay (Firth et al., 2011; Quintela et al., 2017; Mio et al., 2020; Rodan et al., 2021). By contrast, de novo missense variants give rise to severe early-onset epileptic encephalopathies, with developmental delay, hypotonia, ataxia, intellectual disability, and/or autism spectrum disorders (Firth et al., 2011; Iossifov et al., 2014; Rodan et al., 2021). Although not all children with de novo missense variants develop epilepsy, the majority do (≈70% of cases) (Rodan et al., 2021). CACNA1C-associated epilepsies include focal or generalized seizure disorders, as well as severe early-onset epileptic encephalopathies (West syndrome, neonatal epileptic encephalopathies), often evolving toward Lennox-Gastaut syndrome (Bozarth et al., 2018; Rodan et al., 2021). Electroencephalographic (EEG) findings are nonspecific, displaying focal or generalized epileptiform discharges, and brain imaging is usually unremarkable (Rodan et al., 2021).

Figure 46–2.. CaV1.

Figure 46–2.

CaV1.2 and CaV1.3 L-type voltage-gated calcium channels. A. Schematic representation of selected CACNA1C mutations, color-coded according to their associated phenotypes. DEE, developmental epileptic encephalopathies; NDD, neurodevelopmental disorders, (more...)

The functional validation of four de novo missense variants associated with epilepsy revealed either a GoF effect (p.L614P and p.L657F), mostly through a shift of voltage-dependent activation toward more hyperpolarized potentials; or a loss-of-function (LoF) effect (p.L1408V), with reduced current density but unaltered gating kinetics (Rodan et al., 2021). Notably, one variant (p.L614R) had no clear impact on channel function, although this might reflect methodological differences in the associated subunits used to test this particular variant (Rodan et al., 2021). Other potential detrimental impacts leading to cellular toxicity, such as protein misfolding, mistargeting, intracellular accumulation, and ER stress, may contribute to the disease pathogenesis of specific variants (Bidaud and Lory, 2011), as demonstrated also for other VGCC disorders (Mezghrani et al., 2008; Jiang et al., 2019). To better appreciate these potential pathological mechanisms, testing of a larger sample of patient-derived mutations, preferably expressed in neurons, will likely be required (Rodan et al., 2021).

By contrast, some missense CACNA1C mutations present with isolated cardiac phenotypes. For instance, GoF variants have been associated with long QT syndrome (Zhang et al., 2018), while nonsense variants have been associated Brugada syndrome and a risk of sudden death (Antzelevitch et al., 2007), although this is debated (Hosseini et al., 2018). Nonetheless, such system-specific effects may reflect the selective impact of specific variants on different tissue-specific isoforms (Snutch et al., 1991; Chin et al., 1992; Kim et al., 1992; Sinnegger-Brauns et al., 2004; Liao and Soong, 2010; Napolitano and Antzelevitch, 2011; Lipscombe and Andrade, 2015; Bozarth et al., 2018; Clark et al., 2020; Franklin and Laubham, 2021). However, some patients with epilepsy and CACNA1C variants display cardiac conduction disorders, warranting caution in monitoring for potential arrhythmias and a theoretical risk of SUDEP, as some missense variants may affect all channel isoforms (Rodan et al., 2021).

L-Type CaV1.3 Channels: CACNA1D-Associated Disorders in Humans

CaV1.3 L-type channels are expressed in neurons but also in the heart sinoatrial node, pancreatic β islet cells, adrenal chromaffin cells, and cochlear inner ear cells (Bidaud and Lory, 2011). Recessive LoF CACNA1D mutations are associated with a rare syndrome of sinoatrial node dysfunction with deafness (SANDD), without epilepsy or other neurological phenotypes (Baig et al., 2011; Liaqat et al., 2019). By contrast, de novo CACNA1D GoF mutations (Fig. 46–2B) result in a complex disorder, with hyperaldosteronism, developmental delay, spasticity, intellectual disability, and/or autism (Iossifov et al., 2012; O’Roak et al., 2012; Pinggera et al., 2015, 2017). Notably, some of these patients also develop epilepsy with focal and secondary generalized seizures, together with multifocal epileptic discharges on EEG (Scholl et al., 2013; Pinggera et al., 2017; Long et al., 2019; Hofer et al., 2020). The exact frequency of CACNA1D-associated epilepsy is still unknown. This should be assessed in larger cohorts of patients.

Mechanisms of L-Type VGCC-Associated Epilepsy: Dendritic Calcium Transients, Afterhyperpolarization, and Paroxysmal Depolarization Shifts

Physiological Roles of L-Type VGCCs in Neurons

L-type VGCCs are expressed in most neuronal populations, including in pyramidal cells and GABAergic neurons of the neocortex, hippocampus, dentate gyrus, and cerebellum (both granule cells and Purkinje cells) (Chin et al., 1992; Hell et al., 1993). CaV1.2 are the most abundant L-type VGCCs in the brain (90% of all L-type VGCCs) (Sinnegger-Brauns et al., 2009), but both CaV1.2 and CaV1.3 channels are often coexpressed in individual neurons (see review: Zamponi et al., 2015). L-type VGCCs are typically localized at the base of apical and basal dendrites, with some histochemical staining observed on the cell soma (Westenbroek et al., 1990). While CaV1.2 channels form discrete clusters on the dendritic shafts and spines (Westenbroek et al., 1990; Hell et al., 1993), CaV1.3 channels are more diffusely distributed (Westenbroek et al., 1990; Hell et al., 1993), although they can also assemble in clusters on dendrites (Stanika et al., 2016). Interestingly, the somatic distribution of both channels differs between cell types, such that, in the hippocampus, somatic expression of CaV1.2 channels predominates in GABAergic interneurons of the pyramidal cell layer, while CaV1.3 channels are typically found on the cell body of pyramidal cells in CA1 to CA4 and in the dentate gyrus (Kelly et al., 2003). This pattern of expression is notably changed after prolonged seizures and once epilepsy is established, with regional and cell-type redistribution of both channels (Kelly et al., 2003; Xu et al., 2018).

L-type VGCCs play multiple critical roles during physiological brain functions. They mediate long-lasting dendritic calcium transients (Schwartzkroin and Slawsky, 1977; Wong and Prince, 1978; Benardo et al., 1982; Nowycky et al., 1985; Murphy et al., 1995), which help propagate postsynaptic depolarizations resulting from dendritic synaptic inputs (Amitai et al., 1993; Reuveni et al., 1993). They can also be activated by back-propagating action potentials following somatic depolarizations (Christie et al., 1995; Spruston et al., 1995), particularly in the context of tetanic stimulations (Regehr et al., 1989; Chavis et al., 1996). In particular, CaV1.2 channels contribute actively to NMDA-independent long-term potentiation (LTP) following repeated high-frequency stimuli in the hippocampus (Grover and Teyler, 1990, 1992; Kato et al., 1993; Cavus and Teyler, 1996, 1998; Clark et al., 2003) and to theta burst-induced LTP (Sridharan et al., 2020), both being independent of CaV1.3 channels (Clark et al., 2003). The prolonged Ca2+ plateaus generated by L-type VGCCs provide the biological substrate for repeated burst firing upon repeated stimulations. Similar L-type VGCC-mediated somatic plateau potentials can be elicited by muscarinic activation, which drives persistent firing of hippocampal neurons through the activation of nonselective cation conductances (TRPC channels) (Egorov et al., 2002; Zhang and Seguela, 2010; Zhang et al., 2011).

In addition, L-type VGCCs trigger membrane afterhyperpolarization following neuronal discharges by activating Ca2+-dependent K+ channels (Lancaster and Nicoll, 1987; Lancaster et al., 1991; Chavis et al., 1998). This occurs through coupling of CaV1.3 with KCa3.1 channels, which permits repeated firing in hippocampal neurons, in a process that is independent from CaV1.2 (Sahu et al., 2017). Such CaV1.3-mediated afterhyperpolarization is also important for spontaneous pace making, for instance, in dopaminergic neurons (Putzier et al., 2009; Poetschke et al., 2015; Sun et al., 2017), as it activates LVA channels that trigger rebound depolarizations and spontaneous discharges (Pennartz et al., 2002; Guzman et al., 2009).

Importantly, together with NMDA receptors, L-type VGCCs are key regulators of excitation-transcription coupling, a process by which activity triggers gene transcription (Fig. 46–2C; Morgan and Curran, 1986, 1988; Murphy et al., 1991; Greenberg et al., 1992; Misra et al., 1994; Striessnig et al., 2006). This effect is mediated by the activation of Ca2+-dependent (calmoduline/calcineurin) (Dolmetsch et al., 2001; Yap and Greenberg, 2018) and Ca2+-independent (ERK-CREB) pathways (Servili et al., 2018, 2019; Fig. 46–2C). L-type VGCCs thus contribute actively to activity-dependent structural plasticity underlying the maturation, maintenance, and pruning of dendritic spines and network connectivity during normal development and age-related degeneration (Day et al., 2006; Hirtz et al., 2011, 2012; Stanika et al., 2015, 2016). In particular, given their intimate interactions with postsynaptic proteins, including shank and densin-180, CaV1.3 channels control synaptic plasticity and spine morphology following input summation (Zhang et al., 2005; Stanika et al., 2015, 2016; Wang et al., 2017). L-type VGCCs also promote neuronal survival through the induction of NMDA-independent LTP (Kang and Schuman, 2000), driving TrkB and BDNF expression (Ghosh et al., 1994; Cavus and Teyler, 1996).

Further, CaV1.3 channels, expressed in neural stem cells of the dentate gyrus, are critical for adult hippocampal neurogenesis, as well as for the survival and maturation of newborn neurons involved in novel spatial memory formation (Deisseroth et al., 2004; Marschallinger et al., 2015; Kim et al., 2017). By contrast, CaV1.2 channels are expressed in postmitotic neurons and are thus required for the survival, but not the proliferation, of newborn neurons during adult neurogenesis (Lee et al., 2016).

Physiological Roles of L-Type VGCCs during Brain Development

L-type VGCCs act as critical regulators of brain and circuit development, with different splice isoforms expressed during early cortical development (Tang et al., 2009). In the embryo, L-type VGCCs mediate spontaneous nonsynaptic somatic calcium transients, so-called calcium spikes (Crepel et al., 2007; Bortone and Polleux, 2009), as well as calcium transients in neurites that occur independently from somatic spikes (Kamijo et al., 2018). These early calcium transients regulate neuronal proliferation (LoTurco et al., 1995; Malmersjo et al., 2013), differentiation (Gu and Spitzer, 1997), growth cone dynamics (Mattson and Kater, 1987; Kater et al., 1988; Tang et al., 2003), neurite extension, axonal branching, and axonal path finding (Schmitz et al., 2009; Kamijo et al., 2018; Nakagawa-Tamagawa et al., 2021). Calcium transients are thus required for the migration of pyramidal cells (Kamijo et al., 2018; Horigane et al., 2021; Nakagawa-Tamagawa et al., 2021) and GABAergic interneurons (Bortone and Polleux, 2009; Birey et al., 2022). The progressive increase in calcium transient frequency, mediated by both L-type and N-type VGCCs, eventually leads to the arrest of migration and to the onset of dendritogenesis (Bando et al., 2016).

Furthermore, in the early postnatal days in rodents (≈ late gestation in humans), L-type VGCCs generate calcium plateaus that drive spontaneous population burst discharges, the synchronous plateau assemblies (SPAs), in small groups of gap-junction interconnected neurons in the hippocampus (Crepel et al., 2007) and neocortex (Yuste et al., 1995). These coincide in time with the emergence of large-scale synchronous population discharges, the so-called early-network activities (ENOs), recorded in the neonatal hippocampus CA1 (Garaschuk et al., 1998) and neocortex (Garaschuk et al., 2000), reflecting synaptically induced calcium transients on pyramidal cell apical dendrites (Garaschuk et al., 1998), following GABAergic inputs in the hippocampus and glutamatergic inputs in the neocortex (Garaschuk et al., 2000; Allene et al., 2008). These are followed by giant depolarization potentials (GDPs), GABAergic-driven large-scale synchronized population events recorded in vitro in the hippocampus (Ben-Ari et al., 1989; Leinekugel et al., 1997; Crepel et al., 2007) and neocortex (Allene et al., 2008) at the end of the first postnatal week, and likely corresponding to in vivo hippocampal sharp waves (Leinekugel et al., 2002) and cortical spindle bursts (Khazipov et al., 2004). GDP require the action of VGCCs in an NMDA-independent fashion (Leinekugel et al., 1995, 1997). These patterned activities are critical for the proper establishment of hippocampal and cortical circuitry (see review: Cossart and Khazipov, 2022).

Given their early pattern of expression (Kamijo et al., 2018), L-type VGCCs likely contribute to calcium transients observed during ENO and GDPs, as well as to the intrinsic neuronal responses to these patterned activities during the early postnatal days. For instance, L-type VGCCs induce the translocation of axon initial segment components to fine-tune neuronal excitability to the level of ambient activity, in a process of developmental activity-regulated plasticity (Grubb and Burrone, 2010). They also participate in the transition to intrinsic burst-firing modes in neonatal hippocampal neurons, together with N- and T-type VGCCs, through the induction of plateau potentials and Ca2+-dependent spike afterdepolarization (Chen et al., 2005). Similar early L-type VGCC dependent plateau potentials have been observed in a range of neuronal populations (Lo and Erzurumlu, 2002; Lo et al., 2002; Dilger et al., 2011) and are required for the establishment of network connectivity. They increase the excitability of basolateral amygdala neurons in response to stimulation in neonatal (P7) but not older (P21) animals (Zhang et al., 2020). These early changes in neuronal excitability result in long-term enhancement of network excitability, with behavioral relevance (Zhang et al., 2022). L-type VGCCs participate in a range of other experience-dependent plasticity in the neonatal hippocampus: they are essential for the induction of depolarization-induced LTD (spike timing-dependent LTD), which is required for proper maturation of mossy fiber to CA3 pyramidal cell synapses (Ho et al., 2009) and of CA3 to CA1 synapses (Falcon-Moya et al., 2020), as well as for the conversion of silent synapses into active ones following repeated stimulations (Yao et al., 2006). They are also critical for experience and Ca2+-dependent structural plasticity of dendrites during development in both pyramidal cells (Korkotian and Segal, 1999; Schwab et al., 2001; Yamashita et al., 2016), and GABAergic interneurons (Jiang and Swann, 2005), mostly through specific, developmentally regulated CaV1.3 isoforms (Stanika et al., 2016).

Involvement of L-Type VGCCs in Ictogenesis and Epileptogenesis

Given their various roles in regulating neuronal excitability, activity-regulated gene transcription and network reorganization, L-type VGCCs are thought to contribute actively to various aspects of ictogenesis and epileptogenesis. Indeed, L-type VGCCs are upregulated in a variety of pro-epileptogenic contexts, such as ischemia, stroke, fever, trauma, and status epilepticus (Kohr and Mody, 1991; Vreugdenhil and Wadman, 1992, 1994; Beck et al., 1998). This upregulation appears to be region- and channel-specific (Kelly et al., 2003; Xu et al., 2018), such that CaV1.2 channels are upregulated in neuronal somata of pyramidal and granule cell layers of both CA1 and CA3, while they are reduced in the neuropil (Kelly et al., 2003). By contrast, CaV1.3 channels are downregulated in CA3 and hilus (Kelly et al., 2003).

Following pro-convulsive insults, L-type VGCCs trigger the generation of paroxysmal depolarization shifts (PDSs) (Rubi et al., 2013), the intracellular correlate of interictal spikes (see reviews: Speckmann et al., 1993; de Curtis and Avanzini, 2001; Kubista et al., 2019; Fig. 46–2D). PDSs are initiated by AMPA-receptor activity following synaptic input on distal dendrites, which activates local L-type VGCCs that generate the prolonged depolarization characteristic of PDSs (Schiller, 2002). PDSs then propagate toward the cell body through NMDA and L-type VGCC conductances (Schiller, 2002; Meyer et al., 2021; Fig. 46–2D). In the soma, coupling of L-type VGCCs with sodium channels enhances neuronal excitability (Stiglbauer et al., 2017), triggering repetitive firing in CA3 neurons (Ishihara et al., 1993), which is sensitive to L-type VGCC blockers (Rubi et al., 2013; Stiglbauer et al., 2017). Although the targeted activation of L-type VGCCs by BayK is insufficient to trigger PDSs in hippocampal preparations (Rubi et al., 2013), it greatly enhance the frequency and duration of PDSs when applied with other precipitating factors (GABAA block with bicuculline, reducing extracellular Mg+ or Ca2+, increasing NMDA/AMPA signaling, blocking 4AP sensitive K+ channels, oxidative stress, or electrical stimulations) (Walden et al., 1986; Rubi et al., 2013; Meyer et al., 2021). Notably, blocking L-type VGCCs with isradipine prevents PDSs in these contexts, highlighting their importance in the generation of PDSs in pro-convulsive contexts (Rubi et al., 2013; Meyer et al., 2021).

However, L-type VGCCs have also been involved in the termination of epileptic discharges, particularly in the hippocampus (Empson and Jefferys, 2001), while in the neocortex, discharge termination largely depends on N-type VGCCs (Pineda et al., 1998). This occurs through L-type VGCC-mediated Ca2+ influx that activates ryanodine receptors, triggering Ca2+ release from intracellular stores (Chavis et al., 1996), which in turn activates Ca2+-dependent K+ channels (Tanabe et al., 1998) that mediate slow somatic afterhyperpolarizations (IAHP current) that limit CA3 neuronal discharges (Empson and Jefferys, 2001). Thus, acutely blocking L-type VGCCs with low-dose nifedipine results in longer and more frequent epileptiform discharges in disinhibited chronic hippocampal slice culture models (Empson and Jefferys, 2001). However, blocking L-type VGCCs with higher concentrations of nifedipine or verapamil results in a sustained suppression of epileptiform activity following a brief enhancement (Aicardi and Schwartzkroin, 1990; Straub et al., 1992, 1994, 1996, 1997, 2000). Notably, after seizure initiation, L-type VGCC blockers fail to interrupt prolonged epileptiform events (Rubi et al., 2013), suggesting that other mechanisms sustain epileptiform discharges after seizure onset. Nonetheless, the intraventricular infusion of verapamil in the rat cortical penicillamine model of epilepsy reduces interictal spike frequency and spontaneous seizures in vivo (Walden et al., 1985; Walden and Speckmann, 1988), suggesting a net antiseizure effect.

In neocortical slice models, high-dose verapamil also reduces epileptiform discharges in pro-convulsive contexts, once again after a brief enhancement (Straub et al., 1992, 2000). Notably, intra-dendritic injections of verapamil suffice to prevent dendritic depolarizations following tetanic stimulation (Schiller, 2002), in a ryanodine-receptor-independent fashion (Schiller, 2004), confirming the local dendritic origin of PDSs. However, high-dose nifedipine results in a sustained enhancement of epileptiform discharges (Straub et al., 1992, 2000), suggesting different sensitivity to phenylalkylamine-type blockers (nifedipine) and dihydropyridine-type blockers (verapamil) in these circuits, and in different neuronal populations (Straub et al., 2000). Interestingly, recent evidence suggests that PDS generation in neocortical pyramidal cells require CaV1.3, but not CaV1.2 channels (Stiglbauer et al., 2017; Fig. 46–2E). By contrast, coupling of CaV1.2 channels with Ca2+-dependent cation (CAN) channels at the soma induces prolonged depolarizations during long-lasting seizure-like events (Stiglbauer et al., 2017). Different sensitivity of CaV1.2 and CaV1.3 channels to specific VGCC blockers, and their different roles in dendritic or somatic processes, may thus explain the selective pharmacological effects of various agents (Koschak et al., 2001; Zamponi et al., 2015).

Apart from their roles in inducing PDS and acute epileptiform events, L-type VGCCs are thought to contribute actively to the epileptogenic process. Indeed, PDS themselves, occurring shortly after pro-epileptogenic insults and resembling developmental GDP, are thought to actively drive the remodeling of neuronal circuits that ultimately leads to epilepsy, through their various known roles in excitation-transcription coupling and plasticity (Staley et al., 2005, 2011; Staley and Dudek, 2006; Meyer et al., 2021). Thus, blocking L-type VGCCs immediately after a pro-convulsive insult reduces epileptiform activity in vitro and long-term seizure burden in vivo (Walden et al., 1985, 1992), even when the acute seizure threshold is unchanged (Mikati et al., 2004).

By contrast, the implication of L-type VGCCs in promoting epileptiform discharges once epilepsy is established is unclear, and it appears to depend on the underlying etiology. In some genetic models of epilepsy, enhanced L-type VGCCs expression persists long after epilepsy onset, as in the spontaneously epileptic rat (SER) (Sasa et al., 1992; Ishihara et al., 1993; Momiyama et al., 1995; Amano et al., 2001; Yan et al., 2007), in the Wistar Albino Glaxo rats from Rijswwijk (WAG/Rij), the genetically epilepsy prone rat (GEPR), or the seizure-prone gerbil model, in which CaV1.3 but not CaV1.2 channels remain overexpressed in subcortical structures (Park et al., 2003; N’Gouemo et al., 2010; Kanyshkova et al., 2014). Similar enhancement has been reported in chronic epileptogenic lesions (hypothalamic hamartoma) (Simeone et al., 2011). By contrast, other models, such as the pilocarpine model of epilepsy, display increased non-L-type VGCCs mediated calcium currents (Su et al., 2002) and a persistent reduction of both CaV1.2 and CaV1.3 channels in CA1 and CA3 neurons (Xu et al., 2018), while a more selective long-term reduction of CaV1.3 expression is observed in the enthorinal kindled rat model (Blalock et al., 2001). Thus, an early upregulation and/or enhanced activation of L-type VGCCs following an acute insult may facilitate network reorganization, but this may be dispensable during the subsequent phase of established epilepsy in different contexts.

Interestingly, in multiple models of brain injury (trauma, hypomyelination, ischemia, or kainic acid–induced status epilepticus), the chronic upregulation of CaV1.2 channels appears to be mostly restricted to reactive astrocytes rather than to neurons (Westenbroek et al., 1998). In the spontaneously epileptic rat model, an upregulation of CaV1.3 but not CaV1.2 channels has been observed, once again in astrocytes (Xu et al., 2018). The upregulation of CaV1.2 channels in astrocytes has been hypothesized to promote their pro-repair functions (ionic homeostasis, neurotrophin secretion, synaptic plasticity) (Westenbroek et al., 1998), while the upregulation of CaV1.3 channels would drive spontaneous Ca2+-dependent oscillations that trigger neuronal depolarizations and epileptiform discharges (Wetherington et al., 2008; Gomez-Gonzalo et al., 2010; Xu et al., 2018).

Given this complexity, caution seems warranted when considering long-term treatments with L-type VGCC blockers once epilepsy is established. A small randomized cross-over clinical trial of nimodipine as an add-on treatment for refractory epilepsy in humans failed to show benefit (Meyer et al., 1995), and evidence of long-term benefit of flunarizine or amlodipine is lacking at this point (Hasan et al., 2013). Furthermore, studies in adult patients with long-standing refractory epilepsy have shown no benefit from verapamil (Borlot et al., 2014). However, verapamil reduced seizure burden to variable degrees in various small cohorts of children with refractory epilepsy, including temporal lobe epilepsy, Lennox-Gastaut syndrome, or Dravet syndrome (Iannetti et al., 2009; Asadi-Pooya et al., 2013; Nicita et al., 2014, 2016; Narayanan et al., 2016), likely reflecting different etiologies and age-dependent effects. It is also possible that these results reflected other actions of verapamil, such as blocking P-glycoprotein, an active efflux transporter protein involved in drug resistance through the extrusion of antiepileptic drugs (Feldmann and Koepp, 2012). Thus, whether verapamil should be considered in specific genetic epilepsies, particularly in those characterized by a primary GoF of L-type VGCCs, remains to be explored using appropriate preclinical models.

Network Mechanisms of CACNA1C and CACNA1D-Related Epilepsies

A primary LoF of L-type VGCCs might result in prolonged seizures in specific contexts, due to a failure of seizure termination (Empson and Jefferys, 2001), although it might also be protective against the induction of seizures, given the central role of L-type VCC in PDS (Rubi et al., 2013; Stiglbauer et al., 2017). However, whether it would suffice to induce epilepsy in the absence of other pro-convulsive events is uncertain. Indeed, homozygous Cacna1c knockout (KO) mice do not develop spontaneous seizures, although they present with other deficits (deafness and cardiac arrhythmias) (Platzer et al., 2000). Conditional KO mice, carrying a targeted deletion of Cacna1c in the medial prefrontal cortex, display anxiety, and deficits in social interactions and in spatial learning and memory, together with elevated excitation/inhibition (E/I) ratio in frontal cortical circuits (Lee et al., 2012; Kabir et al., 2017). However, they do not present spontaneous seizures or early lethality, and their seizure threshold has not been reported. Similarly, homozygous Cacna1d KO mice do not develop epilepsy (Clark et al., 2003), despite having other systemic deficits (arrhythmias and deafness) (Platzer et al., 2000). Thus, the primary loss of either CaV1.2 or CaV1.3 is unlikely to be sufficient to induce epilepsy in the absence of other pro-epileptogenic insults, in agreement with the observation that most epilepsy-associated CACNA1C or CACNA1D mutations are GoF variants.

Mechanistically, GoF of L-type VGCCs would be expected to enhance neuronal excitability in response to synaptic inputs, to promote burst-firing mode in multiple neuronal populations, to facilitate PDS induction and ictogenesis, and to enhance excitation-transcription coupling and the remodeling of circuits following sustained discharges. A classical mouse model of Timothy syndrome (TS2-neo), carrying the G406R Cav1.2 GoF mutation, displays autism-like phenotypes (repetitive behaviors, altered social interactions, enhanced conditioned fear memory acquisition) (Bader et al., 2011), increased excitability and bursting of chromaffin cells (Calorio et al., 2019), accelerated oligodendrocyte maturation and myelination (Cheli et al., 2018), and aberrant development of GABAergic interneurons (increased number of smaller GABAergic synaptic boutons in cortical layer IV) (Horigane et al., 2020). However, whether this model is more sensitive to pro-convulsive triggers has not been reported. Notably, this model shows an attenuated phenotype, thanks to the persistence of a neomycin cassette likely causing haploinsufficiency of the mutant allele, while an initial knock-in mouse generated without this cassette was not viable (Bader et al., 2011). By contrast, the expression of mutant Cav1.2 channels carrying this same mutation (G406R) in neurons in vitro results in a range of early developmental aberrations, including altered neuronal cell fate in favor of callosal projection neurons (reduced CTIP2+ subcortical projection neurons) (Pasca et al., 2011; Panagiotakos et al., 2019), enhanced activity-dependent dendritic retraction (Krey et al., 2013), and decreased radial migration attributable to prematurely enhanced calcium transients (Kamijo et al., 2018), but also perhaps to a massive increase in baseline activity-independent CREB phosphorylation and gene transcription (Pasca et al., 2011; Servili et al., 2020). Similarly, the in utero electroporation of a mutant version of CaV1.2 channels carrying the Timothy syndrome GoF mutation I1166T impairs radial migration of callosal projection neurons and largely prevents the extension of callosal projections to the contralateral hemisphere (Nakagawa-Tamagawa et al., 2021). In addition, iPSC-derived mixed organoids from Timothy syndrome patients reveal stark deficits in GABAergic interneuron migration, with more frequent but shorter bouts of saltation, reducing net migration (Birey et al., 2017, 2022). Although such deficits have not yet been reported in Timothy syndrome mouse models, impaired migration of GABAergic interneurons would be expected to result in early developmental epilepsies (see review: Jiang et al., 2016). Thus, embryonic or early postnatal lethality of particular Timothy syndrome mice models might preclude the in-depth study of epileptogenic mechanisms.

Furthermore, Cacna1c undergoes extensive splicing, generating more than 40 isoforms of CaV1.2 (Tang et al., 2009). Biasing the expression of specific isoforms, for instance by deleting the alternatively spliced exon 33 of CaV1.2 (typically expressed in the heart and in early brain development; Tang et al., 2009), which results in a channel isoform with increased current density and open probability, results in spontaneous arrhythmias, enhanced LTP, lack of long-term depression (LTD), and impaired spatial learning and sociability/social novelty seeking (Li et al., 2017; Navakkode et al., 2022). Whether these animals present spontaneous seizures or reduced seizure threshold must be investigated. However, given the extensive splicing of CaV1.2, and the tissue selectivity of specific isoforms, together with the fact that only a quarter of patients with Timothy syndrome develop epilepsy (Splawski et al., 2004), the full appraisal of network deficits leading to CACNA1C-related epilepsy will likely require the generation of novel mice models carrying epilepsy-associated mutations from patients with isolated neurological phenotypes, thus bypassing the early cardiac mortality while ensuring an effect on brain isoforms relevant to epilepsy. Such models will be instrumental to decipher disease mechanisms and to test novel therapies.

P/Q-Type CaV2.1 VGCC and Epilepsy

P/Q-Type CaV2.1 Channels: CACNA1A-Associated Disorders in Humans

CaV2.1 channels are expressed broadly in the brain, including in most neuronal populations of the cortex, hippocampus, striatum, amygdala, thalamus, and cerebellum (Westenbroek et al., 1995; Ludwig et al., 1997), as well as in the spinal cord and neuromuscular junctions (Westenbroek et al., 1998). CACNA1A mutations have thus been associated with multiple neurological disorders characterized by prominent motor manifestations, including acetazolamide-responsive episodic ataxia type II (EA2; MIM #108500) (Ophoff et al., 1996), degenerative spinocerebellar ataxia type 6 (SCA6; MIM # 183086) (Zhuchenko et al., 1997), and familial hemiplegic migraines type I (FHM1; MIM #141500) (Ophoff et al., 1996), as reviewed extensively elsewhere (Pietrobon, 2010). Patients with EA2 or FHM1 occasionally present with epilepsy, EA2 being typically associated with absence seizures (Rajakulendran et al., 2010) while FHM1 results in brain edema and focal seizures or status epilepticus following minor head trauma (Chan et al., 2008; Stam et al., 2009). Mechanistically, EA2 is classically associated with CACNA1A LoF mutations (Mantuano et al., 2010; Rajakulendran et al., 2010), while GoF mutations induce FHM1 (Tottene et al., 2009; Pietrobon, 2010), although some exceptions exist (Barrett et al., 2005; Cao and Tsien, 2005). By contrast, triplet expansions result in progressive ataxia (SCA6) with degenerative cerebellar atrophy, mostly due to the accumulation of misfolded protein and ensuing cytotoxic effects, rather than to the loss of CaV2.1 channel function (Mark et al., 2015).

More recently, heterozygous de novo or inherited CACNA1A nonsense mutations (stop gain, frameshift, splice mutations, deletions), causing a presumed LoF, have been associated with developmental epileptic encephalopathies, global developmental delay, limb hypotonia, nystagmus, congenital episodic and progressive ataxia, and cognitive-behavioral impairments ranging from attention deficits and learning disability to frank intellectual disability and/or autism spectrum disorder (Fig. 46–3A; Damaj et al., 2015; Lupien-Meilleur et al., 2021). Other nonepileptic events, such as paroxysmal tonic upward gaze deviation, are frequent and may be confused with focal seizures (Tantsis et al., 2016; Le Roux et al., 2021). Epilepsies associated with nonsense CACNA1A mutations usually start in the first 4 years of life, with absences, generalized tonic-clonic seizures, focal seizures, and/or febrile seizures, with rare instances of infantile spasms (Auvin et al., 2009; Damaj et al., 2015; Le Roux et al., 2021). These epilepsies are accompanied by pathological EEG patterns, including hypsarrhythmia, generalized 3 Hz spike-wave discharges, and slow background with multifocal spikes (Auvin et al., 2009; Damaj et al., 2015; Le Roux et al., 2021). There is a high degree of inter-individual phenotypic variability regarding seizure susceptibility, even with the same mutation. Indeed, in 16 individuals from 4 unrelated families with confirmed CACNA1A LoF variants, epilepsy was observed in 19% of individuals, while isolated febrile seizures occurred in another 40% of individuals (Damaj et al., 2015). Nonsense heterozygous CACNA1A variants otherwise present with isolated neurodevelopmental phenotypes without epilepsy, characterized by global delay, developmental and progressive ataxia, autism spectrum disorder, and/or cognitive impairments (Tonelli et al., 2006; Damaj et al., 2015; Travaglini et al., 2017).

Figure 46–3.. CaV2.

Figure 46–3.

CaV2.1 P/Q-type voltage-gated calcium channels. A. Schematic representation of selected published CACNA1A mutations, color-coded according to their associated phenotype. DEE, developmental epileptic encephalopathy; EA2, episodic ataxia type II; FHM1, (more...)

By contrast, rare cases of recessive biallelic inheritance of a nonsense variant have been described, including four babies with a homozygous variant (p.Arg932*) who presented with early-onset epileptic encephalopathy and premature death within the first 3–6 months of life, and two siblings with compound heterozygous variants (p.Trp1439Arg/p.Ala158Thr fs*6) who developed severe epileptic encephalopathy, progressive neurodegeneration, brain atrophy, optic nerve atrophy, and early lethality (Reinson et al., 2016; Arteche-Lopez et al., 2021). Thus, the homozygous LoF of CACNA1A consistently results in epilepsy, with a severe neurodegeneration phenotype.

Interestingly, some de novo missense mutations have been associated with a drastically more severe phenotype than seen in patients with heterozygous nonsense variants (Fig. 46–3A). These missense variants induce an early-onset developmental epileptic encephalopathy (DEE type 42, MIM# 617106), characterized by severe refractory epilepsy starting in the first few days or months of life, accompanied by profound developmental delay, regression, progressive ataxia and cerebellar atrophy, nystagmus, spasticity, and severe cognitive deficits (moderate to severe intellectual deficiency, severely restricted language abilities) (Myers et al., 2016; Hamdan et al., 2017; Jiang et al., 2019; Le Roux et al., 2021). These patients present with complex seizure disorders, combining myoclonic seizures, epileptic spasms, atypical absences, tonic seizures, tonic-clonic seizures, atonic drop attacks, status epilepticus, and febrile seizures, often reminiscent of Lennox-Gastaut syndrome (Myers et al., 2016; Hamdan et al., 2017; Jiang et al., 2019; Le Roux et al., 2021; Niu et al., 2022). These epilepsies tend to be highly refractory to medical therapies, vagus nerve stimulation, or the ketogenic diet, with some responses to antiepileptic drugs with calcium channel blocking activity (topiramate or valproic acid), and anecdotal reports of benefits from the sodium channel blocker lamotrigine or the L-type VGCC blocker verapamil (Byers et al., 2016; Jiang et al., 2019; Le Roux et al., 2021; Niu et al., 2022). Brain imaging typically reveals progressive brain atrophy, with preferential atrophy of the vermis and cerebellar hemispheres (Jiang et al., 2019; Le Roux et al., 2021), although unilateral cerebral atrophy following prolonged focal status epilepticus has been reported (Niu et al., 2022). Altogether, de novo missense CACNA1A variants underlie ≈1% of cases of severe developmental epileptic encephalopathies in large cohorts of patients, suggesting that it is an important contributor to severe early-onset human epilepsies (Myers et al., 2016; Hamdan et al., 2017). Further, they exacerbate the epilepsy phenotype in patients with SCN1A-related Dravet syndrome (Ohmori et al., 2013).

CACNA1A epilepsy-associated missense mutations are distributed across all domains of the CaV2.1 protein, particularly on the cytoplasmic ends of the S5–S6 segments, on the S4 segment of DIII and IV, and on the DII-III linker loop (Fig. 46–3A), regions involved in voltage-dependent activation and inactivation (Hans et al., 1999; Luvisetto et al., 2004; Catterall et al., 2020). Notably, the DII-DIII cytoplasmic linker loop includes the synaptic protein interaction site (“synprint”), at which various presynaptic proteins involved in vesicular release (SNAP-25 and syntaxin) interact with the α1 subunit, linking CaV2.1 channels to the synaptic release machinery (Rettig et al., 1996). Mutations in this region would thus be expected to impair synaptic release mechanisms.

The functional investigation of four epileptic encephalopathy-associated de novo missense variants revealed that both GoF and dominant-negative LoF mechanisms result in similar clinical phenotypes, with severe early-onset developmental epileptic encephalopathies in the spectrum of Lennox-Gastaut syndrome (Jiang et al., 2019). Two of the studied variants (p.A713T and p.V1396M) resulted in GoF effects, with facilitated voltage-dependent activation (hyperpolarization shift of voltage activation curves), steeper activation curves, a significant delay of channel inactivation, and an increased current density, together with predicted conformational changes on 3D modeling likely mediating these alterations in gating (Fig. 46–3B; Jiang et al., 2019). This was in stark contrast with the effects of a nearby FHM1-associated mutation (p.V714A), characterized by an increased current density and a steeper activation curve, but with intact voltage dependency of activation and inactivation kinetics (Fig. 46–3C; Jiang et al., 2019). Thus, the alteration in channel gating properties likely contributes to the severity of the overall phenotype in patients with GoF mutations causing epileptic encephalopathy compared to those inducing a FHM1 phenotype.

By contrast, two other epileptic encephalopathy- associated de novo missense variants (p.G230V and p.I1357S) resulted in reduced current density, suggesting a LoF effect. These variants also resulted in a mistargeting of the α1 subunit to the plasma membrane, leading to aberrant accumulations in intracytoplasmic inclusions, which may be cytotoxic (Jiang et al., 2019). Notably, when coexpressed with wild-type α1 subunits, mutant α1 subunits sequestered wild-type subunits in these intracellular inclusions, suggesting a dominant- negative (DN) LoF effect (Fig. 46–3D; Jiang et al., 2019). Such DN-LoF mutations are expected to result in greater functional consequences than nonsense heterozygous LoF mutations (haploinsufficiency), since they prevent proper targeting of the wild-type allele to the membrane, resulting in a net decrease of channel expression at the cell surface, while also causing cellular toxicity through pathological accumulation of misfolded proteins, as observed in SCA6-associated mutations (Mark et al., 2015). The cellular and network mechanisms by which both GoD and DN-LoF mutations induce such similar clinical phenotypes will need to be investigated using clinically relevant mouse models expressing epilepsy-associated missense variants.

Furthermore, the characterization of larger patient cohorts may ultimately help identify phenotypic differences between these two functional classes of missense variants, which may ultimately help guide future therapies designed to target specific LoF or GoF mechanisms. For instance, early motor signs, such as tremors, myoclonus, and jitteriness in the first postnatal days, are observed in babies with GoF mutations (Le Roux et al., 2021), likely reflecting enhanced synaptic release at the neuromuscular junctions. Future natural evolution studies may help pinpoint other clinical markers suggesting a GoF or a DN- LoF effects in patients with de novo missense CACNA1A variants.

Physiological Roles of P/Q-Type CaV2.1 Channels

CaV2.1 channels are expressed in most neuronal cell types of the brain, spinal cord, and peripheral nervous system (Westenbroek et al., 1995, 1998; Ludwig et al., 1997), including in a majority of cortical and limbic glutamatergic neurons, as well as in most (but not all) GABAergic interneurons, notably in parvalbumin-positive fast-spiking basket cells and somatostatin-positive interneurons (Rossignol et al., 2013). CaV2.1 channels are most abundant at presynaptic terminals, although low-density staining is observed on the cell soma and dendrites of most cortical, hippocampal and cerebellar neurons (Westenbroek et al., 1995). Multiple brain isoforms have been described (Soong et al., 1993; Woppmann et al., 1994; Westenbroek et al., 1995; Bourinet et al., 1999), with preferential targeting of different isoforms to the soma, dendrites, or presynaptic sites (Rettig et al., 1996; Sakurai et al., 1996).

CaV2.1 channels mediate synaptic release from a range of neuronal cell types (Caddick et al., 1999; Jun et al., 1999; Qian and Noebels, 2000; Cao and Tsien, 2005; Ali and Nelson, 2006; Sasaki et al., 2006; Zaitsev et al., 2007; Kruglikov and Rudy, 2008; Rossignol et al., 2013), and are the principal source of Ca2+ for neurotransmission at most central synapses (see review: Catterall et al., 2013). Other HVA channels, in particular CaV2.2 and CaV2.3 channels, also contribute to synaptic release, and most neurons coexpress multiple HVA VGCCs, often within single release sites (Mintz et al., 1995; Wu et al., 1999; Eltes et al., 2017). Thus, the synaptic impairments due to the loss of CaV2.1 channels can be partially compensated by other VGCCs channels, most notably by N-type CaV2.2 channels, at most synapses (Jun et al., 1999; Qian and Noebels, 2000; Matsushita et al., 2002; Cao and Tsien, 2005; Kruglikov and Rudy, 2008; Rossignol et al., 2013). However, this occurs with variable degrees of efficiency in different cell types, mostly due to the further localization of N-type CaV2.2 channels from the active zone (Sheng et al., 1994; Mintz et al., 1995; Rettig et al., 1996; Wu et al., 1999; Rozov et al., 2001) and to their specific modulation by different intracellular partners and neuromodulators, resulting in different short-term synaptic plasticity rules and adaptation to repeated stimulations (Reyes et al., 1998; Koester and Johnston, 2005; Eltes et al., 2017; Yamamoto and Kobayashi, 2018). Indeed, whereas synaptic release relying on CaV2.1 channels induces short-term synaptic facilitation, the loss of CaV2.1 channels results in short-term synaptic depression or a loss of facilitation (Catterall et al., 2013). This thus leads to greater functional deficits in some neuronal populations, particularly in those discharging at high frequency, such as parvalbumin-positive fast-spiking basket cells, and in afferent populations mediating climbing fiber and parallel fiber excitatory synapses onto Purkinje cells, which may be more sensitive to the immediate availability of presynaptic calcium ions to ensure rapid vesicular release (Pietrobon, 2010; Mark et al., 2011; Rossignol et al., 2013; Jiang et al., 2018).

The exact contributions of CaV2.1 channels to somatic and dendritic physiology, given their expression on the soma and dendritic shafts of many neuronal populations, and thus their potential role in determining neuronal excitability and dendritic responses to inputs, remain to be explored. Nonetheless, the intact intrinsic properties of cortical pyramidal cells and fast-spiking GABAergic interneurons in juvenile mice carrying selective deletions of Cacna1a in cortical pyramidal cells (Rossignol et al., 2013; Bomben et al., 2016) or GABAergic interneurons (Rossignol et al., 2013) suggest that CaV2.1 channels are not required to regulate somatic excitability in these cells and at this age.

Cellular and Circuit Mechanisms of Cacna1a-Associated Epilepsy

Mechanisms of Cacna1a LoF-Associated Epilepsy

Multiple mice lines carrying spontaneous Cacna1a LoF mutations have been described and well characterized over the years, including the tottering Cacna1a  tg/J (Noebels and Sidman, 1979; Fletcher et al., 1996), leaner Cacna1a  tg-la/J (Meier and MacPike, 1971), and Rolling Nogoya mice (Oda, 1973), as well as null mutants (Cacna1a–/–) carrying a targeted gene deletion (Jun et al., 1999). While heterozygous mutants are largely asymptomatic, homozygous LoF and null mutants typically develop gait ataxia and paroxysmal dystonia/dyskinesia, together with generalized spike-wave discharges (SWDs) and absence seizures, starting in the second postnatal week (P10–P14) (Noebels and Sidman, 1979; Hosford et al., 1992; Fletcher et al., 1996; Zwingman et al., 2001; Kodama et al., 2006; Miki et al., 2008). Whereas most homozygous missense mutants survive into adulthood, reflecting partial LoF, homozygous leaner Cacna1ala/la and Cacna1a–/– null mice die from seizures, progressive paralysis, and dehydration before P30 (Fletcher et al., 1996; Jun et al., 1999).

Pathological SWD and absence seizures reflect hypersynchronous discharges within the thalamocortical circuitry (Huguenard and McCormick, 2007; Maheshwari and Noebels, 2014; Crunelli et al., 2020). In this circuit, reticular thalamic nucleus neurons provide GABAergic hyperpolarizing inputs to the thalamic ventrobasal relay neurons, activating LVA VGCCs that trigger burst discharges of relay neurons, which in turn activate the cortex through thalamocortical projections, ultimately resulting in feedback activation of the reticular nucleus (Fig. 46–3E; Huguenard and McCormick, 2007; Crunelli et al., 2020). Notably, reciprocal inhibitory connections within the reticular nucleus desynchronize these oscillations (Huntsman et al., 1999). Thus, minute changes in the excitability or connectivity of reticular nucleus GABAergic neurons, thalamic relay neurons, or corticothalamic neurons may all result in pathological hypersynchrony and contribute to Cacna1a LoF-associated epilepsy. However, the mechanisms by which the homozygous LoF of Cacna1a results in epilepsy, with absences and motor seizures, are complex and involve alterations in multiple circuits and cell types that jointly contribute to epileptogenesis and to the persistence of seizures over time, as summarized below.

First, in homozygous LoF tottering and leaner mice, a functional potentiation of LVA T-type currents in thalamic relay neurons has been proposed to induce pathological thalamocortical oscillations, SWD, and absence seizures (Zhang et al., 2002) (see reviews on thalamocortical synchrony; Huguenard and McCormick, 2007; Maheshwari and Noebels, 2014; Crunelli et al., 2020). Notably, a gain of CaV3.1-mediated T-type currents suffices to induce SWD in CaV3.1 overexpressing mice (Ernst et al., 2009), and CaV3.1 channels are required for the generation of pathological SWD in Cacna1a–/– null mice (Song et al., 2004). The process by which this potentiation takes place is uncertain, but it does not involve LVA channel overexpression (Zhang et al., 2002; Song et al., 2004); this might reflect a lack of desynchronizing inputs from other neuronal populations in the thalamus.

Second, the loss of Cacna1a in the tottering mouse results in synaptic release deficits from glutamatergic ventrobasal thalamic relay neurons, with unaltered GABAergic synaptic events, resulting in a relative imbalance of excitation/inhibition ratio in the thalamus (Caddick et al., 1999). Similar changes occur in the lethargic mutant mice (Cacnb4lh/lh), which lack the α1 interacting site on the associated β4 subunit, leading to reduced Ca2+ influx through CaV2.1 channels, and which display a similar phenotype of epilepsy, ataxia, and dyskinesia (Burgess et al., 1997; Caddick et al., 1999). How this imbalance results in SWD and epilepsy is unclear.

Third, thalamocortical inputs from relay neurons to the somatosensory cortex are functionally impaired in a target-specific manner in the tottering mouse, such that excitatory synapses onto fast-spiking basket cells, which mediate feedforward inhibition of thalamocortical inputs, are reduced in comparison with excitatory inputs onto layer IV pyramidal cells (Sasaki et al., 2006), producing a net aberrant gain of thalamocortical excitation inputs to the somatosensory cortex.

Fourth, in the neocortex, the conditional deletion of Cacna1a in Emx1Cre; Cacna1ac/c mutants reduces synaptic release from cortical layer V pyramidal cells to neighboring pyramidal cells and induces a net reduction of cortical excitability to local extracellular stimulations (Fig. 46–3E, blue box; Rossignol et al., 2013). However, this is insufficient to induce epilepsy in the first 6 weeks of life in Emx1Cre; Cacna1ac/c mutants (Rossignol et al., 2013). By contrast, the selective deletion of Cacna1a in layer VI corticothalamic projection neurons, in Ntsr1Cre; Cacna1a  Loxp/Loxp mutants, suffices to induce SWD and absence seizures, as documented at 8 weeks of life (Bomben et al., 2016). Mechanistically, these absence seizures result from a reduction of synaptic release from glutamatergic corticothalamic neurons onto thalamic relay neurons and reticular nucleus GABAergic neurons, inducing an upregulation of LVA T-type currents in both structures (Fig. 46–3E, green box), and an enhanced bursting behavior of thalamic reticular nucleus neurons (Bomben et al., 2016). This suggests that the involvement of cortical pyramidal cells by Cacna1a LoF may contribute to the emergence of epilepsy in the adult brain. Notably, however, a delayed deletion of Cacna1a at 6 weeks of age in all brain cells, using an inducible Cre line (CAG::Cre-ER; Cacna1a  Loxp/Loxp mice), results in epilepsy, ataxia, and dystonia without enhancing LVA T-type currents or reticular nucleus bursting behavior (Miao et al., 2020), suggesting that other downstream mechanisms may be at play to induce SWD following Cacna1a LoF in the adult brain.

Fifth, the loss of Cacna1a significantly impairs synaptic release from cortical GABAergic fast-spiking parvalbumin-expressing interneurons, as demonstrated in Nkx2.1Cre; Cacna1ac/c carrying a prenatal deletion of Cacna1a in medial ganglionic eminence (MGE)-derived interneurons in the cortex, hippocampus, striatum, and amygdala, but sparing the thalamus (Rossignol et al., 2013). This deletion results in reduced reliability and increased synaptic failures from parvalbumin interneurons to layer V pyramidal cells, despite a partial functional compensation by N-type CaV2.2 VGCCs, resulting in a net deficit of cortical feed-forward inhibition (Fig. 46–3E, red box; Rossignol et al., 2013). Notably, this deletion suffices to induce a severe epilepsy phenotype reminiscent of Lennox-Gastaut syndrome, with absences, tonic-clonic seizures, and tonic seizures starting in the third postnatal week (≈P14–P17) (Rossignol et al., 2013). Furthermore, T-type currents are unchanged in the ventrobasal thalamic relay neurons in these mutants (Rossignol et al., 2013), suggesting that a primary deficit in cortical inhibition suffices to induce a severe and complex generalized epilepsy phenotype reminiscent of the epileptic encephalopathy observed in patients with CACNA1A DN-LOF or homozygous LoF mutations. Notably, this deficit in cortical inhibition is specific to parvalbumin-expressing interneurons, as synaptic release from somatostatin-expressing interneurons is unaffected in SOMCre; Cacna1ac/c mutant mice (Rossignol et al., 2013). Further, the simultaneous deletion of Cacna1a in pyramidal cells and cortical GABAergic interneurons (Nkx2.1Cre; Emx1Cre; Cacna1ac/c) prevents motor seizures, although generalized spike-wave absence seizures and myoclonus are still observed from P17 onward, similar to the time of onset of epilepsy in tottering and Cacna1a–/– mutants (Rossignol et al., 2013). This suggests that motor seizures are triggered by a failure of cortical perisomatic inhibition provided by parvalbumin-expressing interneurons in Cacna1a LoF mutants, but that reduced cortical glutamatergic synaptic release prevents the amplification and spread of motor seizures. Finally, the delayed postnatal deletion of Cacna1a in parvalbumin-expressing GABAergic interneurons (PVCre; Cacna1ac/c) suffices to induce a generalized epilepsy disorder, with frequent absences and rare motor seizures, once again through a prominent disruption of synaptic release from parvalbumin-positive interneurons (Jiang et al., 2018). Interestingly, reduced feed-forward inhibition in PVCre; Cacna1ac/c mutants was partly compensated by a progressive gain of functional inhibitory synapses on the distal dendrites of cortical pyramidal cells, through an exuberant mTOR-dependent sprouting of layer I targeting somatostatin-positive interneurons (presumed Martinotti cells), leading to an increase in cortical di-synaptic and feedback inhibition (Jiang et al., 2018). Notably, blocking this remodeling with rapamycin, an mTORC1 antagonist, results in more frequent tonic-clonic seizures, with comparatively reduced frequency of absence seizures (Jiang et al., 2018). Furthermore, activating cortical somatostatin interneurons selectively using chemogenetic stimulation prevents tonic-clonic motor seizures in a model of kainate-induced seizures (Jiang et al., 2018), supporting the hypothesis that this remodeling of somatostatin interneurons in response to a reduction of perisomatic feed-forward inhibition is clinically relevant and helps reduce motor seizure burden with time. Together, these studies suggest a critical role of reduced cortical perisomatic and feed-forward inhibition in triggering the onset of Cacna1a LoF-associated epilepsy in the early postnatal period, with a progressive homeostatic remodeling of cortical inhibition, resulting in a gain of dendritic inhibition, largely preventing motor seizures with age.

Sixth, a deregulation of cerebellar output also contributes to the emergence or chronicity of SWD and absence seizures in adult Cacna1a LoF mutants. Indeed, the targeted deletion of Cacna1a in Purkinje cells and in sparse populations of forebrain neurons (potentially a subset of GABAergic interneurons), in purky mice (Pcp2Cre; Cacna1ac/c), results in generalized absence seizures, together with ataxia and dystonia, through a reduction of synaptic release and spontaneous activity of Purkinje cells (Mark et al., 2011). Further, a targeted deletion of Cacna1a in rhombic lip-derived neurons, in quirky mice (TgGabra6-Cre; Cacna1ac/c), including cerebellar granule cells and precerebellar nuclei neurons that provide climbing fibers and mossy fibers as glutamatergic inputs to Purkinje cells, causes a disruption of excitatory inputs from parallel fibers to Purkinje cells and induces absence seizures, ataxia, and dyskinesia (Maejima et al., 2013). Thus, Cacna1a LoF reduces the excitability and synaptic release from Purkinje cells, together with reduced inputs from granule cells to Purkinje cells (parallel fibers) (Plomp et al., 2009; Pietrobon, 2010; Mark et al., 2011; Maejima et al., 2013). Further, during SWD in tottering mice, the activity of cerebellar output nuclei is phase-locked to thalamocortical SWD (Kros et al., 2015). The optogenetic activation of deep cerebellar output nuclei providing direct glutamatergic excitation onto thalamocortical projecting neurons in the tottering Cacna1atg/tg mice prevents thalamocortical SWD and absence seizures, largely by desynchronizing thalamocortical oscillations, further suggesting a direct role of cerebellar-thalamic projections in preventing absence seizures (Kros et al., 2015; Eelkman Rooda et al., 2021). Altered cerebellar output in these mice also results in reduced striatal dopaminergic release, causing paroxysmal dyskinesia (Neychev et al., 2008), which may perhaps also contribute to the net disruption of striato-cortical-thalamic oscillations. However, deregulation of cerebellar output might not be critical in the initial phase of the disease, since prenatal recombination of Cacna1a in parvalbumin-expressing Purkinje cells and thalamic reticular nucleus in the PVCre; Cacna1ac/c mutants fails to induce an early seizure phenotype, but rather results in a delayed seizure onset (after P30) only after cortical and hippocampal parvalbumin-expressing neurons have been recombined (Jiang et al., 2018). Furthermore, the combined deletion of Cacna1a in medial ganglionic eminence (MGE)-derived cortical and hippocampal interneurons, and in all parvalbumin-positive populations (including in the reticular nucleus and cerebellum) in Nkx2.1Cre; PVCre; Cacna1ac/c mutants, does not modify the early seizure phenotype compared to that observed in mutants carrying a deletion selectively in MGE-derived interneurons (Nkx2.1Cre; Cacna1ac/c), suggesting that the involvement of cerebellar and reticular thalamic neurons does not significantly contribute to the seizure phenotype in the first few weeks of life (Jiang et al., 2018).

Seventh, the emergence of epilepsy after a postnatal deletion of Cacna1a in all brain cells (Miao et al., 2020) or in targeted cellular populations (Maejima et al., 2013; Mark et al., 2011; Bomben et al., 2016; Jiang et al., 2018) demonstrates a persistent requirement for CaV2.1 channels in the thalamocortical and cerebellar circuitry. However, these models tend to develop prominent absence seizures, with minimal or no motor seizures, which does not fully recapitulate the disease phenotype observed in patients with CACNA1A-related epileptic encephalopathies. By contrast, the embryonic loss of Cacna1a either globally (Jun et al., 1999) or in selected GABAergic populations derived from the MGE is required to reproduce important features of the human disorder (Rossignol et al., 2013), suggesting an important and early developmental implication of GABAergic interneurons in generating the full spectrum of disease manifestations. Of note, however, the C-terminus segment of the Cacna1a gene can be transcribed as a separate product, through a second cistronic transcription initiation site, resulting in the generation of the α1ACT transcription factor (Du et al., 2013). This transcription factor controls the expression of an array of genes involved in brain developmental, and its selective loss alters the development of Purkinje cells and induces cerebellar atrophy, as well as congenital ataxia in mice (Du et al., 2013). Whether this process affects other neuronal cell types involved in epileptogenesis during embryogenesis and early postnatal days remains to be explored.

Finally, the mechanisms underlying the cognitive deficits observed with Cacna1a LoF mutations have also been extensively studied. Although most homozygous LoF mutant mice display significant neurological phenotypes preventing formal cognitive testing (severe ataxia, dyskinesia, epilepsy), heterozygous mutant mice, initially described as being unaffected, have since been revealed to present a range of memory-related impairments. Indeed, aging heterozygous Rolling Nagoya Cacna1atgRol/+ mice were found to develop progressive motor incoordination and spatial learning and memory deficits after 6 months of age (Alonso et al., 2008; Takahashi et al., 2009). Furthermore, heterozygous mutant mice carrying a targeted deletion of Cacna1a in forebrain pyramidal cells, sparing the cerebellum (NexCre; Cacna1ac/+), display late-onset spatial learning and memory deficits, resulting from impaired signaling from Schaffer collaterals to CA1 hippocampal pyramidal cells (Mallmann et al., 2013). However, these models do not recapitulate the full spectrum of cognitive behavioral deficits described in patients with CACNA1A-associated disorders (Damaj et al., 2015). Interestingly, the heterozygous deletion of Cacna1a in parvalbumin-expressing neurons, in PVCre; Cacna1ac/+ mutant mice, results in increased seizure susceptibility to PTZ, reduced anxiety, hyperactivity, and significant deficits in selective attention and cognitive flexibility, without impairments of spatial memory or gross motor function (no ataxia or dyskinesia) (Lupien-Meilleur et al., 2021). Notably, these cognitive behavioral phenotypes largely result from inefficient synaptic transmission from cortical frontal parvalbumin-positive GABAergic interneurons resulting in reduced cortical feed-forward inhibition in the somatosensory, orbitofrontal, and medial prefrontal cortices, similar to the synaptic deficits observed in parvalbumin-positive interneurons of Nkx2.1Cre; Cacna1ac/c mice (Lupien-Meilleur et al., 2021). Importantly, heterozygous Emx1Cre; Cacna1ac/+ mutant mice, carrying a targeted Cacna1a deletion in cortical and limbic pyramidal cells, behaved as wild-types at P60 in all paradigms tested (Lupien-Meilleur et al., 2021). These mutants also displayed unaltered glutamatergic release from cortical pyramidal cells, suggesting that, contrary to the homozygous condition, the heterozygous deletion of Cacna1a is largely a GABAergic interneuron disease in juvenile mice (Lupien-Meilleur et al., 2021). Importantly, the targeted deletion of Cacna1a in the orbitofrontal or the medial prefrontal areas, using AAVCre injections in Cacna1ac/+ mice, induced similar behavioral deficits as those observed in PVCre; Cacna1ac/+ mutants. Indeed, the targeted deletion of Cacna1a in the orbitofrontal cortex impaired cognitive flexibility, while the targeted deletion of Cacna1a in the medial prefrontal cortex resulted in deficits of selective attention (Lupien-Meilleur et al., 2021). Both deficits could be rescued in adult PVCre; Cacna1ac/+ mutants by the targeted chemogenetic activation of parvalbumin-positive interneurons in these specific cortical areas, suggesting that some of the cognitive behavioral deficits in Cacna1a LoF mutants are reversible, even after symptom onset (Lupien-Meilleur et al., 2021). Thus, there appears to be a critical role for the dysfunction of cortical GABAergic parvalbumin-expressing interneurons in driving both the early onset of epilepsy, including the expression of motor seizures, as well as the cognitive deficits associated with Cacna1a LoF mutations. Nonetheless, whether similar deficits exist in constitutive heterozygous LoF mutants and whether targeted therapies designed to enhance cortical inhibition could reverse these deficits in vivo remains to be demonstrated.

Mechanisms of Cacna1a GoF-Associated Epilepsies

CACNA1A GoF mutations typically result in episodic hemiplegic migraines (FHM1), sometimes accompanied by episodes of brain edema, strokes, and seizures following minor head trauma (Chan et al., 2008; Stam et al., 2009). Homozygous knock-in mice for FHM1 mutations (Cacna1aR192Q and Cacna1aS218L mice) display enhanced glutamatergic synaptic transmission but intact cortical inhibition, resulting in aberrant waves of cortical spreading depression (CSD) (Eikermann-Haerter et al., 2009; Tottene et al., 2009; Vecchia et al., 2014, 2015). Notably, homozygous Cacna1aR192Q mice appear phenotypically normal, unless exposed to specific CSD triggers (van den Maagdenberg et al., 2004). By contrast, Cacna1aS218L mice develop spontaneous migraine-like episodes with episodic hemiplegia, while a minority of mutants also develops rare, but severe, tonic-clonic seizures (van den Maagdenberg et al., 2010). Notably, heterozygous mutants of either allele are phenotypically normal. Given our recent finding that some de novo GoF mutations result in severe early onset epileptic encephalopathies with profound intellectual deficiency and ataxia, in the spectrum of Lennox-Gastaut syndrome (LGS) (Jiang et al., 2019), it appears that current GoF models do not fully recapitulate the spectrum of diseases and will not be sufficient to test novel therapies. Further, the network mechanisms by which specific Cacna1a GoF mutations result in such a severe phenotype compared to the well-described FHM1 phenotype are unknown. Importantly, there are currently no efficient therapies for patients with this severe disorder, often resulting in premature lethality. Thus, better models are urgently needed to clarify the disease mechanisms and to test novel therapies for epilepsy-associated Cacna1a GoF mutations.

R-Type CaV2.3 Channels: CACNA1E-Associated Epilepsy

CACNA1E-Associated Disorders in Humans

CaV2.3 channels are expressed broadly in the brain (Soong et al., 1993; Williams et al., 1994; Day et al., 1996; de Borman et al., 1999; Lin et al., 1999), as well as in the peripheral nervous system, cardiac myocytes, sperm, kidneys, pancreas, and gastrointestinal tract (Vajna et al., 1998; Grabsch et al., 1999; Weiergraber et al., 2000; Mitchell et al., 2002). At least six isoforms have been described, with different age-dependent, tissue-specific, and cell-type-specific differences in expression patterns (Pereverzev et al., 1998, 2002; Grabsch et al., 1999; Mitchell et al., 2002). CACNA1E-associated disorders described to date present with prominent neurological and musculoskeletal symptoms, without frank systemic manifestations, potentially reflecting the tissue-selective roles of specific isoforms.

De novo missense CACNA1E variants have been associated with a severe early-onset developmental and epileptic encephalopathy (DEE 69, MIM618285), characterized by neonatal or early infantile seizure onset, with profound global developmental delay, severe axial hypotonia, spastic quadriplegia, congenital joint contractures, movement disorders, and macrocephaly (Helbig et al., 2018). Hyperkinetic movement disorders, including dystonia and chorea, appear to be prominent early signs of the disease (Helbig et al., 2018; Ortiz Cabrera et al., 2021). Although the majority of children with de novo CACNA1E missense variants develop epilepsy (described in 45%–87% of patients) (Helbig et al., 2018; Heyne et al., 2018), others display a range of neurodevelopmental deficits, including global delay, developmental regression, autism spectrum disorder, and/or intellectual disability without epilepsy (Helbig et al., 2018; Heyne et al., 2018; Royer-Bertrand et al., 2021). There is considerable inter-individual phenotypic variability, even with the same mutations (Royer-Bertrand et al., 2021). Notably, nearly 40% of nonepileptic patients with de novo CACNA1E missense variants display pathological epileptiform activity on EEG, potentially reflecting a greater susceptibility to seizures (Royer-Bertrand et al., 2021). CACNA1E-associated epilepsies are complex seizure disorders, often starting with epileptic spasms, with subsequent combinations of myoclonic, tonic, atonic, focal motor, or focal with impaired consciousness and generalized tonic-clonic seizures (Helbig et al., 2018). EEG recordings typically show severely perturbed brain rhythms, with multifocal spike discharges, slow spike-wave discharges, hypsarrhythmia, and/or continuous spike and wave in slow-wave sleep (CSWS) (Helbig et al., 2018). Brain imaging is often unremarkable or reveals nonspecific brain atrophy (Helbig et al., 2018).

Mechanistically, the majority of published de novo CACNA1E missense variants are localized on the cytoplasmic portion of the S6 segments in domains I–IV (Fig. 46–4A), regions involved in channel activation (Raybaud et al., 2006, 2007; Royer-Bertrand et al., 2021), together with the DII S4–S5 linker (Wall-Lacelle et al., 2011). Most described variants induce GoF effects on CaV2.3 channels, through combinations of facilitated voltage-dependent activation (hyperpolarization shift of voltage activation curves), slower fast inactivation, and increased current density (Fig. 46–4B; Helbig et al., 2018). These findings are clinically relevant as 50% of patients with CACNA1E-related epilepsy display partial or full seizure control with topiramate (Helbig et al., 2018), a known CaV2.3 channel blocker (Kuzmiski et al., 2005). De novo LoF mutations (frameshift or indel) have also been described in patients with epilepsy, global developmental delay, and major speech impairments, although the overall phenotype is milder than that caused by GoF missense variants (Helbig et al., 2018). Nonetheless, in large control populations (GnomAD database), CACNA1E haploinsufficiency appears to be poorly tolerated (pLI score = 1.0) (Lek et al., 2016). Thus, as in CACNA1A-associated disorders, both GoF and LoF variants in CACNA1E have been associated with epilepsy, although GoF variants induce more severe disorders.

Figure 46–4.. CaV2.

Figure 46–4.

CaV2.3 R-type voltage-gated calcium channels. A. Schematic representation of selected published CACNA1E mutations, color-coded according to their associated phenotype. DEE, developmental epileptic encephalopathy; NDD, neurodevelopmental disorders, including (more...)

Mechanisms of CACNA1E-Related Epilepsies: CaV2.3-Mediated R-Type Currents Regulate Firing Mode, Plateau Potentials, and Afterhyperpolarization

CaV2.3 channels are expressed broadly in the brain, including in the cortex, hippocampus, amygdala, hypothalamus, basal ganglia, substantia nigra, cerebellum, and brainstem (Soong et al., 1993; Williams et al., 1994; Day et al., 1996; Parajuli et al., 2012). Expression in the thalamus has been debated, likely reflecting methodological differences, with some investigators reporting a lack of expression in the thalamus (de Borman et al., 1999; Lin et al., 1999; Parajuli et al., 2012), while others observed the expression of multiple different isoforms in the thalamus, including in the reticular nucleus (Weiergraber et al., 2006). CaV2.3 channels are found mostly on neuronal dendrites and spines (Bloodgood and Sabatini, 2007; Parajuli et al., 2012), although presynaptic localization has been observed in the interpeduncular nucleus (Parajuli et al., 2012), and somatic expression in nigrostriatal dopaminergic neurons (Benkert et al., 2019).

Human CaV2.3 channels display typical HVA VGCC characteristics: they are rapidly activated upon depolarizations, with a threshold for activation around +5 mV (Williams et al., 1994; Pereverzev et al., 1998). They display rapid inactivation (Williams et al., 1994; Pereverzev et al., 1998), with some isoform-specific differences (Pereverzev et al., 2002). Rodent CaV2.3 channels share similar characteristics, but they activate at more negative membrane potentials (–50 to –10 mV in heterologous systems) and require significant hyperpolarization to deinactivate, giving them unique properties to contribute to neuronal burst discharges (Soong et al., 1993) (although this may reflect methodological difference rather than species-selective differences; Williams et al., 1994). Voltage-dependent channel activation results from displacement of the S4 segment and conformational changes involving the cytoplasmic ends of DI-IV S6 segments and the DII S4–S5 linker (Fig. 46–4A) (Raybaud et al., 2006, 2007, Wall-Lacelle et al., 2011). Fast voltage-dependent inactivation of CaV2.3 channels occurs through global conformational changes that implicate multiple domains of the protein (Spaetgens and Zamponi, 1999), with a critical role of the I-II linker that includes the β-subunit binding site (AID) (Berrou et al., 2001). The clustering of epilepsy-associated CACNA1E mutations in the cytoplasmic end of DI-IV S6 segments and in the I-II linker further supports the critical importance of these regions in channel function and gating.

CaV2.3 channels are notoriously resistant to most VGCC blockers, including DHP, ω-conotoxin GVIA, ω-conotoxin MVIIC, and ω-agatoxin IVA (Catterall et al., 2005). However, they are sensitive to nickel, cadmium (Schneider et al., 1994; Williams et al., 1994; Vajna et al., 2001), and the tarantula venom toxin SNX-482 (Newcomb et al., 1998). They are also sensitive to a subset of antiepileptic drugs, including topiramate (Kuzmiski et al., 2005) and lamotrigine (Hainsworth et al., 2003), which may be relevant in CACNA1E GoF associated disorders.

Insights from various rodent models illustrate the fundamental roles of CaV2.3 channels in hippocampal and neocortical circuits in health and disease. In particular, CaV2.3 channels mediate a large proportion of R-type currents in neocortical, hippocampal, and dopaminergic neurons (Sochivko et al., 2002; Benkert et al., 2019). Other neuronal populations (cerebellar granule cells and dorsal root ganglion cells) rely on a combination of CaV2.3 and other VGCCs to generate R-type currents, with different populations showing differential sensitivity to the tarantula venom toxin SNX-482 (Piedras-Renteria and Tsien, 1998; Wilson et al., 2000). CaV2.3 channels contribute to neurotransmitter release at different central synapses, including at hippocampal glutamatergic synapses (Gasparini et al., 2001; Dietrich et al., 2003) and brainstem neurons (Wu et al., 1998), likely mediating the ≈5%–20% non P/Q- or N-type VGCC-mediated synaptic release from various cell types (Mintz et al., 1995; Wu and Saggau, 1995; Wu et al., 1998; Qian and Noebels, 2000; Sochivko et al., 2002). They are less efficiently coupled to neurotransmitter release than CaV2.1 P/Q-type channels, reflecting their more distant location from the active zone (Wu et al., 1999). Nonetheless, they participate in release mechanisms involving slower calcium signaling, such as LTP and post-tetanic potentiation, largely by facilitating local accumulation of Ca2+ ions at the presynaptic terminal (Dietrich et al., 2003).

Further, at postsynaptic sites in neocortical and hippocampal neurons, R-type currents contribute to the generation of large sustained dendritic calcium transients, the plateau potentials, which shift the spiking mode from single spike to burst-firing mode following synchronized inputs from multiple sources (Schiller et al., 1997; Larkum et al., 1999, 2001; Magee and Carruth, 1999; Waters et al., 2003). Such plateau potentials are observed in CA1 principal cell dendritic tufts following the coactivation of Schaffer collaterals (CA3 inputs) and the perforant pathways (entorhinal inputs), resulting in burst firing and LTP at perforant path synapses, important for coincidence detection and context-dependent memory retrieval (Breustedt et al., 2003; Takahashi and Magee, 2009). Similar dendritic plateau potentials, resembling dendritic depolarizations observed during epileptiform discharges, can be elicited by muscarinic activation of CA1 neurons with carbachol (Fraser and MacVicar, 1996; Kuzmiski et al., 2005). These R-type VGCC-dependent events enhance neuronal excitability and drive repetitive firing (theta bursts) in CA1 principal cells (Fig. 46–4C; Tai et al., 2006). Importantly, muscarinic activation is known to specifically enhance CaV2.3-mediated R-type currents in hippocampal CA1 neurons (Meza et al., 1999; Melliti et al., 2000; Bannister et al., 2004), but not T-type currents (Tai et al., 2006). Thus, R-type CaV2.3 VGCCs are likely involved in ictogenesis following muscarinic activation, for instance, following pilocarpine injections, by enhancing plateau potentials in pyramidal cell dendrites, triggering burst-firing modes. Notably, KO mice for the muscarinic M1 receptor show reduced seizure susceptibility to pilocarpine (Hamilton et al., 1997). Whether this translates into seizure resistance in Cacna1e KO mice, or enhanced seizures susceptibility in future mutant mice carrying epilepsy-related Cacna1e GoF mutations, remains to be explored.

Furthermore, CaV2.3 channels regulate neuronal firing patterns through their functional coupling with a range of K+ channels that mediate post-spike membrane repolarization and afterhyperpolarization (AHP). In particular, in CA1 pyramidal neurons, CaV2.3 channels help constrain excitatory postsynaptic potentials by activating dendritic small-conductance Ca2+-activated potassium (SK) channels (Bloodgood and Sabatini, 2007; Giessel and Sabatini, 2011). They also drive action-potential repolarization and short-term synaptic plasticity by activating big conductance K+ channels (BK) that mediate spike-induced afterhyperpolarization (AHP) (Gutzmann et al., 2019). Furthermore, they control the surface expression of pore-forming KV4.2 K+ channels, which mediate Ca2+-induced and voltage-dependent transient K+ currents (IA) that regulate interspike intervals, dendritic excitability, and the propagation of synaptic inputs along hippocampal principal cell dendrites (Murphy et al., 2022). Notably, blocking K+ currents in the distal dendrites of CA1 neurons with 4AP allows for the back-propagation of action potentials following somatic inputs, which induces regenerative somatic afterdepolarizations (ADPs) that trigger burst firing, in a process sensitive to R-type VGCC antagonists (Fig. 46–4D) (Magee and Carruth, 1999). Thus, mutations affecting the functional coupling between CaV2.3 channels and dendritic K+ channels, or mutations causing significant GoF of CaV2.3 channels by shifting the activation potential or slowing channel inactivation, are expected to promote repeated burst firing and epileptiform discharges in hippocampal CA1 neurons. Nonetheless, the exact network mechanisms by which GoF mutations in CACNA1E result in epilepsy will need to be further explored using appropriate animal model expressing clinically relevant GoF mutations.

Finally, CaV2.3 channels mediate somatic Ca2+ entry and calcium oscillations in spontaneously bursting neuronal populations, such as substantia nigra dopaminergic neurons (Benkert et al., 2019). Such Ca2+ influx in spontaneously bursting cells is typically offset by mitochondrial Ca2+ uptake. However, during aging, these metabolically demanding CaV2.3-dependent Ca2+ oscillations result in progressive somatic Ca2+ overload and excitotoxicity, as seen in Parkinson disease and in rodent models of the disease (Benkert et al., 2019). Notably, such neurodegeneration following neurotoxic exposure is prevented in Cacna1e KO mice (Benkert et al., 2019). Interestingly, similar CaV2.3-dependent mechanisms have been proposed to mediate neuronal cell death following status epilepticus. Indeed, Cacna1e KO mice display a marked resistance to kainic acid–induced excitotoxicity and apoptosis in hippocampal CA3 neurons following status epilepticus (Weiergraber et al., 2007). In the context of epilepsy, a primary GoF of CaV2.3 channel activity might result in similar neurodegeneration over time, particularly following prolonged seizures. This will need to be tested experimentally in future animal models expressing clinically relevant GoF mutations. Notably, the expression of a mutant form of EFHC1, a CaV2.3-binding partner associated with juvenile myoclonic epilepsy, results in massive apoptosis of hippocampal neurons in vitro, which can be prevented by blocking CaV2.3 channels using SNX-482 (Suzuki et al., 2004). Thus, blocking CaV2.3 channels may be of therapeutic value in specific circumstances in which a primary GoF of CaV2.3 channels occurs. Although SNX-482 is not approved for clinical use, given expected systemic side effects, Food and Drug Administration (FDA)-approved antiseizure medications such as lamotrigine or topiramate with known CaV2.3 antagonism may offer interesting translational alternatives. This will require testing in appropriate preclinical models of CACNA1E-associated epilepsy with GoF mutations.

By contrast, the mechanisms by which CACNA1E LoF results in epilepsy are uncertain. One may speculate that the loss of CaV2.3 expression, and the ensuing reduced expression of KV4.2 expression and AI current, would result in enhanced regenerative somatic ADPs and burst firing in hippocampal neurons. However, this LoF would also likely decrease the frequency of plateau potentials, thus reducing the sensitivity to specific ictogenic triggers. Interestingly, data from Cacna1e KO mice carrying a targeted deletion of exon 2 reveal a striking resistance to PTZ-, kainate- and NMDA-induced tonic-clonic seizures, and seizure-induced lethality, although with a relative facilitation of non-motor seizures (reduced latency to Racine stage 2–3 seizures) (Weiergraber et al., 2006, 2007). By contrast, sensitivity to 4AP remains unchanged in these mutants, likely reflecting the different ictogenic mechanisms involved (Weiergraber et al., 2006). Overall, these data suggest that CaV2.3 channels are critical regulators of seizure initiation and propagation (Weiergraber et al., 2007), and that their loss might actually be protective against specific types of seizures, particularly motor seizures. Interestingly, a progressive reduction of CaV2.3 channels expression has been observed in aging models of generalized epilepsy (GAERS and Wistar rats), suggesting a potential homeostatic compensation mechanism (de Borman et al., 1999), although this was not observed in another genetic model of absence epilepsy (lethargic Cacnb4  lh/lh mice) (Lin et al., 1999). The exact physiological correlates of these findings will need to be further explored, with specific investigation of plateau potentials and PDS-like events in neocortical and hippocampal neurons in Cacna1e KO mice. In addition, the differential impact of the loss of CaV2.3 channels in other cellular populations (GABAergic interneurons, nRT neurons, glia) and other circuits not fully assessed in Cacna1a KO mice to date should be further explored. This is particularly relevant given the known expression of Cav2.3 channels in subpopulations of cortical and hippocampal GABAergic interneurons (van de Bovenkamp-Janssen et al., 2004; Weiergraber et al., 2006), and the implication of a primary dysfunction of GABAergic interneurons in early developmental epileptic encephalopathies (see reviews, Rossignol, 2011; Jiang et al., 2016).

Furthermore, the mechanisms by which LoF of Cav2.3 channels induce the significant cognitive comorbidities of CACNA1E-associated epilepsies remain unclear. LoF of Cav2.3 channels would be expected to induce memory impairments through deficits in activity-dependent synaptic plasticity and LTP, given the central role of these channels in these biological processes (Breustedt et al., 2003; Takahashi and Magee, 2009). Indeed, Cacna1e KO mice carrying a targeted deletion of exon 1 display anxiety and reduced spatial memory (Kubota et al., 2001). However, surprisingly, these mice displayed intact theta-burst induced LTP in CA1 neurons, suggesting some degree of functional compensation by other VGCCs in this model (Kubota et al., 2001). Thus, the exact mechanisms by which CACNA1E LoF results in cognitive impairments remain to be explored.

Importantly, the targeting strategy appears to be critical when studying the functional impact of Cacna1e LoF on different neuronal populations, given the existence of multiple tissue- and cell-type-specific isoforms (Piedras-Renteria and Tsien, 1998; Tottene et al., 2000). Indeed, Cacna1e–/– KO mice carrying a targeted intragenic deletion of the DII S2–S6 segments are phenotypically intact, despite the confirmed absence of transcript I mRNA (Wilson et al., 2000). These KO mice display a loss of SNX-482-sensitive R-type currents, but a preservation of SNX-482 insensitive R-type currents in cerebellar and DRG neurons (Wilson et al., 2000), suggesting that the targeting strategy spared other isoforms mediating SNX-482 insensitive currents, or that other VGCCs mediate R-type currents in the neuronal population studied. The specific isoforms expressed in circuits involved in epilepsy and cognition will thus need to be further explored to clarify the underlying disease mechanisms. For instance, the CaV2.3c splice variant, containing exon 19 insert 1 in the cytosolic DII-III linker loop, corresponding to the major splice isoform in hippocampal and neocortical neurons, determines the degree of Ca2+-dependent channel modulation and may alter the susceptibility of these cell types to status epilepticus–induced calcium excitotoxicity (Weiergraber et al., 2006, 2007).

Finally, systemic effects of Cacna1e LoF have been described in various KO models, including impaired stress-induced glucose homeostasis (Pereverzev et al., 2002) and cardiac arrhythmias (Lu et al., 2004) in exon 2-deleted KO mice; as well as decreased responses to inflammatory visceral pain in exon 1-deleted KO mice (Saegusa et al., 2000), which may be clinically relevant for various other systemic effects of CACNA1E LoF mutations.

Conclusions

In summary, HVA VGCCs are essential for a range of physiological processes regulating dendritic and somatic excitability, burst-firing mode, synaptic release, and structural plasticity, required for optimal brain function. They are also critical for multiple fundamental neurodevelopmental processes, including neuronal proliferation, migration, dendritic development, and structural plasticity. Thus, mutations affecting HVA calcium channels result in multiple neurological disorders, including various types of severe early-onset developmental epileptic encephalopathies. Novel cellular and animal models are shedding light on the underlying pathological mechanisms. Furthermore, new models carrying patient-derived mutations will be required to explore these disease mechanisms and to test novel therapies.

Acknowledgments

I am grateful to M. Lavertu-Jolin for illustrations and figure design, with help from A. Lupien-Meilleur, K. Toulouse, and P. K. Raju. I am also grateful to J. Noebels for his timely and critical input on the manuscript. E. Rossignol holds a Canada Research Chair (CRC-II) on the Neurobiology of Epilepsy. She is also the recipient of three active Canadian Institute for Health Research (CIHR) operating grants, including one to support research on the mechanisms of CACNA1A-related epilepsies.

References

  1. Aicardi G, Schwartzkroin PA.  Suppression of epileptiform burst discharges in CA3 neurons of rat hippocampal slices by the organic calcium channel blocker, verapamil.  Exp Brain Res.  1990;81(2):288–96. [PubMed: 2168841]
  2. Ali AB, Nelson C.  Distinct Ca2+ channels mediate transmitter release at excitatory synapses displaying different dynamic properties in rat neocortex.  Cereb Cortex.  2006;16(3):386–93. [PubMed: 15917483]
  3. Allene C, Cattani A, Ackman JB, Bonifazi P, Aniksztejn L, Ben-Ari Y, et al. Sequential generation of two distinct synapse-driven network patterns in developing neocortex.  J Neurosci.  2008;28(48):12851–63. [PMC free article: PMC6671804] [PubMed: 19036979]
  4. Alonso I, Marques JM, Sousa N, Sequeiros J, Olsson IA, Silveira I.  Motor and cognitive deficits in the heterozygous leaner mouse, a Cav2.1 voltage-gated Ca2+ channel mutant.  Neurobiol Aging.  2008;29(11):1733–43. [PubMed: 17513018]
  5. Amano H, Amano T, Matsubayashi H, Ishihara K, Serikawa T, Sasa M.  Enhanced calcium influx in hippocampal CA3 neurons of spontaneously epileptic rats.  Epilepsia.  2001;42(3):345–50. [PubMed: 11442151]
  6. Amitai Y, Friedman A, Connors BW, Gutnick MJ.  Regenerative activity in apical dendrites of pyramidal cells in neocortex.  Cereb Cortex.  1993;3(1):26–38. [PubMed: 8439739]
  7. Antzelevitch C, Pollevick GD, Cordeiro JM, Casis O, Sanguinetti MC, Aizawa Y, et al. Loss-of-function mutations in the cardiac calcium channel underlie a new clinical entity characterized by ST-segment elevation, short QT intervals, and sudden cardiac death. Circulation.  2007;115(4):442–9. [PMC free article: PMC1952683] [PubMed: 17224476]
  8. Arteche-Lopez A, Alvarez-Mora MI, Sanchez Calvin MT, Lezana Rosales JM, Palma Milla C, Gomez Rodriguez MJ, et al. Biallelic variants in genes previously associated with dominant inheritance: CACNA1A, RET and SLC20A2.  Eur J Hum Genet.  2021;29(10):1520–6. [PMC free article: PMC8484357] [PubMed: 34267336]
  9. Asadi-Pooya AA, Razavizadegan SM, Abdi-Ardekani A, Sperling MR.  Adjunctive use of verapamil in patients with refractory temporal lobe epilepsy: a pilot study.  Epilepsy Behav.  2013;29(1):150–4. [PubMed: 23973639]
  10. Auvin S, Holder-Espinasse M, Lamblin MD, Andrieux J.  Array-CGH detection of a de novo 0.7-Mb deletion in 19p13.13 including CACNA1A associated with mental retardation and epilepsy with infantile spasms.  Epilepsia.  2009;50(11):2501–3. [PubMed: 19874387]
  11. Bader PL, Faizi M, Kim LH, Owen SF, Tadross MR, Alfa RW, et al. Mouse model of Timothy syndrome recapitulates triad of autistic traits.  Proc Natl Acad Sci U S A.  2011;108(37):15432–7. [PMC free article: PMC3174658] [PubMed: 21878566]
  12. Baig SM, Koschak A, Lieb A, Gebhart M, Dafinger C, Nurnberg G, et al. Loss of Ca(v)1.3 (CACNA1D) function in a human channelopathy with bradycardia and congenital deafness.  Nat Neurosci.  2011;14(1):77–84. [PubMed: 21131953]
  13. Bando Y, Irie K, Shimomura T, Umeshima H, Kushida Y, Kengaku M, et al. Control of Spontaneous Ca2+ Transients Is Critical for Neuronal Maturation in the Developing Neocortex.  Cereb Cortex.  2016;26(1):106–17. [PubMed: 25112282]
  14. Bannister RA, Melliti K, Adams BA.  Differential modulation of CaV2.3 Ca2+ channels by Galphaq/11-coupled muscarinic receptors.  Mol Pharmacol.  2004;65(2):381–8. [PubMed: 14742680]
  15. Barrett CF, Cao YQ, Tsien RW.  Gating deficiency in a familial hemiplegic migraine type 1 mutant P/Q-type calcium channel.  J Biol Chem.  2005;280(25):24064–71. [PubMed: 15795222]
  16. Barrett CF, Tsien RW.  The Timothy syndrome mutation differentially affects voltage- and calcium-dependent inactivation of CaV1.2 L-type calcium channels.  Proc Natl Acad Sci U S A.  2008;105(6):2157–62. [PMC free article: PMC2538892] [PubMed: 18250309]
  17. Beck H, Steffens R, Elger CE, Heinemann U.  Voltage-dependent Ca2+ currents in epilepsy.  Epilepsy Res.  1998;32(1-2):321–32. [PubMed: 9761331]
  18. Ben-Ari Y, Cherubini E, Corradetti R, Gaiarsa JL.  Giant synaptic potentials in immature rat CA3 hippocampal neurones.  J Physiol.  1989;416:303–25. [PMC free article: PMC1189216] [PubMed: 2575165]
  19. Benardo LS, Masukawa LM, Prince DA.  Electrophysiology of isolated hippocampal pyramidal dendrites.  J Neurosci.  1982;2(11):1614–22. [PMC free article: PMC6564359] [PubMed: 6292377]
  20. Benkert J, Hess S, Roy S, Beccano-Kelly D, Wiederspohn N, Duda J, et al. Cav2.3 channels contribute to dopaminergic neuron loss in a model of Parkinson’s disease.  Nat Commun.  2019;10(1):5094. [PMC free article: PMC6841684] [PubMed: 31704946]
  21. Berrou L, Bernatchez G, Parent L.  Molecular determinants of inactivation within the I-II linker of alpha1E (CaV2.3) calcium channels.  Biophys J.  2001;80(1):215–28. [PMC free article: PMC1301227] [PubMed: 11159396]
  22. Bidaud I, Lory P.  Hallmarks of the channelopathies associated with L-type calcium channels: a focus on the Timothy mutations in Ca(v)1.2 channels.  Biochimie.  2011;93(12):2080–6. [PubMed: 21664226]
  23. Birey F, Andersen J, Makinson CD, Islam S, Wei W, Huber N, et al. Assembly of functionally integrated human forebrain spheroids.  Nature.  2017;545(7652):54–9. [PMC free article: PMC5805137] [PubMed: 28445465]
  24. Birey F, Li MY, Gordon A, Thete MV, Valencia AM, Revah O, et al. Dissecting the molecular basis of human interneuron migration in forebrain assembloids from Timothy syndrome.  Cell Stem Cell.  2022;29(2):248–64 e7. [PubMed: 34990580]
  25. Blalock EM, Chen KC, Vanaman TC, Landfield PW, Slevin JT.  Epilepsy-induced decrease of L-type Ca2+ channel activity and coordinate regulation of subunit mRNA in single neurons of rat hippocampal ‘zipper’ slices.  Epilepsy Res.  2001;43(3):211–26. [PubMed: 11248533]
  26. Bloodgood BL, Sabatini BL.  Nonlinear regulation of unitary synaptic signals by CaV(2.3) voltage-sensitive calcium channels located in dendritic spines.  Neuron.  2007;53(2):249–60. [PubMed: 17224406]
  27. Bomben VC, Aiba I, Qian J, Mark MD, Herlitze S, Noebels JL.  Isolated P/Q Calcium Channel Deletion in Layer VI Corticothalamic Neurons Generates Absence Epilepsy.  J Neurosci.  2016;36(2):405–18. [PMC free article: PMC4710767] [PubMed: 26758833]
  28. Borlot F, Wither RG, Ali A, Wu N, Verocai F, Andrade DM.  A pilot double-blind trial using verapamil as adjuvant therapy for refractory seizures.  Epilepsy Res.  2014;108(9):1642–51. [PubMed: 25223728]
  29. Bortone D, Polleux F.  KCC2 expression promotes the termination of cortical interneuron migration in a voltage-sensitive calcium-dependent manner.  Neuron.  2009;62(1):53–71. [PMC free article: PMC3314167] [PubMed: 19376067]
  30. Bourinet E, Soong TW, Sutton K, Slaymaker S, Mathews E, Monteil A, et al. Splicing of alpha 1A subunit gene generates phenotypic variants of P- and Q-type calcium channels.  Nat Neurosci.  1999;2(5):407–15. [PubMed: 10321243]
  31. Bozarth X, Dines JN, Cong Q, Mirzaa GM, Foss K, Lawrence Merritt J, 2nd, et al. Expanding clinical phenotype in CACNA1C related disorders: From neonatal onset severe epileptic encephalopathy to late-onset epilepsy.  Am J Med Genet A.  2018;176(12):2733–9. [PMC free article: PMC6312477] [PubMed: 30513141]
  32. Breustedt J, Vogt KE, Miller RJ, Nicoll RA, Schmitz D.  Alpha1E-containing Ca2+ channels are involved in synaptic plasticity.  Proc Natl Acad Sci U S A.  2003;100(21):12450–5. [PMC free article: PMC218778] [PubMed: 14519849]
  33. Brugnoni R, Canioni E, Filosto M, Pini A, Tonin P, Rossi T, et al. Mutations associated with hypokalemic periodic paralysis: from hotspot regions to complete analysis of CACNA1S and SCN4A genes.  Neurogenetics.  2021. [PubMed: 34608571]
  34. 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(3):385–92. [PubMed: 9039265]
  35. Byers HM, Beatty CW, Hahn SH, Gospe SM, Jr. Dramatic Response After Lamotrigine in a Patient With Epileptic Encephalopathy and a De NovoCACNA1A Variant.  Pediatr Neurol.  2016;60:79–82. [PubMed: 27212419]
  36. Caddick SJ, Wang C, Fletcher CF, Jenkins NA, Copeland NG, Hosford DA. Excitatory but not inhibitory synaptic transmission is reduced in lethargic (Cacnb4(lh)) and tottering (Cacna1atg) mouse thalami.  J Neurophysiol.  1999;81(5):2066–74. [PubMed: 10322048]
  37. Cain SM, Snutch TP.  Voltage-Gated Calcium Channels in Epilepsy. In: th, Noebels JL, Avoli M, Rogawski MA, Olsen RW, Delgado-Escueta AV, editors. Jasper’s Basic Mechanisms of the Epilepsies. Bethesda (MD); 2012. [PubMed: 22787663]
  38. Calorio C, Gavello D, Guarina L, Salio C, Sassoe-Pognetto M, Riganti C, et al. Impaired chromaffin cell excitability and exocytosis in autistic Timothy syndrome TS2-neo mouse rescued by L-type calcium channel blockers.  J Physiol.  2019;597(6):1705–33. [PMC free article: PMC6418779] [PubMed: 30629744]
  39. Campiglio M, Flucher BE.  The role of auxiliary subunits for the functional diversity of voltage-gated calcium channels.  J Cell Physiol.  2015;230(9):2019–31. [PMC free article: PMC4672716] [PubMed: 25820299]
  40. Cao YQ, Tsien RW.  Effects of familial hemiplegic migraine type 1 mutations on neuronal P/Q-type Ca2+ channel activity and inhibitory synaptic transmission.  Proc Natl Acad Sci U S A.  2005;102(7):2590–5. [PMC free article: PMC548328] [PubMed: 15699344]
  41. Carpenter D, Ringrose C, Leo V, Morris A, Robinson RL, Halsall PJ, et al. The role of CACNA1S in predisposition to malignant hyperthermia.  BMC Med Genet.  2009;10:104. [PMC free article: PMC2770053] [PubMed: 19825159]
  42. Catterall WA.  Structure and regulation of voltage-gated Ca2+ channels.  Annu Rev Cell Dev Biol.  2000;16:521–55. [PubMed: 11031246]
  43. Catterall WA, Few AP. Calcium channel regulation and presynaptic plasticity.  Neuron.  2008;59(6):882–901. [PubMed: 18817729]
  44. Catterall WA, Leal K, Nanou E.  Calcium channels and short-term synaptic plasticity.  J Biol Chem.  2013;288(15):10742–9. [PMC free article: PMC3624454] [PubMed: 23400776]
  45. Catterall WA, Lenaeus MJ, Gamal El-Din TM.  Structure and Pharmacology of Voltage-Gated Sodium and Calcium Channels.  Annu Rev Pharmacol Toxicol.  2020;60:133–54. [PubMed: 31537174]
  46. Catterall WA, Perez-Reyes E, Snutch TP, Striessnig J. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels.  Pharmacol Rev.  2005;57(4):411–25. [PubMed: 16382099]
  47. Catterall WA, Wisedchaisri G, Zheng N.  The chemical basis for electrical signaling.  Nat Chem Biol.  2017;13(5):455–63. [PMC free article: PMC5464002] [PubMed: 28406893]
  48. Cavus I, Teyler T.  Two forms of long-term potentiation in area CA1 activate different signal transduction cascades.  J Neurophysiol.  1996;76(5):3038–47. [PubMed: 8930253]
  49. Cavus I, Teyler TJ.  NMDA receptor-independent LTP in basal versus apical dendrites of CA1 pyramidal cells in rat hippocampal slice.  Hippocampus.  1998;8(4):373–9. [PubMed: 9744422]
  50. Chan YC, Burgunder JM, Wilder-Smith E, Chew SE, Lam-Mok-Sing KM, Sharma V, et al. Electroencephalographic changes and seizures in familial hemiplegic migraine patients with the CACNA1A gene S218L mutation.  J Clin Neurosci.  2008;15(8):891–4. [PubMed: 18313928]
  51. Chavis P, Ango F, Michel JM, Bockaert J, Fagni L.  Modulation of big K+ channel activity by ryanodine receptors and L-type Ca2+ channels in neurons.  Eur J Neurosci.  1998;10(7):2322–7. [PubMed: 9749760]
  52. Chavis P, Fagni L, Lansman JB, Bockaert J.  Functional coupling between ryanodine receptors and L-type calcium channels in neurons.  Nature.  1996;382(6593):719–22. [PubMed: 8751443]
  53. Cheli VT, Santiago Gonzalez DA, Zamora NN, Lama TN, Spreuer V, Rasmusson RL, et al. Enhanced oligodendrocyte maturation and myelination in a mouse model of Timothy syndrome.  Glia.  2018;66(11):2324–39. [PMC free article: PMC6697123] [PubMed: 30151840]
  54. Chen S, Yue C, Yaari Y.  A transitional period of Ca2+-dependent spike afterdepolarization and bursting in developing rat CA1 pyramidal cells.  J Physiol.  2005;567(Pt 1):79–93. [PMC free article: PMC1474172] [PubMed: 15919718]
  55. Chin H, Smith MA, Kim HL, Kim H.  Expression of dihydropyridine-sensitive brain calcium channels in the rat central nervous system.  FEBS Lett.  1992;299(1):69–74. [PubMed: 1312035]
  56. Christie BR, Eliot LS, Ito K, Miyakawa H, Johnston D.  Different Ca2+ channels in soma and dendrites of hippocampal pyramidal neurons mediate spike-induced Ca2+ influx.  J Neurophysiol.  1995;73(6):2553–7. [PubMed: 7666160]
  57. Clark MB, Wrzesinski T, Garcia AB, Hall NAL, Kleinman JE, Hyde T, et al. Long-read sequencing reveals the complex splicing profile of the psychiatric risk gene CACNA1C in human brain.  Mol Psychiatry.  2020;25(1):37–47. [PMC free article: PMC6906184] [PubMed: 31695164]
  58. Clark NC, Nagano N, Kuenzi FM, Jarolimek W, Huber I, Walter D, et al. Neurological phenotype and synaptic function in mice lacking the CaV1.3 alpha subunit of neuronal L-type voltage-dependent Ca2+ channels.  Neuroscience.  2003;120(2):435–42. [PubMed: 12890513]
  59. Cossart R, Khazipov R.  How development sculpts hippocampal circuits and function.  Physiol Rev.  2022;102(1):343–78. [PubMed: 34280053]
  60. Crepel V, Aronov D, Jorquera I, Represa A, Ben-Ari Y, Cossart R.  A parturition-associated nonsynaptic coherent activity pattern in the developing hippocampus.  Neuron.  2007;54(1):105–20. [PubMed: 17408581]
  61. Crunelli V, Lorincz ML, McCafferty C, Lambert RC, Leresche N, Di Giovanni G, et al. Clinical and experimental insight into pathophysiology, comorbidity and therapy of absence seizures.  Brain.  2020;143(8):2341–68. [PMC free article: PMC7447525] [PubMed: 32437558]
  62. Curtis BM, Catterall WA.  Purification of the calcium antagonist receptor of the voltage-sensitive calcium channel from skeletal muscle transverse tubules.  Biochemistry.  1984;23(10):2113–8. [PubMed: 6329263]
  63. Damaj L, Lupien-Meilleur A, Lortie A, Riou E, Ospina LH, Gagnon L, et al. CACNA1A haploinsufficiency causes cognitive impairment, autism and epileptic encephalopathy with mild cerebellar symptoms.  Eur J Hum Genet.  2015;23(11):1505–12. [PMC free article: PMC4613477] [PubMed: 25735478]
  64. Davies NP, Eunson LH, Samuel M, Hanna MG.  Sodium channel gene mutations in hypokalemic periodic paralysis: an uncommon cause in the UK.  Neurology.  2001;57(7):1323–5. [PubMed: 11591859]
  65. Day M, Wang Z, Ding J, An X, Ingham CA, Shering AF, et al. Selective elimination of glutamatergic synapses on striatopallidal neurons in Parkinson disease models.  Nat Neurosci.  2006;9(2):251–9. [PubMed: 16415865]
  66. Day NC, Shaw PJ, McCormack AL, Craig PJ, Smith W, Beattie R, et al. Distribution of alpha 1A, alpha 1B and alpha 1E voltage-dependent calcium channel subunits in the human hippocampus and parahippocampal gyrus.  Neuroscience.  1996;71(4):1013–24. [PubMed: 8684604]
  67. de Borman B, Lakaye B, Minet A, Zorzi W, Vergnes M, Marescaux C, et al. Expression of mRNA encoding alpha1E and alpha1G subunit in the brain of a rat model of absence epilepsy.  Neuroreport.  1999;10(3):569–74. [PubMed: 10208591]
  68. de Curtis M, Avanzini G.  Interictal spikes in focal epileptogenesis.  Prog Neurobiol.  2001;63(5):541–67. [PubMed: 11164621]
  69. Deisseroth K, Singla S, Toda H, Monje M, Palmer TD, Malenka RC.  Excitation-neurogenesis coupling in adult neural stem/progenitor cells.  Neuron.  2004;42(4):535–52. [PubMed: 15157417]
  70. Dembla E, Dembla M, Maxeiner S, Schmitz F.  Synaptic ribbons foster active zone stability and illumination-dependent active zone enrichment of RIM2 and Cav1.4 in photoreceptor synapses.  Scientific reports.  2020;10(1):5957. [PMC free article: PMC7136232] [PubMed: 32249787]
  71. Dick IE, Joshi-Mukherjee R, Yang W, Yue DT.  Arrhythmogenesis in Timothy Syndrome is associated with defects in Ca(2+)-dependent inactivation.  Nat Commun.  2016;7:10370. [PMC free article: PMC4740114] [PubMed: 26822303]
  72. Diebold RJ, Koch WJ, Ellinor PT, Wang JJ, Muthuchamy M, Wieczorek DF, et al. Mutually exclusive exon splicing of the cardiac calcium channel alpha 1 subunit gene generates developmentally regulated isoforms in the rat heart. Proc Natl Acad Sci U S A.  1992;89(4):1497–501. [PMC free article: PMC48478] [PubMed: 1311102]
  73. Dietrich D, Kirschstein T, Kukley M, Pereverzev A, von der Brelie C, Schneider T, et al. Functional specialization of presynaptic Cav2.3 Ca2+ channels.  Neuron.  2003;39(3):483–96. [PubMed: 12895422]
  74. Dilger EK, Shin HS, Guido W.  Requirements for synaptically evoked plateau potentials in relay cells of the dorsal lateral geniculate nucleus of the mouse.  J Physiol.  2011;589(Pt 4):919–37. [PMC free article: PMC3060370] [PubMed: 21173075]
  75. Dolmetsch RE, Pajvani U, Fife K, Spotts JM, Greenberg ME.  Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway.  Science.  2001;294(5541):333–9. [PubMed: 11598293]
  76. Dolphin AC. Calcium channel auxiliary alpha2delta and beta subunits: trafficking and one step beyond.  Nat Rev Neurosci.  2012;13(8):542–55. [PubMed: 22805911]
  77. Du X, Wang J, Zhu H, Rinaldo L, Lamar KM, Palmenberg AC, et al. Second cistron in CACNA1A gene encodes a transcription factor mediating cerebellar development and SCA6.  Cell.  2013;154(1):118–33. [PMC free article: PMC3939801] [PubMed: 23827678]
  78. Eelkman Rooda OHJ, Kros L, Faneyte SJ, Holland PJ, Gornati SV, Poelman HJ, et al. Single-pulse stimulation of cerebellar nuclei stops epileptic thalamic activity.  Brain Stimul.  2021;14(4):861–72. [PubMed: 34022430]
  79. Egorov AV, Hamam BN, Fransen E, Hasselmo ME, Alonso AA.  Graded persistent activity in entorhinal cortex neurons.  Nature.  2002;420(6912):173–8. [PubMed: 12432392]
  80. Eikermann-Haerter K, Dilekoz E, Kudo C, Savitz SI, Waeber C, Baum MJ, et al. Genetic and hormonal factors modulate spreading depression and transient hemiparesis in mouse models of familial hemiplegic migraine type 1.  J Clin Invest.  2009;119(1):99–109. [PMC free article: PMC2613474] [PubMed: 19104150]
  81. Eltes T, Kirizs T, Nusser Z, Holderith N.  Target Cell Type-Dependent Differences in Ca(2+) Channel Function Underlie Distinct Release Probabilities at Hippocampal Glutamatergic Terminals.  J Neurosci.  2017;37(7):1910–24. [PMC free article: PMC6589975] [PubMed: 28115484]
  82. Empson RM, Jefferys JG. Ca(2+) entry through L-type Ca(2+) channels helps terminate epileptiform activity by activation of a Ca(2+) dependent afterhyperpolarisation in hippocampal CA3.  Neuroscience.  2001;102(2):297–306. [PubMed: 11166116]
  83. Ernst WL, Noebels JL.  Expanded alternative splice isoform profiling of the mouse Cav3.1/alpha1G T-type calcium channel.  BMC Mol Biol.  2009;10:53. [PMC free article: PMC2696442] [PubMed: 19480703]
  84. Ernst WL, Zhang Y, Yoo JW, Ernst SJ, Noebels JL.  Genetic enhancement of thalamocortical network activity by elevating alpha 1g-mediated low-voltage-activated calcium current induces pure absence epilepsy.  J Neurosci.  2009;29(6):1615–25. [PMC free article: PMC2660673] [PubMed: 19211869]
  85. Ertel EA, Campbell KP, Harpold MM, Hofmann F, Mori Y, Perez-Reyes E, et al. Nomenclature of voltage-gated calcium channels.  Neuron.  2000;25(3):533–5. [PubMed: 10774722]
  86. Falcon-Moya R, Perez-Rodriguez M, Prius-Mengual J, Andrade-Talavera Y, Arroyo-Garcia LE, Perez-Artes R, et al. Astrocyte-mediated switch in spike timing-dependent plasticity during hippocampal development.  Nat Commun.  2020;11(1):4388. [PMC free article: PMC7463247] [PubMed: 32873805]
  87. Feldmann M, Koepp M.  P-glycoprotein imaging in temporal lobe epilepsy: in vivo PET experiments with the Pgp substrate [11C]-verapamil.  Epilepsia.  2012;53 Suppl 6:60–3. [PubMed: 23134497]
  88. Firth HV, Wright CF, Study DDD. The Deciphering Developmental Disorders (DDD) study.  Dev Med Child Neurol.  2011;53(8):702–3. [PubMed: 21679367]
  89. Fletcher CF, Lutz CM, O’Sullivan TN, Shaughnessy JD, Jr., Hawkes R, Frankel WN, et al. Absence epilepsy in tottering mutant mice is associated with calcium channel defects.  Cell.  1996;87(4):607–17. [PubMed: 8929530]
  90. Flucher BE.  Skeletal muscle CaV1.1 channelopathies.  Pflugers Arch.  2020;472(7):739–54. [PMC free article: PMC7351834] [PubMed: 32222817]
  91. Franklin WH, Laubham M.  Neurologic complications of genetic channelopathies.  Handb Clin Neurol.  2021;177:185–8. [PubMed: 33632437]
  92. Fraser DD, MacVicar BA.  Cholinergic-dependent plateau potential in hippocampal CA1 pyramidal neurons.  J Neurosci.  1996;16(13):4113–28. [PMC free article: PMC6578995] [PubMed: 8753873]
  93. Garaschuk O, Hanse E, Konnerth A.  Developmental profile and synaptic origin of early network oscillations in the CA1 region of rat neonatal hippocampus.  J Physiol.  1998;507(Pt 1):219–36. [PMC free article: PMC2230780] [PubMed: 9490842]
  94. Garaschuk O, Linn J, Eilers J, Konnerth A.  Large-scale oscillatory calcium waves in the immature cortex.  Nat Neurosci.  2000;3(5):452–9. [PubMed: 10769384]
  95. Gasparini S, Kasyanov AM, Pietrobon D, Voronin LL, Cherubini E.  Presynaptic R-type calcium channels contribute to fast excitatory synaptic transmission in the rat hippocampus.  J Neurosci.  2001;21(22):8715–21. [PMC free article: PMC6762258] [PubMed: 11698583]
  96. Ghosh A, Carnahan J, Greenberg ME.  Requirement for BDNF in activity-dependent survival of cortical neurons.  Science.  1994;263(5153):1618–23. [PubMed: 7907431]
  97. Giessel AJ, Sabatini BL.  Boosting of synaptic potentials and spine Ca transients by the peptide toxin SNX-482 requires alpha-1E-encoded voltage-gated Ca channels.  PLoS One.  2011;6(6):e20939. [PMC free article: PMC3111456] [PubMed: 21695265]
  98. Gomez-Gonzalo M, Losi G, Chiavegato A, Zonta M, Cammarota M, Brondi M, et al. An excitatory loop with astrocytes contributes to drive neurons to seizure threshold.  PLoS Biol.  2010;8(4):e1000352. [PMC free article: PMC2854117] [PubMed: 20405049]
  99. Grabsch H, Pereverzev A, Weiergraber M, Schramm M, Henry M, Vajna R, et al. Immunohistochemical detection of alpha1E voltage-gated Ca(2+) channel isoforms in cerebellum, INS-1 cells, and neuroendocrine cells of the digestive system.  J Histochem Cytochem.  1999;47(8):981–94. [PubMed: 10424882]
  100. Greenberg ME, Thompson MA, Sheng M.  Calcium regulation of immediate early gene transcription.  J Physiol Paris.  1992;86(1-3):99–108. [PubMed: 1343600]
  101. Grover LM, Teyler TJ.  Two components of long-term potentiation induced by different patterns of afferent activation.  Nature.  1990;347(6292):477–9. [PubMed: 1977084]
  102. Grover LM, Teyler TJ.  N-methyl-D-aspartate receptor-independent long-term potentiation in area CA1 of rat hippocampus: input-specific induction and preclusion in a non-tetanized pathway.  Neuroscience.  1992;49(1):7–11. [PubMed: 1357588]
  103. Grubb MS, Burrone J. Activity-dependent relocation of the axon initial segment fine-tunes neuronal excitability.  Nature.  2010;465(7301):1070–4. [PMC free article: PMC3196626] [PubMed: 20543823]
  104. Gu X, Spitzer NC.  Breaking the code: regulation of neuronal differentiation by spontaneous calcium transients.  Dev Neurosci.  1997;19(1):33–41. [PubMed: 9078431]
  105. Gutzmann JJ, Lin L, Hoffman DA.  Functional Coupling of Cav2.3 and BK Potassium Channels Regulates Action Potential Repolarization and Short-Term Plasticity in the Mouse Hippocampus.  Front Cell Neurosci.  2019;13:27. [PMC free article: PMC6393364] [PubMed: 30846929]
  106. Guzman JN, Sanchez-Padilla J, Chan CS, Surmeier DJ.  Robust pacemaking in substantia nigra dopaminergic neurons.  J Neurosci.  2009;29(35):11011–9. [PMC free article: PMC2784968] [PubMed: 19726659]
  107. Haeseleer F, Imanishi Y, Maeda T, Possin DE, Maeda A, Lee A, et al. Essential role of Ca2+-binding protein 4, a Cav1.4 channel regulator, in photoreceptor synaptic function.  Nat Neurosci.  2004;7(10):1079–87. [PMC free article: PMC1352161] [PubMed: 15452577]
  108. Hainsworth AH, McNaughton NC, Pereverzev A, Schneider T, Randall AD.  Actions of sipatrigine, 202W92 and lamotrigine on R-type and T-type Ca2+ channel currents.  Eur J Pharmacol.  2003;467(1-3):77–80. [PubMed: 12706458]
  109. Hamdan FF, Myers CT, Cossette P, Lemay P, Spiegelman D, Laporte AD, et al. High Rate of Recurrent De Novo Mutations in Developmental and Epileptic Encephalopathies.  Am J Hum Genet.  2017;101(5):664–85. [PMC free article: PMC5673604] [PubMed: 29100083]
  110. Hamilton SE, Loose MD, Qi M, Levey AI, Hille B, McKnight GS, et al. Disruption of the m1 receptor gene ablates muscarinic receptor-dependent M current regulation and seizure activity in mice.  Proc Natl Acad Sci U S A.  1997;94(24):13311–6. [PMC free article: PMC24305] [PubMed: 9371842]
  111. Hans M, Luvisetto S, Williams ME, Spagnolo M, Urrutia A, Tottene A, et al. Functional consequences of mutations in the human alpha1A calcium channel subunit linked to familial hemiplegic migraine.  J Neurosci.  1999;19(5):1610–9. [PMC free article: PMC6782159] [PubMed: 10024348]
  112. Hasan M, Pulman J, Marson AG.  Calcium antagonists as an add-on therapy for drug-resistant epilepsy.  Cochrane Database Syst Rev.  2013(3):CD002750. [PMC free article: PMC7100543] [PubMed: 23543516]
  113. Helbig KL, Lauerer RJ, Bahr JC, Souza IA, Myers CT, Uysal B, et al. De Novo Pathogenic Variants in CACNA1E Cause Developmental and Epileptic Encephalopathy with Contractures, Macrocephaly, and Dyskinesias.  Am J Hum Genet.  2018;103(5):666–78. [PMC free article: PMC6216110] [PubMed: 30343943]
  114. Hell JW, Westenbroek RE, Warner C, Ahlijanian MK, Prystay W, Gilbert MM, et al. Identification and differential subcellular localization of the neuronal class C and class D L-type calcium channel alpha 1 subunits.  J Cell Biol.  1993;123(4):949–62. [PMC free article: PMC2200142] [PubMed: 8227151]
  115. Heyne HO, Singh T, Stamberger H, Abou Jamra R, Caglayan H, Craiu D, et al. De novo variants in neurodevelopmental disorders with epilepsy.  Nat Genet.  2018;50(7):1048–53. [PubMed: 29942082]
  116. Hirtz JJ, Boesen M, Braun N, Deitmer JW, Kramer F, Lohr C, et al. Cav1.3 calcium channels are required for normal development of the auditory brainstem.  J Neurosci.  2011;31(22):8280–94. [PMC free article: PMC6622878] [PubMed: 21632949]
  117. Hirtz JJ, Braun N, Griesemer D, Hannes C, Janz K, Lohrke S, et al. Synaptic refinement of an inhibitory topographic map in the auditory brainstem requires functional Cav1.3 calcium channels.  J Neurosci.  2012;32(42):14602–16. [PMC free article: PMC6621425] [PubMed: 23077046]
  118. Ho MT, Pelkey KA, Pelletier JG, Huganir RL, Lacaille JC, McBain CJ.  Burst firing induces postsynaptic LTD at developing mossy fibre-CA3 pyramid synapses.  J Physiol.  2009;587(Pt 18):4441–54. [PMC free article: PMC2766649] [PubMed: 19635819]
  119. Hofer NT, Tuluc P, Ortner NJ, Nikonishyna YV, Fernandes-Quintero ML, Liedl KR, et al. Biophysical classification of a CACNA1D de novo mutation as a high-risk mutation for a severe neurodevelopmental disorder.  Mol Autism.  2020;11(1):4. [PMC free article: PMC6950833] [PubMed: 31921405]
  120. Horigane SI, Hamada S, Kamijo S, Yamada H, Yamasaki M, Watanabe M, et al. Development of an L-type Ca(2+) channel-dependent Ca(2+) transient during the radial migration of cortical excitatory neurons.  Neurosci Res.  2021;169:17–26. [PubMed: 32598973]
  121. Horigane SI, Ozawa Y, Zhang J, Todoroki H, Miao P, Haijima A, et al. A mouse model of Timothy syndrome exhibits altered social competitive dominance and inhibitory neuron development.  FEBS Open Bio.  2020;10(8):1436–46. [PMC free article: PMC7396430] [PubMed: 32598571]
  122. Hosford DA, Clark S, Cao Z, Wilson WA, Jr., Lin FH, Morrisett RA, et al. The role of GABAB receptor activation in absence seizures of lethargic (lh/lh) mice.  Science.  1992;257(5068):398–401. [PubMed: 1321503]
  123. Hosseini SM, Kim R, Udupa S, Costain G, Jobling R, Liston E, et al. Reappraisal of Reported Genes for Sudden Arrhythmic Death: Evidence-Based Evaluation of Gene Validity for Brugada Syndrome.  Circulation.  2018;138(12):1195–205. [PMC free article: PMC6147087] [PubMed: 29959160]
  124. Huguenard JR, McCormick DA.  Thalamic synchrony and dynamic regulation of global forebrain oscillations.  Trends Neurosci.  2007;30(7):350–6. [PubMed: 17544519]
  125. Huntsman MM, Porcello DM, Homanics GE, DeLorey TM, Huguenard JR.  Reciprocal inhibitory connections and network synchrony in the mammalian thalamus.  Science.  1999;283(5401):541–3. [PubMed: 9915702]
  126. Iannetti P, Parisi P, Spalice A, Ruggieri M, Zara F.  Addition of verapamil in the treatment of severe myoclonic epilepsy in infancy.  Epilepsy Res.  2009;85(1):89–95. [PubMed: 19303743]
  127. Iossifov I, O’Roak BJ, Sanders SJ, Ronemus M, Krumm N, Levy D, et al. The contribution of de novo coding mutations to autism spectrum disorder.  Nature.  2014;515(7526):216–21. [PMC free article: PMC4313871] [PubMed: 25363768]
  128. Iossifov I, Ronemus M, Levy D, Wang Z, Hakker I, Rosenbaum J, et al. De novo gene disruptions in children on the autistic spectrum.  Neuron.  2012;74(2):285–99. [PMC free article: PMC3619976] [PubMed: 22542183]
  129. Ishihara K, Sasa M, Momiyama T, Ujihara H, Nakamura J, Serikawa T, et al. Abnormal excitability of hippocampal CA3 pyramidal neurons of spontaneously epileptic rats (SER), a double mutant.  Exp Neurol.  1993;119(2):287–90. [PubMed: 8432367]
  130. Jiang M, Swann JW.  A role for L-type calcium channels in the maturation of parvalbumin-containing hippocampal interneurons.  Neuroscience.  2005;135(3):839–50. [PubMed: 16154277]
  131. Jiang X, Lachance M, Rossignol E.  Involvement of cortical fast-spiking parvalbumin-positive basket cells in epilepsy.  Prog Brain Res.  2016;226:81–126. [PMC free article: PMC5218834] [PubMed: 27323940]
  132. Jiang X, Lupien-Meilleur A, Tazerart S, Lachance M, Samarova E, Araya R, et al. Remodeled cortical inhibition prevents motor seizures in generalized epilepsy.  Ann Neurol.  2018;84(3):436–51. [PubMed: 30048010]
  133. Jiang X, Raju PK, D’Avanzo N, Lachance M, Pepin J, Dubeau F, et al. Both gain-of-function and loss-of-function de novo CACNA1A mutations cause severe developmental epileptic encephalopathies in the spectrum of Lennox-Gastaut syndrome.  Epilepsia.  2019;60(9):1881–94. [PubMed: 31468518]
  134. Jun K, Piedras-Renteria ES, Smith SM, Wheeler DB, Lee SB, Lee TG, et al. Ablation of P/Q-type Ca(2+) channel currents, altered synaptic transmission, and progressive ataxia in mice lacking the alpha(1A)-subunit.  Proc Natl Acad Sci U S A.  1999;96(26):15245–50. [PMC free article: PMC24805] [PubMed: 10611370]
  135. Kabir ZD, Che A, Fischer DK, Rice RC, Rizzo BK, Byrne M, et al. Rescue of impaired sociability and anxiety-like behavior in adult cacna1c-deficient mice by pharmacologically targeting eIF2alpha.  Mol Psychiatry.  2017;22(8):1096–109. [PMC free article: PMC5863913] [PubMed: 28584287]
  136. Kamijo S, Ishii Y, Horigane SI, Suzuki K, Ohkura M, Nakai J, et al. A Critical Neurodevelopmental Role for L-Type Voltage-Gated Calcium Channels in Neurite Extension and Radial Migration.  J Neurosci.  2018;38(24):5551–66. [PMC free article: PMC8174135] [PubMed: 29773754]
  137. Kang H, Schuman EM. Intracellular Ca(2+) signaling is required for neurotrophin-induced potentiation in the adult rat hippocampus.  Neurosci Lett.  2000;282(3):141–4. [PubMed: 10717411]
  138. Kanyshkova T, Ehling P, Cerina M, Meuth P, Zobeiri M, Meuth SG, et al. Regionally specific expression of high-voltage-activated calcium channels in thalamic nuclei of epileptic and non-epileptic rats.  Mol Cell Neurosci.  2014;61:110–22. [PubMed: 24914823]
  139. Kater SB, Mattson MP, Cohan C, Connor J.  Calcium regulation of the neuronal growth cone.  Trends Neurosci.  1988;11(7):315–21. [PubMed: 2465636]
  140. Kato K, Clifford DB, Zorumski CF.  Long-term potentiation during whole-cell recording in rat hippocampal slices.  Neuroscience.  1993;53(1):39–47. [PubMed: 8097020]
  141. Kelly KM, Ikonomovic MD, Abrahamson EE, Kharlamov EA, Hentosz TM, Armstrong DM.  Alterations in hippocampal voltage-gated calcium channel alpha 1 subunit expression patterns after kainate-induced status epilepticus in aging rats.  Epilepsy Res.  2003;57(1):15–32. [PubMed: 14706730]
  142. Khan I, Ahmad S, Thomas N.  Differential translation of the alpha-1 isoforms of L-type calcium channel in rat brain and other tissues.  Biochem Mol Biol Int.  1998;45(5):895–904. [PubMed: 9739454]
  143. Khazipov R, Sirota A, Leinekugel X, Holmes GL, Ben-Ari Y, Buzsaki G.  Early motor activity drives spindle bursts in the developing somatosensory cortex.  Nature.  2004;432(7018):758–61. [PubMed: 15592414]
  144. Kim HL, Kim H, Lee P, King RG, Chin H.  Rat brain expresses an alternatively spliced form of the dihydropyridine-sensitive L-type calcium channel alpha 2 subunit.  Proc Natl Acad Sci U S A.  1992;89(8):3251–5. [PMC free article: PMC48844] [PubMed: 1314383]
  145. Kim JB, Kim MH, Lee SJ, Kim DJ, Lee BC.  The genotype and clinical phenotype of Korean patients with familial hypokalemic periodic paralysis.  J Korean Med Sci.  2007;22(6):946–51. [PMC free article: PMC2694642] [PubMed: 18162704]
  146. Kim SH, Park YR, Lee B, Choi B, Kim H, Kim CH.  Reduction of Cav1.3 channels in dorsal hippocampus impairs the development of dentate gyrus newborn neurons and hippocampal-dependent memory tasks.  PLoS One.  2017;12(7):e0181138. [PMC free article: PMC5513490] [PubMed: 28715454]
  147. Kodama T, Itsukaichi-Nishida Y, Fukazawa Y, Wakamori M, Miyata M, Molnar E, et al. A CaV2.1 calcium channel mutation rocker reduces the number of postsynaptic AMPA receptors in parallel fiber-Purkinje cell synapses.  Eur J Neurosci.  2006;24(11):2993–3007. [PubMed: 17156361]
  148. Koester HJ, Johnston D.  Target cell-dependent normalization of transmitter release at neocortical synapses.  Science.  2005;308(5723):863–6. [PubMed: 15774725]
  149. Kohr G, Mody I.  Endogenous intracellular calcium buffering and the activation/inactivation of HVA calcium currents in rat dentate gyrus granule cells.  J Gen Physiol.  1991;98(5):941–67. [PMC free article: PMC2229099] [PubMed: 1662686]
  150. Korkotian E, Segal M.  Release of calcium from stores alters the morphology of dendritic spines in cultured hippocampal neurons.  Proc Natl Acad Sci U S A.  1999;96(21):12068–72. [PMC free article: PMC18413] [PubMed: 10518577]
  151. Koschak A, Fernandez-Quintero ML, Heigl T, Ruzza M, Seitter H, Zanetti L.  Cav1.4 dysfunction and congenital stationary night blindness type 2.  Pflugers Arch.  2021;473(9):1437–54. [PMC free article: PMC8370969] [PubMed: 34212239]
  152. Koschak A, Reimer D, Huber I, Grabner M, Glossmann H, Engel J, et al. alpha 1D (Cav1.3) subunits can form l-type Ca2+ channels activating at negative voltages.  J Biol Chem.  2001;276(25):22100–6. [PubMed: 11285265]
  153. Krey JF, Pasca SP, Shcheglovitov A, Yazawa M, Schwemberger R, Rasmusson R, et al. Timothy syndrome is associated with activity-dependent dendritic retraction in rodent and human neurons.  Nat Neurosci.  2013;16(2):201–9. [PMC free article: PMC3568452] [PubMed: 23313911]
  154. Kros L, Eelkman Rooda OH, Spanke JK, Alva P, van Dongen MN, Karapatis A, et al. Cerebellar output controls generalized spike-and-wave discharge occurrence.  Ann Neurol.  2015;77(6):1027–49. [PMC free article: PMC5008217] [PubMed: 25762286]
  155. Kruglikov I, Rudy B.  Perisomatic GABA release and thalamocortical integration onto neocortical excitatory cells are regulated by neuromodulators.  Neuron.  2008;58(6):911–24. [PMC free article: PMC2572574] [PubMed: 18579081]
  156. Kubista H, Boehm S, Hotka M.  The Paroxysmal Depolarization Shift: Reconsidering Its Role in Epilepsy, Epileptogenesis and Beyond.  Int J Mol Sci.  2019;20(3). [PMC free article: PMC6387313] [PubMed: 30699993]
  157. Kubota M, Murakoshi T, Saegusa H, Kazuno A, Zong S, Hu Q, et al. Intact LTP and fear memory but impaired spatial memory in mice lacking Ca(v)2.3 (alpha(IE)) channel.  Biochem Biophys Res Commun.  2001;282(1):242–8. [PubMed: 11263998]
  158. Kuzmiski JB, Barr W, Zamponi GW, MacVicar BA.  Topiramate inhibits the initiation of plateau potentials in CA1 neurons by depressing R-type calcium channels.  Epilepsia.  2005;46(4):481–9. [PubMed: 15816941]
  159. Lacinova L.  Voltage-dependent calcium channels.  Gen Physiol Biophys.  2005;24 Suppl 1:1–78. [PubMed: 16096350]
  160. Lancaster B, Nicoll RA.  Properties of two calcium-activated hyperpolarizations in rat hippocampal neurones.  J Physiol.  1987;389:187–203. [PMC free article: PMC1192077] [PubMed: 2445972]
  161. Lancaster B, Nicoll RA, Perkel DJ.  Calcium activates two types of potassium channels in rat hippocampal neurons in culture.  J Neurosci.  1991;11(1):23–30. [PMC free article: PMC6575188] [PubMed: 1986065]
  162. Larkum ME, Zhu JJ, Sakmann B.  A new cellular mechanism for coupling inputs arriving at different cortical layers.  Nature.  1999;398(6725):338–41. [PubMed: 10192334]
  163. Larkum ME, Zhu JJ, Sakmann B.  Dendritic mechanisms underlying the coupling of the dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal neurons.  J Physiol.  2001;533(Pt 2):447–66. [PMC free article: PMC2278642] [PubMed: 11389204]
  164. Le Roux M, Barth M, Gueden S, Desbordes de Cepoy P, Aeby A, Vilain C, et al. CACNA1A-associated epilepsy: Electroclinical findings and treatment response on seizures in 18 patients.  Eur J Paediatr Neurol.  2021;33:75–85. [PubMed: 34102571]
  165. Lee AS, De Jesus-Cortes H, Kabir ZD, Knobbe W, Orr M, Burgdorf C, et al. The Neuropsychiatric Disease-Associated Gene cacna1c Mediates Survival of Young Hippocampal Neurons.  eNeuro.  2016;3(2). [PMC free article: PMC4819284] [PubMed: 27066530]
  166. Lee AS, Gonzales KL, Lee A, Moosmang S, Hofmann F, Pieper AA, et al. Selective genetic deletion of cacna1c in the mouse prefrontal cortex.  Mol Psychiatry.  2012;17(11):1051. [PMC free article: PMC5837808] [PubMed: 23096954]
  167. Lee AS, Ra S, Rajadhyaksha AM, Britt JK, De Jesus-Cortes H, Gonzales KL, et al. Forebrain elimination of cacna1c mediates anxiety-like behavior in mice.  Mol Psychiatry.  2012;17(11):1054–5. [PMC free article: PMC3481072] [PubMed: 22665262]
  168. Leinekugel X, Khazipov R, Cannon R, Hirase H, Ben-Ari Y, Buzsaki G. Correlated bursts of activity in the neonatal hippocampus in vivo. Science.  2002;296(5575):2049–52. [PubMed: 12065842]
  169. Leinekugel X, Medina I, Khalilov I, Ben-Ari Y, Khazipov R.  Ca2+ oscillations mediated by the synergistic excitatory actions of GABA(A) and NMDA receptors in the neonatal hippocampus.  Neuron.  1997;18(2):243–55. [PubMed: 9052795]
  170. Leinekugel X, Tseeb V, Ben-Ari Y, Bregestovski P.  Synaptic GABAA activation induces Ca2+ rise in pyramidal cells and interneurons from rat neonatal hippocampal slices.  J Physiol.  1995;487(Pt 2):319–29. [PMC free article: PMC1156575] [PubMed: 8558466]
  171. Lek M, Karczewski KJ, Minikel EV, Samocha KE, Banks E, Fennell T, et al. Analysis of protein-coding genetic variation in 60,706 humans.  Nature.  2016;536(7616):285–91. [PMC free article: PMC5018207] [PubMed: 27535533]
  172. Le Roux M, Barth M, Gueden S, Desbordes de Cepoy P, Aeby A, Vilain C, et al. CACNA1A-associated epilepsy: Electroclinical findings and treatment response on seizures in 18 patients.  Eur J Paediatr Neurol.  2021;33:75–85. [PubMed: 34102571]
  173. Li G, Wang J, Liao P, Bartels P, Zhang H, Yu D, et al. Exclusion of alternative exon 33 of CaV1.2 calcium channels in heart is proarrhythmogenic.  Proc Natl Acad Sci U S A.  2017;114(21):E4288–E95. [PMC free article: PMC5448171] [PubMed: 28490495]
  174. Liao P, Soong TW.  Understanding alternative splicing of Cav1.2 calcium channels for a new approach towards individualized medicine.  J Biomed Res.  2010;24(3):181–6. [PMC free article: PMC3596553] [PubMed: 23554629]
  175. Liaqat K, Schrauwen I, Raza SI, Lee K, Hussain S, Chakchouk I, et al. Identification of CACNA1D variants associated with sinoatrial node dysfunction and deafness in additional Pakistani families reveals a clinical significance.  J Hum Genet.  2019;64(2):153–60. [PMC free article: PMC6561484] [PubMed: 30498240]
  176. Lin F, Barun S, Lutz CM, Wang Y, Hosford DA. Decreased (45)Ca(2)(+) uptake in P/Q-type calcium channels in homozygous lethargic (Cacnb4lh) mice is associated with increased beta3 and decreased beta4 calcium channel subunit mRNA expression.  Brain Res Mol Brain Res.  1999;71(1):1–10. [PubMed: 10407181]
  177. Lipscombe D, Andrade A. Calcium Channel CaValpha(1) Splice Isoforms—Tissue Specificity and Drug Action.  Curr Mol Pharmacol.  2015;8(1):22–31. [PMC free article: PMC5441838] [PubMed: 25966698]
  178. Lo FS, Erzurumlu RS.  L-type calcium channel-mediated plateau potentials in barrelette cells during structural plasticity.  J Neurophysiol.  2002;88(2):794–801. [PMC free article: PMC3686508] [PubMed: 12163531]
  179. Lo FS, Ziburkus J, Guido W.  Synaptic mechanisms regulating the activation of a Ca(2+)-mediated plateau potential in developing relay cells of the LGN.  J Neurophysiol.  2002;87(3):1175–85. [PubMed: 11877491]
  180. Long S, Zhou H, Li S, Wang T, Ma Y, Li C, et al. The Clinical and Genetic Features of Co-occurring Epilepsy and Autism Spectrum Disorder in Chinese Children.  Front Neurol.  2019;10:505. [PMC free article: PMC6527735] [PubMed: 31139143]
  181. Lory P, Nicole S, Monteil A.  Neuronal Cav3 channelopathies: recent progress and perspectives.  Pflugers Arch.  2020;472(7):831–44. [PMC free article: PMC7351805] [PubMed: 32638069]
  182. LoTurco JJ, Owens DF, Heath MJ, Davis MB, Kriegstein AR.  GABA and glutamate depolarize cortical progenitor cells and inhibit DNA synthesis.  Neuron.  1995;15(6):1287–98. [PubMed: 8845153]
  183. Lu ZJ, Pereverzev A, Liu HL, Weiergraber M, Henry M, Krieger A, et al. Arrhythmia in isolated prenatal hearts after ablation of the Cav2.3 (alpha1E) subunit of voltage-gated Ca2+ channels.  Cell Physiol Biochem.  2004;14(1-2):11–22. [PubMed: 14976402]
  184. Ludwig A, Flockerzi V, Hofmann F.  Regional expression and cellular localization of the alpha1 and beta subunit of high voltage-activated calcium channels in rat brain.  J Neurosci.  1997;17(4):1339–49. [PMC free article: PMC6793722] [PubMed: 9006977]
  185. Lupien-Meilleur A, Jiang X, Lachance M, Taschereau-Dumouchel V, Gagnon L, Vanasse C, et al. Reversing frontal disinhibition rescues behavioural deficits in models of CACNA1A-associated neurodevelopment disorders.  Mol Psychiatry.  2021. [PubMed: 34127816]
  186. Luvisetto S, Fellin T, Spagnolo M, Hivert B, Brust PF, Harpold MM, et al. Modal gating of human CaV2.1 (P/Q-type) calcium channels: I. The slow and the fast gating modes and their modulation by beta subunits.  J Gen Physiol.  2004;124(5):445–61. [PMC free article: PMC2234000] [PubMed: 15504896]
  187. Maejima T, Wollenweber P, Teusner LU, Noebels JL, Herlitze S, Mark MD.  Postnatal loss of P/Q-type channels confined to rhombic-lip-derived neurons alters synaptic transmission at the parallel fiber to purkinje cell synapse and replicates genomic Cacna1a mutation phenotype of ataxia and seizures in mice.  J Neurosci.  2013;33(12):5162–74. [PMC free article: PMC3641643] [PubMed: 23516282]
  188. Magee JC, Carruth M.  Dendritic voltage-gated ion channels regulate the action potential firing mode of hippocampal CA1 pyramidal neurons.  J Neurophysiol.  1999;82(4):1895–901. [PubMed: 10515978]
  189. Maheshwari A, Noebels JL.  Monogenic models of absence epilepsy: windows into the complex balance between inhibition and excitation in thalamocortical microcircuits.  Prog Brain Res.  2014;213:223–52. [PubMed: 25194492]
  190. Mallmann RT, Elgueta C, Sleman F, Castonguay J, Wilmes T, van den Maagdenberg A, et al. Ablation of Ca(V)2.1 voltage-gated Ca(2)(+) channels in mouse forebrain generates multiple cognitive impairments.  PLoS One.  2013;8(10):e78598. [PMC free article: PMC3814415] [PubMed: 24205277]
  191. Malmersjo S, Rebellato P, Smedler E, Planert H, Kanatani S, Liste I, et al. Neural progenitors organize in small-world networks to promote cell proliferation.  Proc Natl Acad Sci U S A.  2013;110(16):E1524–32. [PMC free article: PMC3631687] [PubMed: 23576737]
  192. Mantuano E, Romano S, Veneziano L, Gellera C, Castellotti B, Caimi S, et al. Identification of novel and recurrent CACNA1A gene mutations in fifteen patients with episodic ataxia type 2.  J Neurol Sci.  2010;291(1-2):30–6. [PubMed: 20129625]
  193. Mark MD, Krause M, Boele HJ, Kruse W, Pollok S, Kuner T, et al. Spinocerebellar ataxia type 6 protein aggregates cause deficits in motor learning and cerebellar plasticity.  J Neurosci.  2015;35(23):8882–95. [PMC free article: PMC6605201] [PubMed: 26063920]
  194. Mark MD, Maejima T, Kuckelsberg D, Yoo JW, Hyde RA, Shah V, et al. Delayed postnatal loss of P/Q-type calcium channels recapitulates the absence epilepsy, dyskinesia, and ataxia phenotypes of genomic Cacna1a mutations.  J Neurosci.  2011;31(11):4311–26. [PMC free article: PMC3065835] [PubMed: 21411672]
  195. Markram H, Wang Y, Tsodyks M.  Differential signaling via the same axon of neocortical pyramidal neurons.  Proc Natl Acad Sci U S A.  1998;95(9):5323–8. [PMC free article: PMC20259] [PubMed: 9560274]
  196. Marschallinger J, Sah A, Schmuckermair C, Unger M, Rotheneichner P, Kharitonova M, et al. The L-type calcium channel Cav1.3 is required for proper hippocampal neurogenesis and cognitive functions.  Cell Calcium.  2015;58(6):606–16. [PubMed: 26459417]
  197. Matsushita K, Wakamori M, Rhyu IJ, Arii T, Oda S, Mori Y, et al. Bidirectional alterations in cerebellar synaptic transmission of tottering and rolling Ca2+ channel mutant mice.  J Neurosci.  2002;22(11):4388–98. [PMC free article: PMC6758816] [PubMed: 12040045]
  198. Matthews E, Labrum R, Sweeney MG, Sud R, Haworth A, Chinnery PF, et al. Voltage sensor charge loss accounts for most cases of hypokalemic periodic paralysis. Neurology.  2009;72(18):1544–7. [PMC free article: PMC2848101] [PubMed: 19118277]
  199. Mattson MP, Kater SB.  Calcium regulation of neurite elongation and growth cone motility.  J Neurosci.  1987;7(12):4034–43. [PMC free article: PMC6569087] [PubMed: 3121806]
  200. McDonough SI, Mintz IM, Bean BP.  Alteration of P-type calcium channel gating by the spider toxin omega-Aga-IVA.  Biophys J.  1997;72(5):2117–28. [PMC free article: PMC1184405] [PubMed: 9129813]
  201. Meier H, MacPike AD.  Three syndromes produced by two mutant genes in the mouse. Clinical, pathological, and ultrastructural bases of tottering, leaner, and heterozygous mice.  J Hered.  1971;62(5):297–302. [PubMed: 4941467]
  202. Melliti K, Meza U, Adams B.  Muscarinic stimulation of alpha1E Ca channels is selectively blocked by the effector antagonist function of RGS2 and phospholipase C-beta1.  J Neurosci.  2000;20(19):7167–73. [PMC free article: PMC6772760] [PubMed: 11007872]
  203. Meyer C, Kettner A, Hochenegg U, Rubi L, Hilber K, Koenig X, et al. On the Origin of Paroxysmal Depolarization Shifts: The Contribution of Cav1.x Channels as the Common Denominator of a Polymorphous Neuronal Discharge Pattern.  Neuroscience.  2021;468:265–81. [PubMed: 34015369]
  204. Meyer FB, Cascino GD, Whisnant JP, Sharbrough FW, Ivnik RJ, Gorman DA, et al. Nimodipine as an add-on therapy for intractable epilepsy.  Mayo Clin Proc.  1995;70(7):623–7. [PubMed: 7791383]
  205. Meza U, Bannister R, Melliti K, Adams B.  Biphasic, opposing modulation of cloned neuronal alpha1E Ca channels by distinct signaling pathways coupled to M2 muscarinic acetylcholine receptors.  J Neurosci.  1999;19(16):6806–17. [PMC free article: PMC6782876] [PubMed: 10436038]
  206. Mezghrani A, Monteil A, Watschinger K, Sinnegger-Brauns MJ, Barrere C, Bourinet E, et al. A destructive interaction mechanism accounts for dominant-negative effects of misfolded mutants of voltage-gated calcium channels.  J Neurosci.  2008;28(17):4501–11. [PMC free article: PMC6670939] [PubMed: 18434528]
  207. Miao QL, Herlitze S, Mark MD, Noebels JL.  Adult loss of Cacna1a in mice recapitulates childhood absence epilepsy by distinct thalamic bursting mechanisms.  Brain.  2020;143(1):161–74. [PMC free article: PMC6935748] [PubMed: 31800012]
  208. Mikati MA, Holmes GL, Werner S, Bakkar N, Carmant L, Liu Z, et al. Effects of nimodipine on the behavioral sequalae of experimental status epilepticus in prepubescent rats.  Epilepsy Behav.  2004;5(2):168–74. [PubMed: 15123017]
  209. Miki T, Zwingman TA, Wakamori M, Lutz CM, Cook SA, Hosford DA, et al. Two novel alleles of tottering with distinct Ca(v)2.1 calcium channel neuropathologies.  Neuroscience.  2008;155(1):31–44. [PMC free article: PMC2633778] [PubMed: 18597946]
  210. Miller TM, Dias da Silva MR, Miller HA, Kwiecinski H, Mendell JR, Tawil R, et al. Correlating phenotype and genotype in the periodic paralyses.  Neurology.  2004;63(9):1647–55. [PubMed: 15534250]
  211. Mintz IM, Sabatini BL, Regehr WG.  Calcium control of transmitter release at a cerebellar synapse.  Neuron.  1995;15(3):675–88. [PubMed: 7546746]
  212. Mintz IM, Venema VJ, Swiderek KM, Lee TD, Bean BP, Adams ME.  P-type calcium channels blocked by the spider toxin omega-Aga-IVA.  Nature.  1992;355(6363):827–9. [PubMed: 1311418]
  213. Mio C, Passon N, Baldan F, Bregant E, Monaco E, Mancini L, et al. CACNA1C haploinsufficiency accounts for the common features of interstitial 12p13.33 deletion carriers.  Eur J Med Genet.  2020;63(4):103843. [PubMed: 31953239]
  214. Misra RP, Bonni A, Miranti CK, Rivera VM, Sheng M, Greenberg ME.  L-type voltage-sensitive calcium channel activation stimulates gene expression by a serum response factor-dependent pathway.  J Biol Chem.  1994;269(41):25483–93. [PubMed: 7929249]
  215. Mitchell JW, Larsen JK, Best PM. Identification of the calcium channel alpha 1E (Ca(v)2.3) isoform expressed in atrial myocytes.  Biochim Biophys Acta.  2002;1577(1):17–26. [PubMed: 12151091]
  216. Momiyama T, Ishihara K, Serikawa T, Moritake K, Sasa M.  Effect of nicardipine on abnormal excitability of CA3 pyramidal cells in hippocampal slices of spontaneously epileptic rats.  Eur J Pharmacol.  1995;280(2):119–23. [PubMed: 7589175]
  217. Monnier N, Procaccio V, Stieglitz P, Lunardi J.  Malignant-hyperthermia susceptibility is associated with a mutation of the alpha 1-subunit of the human dihydropyridine-sensitive L-type voltage-dependent calcium-channel receptor in skeletal muscle.  Am J Hum Genet.  1997;60(6):1316–25. [PMC free article: PMC1716149] [PubMed: 9199552]
  218. Morgan JI, Curran T.  Role of ion flux in the control of c-fos expression.  Nature.  1986;322(6079):552–5. [PubMed: 2426600]
  219. Morgan JI, Curran T.  Calcium as a modulator of the immediate-early gene cascade in neurons.  Cell Calcium.  1988;9(5-6):303–11. [PubMed: 3147142]
  220. Murphy JG, Gutzmann JJ, Lin L, Hu J, Petralia RS, Wang YX, et al. R-type voltage-gated Ca(2+) channels mediate A-type K(+) current regulation of synaptic input in hippocampal dendrites.  Cell reports.  2022;38(3):110264. [PMC free article: PMC10496648] [PubMed: 35045307]
  221. Murphy TH, Baraban JM, Wier WG.  Mapping miniature synaptic currents to single synapses using calcium imaging reveals heterogeneity in postsynaptic output.  Neuron.  1995;15(1):159–68. [PubMed: 7619520]
  222. Murphy TH, Worley PF, Baraban JM.  L-type voltage-sensitive calcium channels mediate synaptic activation of immediate early genes.  Neuron.  1991;7(4):625–35. [PubMed: 1657056]
  223. Myers CT, McMahon JM, Schneider AL, Petrovski S, Allen AS, Carvill GL, et al. De novo mutations in SLC1A2 and CACNA1A are important causes of epileptic encephalopathies. Am J Hum Genet.  2016;99(2):287–98. [PMC free article: PMC4974067] [PubMed: 27476654]
  224. N’Gouemo P, Yasuda R, Faingold CL.  Seizure susceptibility is associated with altered protein expression of voltage-gated calcium channel subunits in inferior colliculus neurons of the genetically epilepsy-prone rat.  Brain Res.  2010;1308:153–7. [PMC free article: PMC2793592] [PubMed: 19836362]
  225. Nakagawa-Tamagawa N, Kirino E, Sugao K, Nagata H, Tagawa Y.  Involvement of Calcium-Dependent Pathway and beta Subunit-Interaction in Neuronal Migration and Callosal Projection Deficits Caused by the Cav1.2 I1166T Mutation in Developing Mouse Neocortex.  Front Neurosci.  2021;15:747951. [PMC free article: PMC8692569] [PubMed: 34955712]
  226. Napolitano C, Antzelevitch C.  Phenotypical manifestations of mutations in the genes encoding subunits of the cardiac voltage-dependent L-type calcium channel.  Circ Res.  2011;108(5):607–18. [PMC free article: PMC3056572] [PubMed: 21372292]
  227. Napolitano C, Timothy KW, Bloise R, Priori SG. CACNA1C-Related Disorders. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Gripp KW, et al., editors. GeneReviews((R)). Seattle (WA); 1993. [PMC free article: PMC1403] [PubMed: 20301577]
  228. Narayanan J, Frech R, Walters S, Patel V, Frigerio R, Maraganore DM.  Low dose verapamil as an adjunct therapy for medically refractory epilepsy—An open label pilot study.  Epilepsy Res.  2016;126:197–200. [PubMed: 27513375]
  229. Navakkode S, Zhai J, Wong YP, Li G, Soong TW.  Enhanced long-term potentiation and impaired learning in mice lacking alternative exon 33 of CaV1.2 calcium channel.  Transl Psychiatry.  2022;12(1):1. [PMC free article: PMC8748671] [PubMed: 35013113]
  230. Newcomb R, Szoke B, Palma A, Wang G, Chen X, Hopkins W, et al. Selective peptide antagonist of the class E calcium channel from the venom of the tarantula Hysterocrates gigas.  Biochemistry.  1998;37(44):15353–62. [PubMed: 9799496]
  231. Neychev VK, Fan X, Mitev VI, Hess EJ, Jinnah HA.  The basal ganglia and cerebellum interact in the expression of dystonic movement.  Brain.  2008;131(Pt 9):2499–509. [PMC free article: PMC2724906] [PubMed: 18669484]
  232. Nicita F, Spalice A, Papetti L, Nikanorova M, Iannetti P, Parisi P.  Efficacy of verapamil as an adjunctive treatment in children with drug-resistant epilepsy: a pilot study.  Seizure.  2014;23(1):36–40. [PubMed: 24113539]
  233. Nicita F, Spalice A, Raucci U, Iannetti P, Parisi P.  The possible use of the L-type calcium channel antagonist verapamil in drug-resistant epilepsy.  Expert Rev Neurother.  2016;16(1):9–15. [PubMed: 26567612]
  234. Niu X, Yang Y, Chen Y, Cheng M, Liu M, Ding C, et al. Genotype-phenotype correlation of CACNA1A variants in children with epilepsy.  Dev Med Child Neurol.  2022;64(1):105–11. [PubMed: 34263451]
  235. Noebels JL, Sidman RL.  Inherited epilepsy: spike-wave and focal motor seizures in the mutant mouse tottering.  Science.  1979;204(4399):1334–6. [PubMed: 572084]
  236. Nowycky MC, Fox AP, Tsien RW.  Three types of neuronal calcium channel with different calcium agonist sensitivity.  Nature.  1985;316(6027):440–3. [PubMed: 2410796]
  237. O’Roak BJ, Vives L, Girirajan S, Karakoc E, Krumm N, Coe BP, et al. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations.  Nature.  2012;485(7397):246–50. [PMC free article: PMC3350576] [PubMed: 22495309]
  238. Oda S.  [The observation of rolling mouse Nagoya (rol), a new neurological mutant, and its maintenance (author’s transl)].  Jikken Dobutsu.  1973;22(4):281–8. [PubMed: 4799944]
  239. Ohmori I, Ouchida M, Kobayashi K, Jitsumori Y, Mori A, Michiue H, et al. CACNA1A variants may modify the epileptic phenotype of Dravet syndrome.  Neurobiol Dis.  2013;50:209–17. [PubMed: 23103419]
  240. Ophoff RA, Terwindt GM, Vergouwe MN, van Eijk R, Oefner PJ, Hoffman SM, et al. Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4.  Cell.  1996;87(3):543–52. [PubMed: 8898206]
  241. Ortiz Cabrera NV, Duat Rodriguez A, Fernandez Garoz B, Bernardino Cuesta B, Jimenez Legido M, Cantarin Extremera V, et al. Dystonia and Contractures are Potential Early Signs of CACNA1E-Related Epileptic Encephalopathy.  Mol Syndromol.  2021;12(1):25–32. [PMC free article: PMC7983621] [PubMed: 33776624]
  242. Panagiotakos G, Haveles C, Arjun A, Petrova R, Rana A, Portmann T, et al. Aberrant calcium channel splicing drives defects in cortical differentiation in Timothy syndrome.  Elife.  2019;8. [PMC free article: PMC6964969] [PubMed: 31868578]
  243. Parajuli LK, Nakajima C, Kulik A, Matsui K, Schneider T, Shigemoto R, et al. Quantitative regional and ultrastructural localization of the Ca(v)2.3 subunit of R-type calcium channel in mouse brain.  J Neurosci.  2012;32(39):13555–67. [PMC free article: PMC6621359] [PubMed: 23015445]
  244. Park SK, Hwang IK, An SJ, Won MH, Kang TC. Elevated P/Q type (alpha1A) and L2 type (alpha1D) Purkinje cell voltage-gated calcium channels in the cerebella of seizure prone gerbils.  Mol Cells.  2003;16(3):297–301. [PubMed: 14744018]
  245. Pasca SP, Portmann T, Voineagu I, Yazawa M, Shcheglovitov A, Pasca AM, et al. Using iPSC-derived neurons to uncover cellular phenotypes associated with Timothy syndrome.  Nat Med.  2011;17(12):1657–62. [PMC free article: PMC3517299] [PubMed: 22120178]
  246. Pennartz CM, de Jeu MT, Bos NP, Schaap J, Geurtsen AM.  Diurnal modulation of pacemaker potentials and calcium current in the mammalian circadian clock.  Nature.  2002;416(6878):286–90. [PubMed: 11875398]
  247. Pereverzev A, Klockner U, Henry M, Grabsch H, Vajna R, Olyschlager S, et al. Structural diversity of the voltage-dependent Ca2+ channel alpha1E-subunit.  Eur J Neurosci.  1998;10(3):916–25. [PubMed: 9753159]
  248. Pereverzev A, Leroy J, Krieger A, Malecot CO, Hescheler J, Pfitzer G, et al. Alternate splicing in the cytosolic II-III loop and the carboxy terminus of human E-type voltage-gated Ca(2+) channels: electrophysiological characterization of isoforms.  Mol Cell Neurosci.  2002;21(2):352–65. [PubMed: 12401453]
  249. Pereverzev A, Mikhna M, Vajna R, Gissel C, Henry M, Weiergraber M, et al. Disturbances in glucose-tolerance, insulin-release, and stress-induced hyperglycemia upon disruption of the Ca(v)2.3 (alpha 1E) subunit of voltage-gated Ca(2+) channels.  Mol Endocrinol.  2002;16(4):884–95. [PubMed: 11923483]
  250. Perez-Reyes E.  Molecular physiology of low-voltage-activated t-type calcium channels.  Physiol Rev.  2003;83(1):117–61. [PubMed: 12506128]
  251. Piedras-Renteria ES, Tsien RW.  Antisense oligonucleotides against alpha1E reduce R-type calcium currents in cerebellar granule cells.  Proc Natl Acad Sci U S A.  1998;95(13):7760–5. [PMC free article: PMC22749] [PubMed: 9636224]
  252. Pietrobon D.  Ca(V)2.1 channelopathies.  Pflugers Arch.  2010. [PubMed: 20204399]
  253. Pietrobon D.  Insights into migraine mechanisms and CaV2.1 calcium channel function from mouse models of familial hemiplegic migraine.  J Physiol.  2010;588(Pt 11):1871–8. [PMC free article: PMC2901975] [PubMed: 20194127]
  254. Pineda JC, Waters RS, Foehring RC.  Specificity in the interaction of HVA Ca2+ channel types with Ca2+-dependent AHPs and firing behavior in neocortical pyramidal neurons.  J Neurophysiol.  1998;79(5):2522–34. [PubMed: 9582225]
  255. Pinggera A, Lieb A, Benedetti B, Lampert M, Monteleone S, Liedl KR, et al. CACNA1D de novo mutations in autism spectrum disorders activate Cav1.3 L-type calcium channels.  Biol Psychiatry.  2015;77(9):816–22. [PMC free article: PMC4401440] [PubMed: 25620733]
  256. Pinggera A, Mackenroth L, Rump A, Schallner J, Beleggia F, Wollnik B, et al. New gain-of-function mutation shows CACNA1D as recurrently mutated gene in autism spectrum disorders and epilepsy.  Hum Mol Genet.  2017;26(15):2923–32. [PMC free article: PMC5886262] [PubMed: 28472301]
  257. Pinggera A, Striessnig J.  Cav 1.3 (CACNA1D) L-type Ca(2+) channel dysfunction in CNS disorders.  J Physiol.  2016;594(20):5839–49. [PMC free article: PMC4823145] [PubMed: 26842699]
  258. Platzer J, Engel J, Schrott-Fischer A, Stephan K, Bova S, Chen H, et al. Congenital deafness and sinoatrial node dysfunction in mice lacking class D L-type Ca2+ channels.  Cell.  2000;102(1):89–97. [PubMed: 10929716]
  259. Plomp JJ, van den Maagdenberg AM, Kaja S.  The ataxic Cacna1a-mutant mouse rolling nagoya: an overview of neuromorphological and electrophysiological findings.  Cerebellum.  2009;8(3):222–30. [PMC free article: PMC2734259] [PubMed: 19484318]
  260. Poetschke C, Dragicevic E, Duda J, Benkert J, Dougalis A, DeZio R, et al. Compensatory T-type Ca2+ channel activity alters D2-autoreceptor responses of Substantia nigra dopamine neurons from Cav1.3 L-type Ca2+ channel KO mice. Scientific reports.  2015;5:13688. [PMC free article: PMC4585382] [PubMed: 26381090]
  261. Ptacek LJ, Tawil R, Griggs RC, Engel AG, Layzer RB, Kwiecinski H, et al. Dihydropyridine receptor mutations cause hypokalemic periodic paralysis.  Cell.  1994;77(6):863–8. [PubMed: 8004673]
  262. Putzier I, Kullmann PH, Horn JP, Levitan ES.  Cav1.3 channel voltage dependence, not Ca2+ selectivity, drives pacemaker activity and amplifies bursts in nigral dopamine neurons.  J Neurosci.  2009;29(49):15414–9. [PMC free article: PMC2796195] [PubMed: 20007466]
  263. Qian J, Noebels JL. Presynaptic Ca(2+) influx at a mouse central synapse with Ca(2+) channel subunit mutations.  J Neurosci.  2000;20(1):163–70. [PMC free article: PMC6774110] [PubMed: 10627593]
  264. Quintela I, Eiris J, Gomez-Lado C, Perez-Gay L, Dacruz D, Cruz R, et al. Copy number variation analysis of patients with intellectual disability from North-West Spain.  Gene.  2017;626:189–99. [PubMed: 28506748]
  265. Rajakulendran S, Graves TD, Labrum RW, Kotzadimitriou D, Eunson L, Davis MB, et al. Genetic and functional characterisation of the P/Q calcium channel in episodic ataxia with epilepsy.  J Physiol.  2010;588(Pt 11):1905–13. [PMC free article: PMC2901979] [PubMed: 20156848]
  266. Raybaud A, Baspinar EE, Dionne F, Dodier Y, Sauve R, Parent L.  The role of distal S6 hydrophobic residues in the voltage-dependent gating of CaV2.3 channels.  J Biol Chem.  2007;282(38):27944–52. [PubMed: 17660294]
  267. Raybaud A, Dodier Y, Bissonnette P, Simoes M, Bichet DG, Sauve R, et al. The role of the GX9GX3G motif in the gating of high voltage-activated Ca2+ channels.  J Biol Chem.  2006;281(51):39424–36. [PubMed: 17038321]
  268. Regehr WG, Connor JA, Tank DW.  Optical imaging of calcium accumulation in hippocampal pyramidal cells during synaptic activation.  Nature.  1989;341(6242):533–6. [PubMed: 2797180]
  269. Reichenbach H, Meister EM, Theile H.  [The heart-hand syndrome. A new variant of disorders of heart conduction and syndactylia including osseous changes in hands and feet].  Kinderarztl Prax.  1992;60(2):54–6. [PubMed: 1318983]
  270. Reinson K, Oiglane-Shlik E, Talvik I, Vaher U, Ounapuu A, Ennok M, et al. Biallelic CACNA1A mutations cause early onset epileptic encephalopathy with progressive cerebral, cerebellar, and optic nerve atrophy.  Am J Med Genet A.  2016;170(8):2173–6. [PubMed: 27250579]
  271. Rettig J, Sheng ZH, Kim DK, Hodson CD, Snutch TP, Catterall WA.  Isoform-specific interaction of the alpha1A subunits of brain Ca2+ channels with the presynaptic proteins syntaxin and SNAP-25.  Proc Natl Acad Sci U S A.  1996;93(14):7363–8. [PMC free article: PMC38990] [PubMed: 8692999]
  272. Reuveni I, Friedman A, Amitai Y, Gutnick MJ.  Stepwise repolarization from Ca2+ plateaus in neocortical pyramidal cells: evidence for nonhomogeneous distribution of HVA Ca2+ channels in dendrites.  J Neurosci.  1993;13(11):4609–21. [PMC free article: PMC6576337] [PubMed: 8229187]
  273. Reyes A, Lujan R, Rozov A, Burnashev N, Somogyi P, Sakmann B.  Target-cell-specific facilitation and depression in neocortical circuits.  Nat Neurosci.  1998;1(4):279–85. [PubMed: 10195160]
  274. Rodan LH, Spillmann RC, Kurata HT, Lamothe SM, Maghera J, Jamra RA, et al. Phenotypic expansion of CACNA1C-associated disorders to include isolated neurological manifestations.  Genet Med.  2021;23(10):1922–32. [PMC free article: PMC8488020] [PubMed: 34163037]
  275. Rossignol E.  Genetics and function of neocortical GABAergic interneurons in neurodevelopmental disorders.  Neural Plast.  2011;2011:649325. [PMC free article: PMC3159129] [PubMed: 21876820]
  276. Rossignol E, Kruglikov I, van den Maagdenberg AM, Rudy B, Fishell G.  CaV 2.1 ablation in cortical interneurons selectively impairs fast-spiking basket cells and causes generalized seizures.  Ann Neurol.  2013;74(2):209–22. [PMC free article: PMC3849346] [PubMed: 23595603]
  277. Royer-Bertrand B, Jequier Gygax M, Cisarova K, Rosenfeld JA, Bassetti JA, Moldovan O, et al. De novo variants in CACNA1E found in patients with intellectual disability, developmental regression and social cognition deficit but no seizures.  Mol Autism.  2021;12(1):69. [PMC free article: PMC8547031] [PubMed: 34702355]
  278. Rozov A, Burnashev N, Sakmann B, Neher E.  Transmitter release modulation by intracellular Ca2+ buffers in facilitating and depressing nerve terminals of pyramidal cells in layer 2/3 of the rat neocortex indicates a target cell-specific difference in presynaptic calcium dynamics.  J Physiol.  2001;531(Pt 3):807–26. [PMC free article: PMC2278500] [PubMed: 11251060]
  279. Rubi L, Schandl U, Lagler M, Geier P, Spies D, Gupta KD, et al. Raised activity of L-type calcium channels renders neurons prone to form paroxysmal depolarization shifts.  Neuromolecular Med.  2013;15(3):476–92. [PMC free article: PMC3732764] [PubMed: 23695859]
  280. Saegusa H, Kurihara T, Zong S, Minowa O, Kazuno A, Han W, et al. Altered pain responses in mice lacking alpha 1E subunit of the voltage-dependent Ca2+ channel.  Proc Natl Acad Sci U S A.  2000;97(11):6132–7. [PMC free article: PMC18570] [PubMed: 10801976]
  281. Sahu G, Asmara H, Zhang FX, Zamponi GW, Turner RW. Activity-Dependent Facilitation of CaV1.3 Calcium Channels Promotes KCa3.1 Activation in Hippocampal Neurons.  J Neurosci.  2017;37(46):11255–70. [PMC free article: PMC6596814] [PubMed: 29038242]
  282. Sakurai T, Westenbroek RE, Rettig J, Hell J, Catterall WA.  Biochemical properties and subcellular distribution of the BI and rbA isoforms of alpha 1A subunits of brain calcium channels.  J Cell Biol.  1996;134(2):511–28. [PMC free article: PMC2120867] [PubMed: 8707834]
  283. Sasa M, Ishihara K, Momiyama T, Renming X, Todo N, Amano T, et al. Abnormal calcium current in hippocampal CA3 pyramidal cells of the spontaneously epileptic rat (SER).  Jpn J Psychiatry Neurol.  1992;46(2):526–8. [PubMed: 1331595]
  284. Sasaki S, Huda K, Inoue T, Miyata M, Imoto K.  Impaired feedforward inhibition of the thalamocortical projection in epileptic Ca2+ channel mutant mice, tottering.  J Neurosci.  2006;26(11):3056–65. [PMC free article: PMC6673963] [PubMed: 16540584]
  285. Schartner V, Romero NB, Donkervoort S, Treves S, Munot P, Pierson TM, et al. Dihydropyridine receptor (DHPR, CACNA1S) congenital myopathy.  Acta Neuropathol.  2017;133(4):517–33. [PubMed: 28012042]
  286. Schiller J, Schiller Y, Stuart G, Sakmann B.  Calcium action potentials restricted to distal apical dendrites of rat neocortical pyramidal neurons.  J Physiol.  1997;505(Pt 3):605–16. [PMC free article: PMC1160039] [PubMed: 9457639]
  287. Schiller Y.  Inter-ictal- and ictal-like epileptic discharges in the dendritic tree of neocortical pyramidal neurons.  J Neurophysiol.  2002;88(6):2954–62. [PubMed: 12466421]
  288. Schiller Y.  Activation of a calcium-activated cation current during epileptiform discharges and its possible role in sustaining seizure-like events in neocortical slices.  J Neurophysiol.  2004;92(2):862–72. [PubMed: 15277598]
  289. Schmitz Y, Luccarelli J, Kim M, Wang M, Sulzer D.  Glutamate controls growth rate and branching of dopaminergic axons.  J Neurosci.  2009;29(38):11973–81. [PMC free article: PMC2818361] [PubMed: 19776283]
  290. Schneider T, Wei X, Olcese R, Costantin JL, Neely A, Palade P, et al. Molecular analysis and functional expression of the human type E neuronal Ca2+ channel alpha 1 subunit.  Recept Channels.  1994;2(4):255–70. [PubMed: 7536609]
  291. Scholl UI, Goh G, Stolting G, de Oliveira RC, Choi M, Overton JD, et al. Somatic and germline CACNA1D calcium channel mutations in aldosterone-producing adenomas and primary aldosteronism. Nat Genet.  2013;45(9):1050–4. [PMC free article: PMC3876926] [PubMed: 23913001]
  292. Schramm M, Vajna R, Pereverzev A, Tottene A, Klockner U, Pietrobon D, et al. Isoforms of alpha1E voltage-gated calcium channels in rat cerebellar granule cells--detection of major calcium channel alpha1-transcripts by reverse transcription-polymerase chain reaction.  Neuroscience.  1999;92(2):565–75. [PubMed: 10408605]
  293. Schwab Y, Mouton J, Chasserot-Golaz S, Marty I, Maulet Y, Jover E.  Calcium-dependent translocation of synaptotagmin to the plasma membrane in the dendrites of developing neurones.  Brain Res Mol Brain Res.  2001;96(1-2):1–13. [PubMed: 11731003]
  294. Schwartzkroin PA, Slawsky M.  Probable calcium spikes in hippocampal neurons.  Brain Res.  1977;135(1):157–61. [PubMed: 912429]
  295. Servili E, Trus M, Atlas D.  Ion occupancy of the channel pore is critical for triggering excitation-transcription (ET) coupling.  Cell Calcium.  2019;84:102102. [PubMed: 31733625]
  296. Servili E, Trus M, Maayan D, Atlas D. beta-Subunit of the voltage-gated Ca(2+) channel Cav1.2 drives signaling to the nucleus via H-Ras.  Proc Natl Acad Sci U S A.  2018;115(37):E8624–E33. [PMC free article: PMC6140482] [PubMed: 30150369]
  297. Servili E, Trus M, Sajman J, Sherman E, Atlas D.  Elevated basal transcription can underlie timothy channel association with autism related disorders.  Prog Neurobiol.  2020;191:101820. [PubMed: 32437834]
  298. Sheng ZH, Rettig J, Takahashi M, Catterall WA.  Identification of a syntaxin-binding site on N-type calcium channels.  Neuron.  1994;13(6):1303–13. [PubMed: 7993624]
  299. Simeone KA, Sabesan S, Kim DY, Kerrigan JF, Rho JM, Simeone TA.  L-Type calcium channel blockade reduces network activity in human epileptic hypothalamic hamartoma tissue.  Epilepsia.  2011;52(3):531–40. [PMC free article: PMC3071288] [PubMed: 21269296]
  300. Sinnegger-Brauns MJ, Hetzenauer A, Huber IG, Renstrom E, Wietzorrek G, Berjukov S, et al. Isoform-specific regulation of mood behavior and pancreatic beta cell and cardiovascular function by L-type Ca 2+ channels.  J Clin Invest.  2004;113(10):1430–9. [PMC free article: PMC406526] [PubMed: 15146240]
  301. Sinnegger-Brauns MJ, Huber IG, Koschak A, Wild C, Obermair GJ, Einzinger U, et al. Expression and 1,4-dihydropyridine-binding properties of brain L-type calcium channel isoforms.  Mol Pharmacol.  2009;75(2):407–14. [PubMed: 19029287]
  302. Snutch TP, Tomlinson WJ, Leonard JP, Gilbert MM.  Distinct calcium channels are generated by alternative splicing and are differentially expressed in the mammalian CNS.  Neuron.  1991;7(1):45–57. [PubMed: 1648941]
  303. Sochivko D, Pereverzev A, Smyth N, Gissel C, Schneider T, Beck H. The Ca(V)2.3 Ca(2+) channel subunit contributes to R-type Ca(2+) currents in murine hippocampal and neocortical neurones.  J Physiol.  2002;542(Pt 3):699–710. [PMC free article: PMC2290463] [PubMed: 12154172]
  304. Song I, Kim D, Choi S, Sun M, Kim Y, Shin HS.  Role of the alpha1G T-type calcium channel in spontaneous absence seizures in mutant mice.  J Neurosci.  2004;24(22):5249–57. [PMC free article: PMC6729205] [PubMed: 15175395]
  305. Soong TW, Stea A, Hodson CD, Dubel SJ, Vincent SR, Snutch TP.  Structure and functional expression of a member of the low voltage-activated calcium channel family.  Science.  1993;260(5111):1133–6. [PubMed: 8388125]
  306. Spaetgens RL, Zamponi GW. Multiple structural domains contribute to voltage-dependent inactivation of rat brain alpha(1E) calcium channels.  J Biol Chem.  1999;274(32):22428–36. [PubMed: 10428816]
  307. Speckmann EJ, Straub H, Kohling R.  Contribution of calcium ions to the generation of epileptic activity and antiepileptic calcium antagonism.  Neuropsychobiology.  1993;27(3):122–6. [PubMed: 8232825]
  308. Splawski I, Timothy KW, Sharpe LM, Decher N, Kumar P, Bloise R, et al. Ca(V)1.2 calcium channel dysfunction causes a multisystem disorder including arrhythmia and autism.  Cell.  2004;119(1):19–31. [PubMed: 15454078]
  309. Spruston N, Schiller Y, Stuart G, Sakmann B.  Activity-dependent action potential invasion and calcium influx into hippocampal CA1 dendrites.  Science.  1995;268(5208):297–300. [PubMed: 7716524]
  310. Sridharan PS, Lu Y, Rice RC, Pieper AA, Rajadhyaksha AM.  Loss of Cav1.2 channels impairs hippocampal theta burst stimulation-induced long-term potentiation.  Channels (Austin).  2020;14(1):287–93. [PMC free article: PMC7515572] [PubMed: 32799605]
  311. Staley K, Hellier JL, Dudek FE.  Do interictal spikes drive epileptogenesis? Neuroscientist.  2005;11(4):272–6. [PubMed: 16061513]
  312. Staley KJ, Dudek FE.  Interictal spikes and epileptogenesis.  Epilepsy Curr.  2006;6(6):199–202. [PMC free article: PMC1783497] [PubMed: 17260059]
  313. Staley KJ, White A, Dudek FE.  Interictal spikes: harbingers or causes of epilepsy? Neurosci Lett.  2011;497(3):247–50. [PMC free article: PMC3124147] [PubMed: 21458535]
  314. Stam AH, Luijckx GJ, Poll-The BT, Ginjaar IB, Frants RR, Haan J, et al. Early seizures and cerebral oedema after trivial head trauma associated with the CACNA1A S218L mutation.  Journal of neurology, neurosurgery, and psychiatry.  2009;80(10):1125–9. [PubMed: 19520699]
  315. Stanika R, Campiglio M, Pinggera A, Lee A, Striessnig J, Flucher BE, et al. Splice variants of the CaV1.3 L-type calcium channel regulate dendritic spine morphology.  Scientific reports.  2016;6:34528. [PMC free article: PMC5052568] [PubMed: 27708393]
  316. Stanika RI, Flucher BE, Obermair GJ.  Regulation of Postsynaptic Stability by the L-type Calcium Channel CaV1.3 and its Interaction with PDZ Proteins.  Curr Mol Pharmacol.  2015;8(1):95–101. [PMC free article: PMC5384370] [PubMed: 25966696]
  317. Stewart SL, Hogan K, Rosenberg H, Fletcher JE. Identification of the Arg1086His mutation in the alpha subunit of the voltage-dependent calcium channel (CACNA1S) in a North American family with malignant hyperthermia.  Clin Genet.  2001;59(3):178–84. [PubMed: 11260227]
  318. Stiglbauer V, Hotka M, Ruiss M, Hilber K, Boehm S, Kubista H.  Cav 1.3 channels play a crucial role in the formation of paroxysmal depolarization shifts in cultured hippocampal neurons.  Epilepsia.  2017;58(5):858–71. [PMC free article: PMC7116787] [PubMed: 28295232]
  319. Straub H, Kohling R, Frieler A, Grigat M, Speckmann EJ.  Contribution of L-type calcium channels to epileptiform activity in hippocampal and neocortical slices of guinea-pigs.  Neuroscience.  2000;95(1):63–72. [PubMed: 10619462]
  320. Straub H, Kohling R, Luke A, Fauteck JD, Speckmann EJ, Moskopp D, et al. The effects of verapamil and flunarizine on epileptiform activity induced by bicuculline and low Mg2+ in neocortical tissue of epileptic and primary non-epileptic patients.  Brain Res.  1996;733(2):307–11. [PubMed: 8891316]
  321. Straub H, Kohling R, Speckmann EJ.  Low magnesium induced epileptiform discharges in neocortical slices (guinea pig): increased antiepileptic efficacy of organic calcium antagonist verapamil with elevation of extracellular K+ concentration.  Comp Biochem Physiol C Comp Pharmacol Toxicol.  1992;103(1):57–63. [PubMed: 1360377]
  322. Straub H, Kohling R, Speckmann EJ. Picrotoxin-induced epileptic activity in hippocampal and neocortical slices (guinea pig): suppression by organic calcium channel blockers. Brain Res.  1994;658(1-2):119–26. [PubMed: 7834332]
  323. Straub H, Kohling R, Speckmann EJ.  Strychnine-induced epileptiform activity in hippocampal and neocortical slice preparations: suppression by the organic calcium antagonists verapamil and flunarizine.  Brain Res.  1997;773(1-2):173–80. [PubMed: 9409718]
  324. Striessnig J, Koschak A, Sinnegger-Brauns MJ, Hetzenauer A, Nguyen NK, Busquet P, et al. Role of voltage-gated L-type Ca2+ channel isoforms for brain function.  Biochem Soc Trans.  2006;34(Pt 5):903–9. [PubMed: 17052224]
  325. Su H, Sochivko D, Becker A, Chen J, Jiang Y, Yaari Y, et al. Upregulation of a T-type Ca2+ channel causes a long-lasting modification of neuronal firing mode after status epilepticus.  J Neurosci.  2002;22(9):3645–55. [PMC free article: PMC6758371] [PubMed: 11978840]
  326. Sun Y, Zhang H, Selvaraj S, Sukumaran P, Lei S, Birnbaumer L, et al. Inhibition of L-Type Ca(2+) Channels by TRPC1-STIM1 Complex Is Essential for the Protection of Dopaminergic Neurons.  J Neurosci.  2017;37(12):3364–77. [PMC free article: PMC5373123] [PubMed: 28258168]
  327. Suzuki T, Delgado-Escueta AV, Aguan K, Alonso ME, Shi J, Hara Y, et al. Mutations in EFHC1 cause juvenile myoclonic epilepsy.  Nat Genet.  2004;36(8):842–9. [PubMed: 15258581]
  328. Tai C, Kuzmiski JB, MacVicar BA.  Muscarinic enhancement of R-type calcium currents in hippocampal CA1 pyramidal neurons.  J Neurosci.  2006;26(23):6249–58. [PMC free article: PMC6675200] [PubMed: 16763032]
  329. Takahashi E, Niimi K, Itakura C.  Age-related spatial and nonspatial short-term memory in Cav2.1alpha1 mutant mice, Rolling Nagoya.  Behav Brain Res.  2009;204(1):241–5. [PubMed: 19467269]
  330. Takahashi E, Niimi K, Itakura C.  Motor coordination impairment in aged heterozygous rolling Nagoya, Cav2.1 mutant mice.  Brain Res.  2009;1279:50–7. [PubMed: 19446536]
  331. Takahashi H, Magee JC.  Pathway interactions and synaptic plasticity in the dendritic tuft regions of CA1 pyramidal neurons.  Neuron.  2009;62(1):102–11. [PubMed: 19376070]
  332. Tanabe M, Gahwiler BH, Gerber U.  L-Type Ca2+ channels mediate the slow Ca2+-dependent afterhyperpolarization current in rat CA3 pyramidal cells in vitro.  J Neurophysiol.  1998;80(5):2268–73. [PubMed: 9819242]
  333. Tang F, Dent EW, Kalil K.  Spontaneous calcium transients in developing cortical neurons regulate axon outgrowth.  J Neurosci.  2003;23(3):927–36. [PMC free article: PMC6741922] [PubMed: 12574421]
  334. Tang ZZ, Zheng S, Nikolic J, Black DL.  Developmental control of CaV1.2 L-type calcium channel splicing by Fox proteins.  Mol Cell Biol.  2009;29(17):4757–65. [PMC free article: PMC2725710] [PubMed: 19564422]
  335. Tantsis EM, Gill D, Griffiths L, Gupta S, Lawson J, Maksemous N, et al. Eye movement disorders are an early manifestation of CACNA1A mutations in children.  Dev Med Child Neurol.  2016;58(6):639–44. [PubMed: 26814174]
  336. Tonelli A, D’Angelo MG, Salati R, Villa L, Germinasi C, Frattini T, et al. Early onset, non fluctuating spinocerebellar ataxia and a novel missense mutation in CACNA1A gene.  J Neurol Sci.  2006;241(1-2):13–7. [PubMed: 16325861]
  337. Toppin PJ, Chandy TT, Ghanekar A, Kraeva N, Beattie WS, Riazi S.  A report of fulminant malignant hyperthermia in a patient with a novel mutation of the CACNA1S gene.  Can J Anaesth.  2010;57(7):689–93. [PubMed: 20431982]
  338. Tottene A, Conti R, Fabbro A, Vecchia D, Shapovalova M, Santello M, et al. Enhanced excitatory transmission at cortical synapses as the basis for facilitated spreading depression in Ca(v)2.1 knockin migraine mice.  Neuron.  2009;61(5):762–73. [PubMed: 19285472]
  339. Tottene A, Volsen S, Pietrobon D. alpha(1E) subunits form the pore of three cerebellar R-type calcium channels with different pharmacological and permeation properties.  J Neurosci.  2000;20(1):171–8. [PMC free article: PMC6774111] [PubMed: 10627594]
  340. Travaglini L, Nardella M, Bellacchio E, D’Amico A, Capuano A, Frusciante R, et al. Missense mutations of CACNA1A are a frequent cause of autosomal dominant nonprogressive congenital ataxia.  Eur J Paediatr Neurol.  2017;21(3):450–6. [PubMed: 28007337]
  341. Tsien RW, Lipscombe D, Madison DV, Bley KR, Fox AP.  Multiple types of neuronal calcium channels and their selective modulation.  Trends Neurosci.  1988;11(10):431–8. [PubMed: 2469160]
  342. Vajna R, Klockner U, Pereverzev A, Weiergraber M, Chen X, Miljanich G, et al. Functional coupling between ‘R-type’ Ca2+ channels and insulin secretion in the insulinoma cell line INS-1.  Eur J Biochem.  2001;268(4):1066–75. [PubMed: 11179973]
  343. Vajna R, Schramm M, Pereverzev A, Arnhold S, Grabsch H, Klockner U, et al. New isoform of the neuronal Ca2+ channel alpha1E subunit in islets of Langerhans and kidney--distribution of voltage-gated Ca2+ channel alpha1 subunits in cell lines and tissues.  Eur J Biochem.  1998;257(1):274–85. [PubMed: 9799129]
  344. van de Bovenkamp-Janssen MC, Scheenen WJ, Kuijpers-Kwant FJ, Kozicz T, Veening JG, van Luijtelaar EL, et al. Differential expression of high voltage-activated Ca2+ channel types in the rostral reticular thalamic nucleus of the absence epileptic WAG/Rij rat.  J Neurobiol.  2004;58(4):467–78. [PubMed: 14978724]
  345. van den Maagdenberg AM, Pietrobon D, Pizzorusso T, Kaja S, Broos LA, Cesetti T, et al. A Cacna1a knockin migraine mouse model with increased susceptibility to cortical spreading depression.  Neuron.  2004;41(5):701–10. [PubMed: 15003170]
  346. van den Maagdenberg AM, Pizzorusso T, Kaja S, Terpolilli N, Shapovalova M, Hoebeek FE, et al. High cortical spreading depression susceptibility and migraine-associated symptoms in Ca(v)2.1 S218L mice.  Ann Neurol.  2010;67(1):85–98. [PubMed: 20186955]
  347. Vecchia D, Tottene A, van den Maagdenberg AM, Pietrobon D.  Mechanism underlying unaltered cortical inhibitory synaptic transmission in contrast with enhanced excitatory transmission in CaV2.1 knockin migraine mice.  Neurobiol Dis.  2014;69:225–34. [PMC free article: PMC4107271] [PubMed: 24907493]
  348. Vecchia D, Tottene A, van den Maagdenberg AM, Pietrobon D.  Abnormal cortical synaptic transmission in CaV2.1 knockin mice with the S218L missense mutation which causes a severe familial hemiplegic migraine syndrome in humans.  Front Cell Neurosci.  2015;9:8. [PMC free article: PMC4330891] [PubMed: 25741235]
  349. Vreugdenhil M, Wadman WJ.  Enhancement of calcium currents in rat hippocampal CA1 neurons induced by kindling epileptogenesis.  Neuroscience.  1992;49(2):373–81. [PubMed: 1331856]
  350. Vreugdenhil M, Wadman WJ.  Kindling-induced long-lasting enhancement of calcium current in hippocampal CA1 area of the rat: relation to calcium-dependent inactivation.  Neuroscience.  1994;59(1):105–14. [PubMed: 8190261]
  351. Walden J, Pockberger H, Speckmann EJ, Petsche H.  Paroxysmal neuronal depolarizations in the rat motorcortex in vivo: intracellular injection of the calcium agonist BAY K 8644.  Exp Brain Res.  1986;64(3):607–9. [PubMed: 2433141]
  352. Walden J, Speckmann EJ. Suppression of recurrent generalized tonic-clonic seizure discharges by intraventricular perfusion of a calcium antagonist. Electroencephalogr Clin Neurophysiol.  1988;69(4):353–62. [PubMed: 2450733]
  353. Walden J, Speckmann EJ, Witte OW.  Suppression of focal epileptiform discharges by intraventricular perfusion of a calcium antagonist.  Electroencephalogr Clin Neurophysiol.  1985;61(4):299–309. [PubMed: 2411508]
  354. Walden J, Straub H, Speckmann EJ.  Epileptogenesis: contributions of calcium ions and antiepileptic calcium antagonists.  Acta Neurol Scand Suppl.  1992;140:41–6. [PubMed: 1332361]
  355. Waldner DM, Ito K, Chen LL, Nguyen L, Chow RL, Lee A, et al. Transgenic Expression of Cacna1f Rescues Vision and Retinal Morphology in a Mouse Model of Congenital Stationary Night Blindness 2A (CSNB2A).  Transl Vis Sci Technol.  2020;9(11):19. [PMC free article: PMC7571326] [PubMed: 33117610]
  356. Wall-Lacelle S, Hossain MI, Sauve R, Blunck R, Parent L.  Double mutant cycle analysis identified a critical leucine residue in the IIS4S5 linker for the activation of the Ca(V)2.3 calcium channel.  J Biol Chem.  2011;286(31):27197–205. [PMC free article: PMC3149313] [PubMed: 21652722]
  357. Wang S, Stanika RI, Wang X, Hagen J, Kennedy MB, Obermair GJ, et al. Densin-180 Controls the Trafficking and Signaling of L-Type Voltage-Gated Cav1.2 Ca(2+) Channels at Excitatory Synapses.  J Neurosci.  2017;37(18):4679–91. [PMC free article: PMC5426563] [PubMed: 28363979]
  358. Waters J, Larkum M, Sakmann B, Helmchen F.  Supralinear Ca2+ influx into dendritic tufts of layer 2/3 neocortical pyramidal neurons in vitro and in vivo.  J Neurosci.  2003;23(24):8558–67. [PMC free article: PMC6740370] [PubMed: 13679425]
  359. Weber F, Lehmann-Horn F.  Hypokalemic Periodic Paralysis. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Mirzaa G, et al., editors. GeneReviews((R)). Seattle (WA); 1993. [PMC free article: PMC1338] [PubMed: 20301512]
  360. Weiergraber M, Henry M, Krieger A, Kamp M, Radhakrishnan K, Hescheler J, et al. Altered seizure susceptibility in mice lacking the Ca(v)2.3 E-type Ca2+ channel.  Epilepsia.  2006;47(5):839–50. [PubMed: 16686648]
  361. Weiergraber M, Henry M, Radhakrishnan K, Hescheler J, Schneider T.  Hippocampal seizure resistance and reduced neuronal excitotoxicity in mice lacking the Cav2.3 E/R-type voltage-gated calcium channel.  J Neurophysiol.  2007;97(5):3660–9. [PubMed: 17376845]
  362. Weiergraber M, Pereverzev A, Vajna R, Henry M, Schramm M, Nastainczyk W, et al. Immunodetection of alpha1E voltage-gated Ca(2+) channel in chromogranin-positive muscle cells of rat heart, and in distal tubules of human kidney.  J Histochem Cytochem.  2000;48(6):807–19. [PubMed: 10820154]
  363. Weiss N, Zamponi GW.  Genetic T-type calcium channelopathies.  J Med Genet.  2020;57(1):1–10. [PMC free article: PMC6929700] [PubMed: 31217264]
  364. Westenbroek RE, Ahlijanian MK, Catterall WA.  Clustering of L-type Ca2+ channels at the base of major dendrites in hippocampal pyramidal neurons.  Nature.  1990;347(6290):281–4. [PubMed: 2169591]
  365. Westenbroek RE, Bausch SB, Lin RC, Franck JE, Noebels JL, Catterall WA.  Upregulation of L-type Ca2+ channels in reactive astrocytes after brain injury, hypomyelination, and ischemia.  J Neurosci.  1998;18(7):2321–34. [PMC free article: PMC6793103] [PubMed: 9502793]
  366. Westenbroek RE, Hoskins L, Catterall WA.  Localization of Ca2+ channel subtypes on rat spinal motor neurons, interneurons, and nerve terminals.  J Neurosci.  1998;18(16):6319–30. [PMC free article: PMC6793183] [PubMed: 9698323]
  367. Westenbroek RE, Sakurai T, Elliott EM, Hell JW, Starr TV, Snutch TP, et al. Immunochemical identification and subcellular distribution of the alpha 1A subunits of brain calcium channels.  J Neurosci.  1995;15(10):6403–18. [PMC free article: PMC6578002] [PubMed: 7472404]
  368. Wetherington J, Serrano G, Dingledine R.  Astrocytes in the epileptic brain.  Neuron.  2008;58(2):168–78. [PMC free article: PMC4124883] [PubMed: 18439402]
  369. Wielowieyski PA, Wigle JT, Salih M, Hum P, Tuana BS.  Alternative splicing in intracellular loop connecting domains II and III of the alpha 1 subunit of Cav1.2 Ca2+ channels predicts two-domain polypeptides with unique C-terminal tails.  J Biol Chem.  2001;276(2):1398–406. [PubMed: 11010971]
  370. Williams ME, Brust PF, Feldman DH, Patthi S, Simerson S, Maroufi A, et al. Structure and functional expression of an omega-conotoxin-sensitive human N-type calcium channel.  Science.  1992;257(5068):389–95. [PubMed: 1321501]
  371. Williams ME, Marubio LM, Deal CR, Hans M, Brust PF, Philipson LH, et al. Structure and functional characterization of neuronal alpha 1E calcium channel subtypes.  J Biol Chem.  1994;269(35):22347–57. [PubMed: 8071363]
  372. Wilson SM, Toth PT, Oh SB, Gillard SE, Volsen S, Ren D, et al. The status of voltage-dependent calcium channels in alpha 1E knock-out mice.  J Neurosci.  2000;20(23):8566–71. [PMC free article: PMC6773068] [PubMed: 11102459]
  373. Wong RK, Prince DA.  Participation of calcium spikes during intrinsic burst firing in hippocampal neurons.  Brain Res.  1978;159(2):385–90. [PubMed: 728808]
  374. Woppmann A, Ramachandran J, Miljanich GP.  Calcium channel subtypes in rat brain: biochemical characterization of the high-affinity receptors for omega-conopeptides SNX-230 (synthetic MVIIC), SNX-183 (SVIB), and SNX-111 (MVIIA).  Mol Cell Neurosci.  1994;5(4):350–7. [PubMed: 7804605]
  375. Wu LG, Borst JG, Sakmann B.  R-type Ca2+ currents evoke transmitter release at a rat central synapse.  Proc Natl Acad Sci U S A.  1998;95(8):4720–5. [PMC free article: PMC22557] [PubMed: 9539805]
  376. Wu LG, Saggau P.  Block of multiple presynaptic calcium channel types by omega-conotoxin-MVIIC at hippocampal CA3 to CA1 synapses.  J Neurophysiol.  1995;73(5):1965–72. [PubMed: 7623094]
  377. Wu LG, Westenbroek RE, Borst JG, Catterall WA, Sakmann B.  Calcium channel types with distinct presynaptic localization couple differentially to transmitter release in single calyx-type synapses.  J Neurosci.  1999;19(2):726–36. [PMC free article: PMC6782194] [PubMed: 9880593]
  378. Xu JH, Wang H, Zhang W, Tang FR.  Alterations of L-type voltage dependent calcium channel alpha 1 subunit in the hippocampal CA3 region during and after pilocarpine-induced epilepsy.  Neurochem Int.  2018;114:108–19. [PubMed: 29425964]
  379. Yamamoto K, Kobayashi M.  Opposite Roles in Short-Term Plasticity for N-Type and P/Q-Type Voltage-Dependent Calcium Channels in GABAergic Neuronal Connections in the Rat Cerebral Cortex.  J Neurosci.  2018;38(46):9814–28. [PMC free article: PMC6596248] [PubMed: 30249804]
  380. Yamashita N, Aoki R, Chen S, Jitsuki-Takahashi A, Ohura S, Kamiya H, et al. Voltage-gated calcium and sodium channels mediate Sema3A retrograde signaling that regulates dendritic development.  Brain Res.  2016;1631:127–36. [PubMed: 26638837]
  381. Yan HD, Ishihara K, Hanaya R, Kurisu K, Serikawa T, Sasa M.  Voltage-dependent calcium channel abnormalities in hippocampal CA3 neurons of spontaneously epileptic rats.  Epilepsia.  2007;48(4):758–64. [PubMed: 17326796]
  382. Yao J, Qi J, Chen G. Actin-dependent activation of presynaptic silent synapses contributes to long-term synaptic plasticity in developing hippocampal neurons. J Neurosci.  2006;26(31):8137–47. [PMC free article: PMC6673772] [PubMed: 16885227]
  383. Yap EL, Greenberg ME.  Activity-Regulated Transcription: Bridging the Gap between Neural Activity and Behavior.  Neuron.  2018;100(2):330–48. [PMC free article: PMC6223657] [PubMed: 30359600]
  384. Yuste R, Nelson DA, Rubin WW, Katz LC.  Neuronal domains in developing neocortex: mechanisms of coactivation.  Neuron.  1995;14(1):7–17. [PubMed: 7826643]
  385. Zaitsev AV, Povysheva NV, Lewis DA, Krimer LS.  P/Q-type, but not N-type, calcium channels mediate GABA release from fast-spiking interneurons to pyramidal cells in rat prefrontal cortex.  J Neurophysiol.  2007;97(5):3567–73. [PubMed: 17329622]
  386. Zamponi GW, Striessnig J, Koschak A, Dolphin AC.  The Physiology, Pathology, and Pharmacology of Voltage-Gated Calcium Channels and Their Future Therapeutic Potential.  Pharmacol Rev.  2015;67(4):821–70. [PMC free article: PMC4630564] [PubMed: 26362469]
  387. Zhang H, Maximov A, Fu Y, Xu F, Tang TS, Tkatch T, et al. Association of CaV1.3 L-type calcium channels with Shank.  J Neurosci.  2005;25(5):1037–49. [PMC free article: PMC6725973] [PubMed: 15689539]
  388. Zhang Q, Chen J, Qin Y, Wang J, Zhou L.  Mutations in voltage-gated L-type calcium channel: implications in cardiac arrhythmia.  Channels (Austin).  2018;12(1):201–18. [PMC free article: PMC6104696] [PubMed: 30027834]
  389. Zhang Y, Garcia E, Sack AS, Snutch TP.  L-type calcium channel contributions to intrinsic excitability and synaptic activity during basolateral amygdala postnatal development.  J Neurophysiol.  2020;123(3):1216–35. [PubMed: 31967931]
  390. Zhang Y, Mori M, Burgess DL, Noebels JL.  Mutations in high-voltage-activated calcium channel genes stimulate low-voltage-activated currents in mouse thalamic relay neurons.  J Neurosci.  2002;22(15):6362–71. [PMC free article: PMC6758149] [PubMed: 12151514]
  391. Zhang Y, Sack AS, Jones KL, Yang Y, Garcia E, Snutch TP.  Late-Onset Behavioral and Synaptic Consequences of L-Type Ca(2+) Channel Activation in the Basolateral Amygdala of Developing Rats. eNeuro.  2022;9(1). [PMC free article: PMC8868026] [PubMed: 35064022]
  392. Zhang Z, Reboreda A, Alonso A, Barker PA, Seguela P.  TRPC channels underlie cholinergic plateau potentials and persistent activity in entorhinal cortex.  Hippocampus.  2011;21(4):386–97. [PubMed: 20082292]
  393. Zhang Z, Seguela P.  Metabotropic induction of persistent activity in layers II/III of anterior cingulate cortex.  Cereb Cortex.  2010;20(12):2948–57. [PubMed: 20348157]
  394. Zhuchenko O, Bailey J, Bonnen P, Ashizawa T, Stockton DW, Amos C, et al. Autosomal dominant cerebellar ataxia (SCA6) associated with small polyglutamine expansions in the alpha 1A-voltage-dependent calcium channel.  Nat Genet.  1997;15(1):62–9. [PubMed: 8988170]
  395. Zwingman TA, Neumann PE, Noebels JL, Herrup K.  Rocker is a new variant of the voltage-dependent calcium channel gene Cacna1a.  J Neurosci.  2001;21(4):1169–78. [PMC free article: PMC6762232] [PubMed: 11160387]
Copyright Notice

This is an open access publication, available online and distributed under the terms of a Creative Commons Attribution-Non Commercial-No Derivatives 4.0 International licence (CC BY-NC-ND 4.0), a copy of which is available at https://creativecommons.org/licenses/by-nc-nd/4.0/. Subject to this license, all rights are reserved.

Bookshelf ID: NBK609834PMID: 39637109DOI: 10.1093/med/9780197549469.003.0046

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this title (283M)

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...