3,654
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Structural biology of voltage-gated calcium channels

ORCID Icon, ORCID Icon & ORCID Icon
Article: 2290807 | Received 28 Aug 2023, Accepted 27 Nov 2023, Published online: 07 Dec 2023

ABSTRACT

Voltage-gated calcium (Cav) channels mediate Ca2+ influx in response to membrane depolarization, playing critical roles in diverse physiological processes. Dysfunction or aberrant regulation of Cav channels can lead to life-threatening consequences. Cav-targeting drugs have been clinically used to treat cardiovascular and neuronal disorders for several decades. This review aims to provide an account of recent developments in the structural dissection of Cav channels. High-resolution structures have significantly advanced our understanding of the working and disease mechanisms of Cav channels, shed light on the molecular basis for their modulation, and elucidated the modes of actions (MOAs) of representative drugs and toxins. The progress in structural studies of Cav channels lays the foundation for future drug discovery efforts targeting Cav channelopathies.

This article is part of the following collections:
Ion Channel Structure

Introduction

Cav channels constitute a group of integral membrane proteins that facilitate the selective influx of Ca2+, a second messenger involved in numerous cellular events, into the cytosol upon membrane depolarization. These proteins exhibit diverse tissue distributions and play pivotal roles in a wide array of physiological processes, including muscle contraction, neurotransmitter release, hormone secretion, and cell death [Citation1,Citation2]. The history of Cav channel discovery has been comprehensively reviewed by Tsien, Barrett and Dolphin [Citation3,Citation4]. In mammals, the ten primary subtypes of Cav channels are classified into three subfamilies, Cav1 (Cav1.1-Cav1.4), Cav2 (Cav2.1-Cav2.3) and Cav3 (Cav3.1-Cav3.3), based on sequence similarities of the α1 subunit () [Citation5–10]. As implied by the family name, the open probability of Cav channels is regulated by membrane potential. Cav1 and Cav2 are recognized as high-voltage activated Ca2+ (HVA) channels, given their activation at more depolarized membrane potential [Citation1]. In contrast, Cav3 channels form a subset of low-voltage activated Ca2+ channels (LVA) that can be activated at depolarization only slightly above the resting membrane potential and, in some cases, may even require conditioning hyperpolarization for priming [Citation11,Citation12].

Figure 1. Classification, pharmacology, and topology of Cav channels. (a) the ten primary subtypes of Cav channels are classified into three groups-Cav1, Cav2, and Cav3-based on the sequence similarities of the α1 subunits. The Cav1 and Cav2 channels belong to high-voltage activated (HVA) channels, while Cav3 members are low-voltage activated (LVA) channels. Also shown are the genes that encode the α1 subunits and the current type characterized by electrophysiological experiments. The representative drugs of each isoform are listed. (b) chemical structures of small molecule drugs whose MOAs have been determined by high-resolution structures. Please refer to table 1 for PDB codes and additional details. (c) the HVA channels consist of at least three subunits: the transmembrane α1 subunit, the extracellular α2δ subunit, and the cytosolic β subunit. The α1 subunit comprises four homologous repeats (I-IV), each containing six helices (S1-S6). The S1-S4 helices of each repeat constitute the voltage-sensing domain (VSD), and four sets of S5 and S6 helices form the ion-conducting pore domain (PD). The S4 segments, carrying repetitive positively charged residues, are highlighted as yellow cylinders. The auxiliary α2δ subunit is composed of two polypeptides (α2 and δ) associated by an inter-subunit disulphide bond. The glycosyl phosphatidylinositol (GPI) group at the C-terminus of δ anchors it to the outer leaflet of the cell membrane. The α2δ subunit contains one von willebrand factor-A (VWA) domain and four cache domains. The two conserved domains of the β subunit, src homology 3 (SH3) and a guanylate kinase­-like domain (GK), are also shown. GK interacts with the α1-interacting domain (AID) motif located at the intracellular loop between repeat I and II.

Figure 1. Classification, pharmacology, and topology of Cav channels. (a) the ten primary subtypes of Cav channels are classified into three groups-Cav1, Cav2, and Cav3-based on the sequence similarities of the α1 subunits. The Cav1 and Cav2 channels belong to high-voltage activated (HVA) channels, while Cav3 members are low-voltage activated (LVA) channels. Also shown are the genes that encode the α1 subunits and the current type characterized by electrophysiological experiments. The representative drugs of each isoform are listed. (b) chemical structures of small molecule drugs whose MOAs have been determined by high-resolution structures. Please refer to table 1 for PDB codes and additional details. (c) the HVA channels consist of at least three subunits: the transmembrane α1 subunit, the extracellular α2δ subunit, and the cytosolic β subunit. The α1 subunit comprises four homologous repeats (I-IV), each containing six helices (S1-S6). The S1-S4 helices of each repeat constitute the voltage-sensing domain (VSD), and four sets of S5 and S6 helices form the ion-conducting pore domain (PD). The S4 segments, carrying repetitive positively charged residues, are highlighted as yellow cylinders. The auxiliary α2δ subunit is composed of two polypeptides (α2 and δ) associated by an inter-subunit disulphide bond. The glycosyl phosphatidylinositol (GPI) group at the C-terminus of δ anchors it to the outer leaflet of the cell membrane. The α2δ subunit contains one von willebrand factor-A (VWA) domain and four cache domains. The two conserved domains of the β subunit, src homology 3 (SH3) and a guanylate kinase­-like domain (GK), are also shown. GK interacts with the α1-interacting domain (AID) motif located at the intracellular loop between repeat I and II.

Dysfunction of Cav channels is implicated in a range of disorders, including cardiac arrhythmias, primary aldosteronism, ataxia, migraine, cognitive anomalies, and autism. The physiology, pathology, and pharmacology of Cav channels have been extensively documented by Zamponi, Dolphin, et al [Citation2,Citation13–15]. For instance, small molecule drugs, e.g. 1,4-dihydropyridines (DHP, e.g. nifedipine), benzothiazepines (e.g. diltiazem) and phenylalkylamines (e.g. verapamil), that inhibit the cardiovascular L-type Cav channels (LTCCs), have been widely used in clinical practice for decades to treat hypertension and cardiac arrhythmias () [Citation16,Citation17]. Simultaneously, peptide toxins from spiders, snakes, and marine creatures, which selectively bind to specific regions of ion channels, have historically been employed to investigate the subtype-specific inhibitory effects [Citation18,Citation19]. Achieving a precise understanding of the modes of actions (MOAs) of Cav-targeting drugs requires the structural elucidation of drug-bound channels, a daunting challenge that was finally overcome thanks to the resolution revolution of single particle cryo-electron microscopy (cryo-EM).

The pore-forming α1 subunits of Cav channels exhibit nearly identical topologies, with approximately 2000 residues arranged into four homologous repeats (designated I-IV), each containing six transmembrane helices (S1-S6) (). The S5 and S6 segments from all four repeats collectively enclose the central pore domain (PD), forming a specific pathway for the selective passage of Ca2+ ions across the cell membrane. Meanwhile, the S1-S4 segments within each repeat constitute the voltage-sensing domain (VSD), where the S4 segment carries repetitive positively charged residues (arginine or lysine) as gating charges. The four sets of VSDs encircle PD in a domain-swapped manner, cooperatively coupling the fluctuations in membrane potential to the pore gating of ion conduction [Citation5,Citation20–23].

The core α1 subunit is self-sufficient for the autonomous function of Cav3 channels, whereas the Cav1 and Cav2 subfamilies necessitate auxiliary extracellular α2δ and intracellular β subunits for proper membrane trafficking and physiological functions () [Citation11,Citation24–26]. The Cav1.1 channels, specialized for skeletal muscle, also associate with a transmembrane γ subunit sharing the same folding pattern with claudins [Citation27,Citation28]. Within the Cav1 and Cav2 subfamilies, the α1 subunits interact with the α2δ subunits through the extracellular segments, while interacting with the β subunits on the cytosolic side (). Four mammalian α2δ genes, namely CACNA2D1 to D4, encode the extracellular subunits α2δ-1 to −4, respectively [Citation29]. These gene products are initially the preproteins for the α2δ subunits, which will undergo post-translational proteolysis into α2 and δ proteins as the mature forms. Nonetheless, these two proteins remain interconnected due to inter-subunit disulfide bond formation before proteolytic cleavage [Citation30–33]. The mature α2δ subunits are highly glycosylated proteins that aid in trafficking the α1 subunits to the cell membrane, enhancing channel expression, and modulating channel properties [Citation26]. In addition, four subtypes of β subunits, β1-β4 with multiple splice isoforms, contribute to channel trafficking, modulate gating properties, and interact with intracellular signaling molecules [Citation24,Citation34–38]. The specific assortment of subunit constituents varies depending on the particular type of Cav channel and the specific tissue or cell type in which it is located.

Obtaining accurate 3D structures for Cav channels lays the foundation for unveiling the working mechanisms and advancing drug discovery endeavors. The first cryo-EM structure of Cav channels was reported by our group in 2015 [Citation28]. Following that, various subtypes of Cav channels alone and in complex with small molecule and peptide ligands have been extensively characterized (). This review will provide a comprehensive overview of recent advancements in the structural architecture and pharmacology of Cav channels, aiming to facilitate future physiological investigations and the development of innovative drugs.

Table 1. Published cryo-EM structures of Cav channels.

Structural architecture of Cav channels

Cav channels were first purified from the T-tubule membranes of rabbit skeletal muscle and later from the cardiac and brain membranes [Citation53–58]. The endogenously purified rabbit Cav1.1 channel has historically served as the prototype for the structural analysis of working and drug modulation mechanisms [Citation27,Citation28,Citation39,Citation40]. More recently, the establishment of a heterologous expression system for Cav channels has enabled the generation of sufficient samples encompassing diverse subtypes, facilitating the structural elucidation of a broad range of Cav channels [Citation44]. In this section, we will delve into a comprehensive review of the precise structural architectures of representative members within each Cav subfamily.

Subunit architecture of Cav channels

The endogenously purified rabbit Cav1.1 and recombinantly expressed Cav2.2 channels both comprise a transmembrane α1 subunit, along with auxiliary α2δ-1 and β subunits. The rabbit Cav1.1 was also recognized for its interaction with the γ subunit (, left). Examining the overall structures of representative members within the three Cav subfamilies provides detailed insight into their subunit architectures ().

Figure 2. The subunit architecture of Cav channels. (a) the overall structures of rabbit Cav1.1 (PDB: 5GJV), human Cav2.2 (PDB: 7MIY) and human Cav3.1 (PDB: 6KZO). All structures are domain-colored. In Cav2.2, the β3 subunit is sandwiched between AID and the S6II cytosolic segment (S6IIC). The Cav2-specifc cytosolic helix in repeat II (CH2II) is also labeled. The lipid molecule, phosphatidylinositol 4,5-bisphosphate (PIP2, black stick), is coordinated with the interface between down VSDII and PD. CTD, C-terminal domain (forest cartoon). ECL, extracellular loop. (b) structural topology of the α2δ-1 subunit. The VWA domain and four cache domains (domain-colored) are intertwined in the primary sequence. The α1 subunit interacts with VWA (palecyan), Cache1 (light pink) and the intervening loop of the Cache2 domain (light blue) of the α2δ-1 subunit. The interface between VWA and α1 subunit involves the coordination of a Ca2+ ion (green sphere), facilitated by the metal ion-dependent adhesion site (MIDAS) site and the Asp residue on the S1–2 loop of repeat I.

Figure 2. The subunit architecture of Cav channels. (a) the overall structures of rabbit Cav1.1 (PDB: 5GJV), human Cav2.2 (PDB: 7MIY) and human Cav3.1 (PDB: 6KZO). All structures are domain-colored. In Cav2.2, the β3 subunit is sandwiched between AID and the S6II cytosolic segment (S6IIC). The Cav2-specifc cytosolic helix in repeat II (CH2II) is also labeled. The lipid molecule, phosphatidylinositol 4,5-bisphosphate (PIP2, black stick), is coordinated with the interface between down VSDII and PD. CTD, C-terminal domain (forest cartoon). ECL, extracellular loop. (b) structural topology of the α2δ-1 subunit. The VWA domain and four cache domains (domain-colored) are intertwined in the primary sequence. The α1 subunit interacts with VWA (palecyan), Cache1 (light pink) and the intervening loop of the Cache2 domain (light blue) of the α2δ-1 subunit. The interface between VWA and α1 subunit involves the coordination of a Ca2+ ion (green sphere), facilitated by the metal ion-dependent adhesion site (MIDAS) site and the Asp residue on the S1–2 loop of repeat I.

The extracellular α2δ-1 subunit comprises a single von Willebrand factor-A (VWA) domain and four tandem Cache domains (Cache1 to 4), which are intertwined in the primary sequence () [Citation27,Citation28]. Despite undergoing post-translational proteolysis, the α2δ-1 subunit remains covalently tethered through the disulfide bond between VWA domain and δ subunit. As observed in the structures, the VWA, Cache1, and the intervening loop of Cache2 domains are prominently engaged in interactions with the α1 subunit, particularly with the extracellular loops of repeats I to III (ECLI to ECLIII) and the intervening loop between S1 and S2 of repeat I (S1–2I loop) (). Furthermore, the interface between VWA and α1 subunit involves the coordination of a Ca2+ ion, facilitated by the metal ion-dependent adhesion site (MIDAS) motif and the Asp residue on the S1–2I loop of the α1 subunit.

The cytosolic β subunits have been identified to interact with the α1-interacting domain (AID), an 18-aa motif at the proximal region of the intracellular loop between domain I and II [Citation59]. All β subunits feature a conserved src homology 3 (SH3) and a guanylate kinase­-like domain (GK), where the AID motif binds to the deep groove on the GK domain (). Besides the well-characterized interface, the β3 subunit was recognized for its additional interactions with the extended cytosolic segments of S6II (S6IIC) in the Cav2.2 structure (, middle) [Citation46].

In the Cav1.1 structure, the γ subunit, which shares the same fold as claudins, interacts with VSDIV through both transmembrane segments and cytosolic loops. The physical contacts between the γ subunit and the α1 subunit may potentially impact the conformational alterations of VSDIV during voltage-dependent activation or inactivation. The interactions provide the molecular basis for the diverse modulation effects, including antagonistic influences, exerted by the γ subunit on channel properties (, left) [Citation60,Citation61].

Diverse PD conformations in the structures of Cav channels

Four sets of transmembrane S5 and S6 along with the supporting helices P1 and P2 collectively form the central PD, serving as the high-throughput and highly selective pathway for the influx of Ca2+ ions. Inquiries into the precise mechanisms by which Cav channels specifically bind to Ca2+ ions and swiftly release them have held substantial importance for decades [Citation21,Citation62–64]. The high-resolution 3D structures provide the opportunity to reassess previous biophysical measurements, advancing our understanding of the selectivity and permeation of Cav channels.

In the resolved structures, the EEEE/EEDD side chains protrude into the pore lumen, constituting the intrachannel binding site for Ca2+ ion (). The EM densities within the selectivity filter (SF) vestibule can be deconvoluted as two closely spaced Ca2+ ions: one situated directly at the EEEE/EEDD locus, and the other positioned in proximity to the carbonyl groups of the − 1 residues at a Ca2+ concentration of 10 mM (, left and middle). Moreover, the EEEE locus is coordinated with a single Ca2+ ion in the presence of 0.5 mM Ca2+ (, right). It was postulated that Cav channel comprised an intrapore Ca2+ binding site of low affinity (K1/2 ~10 mM), and an external binding site of high affinity (K1/2 ~0.3 μM) [Citation65]. The presence of intrapore Ca2+ density at sub-millimolar Ca2+ concentrations supersedes the initial model of the low-affinity binding site within pore. Nonetheless, mutational studies have failed to reveal an additional high-affinity Cav binding site other than the one formed by EEEE residues [Citation66,Citation67]. Structural and mutational analyses suggest that the EEEE motif tightly binds a single Ca2+ to impede Na+ influx at low concentrations, while undergoing spatial rearrangements to permit Ca2+ influx as concentration increases [Citation68].

Figure 3. The structures of PD. (a) the putative Ca2+ coordination in SF vestibule. Ca2+ ions are coordinated by the EEEE residues for Cav1 and Cav2 channels, contrasting with EEDD residues in Cav3 channels. The distinct features of Ca2+ ion densities (purple and brown meshes) are observed at concentrations of 0.5 and 10 mM, respectively. (b) the resolved structures of PD can be classified into three conformations, tight PD (represented by Cav2.2, PDB: 7MIY), relaxed PD (represented by Cav1.1, PDB: 5GJV), and loose PD (represented by Cav1.2, PDB: 8WEA). The variations include the volume of the central cavity, the composition and size of the intracellular gate, and the contour of SF. The α1 structures are domain colored (grey for repeat I, cyan for repeat II, yellow for repeat III, and palegreen for repeat IV). The Cav2-specific CH2II segment is shown as orange cartoon, with the conserved Trp residue shown as orange ball-and-stick. The hydrophobic residues constituting intracellular gate are depicted as sticks. The detergent molecule, glyco-diosgenin (GDN), is nestle into the expanded intracellular gate, as seen in the loose PD. Panel a was adapted from our published papers [Citation27,Citation50] with minor modifications.

Figure 3. The structures of PD. (a) the putative Ca2+ coordination in SF vestibule. Ca2+ ions are coordinated by the EEEE residues for Cav1 and Cav2 channels, contrasting with EEDD residues in Cav3 channels. The distinct features of Ca2+ ion densities (purple and brown meshes) are observed at concentrations of 0.5 and 10 mM, respectively. (b) the resolved structures of PD can be classified into three conformations, tight PD (represented by Cav2.2, PDB: 7MIY), relaxed PD (represented by Cav1.1, PDB: 5GJV), and loose PD (represented by Cav1.2, PDB: 8WEA). The variations include the volume of the central cavity, the composition and size of the intracellular gate, and the contour of SF. The α1 structures are domain colored (grey for repeat I, cyan for repeat II, yellow for repeat III, and palegreen for repeat IV). The Cav2-specific CH2II segment is shown as orange cartoon, with the conserved Trp residue shown as orange ball-and-stick. The hydrophobic residues constituting intracellular gate are depicted as sticks. The detergent molecule, glyco-diosgenin (GDN), is nestle into the expanded intracellular gate, as seen in the loose PD. Panel a was adapted from our published papers [Citation27,Citation50] with minor modifications.

Additionally, the structural gallery of Cav channels offers a glimpse into diverse conformations for the inactivated states. We propose classifying these states as “tight,” “relaxed,” and “loose” conformations to differentiate the structural variations within PD of the inactivated Cav channels (). The “tight” PD, evident in the structures of Cav2.2 and Cav2.3 exhibiting voltage-dependent closed-state inactivation (CSI) [Citation69], features a smaller central cavity volume with only a minor fenestration (, left). The intracellular gate is notably thick and tightened, facilitated by the presence of a Cav2-specific cytosolic helix at the II-III linker (designated as CH2II) [Citation46–48]. Supporting the structural analysis of CH2II’s role in stabilizing the inactivated state, deletion of the CH2II helix or mutation of the conserved Trp residue both exerted a substantial impact on channel inactivation, shifting the steady-state inactivation curves toward less negative voltages. Significantly, these mutations eliminated cumulative inactivation during action potential trains, indicating the crucial role of the CH2II segment as a structural determinant for the CSI of Cav2 channels [Citation47]. The slightly “relaxed” PD, observed in most Cav1 and Cav3 channel structures, displays four fenestrations and a gate radius of approximately 1 Å (, middle). In these two types of conformations, the structure of SF remains relatively consistent. The “loose” PD configuration was recently identified in the structure of Cav1.2 in the presence of an atypical calcium antagonist [Citation44]. In this structure, the SF became dilated, and the gate, while not permeable, was sufficiently expanded to accommodate a detergent molecule (, right).

The structural arrangement of VSD

It was suggested that the conformational changes in VSDs in response to membrane depolarization were initially attributed to the movement of gating charge residues on S4 segments across the membrane electric field [Citation23,Citation70,Citation71]. As described in the sliding helix model, the transfer of gating charge residues through the hydrophobic constriction site is facilitated by the transient formation of ion pairs with countercharged residues on the opposite helices [Citation71,Citation72]. As a result, the gating charge residues are thought to “slide” through the lipid bilayer, facilitating the conformational changes in the VSDs that ultimately lead to the opening or closing of the ion channel pore.

The resolved structures of VSDs can be categorized into two classes. In the structures of Cav1.1, Cav1.3, and Cav3 channels, all four VSDs are characterized as depolarized or up state. As depicted in the structure of Cav1.1, the gating charge residues R1-R4 within the four VSDs are positioned above the conserved occluding Phe in the charge transfer center. Meanwhile, R5 and R6 are positioned below (). In contrast, the structures of Cav2.2, Cav2.3 and Cav1.2 in the apo state reveal a down/deactivated state of VSDII, with the other three VSDs (VSDI, VSDIII, and VSDIV) remaining in the depolarized state (). In the down VSDII, the entire S4 segment transforms into a 310 helix, placing the gating charge residues on the same side. As shown in the structure of Cav2.2, only R2 is situated above the occluding Phe, while R3-K6 are situated below it. Upon superimposing the four VSDs, S1-S3 are relatively well aligned, whereas S4II shifts significantly downward compared to the other S4 segments.

Figure 4. The structures of VSDs. (a) the depolarized conformations of the four VSDs, as seen in the Cav1.1 structure (PDB: 5GJV). The gating charge residues on S4, An1 and An2 (conserved acidic or polar residues on S2), as well as the occluding Phe, are shown as sticks. The gating charge residues above and below the occluding Phe are labelled cyan and brown, respectively. In the depolarized conformation, R1-R4 are positioned above the occluding Phe. (b) the structure of the four VSDs in human Cav2.2 (PDB: 7MIX). VSDII (brown labeled) is in a down/deactivated conformation, while the other three VSDs remain in the depolarized state. In the down VSDII, the gating charge residues R3-K6 are below the occluding Phe (brown labeled), with only R2 above it (cyan labeled). K6 of VSDII is on the opposite side to the other four gating charge residues and projects into the cytosol. Panel b was adapted from our published paper [Citation46] with minor modifications.

Figure 4. The structures of VSDs. (a) the depolarized conformations of the four VSDs, as seen in the Cav1.1 structure (PDB: 5GJV). The gating charge residues on S4, An1 and An2 (conserved acidic or polar residues on S2), as well as the occluding Phe, are shown as sticks. The gating charge residues above and below the occluding Phe are labelled cyan and brown, respectively. In the depolarized conformation, R1-R4 are positioned above the occluding Phe. (b) the structure of the four VSDs in human Cav2.2 (PDB: 7MIX). VSDII (brown labeled) is in a down/deactivated conformation, while the other three VSDs remain in the depolarized state. In the down VSDII, the gating charge residues R3-K6 are below the occluding Phe (brown labeled), with only R2 above it (cyan labeled). K6 of VSDII is on the opposite side to the other four gating charge residues and projects into the cytosol. Panel b was adapted from our published paper [Citation46] with minor modifications.

To identify the determinants governing the conformational transition of VSDII, we suggest comprehensive investigations that include constructing Cav1 and Cav2 chimeras and performing systematic mutagenesis analysis. The endeavors will significantly enhance our comprehension of the electromechanical coupling mechanism.

The molecular basis for Cav channels modulated by endogenous components

The activity of Cav channels is subject to modulations by a variety of endogenous components, such as calmodulin [Citation73], G proteins [Citation74], lipids [Citation75], and synaptic associated proteins [Citation76]. Recent breakthroughs in elucidating the specific interactions involving calmodulin [Citation77,Citation78], components of muscle excitation-contraction coupling (STAC and junctophilin) [Citation79,Citation80], phosphatidylinositol 4,5-bisphosphate (PIP2) [Citation46,Citation48], as well as chaperones in the assembly of Cav channels [Citation81], have significantly deepened our understanding of Cav channel modulations.

Calmodulin functions as the Ca2+ sensor for calcium-dependent inactivation (CDI) of Cav channels [Citation82]. We ever incubated human Cav3.1 protein with calmodulin, but the final cryo-EM map revealed the absence of calmodulin and a large portion of C-terminal domain (CTD) [Citation50]. Alternatively, insights into calmodulin-Cav interactions were gained through X-ray and NMR structural analyses using only isoleucine-glutamine (IQ) motif of Cav. The IQ motif interacts with both N-lobe and C-lobe of Ca2+/calmodulin, with C-lobe showing higher affinity [Citation77]. A recent NMR structure of Ca2+-free calmodulin (apo calmodulin) revealed that the IQ peptide interacted with the residues in C-lobe and adopted an orientation opposite to that in the structure with Ca2+/calmodulin (). The distinctive structural observations in Ca2+ or Ca2+-free calmodulin may provide insights into CDI regulated by the Ca2+ sensor protein [Citation78].

Figure 5. Structural basis for Cav channels modulated by endogenous components. (a) different binding modes of the Cav IQ motif (cyan and orange cartoon) with Ca2+/calmodulin (palegreen, PDB: 2BE6) or with apo calmodulin (wheat, PDB: 6CTB). Green spheres indicate Ca2+ ions. (b) binding details of PIP2 in the Cav2.2 structure (PDB: 7MIY). Left: the PIP2 molecule is coordinated at the interface between down VSDII and PD. right: the PIP2 molecule is coordinated with the polar residues in Cav2.2. Hydrogen bond interactions are shown as blue dashes. (c) structure of the EMC chaperone-Cav assembly intermediate. The EMC – Cav is an approximately 0.6 MDa complex with dimensions of about 220 Å normal to the membrane plane and around 100 Å × 130 Å parallel to the membrane plane (PDB: 8EOI). EMC1–8 and 10, and the α1 and β3 subunits of Cav1.2 are shown. The EMC extensively interacts with the α1 and β3 at the transmembrane (TM dock) and cytosolic regions (cyto dock).

Figure 5. Structural basis for Cav channels modulated by endogenous components. (a) different binding modes of the Cav IQ motif (cyan and orange cartoon) with Ca2+/calmodulin (palegreen, PDB: 2BE6) or with apo calmodulin (wheat, PDB: 6CTB). Green spheres indicate Ca2+ ions. (b) binding details of PIP2 in the Cav2.2 structure (PDB: 7MIY). Left: the PIP2 molecule is coordinated at the interface between down VSDII and PD. right: the PIP2 molecule is coordinated with the polar residues in Cav2.2. Hydrogen bond interactions are shown as blue dashes. (c) structure of the EMC chaperone-Cav assembly intermediate. The EMC – Cav is an approximately 0.6 MDa complex with dimensions of about 220 Å normal to the membrane plane and around 100 Å × 130 Å parallel to the membrane plane (PDB: 8EOI). EMC1–8 and 10, and the α1 and β3 subunits of Cav1.2 are shown. The EMC extensively interacts with the α1 and β3 at the transmembrane (TM dock) and cytosolic regions (cyto dock).

Excitation-contraction coupling (ECC) in skeletal muscle necessitates functional and mechanical coordination between Cav1.1 and the ryanodine receptor (RyR1) [Citation83,Citation84]. Recently, the Van Petegem group determined the structures of Cav1.1 peptides bound to STAC2 (Cav1.1 II-III loop) [Citation80] and junctophilin 2 (Cav1.1 CTD) [Citation79], revealing the interactions of Cav1.1 with the proposed auxiliary proteins involved in ECC.

PIP2 was initially observed to decelerate the “rundown” process and modify the voltage-dependence of Cav2.1 channels, a finding later corroborated in Cav2.2 channels [Citation75,Citation85,Citation86]. In the structure of Cav2.2 and Cav2.3 channels, PIP2 is coordinated at the interface between the down VSDII and PD (, left). The head group of PIP2 wedges into the cytosolic cavity within VSDII through the interplay of S3 and S4, while the hydrophobic residues on S3 to S6 in repeat II, as well as S5 and S6 in repeat III, coordinate the tails of PIP2. Gating charge residues R4 and K5 interact with the 5-phosphate group of PIP2 (, right). Moreover, the Cav2-specific CH2II helix was noted to affect the coordination of PIP2 [Citation48]. In the CH2II-deleted structure, the elevated shifts of S4–5II and S5II afford a more sharpened angle at the location where PIP2 was observed, rendering it incompatible for PIP2 binding.

The precise biogenesis processes play a pivotal role in ensuring the proper physiological functions of Cav channels ,[Citation87]. In eukaryotes, the conserved large nine-protein complex, termed endoplasmic reticulum membrane protein complex (EMC), has been shown to aid in efficient membrane insertion, and serve as a holdase for partially assembled membrane proteins [Citation88]. Various structural and mutagenesis studies have outlined the functional roles of the EMC and implied diverse interactions with client proteins [Citation89][Citation90][Citation91]. The recent structure of Cav channel bound to EMC revealed extensive interactions at transmembrane and cytosolic docking sites () [Citation43]. Guided by the structural analysis, mutagenesis at the EMC-Cav interface, particularly at the Cavβ-Cyto dock, resulted in a substantial reduction in Cav channel currents. The structure-function analysis underscores the role of EMC as a channel holdase to facilitate the assembly and functional expression of Cav channels, thereby advancing our comprehension of the ion channel assembly intermediate.

Structural pharmacology of Cav channels

In this section, we will present an overview of the recent advancements in the structural pharmacology of Cav channels. The structural gallery, revealing the diverse MOAs of small molecule drugs and peptide toxins, is expected to greatly accelerate the future drug discovery efforts targeting Cav channelopathies.

Structural pharmacology of DHP drugs

DHP drugs, such as nifedipine, and amlodipine, have been extensively utilized in clinical settings for the effective management of hypertension. By inhibiting LTCCs in cardiovascular tissue cells, DHP drugs reduce peripheral vascular resistance and lower blood pressure [Citation16,Citation92,Citation93].

The molecular mechanisms underlying the modulation of Cav channels by representative DHP drugs, including nifedipine, amlodipine, (R)-(+)-Bay K8644 and (S)-(-)-Bay K8644, have been elucidated using rabbit Cav1.1 as the prototype [Citation39,Citation40]. The DHP drugs bind to the fenestration site enclosed by the pore-forming segments in Repeat III and IV (, DHP). Consequently, these drugs regulate Cav channels through an allosteric modulation mechanism instead of directly obstructing the ion-permeation path. The conserved N1 amine of the DHP core plays a crucial role by forming a hydrogen bond with surrounding residues [Citation94,Citation95]. Furthermore, the two oxygens of the C3-ester group, which appear to be crucial for antagonistic activities, are also H-bonded to the polar residues conserved in the Cav1 subfamily [Citation96,Citation97].

Figure 6. Overview of the structural pharmacology of Cav channels. (a) structural mapping of the drug binding sites in the Cav structures. Representative DHP drug (nifedipine, black ball-and-stick) and pore blocker (Diltiazem, magenta ball-and-stick) are shown. key interactions of representative drugs are summarized below. Hydrogen bonds are indicated by red dashes. Nifedipine: the N1 amine is hydrogen-bonded (H-bonded) to the hydroxyl group of Ser1011 on P1III, the two oxygen groups of the C3-ester are each H-bonded to Thr935 and Gln939 on S5III, the nitrophenyl ring is situated within a hydrophobic pocket formed by Val932 on S5III as well as Met1057 and Phe1060 on S6III, the DHP backbone and the methyl groups are surrounded by Phe1008 on P1III as well as Tyr1365 and Met1366 on S6IV [Citation39]; Diltiazem: interacts with the hydrophobic residues on S6I, P1III, S6III, and S6IV. Replacement of Tyr1365, Ala1369, and Ile1372 on S6IV with non-LTCC residues was shown to significantly decrease the sensitivity to the drug [Citation39]; Drug-drug interaction: AMIO (chocolate ball-and-stick) binds within the III-IV fenestration and mainly interacts with the hydrophobic residues on S6III and S6IV. AMIO anchors the accommodation of SOF (pink ball-and-stick) into the cental cavity through hydrophobic contacts (black dashes) and a polar interaction (cyan dashes) between the phosphate group of SOF and the tertiary amine of AMIO [Citation41]; Ziconotide (chocolate): Arg10 and Tyr13 of ziconotide are H-bonded to Asp664 on P2II, Ser19 is H-bonded to Glu1659 on P2IV, Thr17 is H-bonded with Asp1345 on ECLIII, Arg21 and Lys4 interact with Asp1628 and Asp1629 on ECLIV, respectively. Four of the eight ziconotide-coordinating residues in Cav2.2, Thr643, Asp1345, Lys1372 and Asp1629 (blue label), are not conserved in other Cav channels [Citation46]; Calciseptine (light purple): Arg31 and Gln49 of CaS are H-bonded to Trp1111 (O) and Asn1113 (N) on ECLIII, Arg28 is H-bonded to Ser1114 and Asp1117 on ECLIII, Gln30 interacts with Gly1476 (O), Trp46 is surrounded by hydrophobic residues Phe1116 on ECLIII, Met1126 on P2III, and Leu1471 on P2IV. Replacement of Cav1.2-specific residues at the channel-toxin interface with their counterparts in Cav1.1 or Cav2.2 (D1117H, V1501Y, N1113K, and A1123W) exhibited reduced sensitivity to calciseptine, revealing the molecular basis for the subtype specificity [Citation44]. (b) the diverse configurations of Cav channels in complex with various drugs and toxins. Pinaverium bromide (PIN) captures Cav1.2 in an inactivated state with a loose PD. Upon PIN binding, the transverse S4–5III helix combines with the S5III segment to form a single straight helix (orange arrow). P2III helix is missing in the cryo-EM map (yellow dashes). The dilated intracellular gate is fitted with a GDN molecule. Shown below is a schematic cartoon illustrating different conformations for the inactivated states. The drugs and toxins that prefer “tight,” “relaxed” and “loose” PD, as revealed by cryo-EM structures, are indicated. The binding details for ziconotide in panle A, and panel B were adapted from our published paper [Citation44,Citation46] with modifications.

Figure 6. Overview of the structural pharmacology of Cav channels. (a) structural mapping of the drug binding sites in the Cav structures. Representative DHP drug (nifedipine, black ball-and-stick) and pore blocker (Diltiazem, magenta ball-and-stick) are shown. key interactions of representative drugs are summarized below. Hydrogen bonds are indicated by red dashes. Nifedipine: the N1 amine is hydrogen-bonded (H-bonded) to the hydroxyl group of Ser1011 on P1III, the two oxygen groups of the C3-ester are each H-bonded to Thr935 and Gln939 on S5III, the nitrophenyl ring is situated within a hydrophobic pocket formed by Val932 on S5III as well as Met1057 and Phe1060 on S6III, the DHP backbone and the methyl groups are surrounded by Phe1008 on P1III as well as Tyr1365 and Met1366 on S6IV [Citation39]; Diltiazem: interacts with the hydrophobic residues on S6I, P1III, S6III, and S6IV. Replacement of Tyr1365, Ala1369, and Ile1372 on S6IV with non-LTCC residues was shown to significantly decrease the sensitivity to the drug [Citation39]; Drug-drug interaction: AMIO (chocolate ball-and-stick) binds within the III-IV fenestration and mainly interacts with the hydrophobic residues on S6III and S6IV. AMIO anchors the accommodation of SOF (pink ball-and-stick) into the cental cavity through hydrophobic contacts (black dashes) and a polar interaction (cyan dashes) between the phosphate group of SOF and the tertiary amine of AMIO [Citation41]; Ziconotide (chocolate): Arg10 and Tyr13 of ziconotide are H-bonded to Asp664 on P2II, Ser19 is H-bonded to Glu1659 on P2IV, Thr17 is H-bonded with Asp1345 on ECLIII, Arg21 and Lys4 interact with Asp1628 and Asp1629 on ECLIV, respectively. Four of the eight ziconotide-coordinating residues in Cav2.2, Thr643, Asp1345, Lys1372 and Asp1629 (blue label), are not conserved in other Cav channels [Citation46]; Calciseptine (light purple): Arg31 and Gln49 of CaS are H-bonded to Trp1111 (O) and Asn1113 (N) on ECLIII, Arg28 is H-bonded to Ser1114 and Asp1117 on ECLIII, Gln30 interacts with Gly1476 (O), Trp46 is surrounded by hydrophobic residues Phe1116 on ECLIII, Met1126 on P2III, and Leu1471 on P2IV. Replacement of Cav1.2-specific residues at the channel-toxin interface with their counterparts in Cav1.1 or Cav2.2 (D1117H, V1501Y, N1113K, and A1123W) exhibited reduced sensitivity to calciseptine, revealing the molecular basis for the subtype specificity [Citation44]. (b) the diverse configurations of Cav channels in complex with various drugs and toxins. Pinaverium bromide (PIN) captures Cav1.2 in an inactivated state with a loose PD. Upon PIN binding, the transverse S4–5III helix combines with the S5III segment to form a single straight helix (orange arrow). P2III helix is missing in the cryo-EM map (yellow dashes). The dilated intracellular gate is fitted with a GDN molecule. Shown below is a schematic cartoon illustrating different conformations for the inactivated states. The drugs and toxins that prefer “tight,” “relaxed” and “loose” PD, as revealed by cryo-EM structures, are indicated. The binding details for ziconotide in panle A, and panel B were adapted from our published paper [Citation44,Citation46] with modifications.

Among DHP drugs, amlodipine demonstrates remarkable potency and a pH-dependent inhibitory impact [Citation98]. The ethylamine side chain has been noted to extend toward the ion pore, interacting with a phospholipid molecule that transverses the central cavity. This unique MOA may distinguish the efficacy of amlodipine from that of other DHP drugs [Citation40].

Structural pharmacology of small molecule pore blockers

In addition to DHP drugs, a variety of small molecule drugs, exemplified by diltiazem and verapamil, are commonly used to manage cardiovascular conditions. Unlike DHP compounds, these drugs inhibit Cav activities by directly obstructing the ion-permeation path, classifying them as pore blockers. While the pore blockers accommodate into the proximal central cavity pockets, there are slight differences in the coordination details [Citation39,Citation45,Citation47,Citation51].

Diltiazem and verapamil serve as the prototypes for benzothiazepine and phenylalkylamine drugs, respectively, which are commonly used to treat hypertension and cardiac arrhythmias. In the resolved structures, diltiazem predominantly engages with numerous hydrophobic residues on S6I, P1III, S6III, and S6IV (, Diltiazem) [Citation39]. In contrast, verapamil potentially adopts two distinct binding modes [Citation39]. In the first mode, verapamil is primarily coordinated by the hydrophobic residues on P1II, S6II and S6III. The second mode, where verapamil interacts mainly with the hydrophobic residues on S6IV, aligns with the previous biophysical characterizations [Citation99].

Cinnarizine functions as a nonselective Cav blocker employed in the treatment of motion sickness [Citation100]. It was observed that cinnarizine binds within the central cavity of PD, with the styrene group interacting with the residues on SF vestibule [Citation45]. The core scaffold, diphenylmethylpiperazine group, fits into the cavity formed by the hydrophobic residues on repeat II and III, thereby stabilizing the overall binding of the molecule. Beside steric blockage, local structural shifts resulting from cinnarizine binding induce an α→π transition of the helical turn constituting the intracellular gate, further constricting the ion-permeation pore.

The structural basis for pore blockage of LVA channels, T-type Cav channels, has also been elucidated [Citation50,Citation51]. Z944 exhibits characteristics of both a pore blocker and an allosteric antagonist [Citation50]. In the high-resolution structure, Z944 interacts with the II-III fenestration rather than the typical DHP binding pocket at the III-IV fenestration [Citation39]. The phenyl ring is positioned into the fenestration site, while the tert-butyl group is directed toward the ion pore, impeding the permeation of ions. Similar to the cinnarizine-bound structure, Z944 binding induces an α→π transition in S6II. A recent study has further unveiled analogous binding modes observed in Cav3.3 channels complexed with mibefradil, pimozide, and otilonium bromide [Citation51].

Two inhibitors, PD17312 and Cav2.2 blocker 1, were observed to block Cav2.2 channels through a dual mechanism involving both pore blockage and allosteric modulation [Citation47]. Despite having distinct chemical structures, the two molecules occupy the same binding site (III-IV fenestration) within Cav2.2. Unveiling the structural details of the III-IV fenestration site will facilitate the drug discovery efforts aimed at targeting Cav2.2 for the treatment of chronic pain.

Structural basis for the adverse interactions of sofosbuvir and amiodarone on LTCCs

Sofosbuvir, which targets NS5B polymerase of the hepatitis C virus, has achieved almost 100% cure rates for hepatitis C [Citation101–104]. Nevertheless, co-administration with amiodarone, an antiarrhythmic drug that primarily inhibits diverse ion channels in heart, leads to a drug-drug interaction causing severe bradycardia [Citation105–107]. Analysis of the cryo-EM structures of LTCCs in the presence of the two drugs uncovered a direct drug-drug interaction within the channels, shedding light on the concerning clinical observation [Citation41].

Amiodarone binds to the III-IV fenestration of the LTCC through extensive hydrophobic interactions (, Drug-drug interaction). However, when administered alone, sofosbuvir (or its analogue MNI-1) is not present in the structure. The synergistic inhibition of Cav channels by amiodarone and sofosbuvir (or its analogue MNI-1) is clearly elucidated by the two drug-bound structures. Amiodarone anchors sofosbuvir/MNI-1 to the central cavity of LTCC, effectively obstructing the ion-conducting pathway. The drug-drug interaction is facilitated by the hydrophobic contacts and notably enhanced by a polar interaction between the phosphate group of sofosbuvir/MNI-1 and the tertiary amine of amiodarone. Observing the direct physical and pharmacodynamic interaction of amiodarone and sofosbuvir/MNI-1 on the scaffold of LTCCs underscores the clinically pertinent, potentially life-threatening nature of this drug-drug interaction [Citation41].

Gabapentinoids bind to the α2δ-1 subunit

Gabapentinoid drugs are widely used to treat epilepsy, post-herpetic neuralgia, diabetic neuropathy, fibromyalgia, and anxiety disorder. Unlike the above-mentioned MOAs, gabapentinoids are recognized for interacting with the α2δ subunit, specifically the α2δ-1 and α2δ-2 isoforms, thereby reducing the membrane expression of the neuronal Cav channels [Citation108–113].

The molecular recognition of the gabapentinoid drugs, gabapentin and mirogabalin, by the α2δ-1 subunit have been investigated, but the structural observations are somewhat ambiguous [Citation42,Citation52]. Gabapentinoid drugs were supposed to occupy a pocket identical to the L-leucine binding site at the Cache1 domain (). No discernible conformational change was observed in the α2δ-1 subunit upon binding to gabapentin or mirogabalin. However, our group ever refrained from conclusively assigning densities in the same site to gabapentinoid (pregabalin) due to its similar size and shape to the endogenous L-leucine ligand [Citation27]. It remains to be investigated how gabapentinoids regulate the function of neuronal Cav channels if there is no conformational alternation for the α2δ-1 subunit.

Structural basis for Cav channels blocked by peptide toxins

Peptide toxins have historically been valuable tools for understanding the physiological functions of ion channels. Moreover, ongoing advancements in peptide chemistry, drug delivery, and formulation are making peptide toxins an increasingly attractive avenue for the development of ion channel drugs [Citation114].

Ziconotide, derived from ω-conotoxin MVIIA, acts as a selective blocker for Cav2.2 channels and received FDA approval in 2004 for the treatment of severe chronic pain [Citation18,Citation19,Citation115]. Structural studies have unveiled the molecular underpinnings of the specific pore blockage of ziconotide on human Cav2.2 [Citation46,Citation47]. Ziconotide is situated in the electronegative pocket surrounding the entrance to SF (, Ziconotide). To accommodate ziconotide, ECLIII must move upward together with the α2δ-1 subunit. The positively charged residues in ziconotide effectively neutralize the negatively charged pocket that attracts ions into the ion entrance, creating a spatial obstruction to prevent Ca2+ influx into the pore. Further sequence comparison revealed that half of the crucial residues required for ziconotide coordination were not conserved among Cav channels, thus elucidating the molecular basis for the selective pore blockade of ziconotide.

Calciseptine, a member of the three-finger toxin isolated from the venom of black mamba, consists of 60 amino acids and features four pairs of disulfide bonds. With the ability to selectively block Cav1.2 and Cav1.3 channels, calciseptine effectively inhibits the contraction of smooth and cardiac muscles [Citation116–121]. In contrast to other VSD or PD-binding toxins, calciseptine is positioned on the PD shoulder and thoroughly interacts with the P2 pore helices and the ECLs in repeats III and IV (, Calciseptine) [Citation44]. Mutation assays have highlighted the role of Cav1.2-specific residues, mapped at the protein-toxin interface, in determining the subtype specificity. The unexpected binding mode of calciseptine provides a valuable template and induces novel concepts for the development of innovative antihypertensive medications.

Pinaverium bromide traps Cav1.2 in an inactivated state with a loose PD

Studies in structural pharmacology not only clarify the MOAs of various drugs, but also have the potential to capture additional functional states of Cav channels. However, most investigations have yielded the conformations similar to the apo state. For instance, the structures of Cav.2.2/Cav2.3 in the apo state or in complex with drugs often exhibit an inactivated state with a tight PD (, left). Additionally, the structures of most drug bound LTCCs show a relaxed PD, similar to the apo structure (, middle). Our group recently revealed that pinaverium bromide (PIN), an antispasmodic agent used to alleviate symptoms of irritable bowel syndrome, could trap Cav1.2 in a conformation distinct to the apo state [Citation44].

PIN potentially wedges into the corner enclosed by S4–5III, S6III and S6IV. Upon binding to PIN, Cav1.2 undergoes a substantial range of conformational changes, including a structural transition of VSDII from a down to an up state, a significant reshaping of SF vestibule, and dilation of intracellular gate (, right, Loose PD). Despite the considerable expansion of the ion-permeation path, molecular dynamics simulations reveal that the PIN-bound state remains non-conductive for Ca2+ ions, thereby supporting the antagonistic activities of PIN for Cav1.2.

Conclusion

Cav channels are of great physiological importance across cardiovascular, neuronal, and endocrine systems, serving as primary targets for treating diseases such as hypertension, cardiac arrhythmia, chronic pain, and epilepsy. In this review, we have summarized the strides made and novel insights achieved through the structural analyses of different Cav subtypes. The structure gallery establishes the foundation for dissecting the physiological functions of Cav channels and underlying the electromechanical coupling mechanisms. Moreover, the elucidation of structures in complex with clinically relevant drugs and promising ligands offers detailed insights into the structural pharmacology of Cav channels. These studies advance our comprehension of the working mechanisms, and pave the way for prospective drug discovery against Cav channelopathies.

Disclosure statement

No potential conflict of interest was reported by the authors.

Data availability statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Additional information

Funding

This work was not supported by any funding.

References

  • Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16(1):521–19. doi: 10.1146/annurev.cellbio.16.1.521
  • Zamponi GW, Striessnig J, Koschak A, et al. The physiology, pathology, and pharmacology of voltage-gated calcium channels and their future therapeutic potential. Pharmacol Rev. 2015 Oct;67(4):821–870. doi: 10.1124/pr.114.009654
  • Tsien RW, Barrett CF. A Brief history of calcium channel discovery. Voltage-gated calcium channels. Boston, MA: Springer US; 2005. pp. 27–47.
  • Dolphin AC. A short history of voltage-gated calcium channels. Br J Pharmacol. 2006 Jan;147(Suppl S1):S56–62. doi: 10.1038/sj.bjp.0706442
  • Ertel EA, Campbell KP, Harpold MM, et al. Nomenclature of voltage-gated calcium channels. Neuron. 2000 Mar;25(3):533–535. doi: 10.1016/S0896-6273(00)81057-0
  • Tanabe T, Takeshima H, Mikami A, et al. Primary structure of the receptor for calcium channel blockers from skeletal muscle. Nature. 1987 Jul 23-29;328(6128):313–318. doi: 10.1038/328313a0
  • Mikami A, Imoto K, Tanabe T, et al. Primary structure and functional expression of the cardiac dihydropyridine-sensitive calcium channel. Nature. 1989 Jul 20;340(6230):230–233. doi: 10.1038/340230a0
  • Mori Y, Friedrich T, Kim MS, et al. Primary structure and functional expression from complementary DNA of a brain calcium channel. Nature. 1991 Apr 4;350(6317):398–402. doi: 10.1038/350398a0
  • Dubel SJ, Starr TV, Hell J, et al. Molecular cloning of the alpha-1 subunit of an omega-conotoxin-sensitive calcium channel. Proc Natl Acad Sci U S A. 1992 Jun 1;89(11):5058–5062. doi: 10.1073/pnas.89.11.5058
  • Soong TW, Stea A, Hodson CD, et al. Structure and functional expression of a member of the low voltage-activated calcium channel family. Science. 1993 May 21;260(5111):1133–1136. doi: 10.1126/science.8388125
  • Perez-Reyes E. Molecular physiology of low-voltage-activated t-type calcium channels. Physiol Rev. 2003 Jan;83(1):117–161. doi: 10.1152/physrev.00018.2002
  • Carbone E, Lux HD. A low voltage-activated, fully inactivating Ca channel in vertebrate sensory neurones. Nature. 1984 Aug 9-15;310(5977):501–502. doi: 10.1038/310501a0
  • Iftinca MC, Zamponi GW. Regulation of neuronal T-type calcium channels. Trends Pharmacol Sci. 2009 Jan;30(1):32–40. doi: 10.1016/j.tips.2008.10.004
  • Schroeder CI, Doering CJ, Zamponi GW, et al. N-type calcium channel blockers: novel therapeutics for the treatment of pain. Med Chem. 2006 Sep;2(5):535–543. doi: 10.2174/157340606778250216
  • Zamponi GW. Targeting voltage-gated calcium channels in neurological and psychiatric diseases. Nat Rev Drug Discov. 2016 Jan;15(1):19–34. doi: 10.1038/nrd.2015.5
  • Sorkin EM, Clissold SP, Brogden RN. Nifedipine a review of its Pharmacodynamic and pharmacokinetic properties, and therapeutic efficacy, in ischaemic heart disease, hypertension and related cardiovascular disorders. Drugs. 1985 Sep;30(3):182–274. doi: 10.2165/00003495-198530030-00002
  • Fleckenstein A. History of calcium antagonists. Circ Res. 1983 Feb;52(2 Pt 2):I3–16.
  • Bowersox SS, Luther R. Pharmacotherapeutic potential of omega-conotoxin MVIIA (SNX-111), an N-type neuronal calcium channel blocker found in the venom of conus magus. Toxicon. 1998 Nov;36(11):1651–1658. doi: 10.1016/S0041-0101(98)00158-5
  • Miljanich GP. Ziconotide: neuronal calcium channel blocker for treating severe chronic pain. Curr Med Chem. 2004 Dec;11(23):3029–3040. doi: 10.2174/0929867043363884
  • Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol. 2011 Aug 1;3(8):a003947. doi: 10.1101/cshperspect.a003947
  • Tsien RW, Hess P, McCleskey EW, et al. Calcium channels: mechanisms of selectivity, permeation, and block. Ann Rev Biophys Biophys Chem. 1987;16(1):265–290.
  • Corry B, Allen TW, Kuyucak S, et al. Mechanisms of permeation and selectivity in calcium channels. Biophys J. 2001 Jan;80(1):195–214. doi: 10.1016/S0006-3495(01)76007-9
  • Tombola F, Pathak MM, Isacoff EY. How does voltage open an ion channel? Annu Rev Cell Dev Biol. 2006;22(1):23–52. doi: 10.1146/annurev.cellbio.21.020404.145837
  • Buraei Z, Yang J. The ß subunit of voltage-gated Ca2+ channels. Physiol Rev. 2010 Oct;90(4):1461–506. doi: 10.1152/physrev.00057.2009
  • Dolphin AC. Calcium channel auxiliary α2δ and β subunits: trafficking and one step beyond. Nat Rev Neurosci. 2012 Jul 18;13(8):542–555. doi: 10.1038/nrn3311
  • Dolphin AC. Voltage-gated calcium channels and their auxiliary subunits: physiology and pathophysiology and pharmacology. J Physiol. 2016 Oct 1;594(19):5369–5390. doi: 10.1113/JP272262
  • Wu J, Yan Z, Li Z, et al. Structure of the voltage-gated calcium channel Ca(v)1.1 at 3.6 Å resolution. Nature. 2016 Sep 8;537(7619):191–196. doi: 10.1038/nature19321
  • Wu J, Yan Z, Li Z, et al. Structure of the voltage-gated calcium channel Cav1.1 complex. Science. 2015 Dec 18;350(6267):aad2395. doi: 10.1126/science.aad2395
  • Davies A, Hendrich J, Van Minh AT, et al. Functional biology of the alpha(2)delta subunits of voltage-gated calcium channels. Trends Pharmacol Sci. 2007 May;28(5):220–8. doi: 10.1016/j.tips.2007.03.005
  • De Jongh KS, Warner C, Catterall WA. Subunits of purified calcium channels. Alpha 2 and delta are encoded by the same gene. J Biol Chem. 1990 Sep 5;265(25):14738–14741. doi: 10.1016/S0021-9258(18)77174-3
  • Jay SD, Sharp AH, Kahl SD, et al. Structural characterization of the dihydropyridine-sensitive calcium channel alpha 2-subunit and the associated delta peptides. J Biol Chem. 1991 Feb 15;266(5):3287–3293. doi: 10.1016/S0021-9258(18)49986-3
  • Davies A, Kadurin I, Alvarez-Laviada A, et al. The α 2 δ subunits of voltage-gated calcium channels form GPI-anchored proteins, a posttranslational modification essential for function. Proc Natl Acad Sci U S A. 2010 Jan 26;107(4):1654–9. doi: 10.1073/pnas.0908735107
  • Calderón-Rivera A, Andrade A, Hernández-Hernández O, et al. Identification of a disulfide bridge essential for structure and function of the voltage-gated Ca2+ channel α2δ-1 auxiliary subunit. Cell Calcium. 2012 Jan;51(1):22–30. doi: 10.1016/j.ceca.2011.10.002
  • Lacerda AE, Kim HS, Ruth P, et al. Normalization of current kinetics by interaction between the alpha 1 and beta subunits of the skeletal muscle dihydropyridine-sensitive Ca2+ channel. Nature. 1991 Aug 8;352(6335):527–30. doi: 10.1038/352527a0
  • Shistik E, Ivanina T, Puri T, et al. Ca2+ current enhancement by alpha 2/delta and beta subunits in xenopus oocytes: contribution of changes in channel gating and alpha 1 protein level. J Physiol. 1995 Nov 15;489(Pt 1):55–62. doi: 10.1113/jphysiol.1995.sp021029
  • Varadi G, Lory P, Schultz D, et al. Acceleration of activation and inactivation by the beta subunit of the skeletal muscle calcium channel. Nature. 1991 Jul 11;352(6331):159–62. doi: 10.1038/352159a0
  • Josephson IR, Varadi G. The beta subunit increases Ca2+ currents and gating charge movements of human cardiac L-type Ca2+ channels. Biophys J. 1996 Mar;70(3):1285–1293. doi: 10.1016/S0006-3495(96)79685-6
  • Neely A, Wei X, Olcese R, et al. Potentiation by the beta subunit of the ratio of the ionic current to the charge movement in the cardiac calcium channel. Science. 1993 Oct 22;262(5133):575–8. doi: 10.1126/science.8211185
  • Zhao Y, Huang G, Wu J, et al. Molecular basis for ligand modulation of a Mammalian voltage-gated Ca(2+) channel. Cell. 2019 May 30;177(6):1495–1506 e12. doi: 10.1016/j.cell.2019.04.043
  • Gao S, Yan N. Structural Basis of the Modulation of the Voltage-Gated Calcium Ion Channel Ca(v)1.1 by Dihydropyridine Compounds**. Angew Chem-Int Ed. 2021 Feb;60(6):3131–3137. doi: 10.1002/anie.202011793
  • Yao X, Gao S, Wang J, et al. Structural basis for the severe adverse interaction of sofosbuvir and amiodarone on L-type Ca(v) channels. Cell. 2022 Dec 8;185(25):4801–4810.e13. doi: 10.1016/j.cell.2022.10.024
  • Chen Z, Mondal A, Minor DLJr. Structural basis for Ca(V)α(2)δ: gabapentin binding. Nat Struct Mol Biol. 2023 Jun;30(6):735–739. doi: 10.1038/s41594-023-00951-7
  • Chen Z, Mondal A, Abderemane-Ali F, et al. EMC chaperone–CaV structure reveals an ion channel assembly intermediate. Nature. 2023 May 17;619(7969):410–419. doi: 10.1038/s41586-023-06175-5
  • Gao S, Yao X, Chen J, et al. Structural basis for human Ca(v)1.2 inhibition by multiple drugs and the neurotoxin calciseptine. Cell. 2023 Nov 8;186(24):5363–5374.e16. doi: 10.1016/j.cell.2023.10.007
  • Yao X, Gao S, Yan N. Structural basis for pore blockade of human voltage-gated calcium channel Ca(v)1.3 by motion sickness drug cinnarizine. Cell Res. 2022 Oct;32(10):946–948. doi: 10.1038/s41422-022-00663-5
  • Gao S, Yao X, Yan N. Structure of human Cav2.2 channel blocked by the painkiller ziconotide. Nature. 2021 Aug;596(7870):143–147. doi: 10.1038/s41586-021-03699-6
  • Dong Y, Gao Y, Xu S, et al. Closed-state inactivation and pore-blocker modulation mechanisms of human Ca(V)2.2. Cell Rep. 2021 Nov 2;37(5):109931. doi: 10.1016/j.celrep.2021.109931
  • Yao X, Wang Y, Wang Z, et al. Structures of the R-type human Ca(v)2.3 channel reveal conformational crosstalk of the intracellular segments. Nat Commun. 2022 Nov 30;13(1):7358. doi: 10.1038/s41467-022-35026-6
  • Gao Y, Xu S, Cui X, et al. Molecular insights into the gating mechanisms of voltage-gated calcium channel Ca(V)2.3. Nat Commun. 2023 Jan 31;14(1):516. doi: 10.1038/s41467-023-36260-2
  • Zhao Y, Huang G, Wu Q, et al. Cryo-EM structures of apo and antagonist-bound human Ca(v)3.1. Nature. 2019 Dec;576(7787):492–497. doi: 10.1038/s41586-019-1801-3
  • He L, Yu Z, Geng Z, et al. Structure, gating, and pharmacology of human Ca(V)3.3 channel. Nat Commun. 2022 Apr 19;13(1):2084. doi: 10.1038/s41467-022-29728-0
  • Kozai D, Numoto N, Nishikawa K, et al. Recognition Mechanism of a Novel Gabapentinoid Drug, Mirogabalin, for Recombinant Human alpha(2)delta1, a Voltage-Gated Calcium Channel Subunit. J Mol Biol. 2023 May 15;435(10):168049. doi: 10.1016/j.jmb.2023.168049
  • Curtis BM, Catterall WA. Purification of the calcium antagonist receptor of the voltage-sensitive calcium channel from skeletal muscle transverse tubules. Biochemistry. 1984 May 8;23(10):2113–2118. doi: 10.1021/bi00305a001
  • Chang FC, Hosey MM. Dihydropyridine and phenylalkylamine receptors associated with cardiac and skeletal muscle calcium channels are structurally different. J Biol Chem. 1988 Dec 15;263(35):18929–18937. doi: 10.1016/S0021-9258(18)37371-X
  • Kuniyasu A, Oka K, Ide-Yamada T, et al. Structural characterization of the dihydropyridine receptor-linked calcium channel from porcine heart. J Biochem. 1992 Aug;112(2):235–42. doi: 10.1093/oxfordjournals.jbchem.a123883
  • Schneider T, Hofmann F. The bovine cardiac receptor for calcium channel blockers is a 195-kDa protein. Eur J Biochem. 1988 Jun 1;174(2):369–375. doi: 10.1111/j.1432-1033.1988.tb14107.x
  • McEnery MW, Snowman AM, Sharp AH, et al. Purified omega-conotoxin GVIA receptor of rat brain resembles a dihydropyridine-sensitive L-type calcium channel. Proc Natl Acad Sci U S A. 1991 Dec 15;88(24):11095–11099. doi: 10.1073/pnas.88.24.11095
  • Witcher DR, De Waard M, Sakamoto J, et al. Subunit identification and reconstitution of the N-type Ca2+ channel complex purified from brain. Science. 1993 Jul 23;261(5120):486–489. doi: 10.1126/science.8392754
  • Pragnell M, De Waard M, Mori Y, et al. Calcium channel beta-subunit binds to a conserved motif in the I-II cytoplasmic linker of the alpha 1-subunit. Nature. 1994 Mar 3;368(6466):67–70. doi: 10.1038/368067a0
  • Kang MG, Campbell KP. Gamma subunit of voltage-activated calcium channels. J Biol Chem. 2003 Jun 13;278(24):21315–8. doi: 10.1074/jbc.R300004200
  • Andronache Z, Ursu D, Lehnert S, et al. The auxiliary subunit gamma 1 of the skeletal muscle L-type Ca2+ channel is an endogenous Ca2+ antagonist. Proc Natl Acad Sci U S A. 2007 Nov 6;104(45):17885–90. doi: 10.1073/pnas.0704340104
  • Tsien RW. Calcium Signals. IOP Publishing; 2023. Calcium channel selectivity and permeation: function meets 3D structure; 2-1-2–20. doi: 10.1088/978-0-7503-2009-2ch2
  • Hess P, Tsien RW. Mechanism of ion permeation through calcium channels. Nature. 1984 May 31-Jun 6;309(5967):453–6. doi: 10.1038/309453a0
  • Sather WA, McCleskey EW. Permeation and selectivity in calcium channels. Annu Rev Physiol. 2003;65(1):133–159. doi: 10.1146/annurev.physiol.65.092101.142345
  • Kostyuk PG, Mironov SL, Shuba YM. 2 ION-SELECTING FILTERS IN THE CALCIUM-CHANNEL OF THE SOMATIC MEMBRANE OF MOLLUSK NEURONS. J Membrain Biol. 1983;76(1):83–93. Article. doi: 10.1007/BF01871455
  • Ellinor PT, Yang J, Sather WA, et al. Ca2+ channel selectivity at a single locus for high-affinity Ca2+ interactions. Neuron. 1995 Nov;15(5):1121–1132. doi: 10.1016/0896-6273(95)90100-0
  • Yang J, Ellinor PT, Sather WA, et al. Molecular determinants of Ca2+ selectivity and ion permeation in L-type Ca2+ channels. Nature. 1993 Nov 11;366(6451):158–61. doi: 10.1038/366158a0
  • Shuba YM. Models of calcium permeation through T-type channels. Pflugers Arch - Eur J Physiol. 2014 Apr;466(4):635–644. doi: 10.1007/s00424-013-1437-3
  • Patil PG, Brody DL, Yue DT. Preferential closed-state inactivation of neuronal calcium channels. Neuron. 1998 May;20(5):1027–1038. doi: 10.1016/S0896-6273(00)80483-3
  • Hering S, Zangerl-Plessl EM, Beyl S, et al. Calcium channel gating. Pflugers Arch - Eur J Physiol. 2018 Sep;470(9):1291–1309. doi: 10.1007/s00424-018-2163-7
  • Catterall WA, Wisedchaisri G, Zheng N. The chemical basis for electrical signaling. Nat Chem Biol. 2017 Apr 13;13(5):455–463. doi: 10.1038/nchembio.2353
  • Yarov-Yarovoy V, DeCaen PG, Westenbroek RE, et al. Structural basis for gating charge movement in the voltage sensor of a sodium channel. Proc Natl Acad Sci U S A. 2012 Jan 10;109(2):E93–102. doi: 10.1073/pnas.1118434109
  • Saimi Y, Kung C. Calmodulin as an ion channel subunit. Annu Rev Physiol. 2002;64(1):289–311. doi: 10.1146/annurev.physiol.64.100301.111649
  • Dolphin AC. G protein modulation of voltage-gated calcium channels. Pharmacol Rev. 2003 Dec;55(4):607–627. doi: 10.1124/pr.55.4.3
  • Wu L, Bauer CS, Zhen XG, et al. Dual regulation of voltage-gated calcium channels by PtdIns(4,5)P2. Nature. 2002 Oct 31;419(6910):947–952. doi: 10.1038/nature01118
  • Zamponi GW. Regulation of presynaptic calcium channels by synaptic proteins. J Pharmacol Sci. 2003 Jun;92(2):79–83. doi: 10.1254/jphs.92.79
  • Van Petegem F, Chatelain FC, Minor DL. Insights into voltage-gated calcium channel regulation from the structure of the CaV1.2 IQ domain-Ca2+/calmodulin complex. Nat Struct Mol Biol. 2005 Dec;12(12):1108–15. doi: 10.1038/nsmb1027
  • Turner M, Anderson DE, Bartels P, et al. Alpha-actinin-1 promotes activity of the L-type Ca(2+) channel Ca(v) 1.2. EMBO J. 2020 Mar 2;39(5):e102622. doi: 10.15252/embj.2019102622
  • Yang ZF, Panwar P, McFarlane CR, et al. Structures of the junctophilin/voltage-gated calcium channel interface reveal hot spot for cardiomyopathy mutations. Proc Natl Acad Sci U S A. 2022 Mar 8;119(10):e2120416119. doi: 10.1073/pnas.2120416119
  • King Yuen SM W, Campiglio M, Tung CC, et al. Structural insights into binding of STAC proteins to voltage-gated calcium channels. Proc Natl Acad Sci U S A. 2017 Nov 7;114(45):E9520–E9528. doi: 10.1073/pnas.1708852114
  • Chen Z, Mondal A, Ali FA, et al. EMC chaperone–CaV structure reveals an ion channel assembly intermediate. Nature. 2023 May 17;619(7969):410–419. doi: 10.1038/s41586-023-06175-5
  • Peterson BZ, DeMaria CD, Adelman JP, et al. Calmodulin is the Ca2+ sensor for Ca2+ -dependent inactivation of L-type calcium channels. Neuron. 1999 Mar;22(3):549–58. doi: 10.1016/S0896-6273(00)80709-6
  • Meissner G, Lu X. Dihydropyridine receptor-ryanodine receptor interactions in skeletal muscle excitation-contraction coupling. Biosci Rep. 1995 Oct;15(5):399–408. doi: 10.1007/BF01788371
  • Takeshima H, Iino M, Takekura H, et al. Excitation-contraction uncoupling and muscular degeneration in mice lacking functional skeletal muscle ryanodine-receptor gene. Nature. 1994 Jun 16;369(6481):556–9. doi: 10.1038/369556a0
  • Gamper N, Reznikov V, Yamada Y, et al. Phosphatidylinositol [correction] 4,5-bisphosphate signals underlie receptor-specific Gq/11-mediated modulation of N-type Ca2+ channels. J Neurosci. 2004 Dec 1;24(48):10980–92. doi: 10.1523/JNEUROSCI.3869-04.2004
  • Wu X, Kushwaha N, Albert PR, et al. A critical protein kinase C phosphorylation site on the 5-HT(1A) receptor controlling coupling to N-type calcium channels. J Physiol. 2002 Jan 1;538(Pt 1):41–51. doi: 10.1113/jphysiol.2001.012668
  • Deutsch C. (2003). The Birth of a Channel. Neuron, 40(2), 265–276. 10.1016/S0896-6273(03)00506-3
  • Hegde R S. (2022). The Function, Structure, and Origins of the ER Membrane Protein Complex. Annu. Rev. Biochem., 91(1), 651–678. 10.1146/annurev-biochem-032620-104553
  • O'Donnell J P, Phillips B P, Yagita Y, Juszkiewicz S, Wagner A, Malinverni D, Keenan R J, Miller E A and Hegde R S. (2020). The architecture of EMC reveals a path for membrane protein insertion. eLife, 9 10.7554/eLife.57887
  • Pleiner T, Tomaleri G Pinton, Januszyk K, Inglis A J, Hazu M and Voorhees R M. (2020). Structural basis for membrane insertion by the human ER membrane protein complex. Science, 369(6502), 433–436. 10.1126/science.abb5008
  • Miller-Vedam L E et al . (2020). Structural and mechanistic basis of the EMC-dependent biogenesis of distinct transmembrane clients. eLife, 9 10.7554/eLife.62611
  • Kiowski W, Erne P, Bühler FR. Use of nifedipine in hypertension and Raynaud’s phenomenon. Cardiovasc Drugs Ther. 1990 Aug;4(5):935–40. doi: 10.1007/BF02018296
  • Hockerman GH, Peterson BZ, Johnson BD, et al. Molecular determinants of drug binding and action on L-type calcium channels. Annu Rev Pharmacol Toxicol. 1997;37(1):361–396.
  • Langs DA, Strong PD, Triggle DJ. Receptor model for the molecular basis of tissue selectivity of 1, 4-dihydropyridine calcium channel drugs. J Comput Aided Mol Des. 1990 Sep;4(3):215–230. doi: 10.1007/BF00125011
  • Langs DA, Kwon YW, Strong PD, et al. Molecular level model for the agonist/antagonist selectivity of the 1,4-dihydropyridine calcium channel receptor. J Comput Aided Mol Des. 1991 Apr;5(2):95–106. doi: 10.1007/BF00129749
  • Mitterdorfer J, Wang Z, Sinnegger MJ, et al. Two amino acid residues in the IIIS5 segment of L-type calcium channels differentially contribute to 1,4-dihydropyridine sensitivity. J Biol Chem. 1996 Nov 29;271(48):30330–30335. doi: 10.1074/jbc.271.48.30330
  • Yamaguchi S, Okamura Y, Nagao T, et al. Serine residue in the IIIS5-S6 linker of the L-type Ca2+ channel alpha 1C subunit is the critical determinant of the action of dihydropyridine Ca2+ channel agonists. J Biol Chem. 2000 Dec 29;275(52):41504–11. doi: 10.1074/jbc.M007165200
  • Kass RS, Arena JP. Influence of pHo on calcium channel block by amlodipine, a charged dihydropyridine compound. Implications for location of the dihydropyridine receptor. J Gen Physiol. 1989 Jun;93(6):1109–1127. doi: 10.1085/jgp.93.6.1109
  • Döring F, Degtiar VE, Grabner M, et al. Transfer of L-type calcium channel IVS6 segment increases phenylalkylamine sensitivity of α1A(*). J Biol Chem. 1996 05 17;271(20):11745–11749. doi: 10.1074/jbc.271.20.11745
  • Kirtane MV, Bhandari A, Narang P, et al. Cinnarizine: A Contemporary Review. Indian J Otolaryngol Head Neck Surg. 2019 Nov;71(Suppl 2):1060–1068. doi: 10.1007/s12070-017-1120-7
  • Afdhal N, Zeuzem S, Kwo P, et al. Ledipasvir and sofosbuvir for untreated HCV genotype 1 infection. N Engl J Med. 2014 May 15;370(20):1889–1898. doi: 10.1056/NEJMoa1402454
  • Jacobson IM, Gordon SC, Kowdley KV, et al. Sofosbuvir for hepatitis C genotype 2 or 3 in patients without treatment options. N Engl J Med. 2013 May 16;368(20):1867–1877. doi: 10.1056/NEJMoa1214854
  • Lawitz E, Mangia A, Wyles D, et al. Sofosbuvir for previously untreated chronic hepatitis C infection. N Engl J Med. 2013 May 16;368(20):1878–1887. doi: 10.1056/NEJMoa1214853
  • Mangia A, Milligan S, Khalili M, et al. Global real-world evidence of sofosbuvir/velpatasvir as simple, effective HCV treatment: analysis of 5552 patients from 12 cohorts. Liver Int. 2020 Aug;40(8):1841–1852. doi: 10.1111/liv.14537
  • Lagrutta A, Zeng H, Imredy J, et al. Interaction between amiodarone and hepatitis-C virus nucleotide inhibitors in human induced pluripotent stem cell-derived cardiomyocytes and HEK-293 Cav1.2 over-expressing cells. Toxicol Appl Pharmacol. 2016 Oct 1;308:66–76. doi: 10.1016/j.taap.2016.08.006
  • Lagrutta A, Regan CP, Zeng H, et al. Cardiac drug-drug interaction between HCV-NS5B pronucleotide inhibitors and amiodarone is determined by their specific diastereochemistry. Sci Rep. 2017 Mar 22;7(1):44820. doi: 10.1038/srep44820
  • Millard DC, Strock CJ, Carlson CB, et al. Identification of drug-drug interactions in vitro: a case study evaluating the effects of sofosbuvir and amiodarone on hiPSC-Derived cardiomyocytes. Toxicol Sci. 2016 Nov;154(1):174–182. doi: 10.1093/toxsci/kfw153
  • Field MJ, Cox PJ, Stott E, et al. Identification of the alpha2-delta-1 subunit of voltage-dependent calcium channels as a molecular target for pain mediating the analgesic actions of pregabalin. Proc Natl Acad Sci U S A. 2006 Nov 14;103(46):17537–42. doi: 10.1073/pnas.0409066103
  • Domon Y, Arakawa N, Inoue T, et al. Binding characteristics and analgesic effects of Mirogabalin, a novel ligand for the α(2)δ subunit of voltage-gated calcium channels. J Pharmacol Exp Ther. 2018 Jun;365(3):573–582. doi: 10.1124/jpet.117.247551
  • Gee NS, Brown JP, Dissanayake VU, et al. The novel anticonvulsant drug, gabapentin (Neurontin), binds to the alpha2delta subunit of a calcium channel. J Biol Chem. 1996 Mar 8;271(10):5768–76. doi: 10.1074/jbc.271.10.5768
  • Bauer CS, Nieto-Rostro M, Rahman W, et al. The increased trafficking of the calcium channel subunit alpha2delta-1 to presynaptic terminals in neuropathic pain is inhibited by the alpha2delta ligand pregabalin. J Neurosci. 2009 Apr 1;29(13):4076–88. doi: 10.1523/JNEUROSCI.0356-09.2009
  • Cassidy JS, Ferron L, Kadurin I, et al. Functional exofacially tagged N-type calcium channels elucidate the interaction with auxiliary α2δ-1 subunits. Proc Natl Acad Sci U S A. 2014 Jun 17;111(24):8979–8984. doi: 10.1073/pnas.1403731111
  • Tran-Van-Minh A, Dolphin AC. The alpha2delta ligand gabapentin inhibits the Rab11-dependent recycling of the calcium channel subunit alpha2delta-2. J Neurosci. 2010 Sep 22;30(38):12856–67. doi: 10.1523/JNEUROSCI.2700-10.2010
  • Ramírez D, Gonzalez W, Fissore RA, et al. Conotoxins as tools to understand the physiological function of voltage-gated calcium (Ca(v)) channels. Mar Drugs. 2017 Oct 13;15(10):313. doi: 10.3390/md15100313
  • Schmidtko A, Lötsch J, Freynhagen R, et al. Ziconotide for treatment of severe chronic pain. Lancet. 2010 May 1;375(9725):1569–1577. doi: 10.1016/S0140-6736(10)60354-6
  • de Weille JR, Schweitz H, Maes P, et al. Calciseptine, a peptide isolated from black mamba venom, is a specific blocker of the L-type calcium channel. Proc Natl Acad Sci U S A. 1991 Mar 15;88(6):2437–40. doi: 10.1073/pnas.88.6.2437
  • Kuroda H, Chen YN, Watanabe TX, et al. Solution synthesis of calciseptine, an L-type specific calcium channel blocker. Pept Res. 1992 Sep-Oct;5(5):265–268.
  • Yasuda O, Morimoto S, Chen Y, et al. Calciseptine binding to a 1,4-dihydropyridine recognition site of the L-type calcium channel of rat synaptosomal membranes. Biochem Biophys Res Commun. 1993 Jul 30;194(2):587–594. doi: 10.1006/bbrc.1993.1862
  • Schleifer KJ. Comparative molecular modelling study of the calcium channel blockers nifedipine and black mamba toxin FS2. J Comput Aided Mol Des. 1997 Sep;11(5):491–501. doi: 10.1023/A:1007974124426
  • Kini RM, Caldwell RA, Wu QY, et al. Flanking proline residues identify the L-type Ca2+ channel binding site of calciseptine and FS2. Biochemistry. 1998 Jun 23;37(25):9058–9063. doi: 10.1021/bi9802723
  • Chen C-C, Gao S, Ai H-S, et al. Racemic X-ray structure of L-type calcium channel antagonist calciseptine prepared by total chemical synthesis. Sci China Chem. 2018 06 1;61(6):702–707. doi: 10.1007/s11426-017-9198-y