Blog

Closed-state inactivation and pore-blocker modulation mechanisms of human CaV2.2

Closed-state inactivation and pore-blocker modulation mechanisms of human CaV2.2

By Yanli Dong, Yiwei Gao, Shuai Xu, Yuhang Wang, Zhuoya Yu, Yue Li, Bin Li, Tian Yuan, Bei Yang, Xuejun Cai Zhang, Daohua Jiang, Zhuo Huang, and Yan Zhao

Excerpt from the article published in Cell Reports, Volume 37, Issue 5, 2021, 109931, ISSN 2211-1247 https://doi.org/10.1016/j.celrep.2021.109931

Editor’s Highlights

  • N-type voltage-gated calcium (CaV) channels mediate Ca2+ influx at presynaptic terminals in response to action potentials and play vital roles in synaptogenesis, and release of neurotransmitters.
  • The CaV2.2 is a High-voltage activated (HVA) channel that is exclusively expressed in central and peripheral neurons.
  • The structure of the CaV2.2 complex contains α1, α2δ1, and β1 subunits.
  • The α2δ1 subunit is divided into α2 and δ subunits, which are linked by a disulfide bond and associate with the α subunit by interacting with extracellular loops (ECL).
  • The α2δ1 subunits in both CaV2.2 and CaV1.1 complexes display similar binding geometry.
  • The α2δ1 exhibits a little tilt relative to the remaining complex because the ECL2 and ECL3 are directly packed with the α2δ1 subunit.

Authors’ Highlights

  • W768 on W-helix is a structural determinant for CSI of CaV2.2 channel
  • Voltage-sensing domain II was trapped in the resting state by a PIP2molecule
  • Ziconotide is located on the top of the selectivity filter blocking ion influx
  • Small molecules, PD173212 and CaV2.2 blocker 1, occupy the DIII–DIVfenestration site

Summary

N-type voltage-gated calcium (CaV) channels mediate Ca2+ influx at presynaptic terminals in response to action potentials and play vital roles in synaptogenesis, release of neurotransmitters, and nociceptive transmission. Here, we elucidate a cryo-electron microscopy (cryo-EM) structure of the human CaV2.2 complex in apo, ziconotide-bound, and two CaV2.2-specific pore blockers-bound states. The second voltage-sensing domain (VSD) is captured in a resting-state conformation, trapped by a phosphatidylinositol 4,5-bisphosphate (PIP2) molecule, which is distinct from the other three VSDs of CaV2.2, as well as activated VSDs observed in previous structures of CaV channels. This structure reveals the molecular basis for the unique inactivation process of CaV2.2 channels, in which the intracellular gate formed by S6 helices is closed and a W-helix from the domain II–III linker stabilizes closed-state inactivation. The structures of this inactivated, drug-bound complex lay a solid foundation for developing new state-dependent blockers for treatment of chronic pain.

Graphical abstract

Introduction

Voltage-gated calcium (CaV) channels are essential mediators to convert action potentials (APs) into an influx of Ca2+ ions, a crucial secondary messenger to regulate a variety of types of cellular event, such as muscle contraction, secretion of neurotransmitters, cell division, differentiation, and apoptosis (Catterall, 1991Catterall and Few, 2008Reuter, 1979Tsien et al., 1988). CaV channels are generally categorized into two groups according to their activation threshold, i.e., high-voltage activated (HVA) and low-voltage activated (LVA) CaV channels. Based on their sequence homology, CaV channels in mammals contain 10 members, which are further classified into three subfamilies (CaV1, CaV2, and CaV3) that conduct six types of Ca2+ current (L-, P/Q-, N-, R-, T-types) (Catterall et al., 2005Nowycky et al., 1985Snutch and Reiner, 1992). CaV2.2 is a HVA channel that is exclusively expressed in central and peripheral neurons (Olivera et al., 1994). It is predominantly located in the presynaptic terminals and mediates the Ca2+ influx that triggers neurotransmitter release at the fast synapses (Westenbroek et al., 19921998). Dysfunctions of CaV2.2 channels alter neuronal functions and lead to diseases, such as myoclonus-dystonia-like syndrome (Weiss, 2015). Moreover, the CaV2.2 channel plays a critical role in spinal nociceptive signaling, and thus it has become an important drug target for chronic pain treatment (Patel et al., 2018).

Molecular mechanisms of the CaV channels have been studied extensively for decades, including recent structural studies of L-type CaV1.1 isolated from rabbit skeletal muscle (Gao and Yan, 2021Wu et al., 20152016Zhao et al., 2019a) and human T-type CaV3.1 (Zhao et al., 2019b). These studies revealed structural features of the CaV channels in their inactivated state with all the voltage-sensing domains (VSDs; VSDI–VSDIV) adopting an activated “up” conformation. The structures of the CaV1.1 complex also elucidate the assembly of CaV channel with auxiliary β and α2δ subunits. A range of ligands was determined in complexes with CaV1.1 or CaV3.1, and these structural studies provide molecular bases for ligand recognition and facilitate further rational drug development targeting CaV channels. Despite these advances in structural studies on the CaV channels, more structural insight is desirable to fully understand the molecular mechanism of the CaV2.2 channel from the structural untouched CaV2 subfamily, because it shares only low sequence identity with the available structures from the other two subfamilies, and some fundamental mechanistic questions remain to be addressed at the structural level. For example, although the available structures of CaV channels have activated VSDs (Wu et al., 20152016Zhao et al., 2019a2019b), structures of these VSDs in their resting state are required to fully understand the gating mechanism. Moreover, CaV2.2 inactivates more rapidly than the closely related CaV2.1 isoform, exhibits a distinct closed-state inactivation (CSI) mechanism during repolarization (Jones et al., 1999Patil et al., 1998Thaler et al., 2004), and is modulated by phosphatidylinositol 4,5-bisphosphate (PIP2) molecules (Hille et al., 2015Keum et al., 2014Rodríguez-Menchaca et al., 2012bSuh et al., 2012Vivas et al., 2013Wu et al., 2002). Further structural analysis is required to understand the underlying molecular mechanisms for these unique functional properties of CaV2.2, which are essential to its key physiological roles in neurotransmitter release and regulation of synaptic transmission.

Here, we purified the recombinant human CaV2.2 (also called the α1 subunit) in complex with the β1 and α2δ1 subunits and determined their complex structure using the single-particle cryo-electron microscopy (cryo-EM) method. This structure reveals the asymmetric activation of the four VSDs with the VSDII in its resting state, evidently stabilized by binding of a PIP2 lipid molecule. In addition, we identify a unique α helix that locks the intracellular gate of CaV2.2 in closed conformation, contributing to the CSI of CaV2.2. Specific binding of a state-dependent channel blocker in the pore provides a template for future development of next-generation, nonaddictive analgesic drugs. These results provide insights into the activation and inactivation mechanisms of human CaV channels and point the way to future drug discovery. During the preparation of this manuscript, the structure of human CaV2.2 complex was published (Gao et al., 2021). Despite high similarity between the CaV2.2 complex structures, we provide more mechanistic interpretation about CSI of the CaV2.2 validated by electrophysiological experiment and also elucidate a binding site for small-molecule inhibitors targeting CaV2.2.

Results and Discussion

Structure determination and architecture of the CaV2.2 complex

To investigate the architecture of the N-type CaV2.2 complex, we co-expressed human N-terminal truncated CaV2.2, full-length wild-type (WT) α2δ1, and full-length WT β1 in HEK293 cells. To monitor the expression of the complex during purification, we fused CaV2.2 with a C-terminal GFP-Twinstrep tag. In order to enhance the expression level, we truncated the intrinsically disordered N-terminal region (residues 1–64), and we denote this construct as CaV2.2EM. We carried out whole-cell patch-clamp experiments in HEK293T cells to characterize the channel properties of both human WT full-length CaV2.2 and CaV2.2EM constructs in the presence of auxiliary β1 and α2δ1 subunits. The CaV2.2EM complex shows indistinguishable gating properties compared with the full-length CaV2.2, in terms of voltage-dependent activation and steady-state inactivation (Figures 1A, 1B, and S1A–S1C). Thus, the CaV2.2EM construct was subjected to further structural and functional studies. The purified CaV2.2EM-α2δ1-β1 complex (CaV2.2 complex) displaying monodisperse peaks on both size-exclusion chromatography (SEC) and SDS-PAGE further confirmed that all three subunits were present in the purified complex sample (Figures S1D and S1E).

Figure 1. Architecture of the CaV2.2 complex
(A) Normalized conductance-voltage (G/V) relationship and the Boltzmann fits for the N-terminal truncated CaV2.2EM complex and full-length wild-type (WT) CaV2.2 complex. HEK293T cells expressing the complex were stimulated with 200-ms depolarizing pulses between −60 and 50 mV in steps of 10 mV from a holding potential of −100 mV.
(B) Steady-state inactivation of the WT CaV2.2 complex and the CaV2.2EM complex. Cells were stepped from a holding potential of −100 mV to pre-pulse potentials between −100 and −15 mV in 5-mV increments for 10 s. Black, CaV2.2EM complex; red, full-length WT CaV2.2 complex.
(C and D) The density map and model of the CaV2.2EM complex as seen in parallel to the membrane plane. The α2δ1 and β1 subunit, C-terminal domain (CTD), extracellular loops (ECLs), alpha-interacting domain (AID), and transmembrane helices S1II–S4II in VSDII were labeled. The α1 subunit is colored in light blue (DI), violet (DII), deep blue (DIII), magenta (DIV), and gray (CTD), respectively. The β1, α2, and δ1 subunits are colored in light gray, tomato red, and turquoise, respectively. N-glycans are displayed and colored in gold.

To elucidate its architecture, we carried out cryo-EM study of the CaV2.2 complex and obtained the final reconstruction at 2.8-Å resolution for the CaV2.2 complex (Figures 1C and S1F–S1J; Table S1). Cryo-EM map of the CaV2.2 is rich in structural features, including densities for side chains, N-glycans, disulfide bonds, and associated lipid molecules, which enabled us to unambiguously build de novoatomic models of the CaV2.2 complex (Figures 1D and S1J). Within its overall size of approximately 118 Å × 113 Å × 169 Å, the structure of the CaV2.2 complex contains α1, α2δ1, and β1 subunits and closely resembles the classic shape of CaV1.1.

The α subunit is composed of four transmembrane domains, DI–DIV, which form an ion conducting pore in a domain-swapped fashion. Each domain is composed of six helices (S1–S6), of which S1−S4 helices form the VSD. Four VSDs (VSDI–VSDIV) encircle the central pore formed by S5 and S6 helices from all four domains. The β1 subunit comprises a Src homology 3 (SH3) domain and a guanylate kinase (GK) domain. The GK domain, through its alpha-interacting domain (AID) helix, interacts with the CaV2.2 (i.e., the α subunit). Nevertheless, the GK domain was not well resolved in our cryo-EM map, presumably because of conformational heterogeneity (Figures 1C and 1D). The α2δ1 subunit is divided into α2 and δ subunits, which are linked by a disulfide bond (C404-C1059) and associate with the α subunit by interacting with extracellular loops (ECLs; in particular the loops between S5 and S6 helices) ECLI, ECLII, and ECLIII, and the S1−S2 loop from VSDI. The structure of human α2δ1 in our complex is nearly identical with rabbit α2δ1 resolved in the CaV1.1 complex, with root-mean-square deviation (RMSD) of ∼1.1 Å.

Ion conduction pore of the CaV2.2 complex

The central pore is made up of helices S5 and S6 from all four domains of the α subunit (Figures 2A and 2B). In each domain, a re-entrant P loop is located between the S5 and S6 helices and contains two short helices, P1 and P2, connected by a short linker. Four P loops form a funnel-like shape and line the outer entry to the pore. Four highly conserved glutamate residues, E314, E663, E1365, and E1655, form a selectivity-filter ring at the bottom of the funnel, which is one of the narrowest segments of the central pore (Figure 2C) (Smart et al., 1996). These four acidic residues, together with surrounding negatively charged residues, create a strong negative electric field strength to attract cations and repel anions. This pore region also functions as a selectivity filter to discriminate Ca2+ from other cation ions. We did identify a strong density within the selectivity filter and close to the glutamate cluster (Figures 2D and 2E). Presumably, this piece of density represents a calcium ion, consistent with observations in previous structures of CaV channels (Gao and Yan, 2021Wu et al., 20152016Zhao et al., 2019a2019b). Thus, on the extracellular side, the S5 and S6 helices and the P loops contribute to the formation and stability of the selectivity filter.

Figure 2. Ion conduction pore of the CaV2.2
(A and B) Ion permeation pathways calculated by the program HOLE were shown in dots. The selectivity filter and S5–S6 helices are shown in cartoon and viewed in parallel to the membrane plane.
(C) Plot of pore radii for CaV2.2 complex. Vertical dashed line marks 1.0-Å pore radius.
(D and E) “Top-down” and side view of the selectivity filter. The selectivity filter ring of four glutamate residues from the four domains of CaV channel were shown in sticks. A cation ion is shown as a gray sphere, overlaid with corresponding EM density colored in marine.
(F) The intracellular gate formed by four S6 helices viewed from intracellular side. Hydrophobic residues are shown in sticks. The segments from DI, DII, DIII, and DIV in the CaV2.2 complex are colored in light blue, violet, deep blue, and magenta, respectively.

On the intracellular side, the pore-lining S6 helices comprise an intracellular gate to control ion permeation (Figure 2F). The CaV2.2 complex was determined in the absence of membrane potential and thus is likely to represent a depolarized state. The cytoplasmic ends of the S6 helices have converged to form a hydrophobic seal through a cluster of hydrophobic residues (Figure 2F). In each of the four domains, the S5 helix is positioned proximal to the S6 helix and connects to S4 of VSD through an amphipathic horizontal helix (termed S4–S5 helix, ∼16 residues). This S4–S5 helix interacts with the S6 helix and thus couples S4 movement in response to depolarization with pore opening.

The central pore of the CaV2.2 channel is further compared with those from the CaV1.1 (PDB: 5GJW) and CaV3.1 (PDB: 6KZP) channels. The S5–S6 ECLs exhibit significant conformational difference among these structures (Figures S2A, S2B, S2D, and S2E). The CaV2.2 α subunit possesses shorter ECLI, ECLII, and ECLIII, but this structural difference does not hamper binding of the α2δ1 subunit. The S1–S2 loops of VSDII, also involved in interactions with the α2δ1 subunit, are nearly identical in both CaV2.2 and CaV1.1. Therefore, the α2δ1 subunits in both CaV2.2 and CaV1.1 complexes display similar binding geometry. In contrast, the ECLs of the CaV3.1 channel exhibit remarkable structural differences from the corresponding loops of CaV1.1 and CaV2.2, thus giving rise to incompatibility of the α2δ1 subunit with CaV3.1 (Figures S2B and S2E). Despite these structural differences occurring at ECLs, the transmembrane helices S5 and S6, as well as the selectivity filter of CaV2.2, are superimposable with CaV1.1 and CaV3.1, yielding RMSD of ∼1.8 Å for CaV1.1 (for 332 Cα atom pairs) and ∼1.6 Å for CaV3.1 (for 300 Cα atom pairs), respectively (Figures S2C and S2F). In particular, the intracellular gate formed by the four-helix bundle of S6 helices is nearly identical in all three structures, assuming a closed form (Figure S2G). The S6II helix of the CaV2.2 extends into the cytoplasm and is longer than that of the CaV1.1. Consequently, the C terminus of the S6II helix forms a close contact with the β1 subunit in the CaV2.2 complex, which is not observed in the structure of CaV1.1. Furthermore, the cytoplasmic part of the S6 helix bends by 13° compared with that in CaV1.1 (Figure S2H), which may affect gate opening and represent a potential regulatory mechanism of the β subunit.

Voltage sensor of the N-type CaV channel

A hallmark feature of the CaV channel is to open or close the channel in response to changes in membrane potential. Surrounding the central pore, the four VSDs play essential roles in converting the electrostatic signal into a conformational change of the intracellular gate. Each VSD is composed of S1–S4 transmembrane helices. Each S4 helix contains several positively charged residues, either arginine or lysine, spaced at intervals of three. CaV2.2 is a depolarization-activated channel. Upon depolarization of the membrane potential, the gating charges on the S4 helix move toward the extracellular cavity of the VSD. The two cavities are separated by a highly conserved hydrophobic constriction site. When the gating charges pass through this hydrophobic seal, existing interactions of the gating charges with hydrophilic and negatively charged residues on surrounding helices are disrupted on one side, and new interactions are formed on the other side. The displacement of the S4 helices across the electrostatic field of the membrane potential would induce lateral movement of the S4–S5 linkers, resulting in disengagement between S5 and S6 helices and thus the channel opening.

In our structure, the S4 helices adopt a 310-helix conformation, which is consistent with observations from previous structures of voltage-gated ion channel, including CaV and NaV channels, allowing the side chains of gating-charge residues to be aligned on the same side of the helix surface. Using the central pore as a reference, superimposition of the structure of CaV2.2 onto that of CaV1.1 shows close overlap of VSDI, VSDIII, and VSDIV, demonstrating that they are stabilized in the same presumably activated state (Figures S3A–S3E). In contrast, VSDII of CaV2.2 exhibits a strikingly different conformational arrangement compared with that from the CaV1.1 structure. Taking a closer look at the four VSDs, we found that four, four, and three gating charges are located at the extracellular aqueous cavities of VSDI, VSDIII, and VSDIV, respectively, and interact with polar residues residing on nearby S1–S3 helices, consistent with observations from the available structures of CaV channels (Wu et al., 20152016Zhao et al., 2019a2019b) (Figures 3A, 3C, 3D, and S3C–S3E). Strikingly, in VSDII, only one arginine (R578) on the S4 helix is located on the extracellular side of the hydrophobic seal, but the other four conserved gating charges (R581, R584, K587, and K590) are located in the intracellular cavity, indicating that VSDII is stabilized in the resting state. Among these basic residues, K590 is located at a position where S4 becomes unwound and is completely exposed to the solvent (Figure 3B). The asymmetric activation of CaV2.2 VSDs suggests that each individual VSD may sense membrane potential asynchronously, which is reported to be important for eukaryotic NaV activation and inactivation (Chanda and Bezanilla, 2002).

Figure 3. Structural analysis of the voltage-sensing domains
(A–D) Voltage-sensing domains from DI (A), DII (B), DIII (C), and DIV (D) are shown in cartoon. The gating-charge residues (R1–R6) on S4 helix and residues from surrounding helices are shown in sticks.
(E and F) Structural comparison of the VSDII at resting state from CaV2.2 complex and activated state from CaV1.1 complex, viewed parallel (E) or perpendicular (F) to the membrane plane. The gating-charge residues are shown as spheres, colored in deep blue (CaV2.2) or light blue (CaV1.1).
(G and H) Superimposition of the VSDIIR and VSDIIA using S5–S6 helices as a reference, viewed perpendicular (G) or parallel (H) to the membrane plane.
(I) Domain II of the CaV2.2 complex, overlaid with an electrostatic potential surface on the VSDII. The PIP2 molecular is shown as spheres.

To investigate the conformational change of the VSDII of CaV2.2 upon depolarization, we overlaid its structure in the resting state (VSDIIR) onto that of CaV1.1 in the activated state (VSDIIA) and found that these VSDII structures are highly superimposable in helices S1–S3 (Figures 3E and 3F). In contrast, the S4 helix slides to the intracellular side in the resting state. Consequently, the extracellular half of the S3 helix bends ∼15° toward S4 in the resting state (Figure 3E). Using the S5 and S6 helices as a superimposition reference, the VSDII undergoes substantial conformational change relative to the central pore between these two states. In particular, in transition from the activated state to the resting state, the cytoplasmic sides of both S1 and S2 helices in CaV2.2 rotate toward the central pore (Figure 3G). Consequently, the intracellular and extracellular termini of the S3 helices are remarkably displaced, moving toward the central pore by 4 and 9 Å, respectively. In addition to sliding inward, the S4 helices also slightly shift toward the pore domain (Figure 3H). Interestingly, even though distinguishable conformational change occurs in VSDII, the extracellular end of the S1–S2 helix hairpin forms similar interactions with helices S5 and P1 from domain III. It appears that this interaction is stabilized by a cholesterol derivative, cholesteryl hemisuccinate (CHS), which is clearly resolved in both our structure of CaV2.2 and the structure of CaV3.1 (Figure S3F), consistent with earlier reports suggesting that cholesterol is important to regulating activity of CaV channels (Purcell et al., 2011Xia et al., 2008).

Compared with S4 in the activated VSDII from CaV1.1, the S4 helix in the resting state undergoes a sliding movement by two helical turns, ∼13 Å, toward intracellular side (Figure 3H). The displacement of S4 is comparable with previous observations in structures of TPC1 channel and NaVAb channel (Figures S3G and S3H) (Guo et al., 2016Kintzer and Stroud, 2016Wisedchaisri et al., 2019). With the exception of R5 (K590) of CaV2.2, all gating charges are facing to one side, and no rotation about the helix axis is observed, which support a sliding-helix model of voltage-dependent gating (Wisedchaisri et al., 2019Zhang et al., 2018). As a consequence of the S4 sliding inward, the S4–S5 linker is tilted by 11° toward the intracellular side, and this movement generates more contacts between the S4–S5 linker and S6 helix and thus stabilizes the inner gate in its closed state.

An extra density was identified laying above the S4–S5 linker of domain II in the cryo-EM map, which presumably represents a lipid with two hydrophobic tails and a “palm”-like head group. Its head group is located at the intracellular cavity of the VSDII and flanked by S3 and S4 helices (Figure 3I). Its hydrophobic tails project toward the extracellular side of the membrane (Figure S3I). This lipid appears to act as a plug and would block the S4 helix from sliding upward, reminiscent of the PIP2that attenuates opening of the KV1.2 channel by interacting with the S4–S5 linker (Rodriguez-Menchaca et al., 2012a). We attempted to fit a PIP2 into the density. Whereas its hydrophobic tails and inositol ring agree well with the density, the two phosphate groups attached to the inositol ring could not be well resolved (Figure S3I). Considering that PIP2 is predominant in the inner leaflet of the plasma membrane and reported to positively shift the activation curve of CaV2.2, which are consistent with our structural observations, we postulate that this lipid molecule is a PIP2. In a closer look at the PIP2 lipid binding site, several positively charged residues from helices S0, S4, and S4–S5 linker provide a highly positively charged environment for PIP2 binding (Figures S3J and S3K). Similar local environment is not observed in the other three VSDs (Figure S3J), potentially providing a mechanism by which PIP2 selectively binds to VSDII. In addition, the inositol ring of PIP2 is positioned proximal to the AID helix of the β1 subunit, and this observation is consistent with previous investigations indicating that regulation of the CaV channel by PIP2 depends on type of β subunit present (Suh et al., 2012).

CSI mechanism of the CaV2.2 channels

The N-type CaV channel bears a preferential voltage-dependent CSI (Jones et al., 1999Patil et al., 1998), featuring Ca2+ insensitive U-shaped inactivation curve; substantial inactivation accumulated during interpulse in a two-pulse protocol and cumulative inactivation during voltage-clamped AP trains (Patil et al., 1998). In our structure, the intracellular gate is comparable with that in structures of the CaV1.1 and CaV3.1, and it appears to be stabilized at a closed state (Figure S2G). Strikingly, an additional α helix was unexpectedly revealed in our structure lying underneath the intracellular gate and flanked by the four S6-helix bundle, forming a 36° angle to the membrane plane (Figure 4A). This structural element has not been observed previously in CaV channels. The high-resolution cryo-EM map allowed us to align the amino acid sequence with the model with high confidence. This helix is composed of amino acid residues A764 to S783, which are located in the DII-IIIlinker (Figure 4B). The side chain of W768 points upward into the intracellular gate, forms extensive hydrophobic interactions with residues from S6 helices (Figures 4B and 4C), and thus stabilizes the intracellular gate in its closed state. Thus, we term this six-turn helix as the W-helix. In addition to W768 serving as a plug inserting into and blocking the intracellular gate, the other residues from the W-helix extensively interact with the S6 helices, including hydrogen bond and electrostatic interactions, to further strengthen the association of the W-helix with the intracellular gate (Figures 4B and 4C). We speculate that the newly discovered W-helix functions as a blocking lid, similar to the “hinged lid” model for fast sodium channel inactivation, and is important for the stability of the intracellular gate in the closed/inactivated state of CaV2.2.

Figure 4. Closed-state inactivation mechanism of CaV2.2 complex
(A) The W-helix is composed of 764ARSVWEQRASQLRLQNLRAS783 and determined underneath the intracellular gate of the CaV2.2. The CaV2.2 is shown in ribbon and overlaid with transparent surfaces. The S6 helices are shown in cartoon and highlighted. The W-helix is shown in cartoon and overlaid with a transparent surface, colored in purple.
(B) Interactions between the W-helix and the intracellular gate viewed from the intracellular side. The critical residues involved in the interactions were shown in sticks. The N and C termini of the W-helix is indicated.
(C) The W768 from W-helix stabilized the intracellular gate at closed state. The S6 helices are shown as ribbon. The W-helix is shown in cartoon and overlaid with a transparent gray surface.
(D) Steady-state activation and inactivation for the CaV2.2EM complex and its mutants. Activation: EM, n = 13; ΔW-helix, n = 8; W768Q, n = 6; inactivation: EM, n = 8; ΔW-helix, n = 6; W768Q, n = 6.
(E) Recovery of close-state inactivation for the CaV2.2EM complex and its mutants. Cells were held to –40 mV for 200 ms and then stepped to –100 mV for indicated time delay (4–512 ms), followed by a +10 mV test pulse (35 ms) (EM, n = 9; ΔW-helix, n = 7; W768Q, n = 7).
(F) Ratio of Cav2.2 channels inactivation calculated from the first spike eliciting maximal current to the other spikes in the AP trains (EM, n = 8; ΔW-helix, n = 6; W768Q, n = 8). Black, CaV2.2EM complex; purple, W768Q; orange, ΔW-helix. p < 0.05.

To investigate the functional role of the W-helix and to validate our hypothesis, we designed two mutants, including a deletion of the W-helix (CaV2.2ΔW-Helix) and a point mutation, W768Q (CaV2.2W768Q), and we performed whole-cell patch-clamp recording experiments with Ba2+ as charge carrier. The voltage activation curves of these two mutants were indistinguishable from the WT CaV2.2 complex, suggesting the W-helix does not participate in channel activation. In sharp contrast, the steady-state inactivation curves of these two variants exhibit a positive shift of ∼24 mV (Figures 4D, S4A, and S4B), indicating the W-helix and W768 are critical for the CSI. To investigate recovery from CSI, the membrane potential of pre-pulse was held at –40 mV, where the channel has not opened yet. In this protocol, recovery of the CaV2.2ΔW-Helix variant is significantly faster than that of the WT complex (Figures 4E and S4C), which further supports the conclusion that the W-helix plays a vital role in CSI of the CaV2.2 channel. These results are also in line with a previous functional characterization showing that the domain II–III linker is crucial for CSI of CaV2.2 (Kaneko et al., 2002Thaler et al., 2004).

To further validate our finding in a more physiologically relevant condition, we repetitively activated the CaV2.2 channels with AP trains that were recorded in hippocampal CA1 neurons in whole-cell current-clamp configuration. As observed in previous reports (Patil et al., 1998Thaler et al., 2004), we found that WT CaV2.2 channels exhibited cumulative inactivation in response to AP trains (Figures 4F and S4D). Interestingly, CaV2.2ΔW-Helix variant inactivated more slowly, suggesting that W-helix-regulated CSI contributes to CaV2.2 channel inactivation during the AP trains and may play an important role in short-term synaptic plasticity (Catterall et al., 2013). Collectively, our results show that the W-helix is essential for CSI of the CaV2.2 channel. Interestingly, the W-helix is conserved only in the CaV2 subfamily (Figure S4E), suggesting that CaV2.1 and CaV2.3 channels may have similar inactivation mechanisms.

Recognition and specificity of ziconotide to the CaV2.2 channel

The ziconotide derived from ω-conotoxin MVIIA is able to potently inhibit the CaV2.2 with high selectivity and used to treat severe chronic pain. To gain insight into how the ziconotide blocks the CaV2.2 complex, we determined the ziconotide-bound CaV2.2 complex at 3.0-Å resolution (CaV2.2MVIIA) (Figures S5A and S5D). The ziconotide is positioned right above the selectivity filter and directly blocks the outer entry of calcium, in line with previous speculation that the ziconotide acts as a channel blocker (Figure 5A). The ziconotide is composed of 25 amino acids with six cysteine residues. A serial of structural analysis by NMR spectroscopy revealed that the ziconotide is compact folded, primarily defined by three disulfide bonds, including C1-C16, C8-C20, and C15-C25 (Kohno et al., 1995), which also present in ziconotide from the structure of Cav2.2MVIIA complex (Figure 5B). Interestingly, upon association with the CaV2.2 channel, the main chain and side chains of the ziconotide undergoes substantial conformational changes compared with the NMR structure (PDB ID: 1OMG), suggested by RMSD of ∼1.4 Å for 100 main-chain atom pairs and ∼2.8 Å for 178 side-chain included atom pairs (Figure S6A). We speculate that such conformational change is essential for ziconotide to associate with CaV2.2 and thus block channel activity. Taking a close look at the contacts between the ziconotide and CaV2.2, we found interactions are formed in electrostatic and shape complementary manners. The structural elements of P1–P2 helices and ECLs play important roles to stabilize the ziconotide (Figures 5C–5E). In particular, the R10 of the ziconotide (R10Z) is posed proximal to the selectivity-filter ring of glutamate and is important for the blockage of the ion pathway (Figure 5C). The residues K4Z, R21Z, and R24Z assume extensive electrostatic interactions with D1627 and D1629 from ECL4. In an earlier report, substitution of K2Z or Y13Z by Ala results in a remarkable reduction in activity (Kim et al., 1995). We found that the K2Z contacts with ECL2 and ECL3 by electrostatic interaction with E640, E1314, and E1330 in our structure. The hydroxyl group of Y13Z engages in hydrogen bonds with Q658 and D664, and the benzene ring of Y13Z, together with L11Z and M12Z, forms extensive van der Waals interaction with surrounding side chain and backbone from the CaV2.2 channel (Figure 5E).

Figure 5. Structure basis for blockade of CaV2.2 by ziconotide
(A) Overall structure of the CaV2.2-ziconotide complex. Ziconotide shown in ribbon and half-transparent surface colored in cyan.
(B) The structure of ziconotide. Amino acid sequence of ziconotide with three pairs of disulfide bonds linked (top), and the structure of ziconotide showed side chains of surface residues in sticks (bottom).
(C) Ziconotide blocks selectivity filter of CaV2.2. The positively charged residues of ziconotide and negatively charged residues of CaV2.2 with side chains shown in sticks; calcium ion is shown in green spheres.
(D and E) Detailed interactions between ziconotide and CaV2.2. ECLIII mediates ziconotide binding (D), and ziconotide binding is stabilized by P loops of CaV2.2 (E). Key residues shown as side chains in sticks.

We also compared structure of the Cav2.2MVIIA complex with apo-form CaV2.2 complex. It turns out the CaV2.2 undergoes only subtle conformational change upon the ziconotide association. The majority of the model, including VSDs and helices S5–S6, are fairly superimposable between these two structures. The ECLs, including ECL1, ECL2, and ECL3, surrounding the ziconotide are slightly expanded to enlarge the ziconotide binding cavity (Figures S6B and S6C). Consequently, the α2δ1 exhibits a little tilt relative to the remaining complex, because the ECL2 and ECL3 are directly packed with the α2δ1 subunit (Figure S6D). To assess the binding specificity of the ziconotide, the structures of L-type CaV1.1 and T-type CaV3.1 are superimposed on the structure of Cav2.2MVIIA complex. We found the ECL1 of CaV1.1 and CaV3.1 is longer than that in CaV2.2, pointing to the ziconotide binding cavity and resulting in direct steric clashes with the ziconotide in CaV2.2 (Figures S6E and S6F), suggesting the ECLs are critical for selectivity of the ziconotide.

Blockage of CaV2.2 by PD173212

The PD173212 is a potent and selective channel blocker of CaV2.2, consisting of phenyl, tyrosinamide, and 4-tertbutylbenzyl groups, which has been reported to be efficacious in the audiogenic seizure mouse model (Hu et al., 1999) (Figure 6A). We determined the structure of the CaV2.2 in complex with the PD173212 at 3.0-Å resolution (Figures S5B and S5E), which clearly shows two connected strip-shaped densities located in the central cavity underneath the selectivity filter and nearly parallel to the membrane plane, which are compatible with the PD173212 structure (Figures 6B–6D). Taking a close look at the PD173212 binding site, the interaction is primarily mediated by extensive hydrophobic interaction and van der Waals interactions. The tert-butyl and isobutyl groups of tyrosinamide project into the central cavity and are stabilized by interactions with hydrophobic residues from surrounding helices, such as F345, L352, V1412, L1700, F1703, etc. These two groups are positioned right underneath the selectivity filter, implying their important role in blocking ion influx. The 4-tertbutylbenzyl group occupies the DIII-DIVfenestration, and its benzene ring is stabilized by forming a π-π interaction with the surrounding aromatic residues (Figure 6E). The phenyl group also contributes to stabilize the C-terminal 4-tertbutylbenzyl group by a π-π stacking. Upon PD173212 association, the conformation of the pore module and majority of side chains involved in the ligand binding are unaltered, except the three aromatic residues, including Y1289, F1411, and F1693, which are around the DIII-DIV fenestration site (Figure 6F). The side chains of all three residues undergo discernable conformational change and thus contribute to stabilize the phenyl group of the PD173212. Strikingly, two of the three conformationally changed residues, Y1289 and F1693, are substituted by T935-M1366 and C1396-L1813 in CaV1.1 and CaV3.1, respectively (Figures S7C and S7D). Considering the side fenestration is critical for association of pore blocker with channel (Gamal El-Din et al., 2018), these substitutions suggest a possible mechanism through which PD173212 could selectively inhibit CaV2.2. In contrast with PD173212, the pore blockers diltiazemand verapamil for CaV1.1 bind in the middle of the central cavity, the pore blocker Z944 for CaV3.1 extends into the DII-DIII fenestration, and the binding position of PD173212 extends toward the DIII-DIV fenestration (Figure S7B), providing a site for rational drug design to minimize off-target side effects.

Figure 6. Structure basis for inhibition of CaV2.2 by blocker PD173212
(A) Chemical structure of PD173212.
(B) The cryo-EM density shown in blue mesh for PD173212 in sticks.
(C) PD173212 binding site in CaV2.2. Top-down (left panel) and side (right panel) view of CaV2.2 with PD173212 shown in spheres. The yellow and black dashed squares indicate the areas shown in (D) and (E), respectively.
(D and E) Detailed binding site for PD173212 viewed from extracellular and side, respectively. Key residues shown as side chains in sticks.
(F) Comparison of the PD173212 binding site with apo CaV2.2 (colored in gray) structure. Residues with conformational changes are labeled in red.

Blockage of CaV2.2 by CaV2.2 blocker 1

A novel and potent pyrazolyltetrahydropyran N-type calcium channel blocker, termed as CaV2.2 blocker 1, has been identified recently, and oral administration of CaV2.2 blocker 1 in a rat model showed promising efficacy in reducing inflammatory and neuropathic pain (Wall et al., 2018). The CaV2.2 blocker 1 consisted of 4-ketopyran, pyrazole, and two modified benzene groups (Figure 7A). Our cryo-EM map clearly defined the binding site of the CaV2.2 blocker 1 within the CaV2.2 channel (Figures 7B, S5C, and S5F). Its 4-ketopyran group points to the central cavity stabilized by direct contacts with adjacent hydrophobic residues and thus blocks the ion pathway. One of the benzene groups is extended to the DIII-DIVfenestration and stabilized by extensive hydrophobic interactions (Figure 7C). Addition of the methyl group enlarges the size of the benzene group, consequently resulting in close pack with residues from S5III (Figures 7D and 7E), supporting that this methyl group is critical for potency of the CaV2.2 blocker 1 (Wall et al., 2018). Interestingly, although the chemical structures of the PD173212 and CaV2.2 blocker 1 are different, they share an overlapped binding site close to the DIII-DIVfenestration, suggesting this fenestration site could be a hotspot for future drug designment (Figure S7E). Similar to the PD173212, association of the CaV2.2 blocker 1 does not induce significant conformational change of the CaV2.2 channel, and only the side chain of the F1693 exhibits dramatic conformational change (Figure 7F), suggesting that F1693 acts as a “door,” a critical residue for recognizing and stabilizing N-type CaV channel blocker.

Figure 7. Structure basis for inhibition of CaV2.2 by CaV2.2 blocker 1
(A) Chemical structure of CaV2.2 blocker 1.
(B) The cryo-EM density is shown in blue mesh for CaV2.2 blocker 1 (green) in sticks.
(C) CaV2.2 blocker 1 binding site in CaV2.2. Top-down (left panel) and side (right panel) view of CaV2.2 with CaV2.2 blocker 1 shown in green spheres. The yellow and black dashed squares indicate the areas shown in (D) and (E), respectively.
(D and E) Detailed binding site for CaV2.2 blocker 1 viewed from extracellular and side, respectively. Key interaction residues shown as side chains in sticks.
(F) Comparison of the CaV2.2 blocker 1 binding site with apo CaV2.2 (colored in gray) structure. Residues with conformational changes are labeled in red.

Concluding remarks

Here, we elucidated the structures of the human CaV2.2 complex in apo, ziconotide-bound, PD173212-bound, and CaV2.2 blocker 1-bound states. These structures show the VSDII is trapped at resting state by a PIP2 molecule, featuring only one gating charge located at the extracellular cavity. A W-helix is determined underneath the closed intracellular gate formed by S6 helices, with the W768 inserting into and stabilizing the closed gate. Our electrophysiological experiments demonstrated that the W768 and W-helix are fundamentally important for the CSI of CaV2.2. Moreover, the structure of the ziconotide-bound complex illustrated that ziconotide is located above the selectivity filter and acts as channel blocker to inhibit CaV2.2. By comparing with structures of the CaV1.1 and CaV3.1, we conclude that the ECLs are structural determinants for the ziconotide to specifically bind with the CaV2.2 channel. Furthermore, the complex structures of CaV2.2 with small molecular ligands PD173212 and CaV2.2 blocker 1 reveal the DIII-DIV fenestration site is critical for pore blocker binding and could be a crucial position for further designment of a CaV2.2-specific drug to treat chronic pain.