Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Feb 1.
Published in final edited form as: Nature. 2021 Jul 7;596(7870):143–147. doi: 10.1038/s41586-021-03699-6

Structure of human Cav2.2 channel blocked by the pain killer ziconotide

Shuai Gao 1,2, Xia Yao 1,2, Nieng Yan 1,3
PMCID: PMC8529174  NIHMSID: NIHMS1744430  PMID: 34234349

Abstract

The neuronal-type (N-type) voltage-gated calcium (Cav) channels, which are designated Cav2.2, have an important role in the release of neurotransmitters13. Ziconotide is a Cav2.2-specific peptide pore blocker that has been clinically used for treating intractable pain46. Here we present cryo-electron microscopy structures of human Cav2.2 (comprising the core α1 and the ancillary α2δ−1 and β3 subunits) in the presence or absence of ziconotide. Ziconotide is thoroughly coordinated by helices P1 and P2, which support the selectivity filter, and the extracellular loops (ECLs) in repeats II, III and IV of α1. To accommodate ziconotide, the ECL of repeat III and α2δ−1 have to tilt upward concertedly. Three of the voltage-sensing domains (VSDs) are in a depolarized state, whereas the VSD of repeat II exhibits a down conformation that is stabilized by Cav2-unique intracellular segments and a phosphatidylinositol 4,5-bisphosphate molecule. Our studies reveal the molecular basis for Cav2.2-specific pore blocking by ziconotide and establish the framework for investigating electromechanical coupling in Cav channels.


Cav2.2 channels are also known as the N-type Cav channels for their role in transmitter release in the central and peripheral nervous systems. Combining different splice forms of the voltage-dependent Ca2+ conducting core subunit α1 with auxiliary subunits—including the extracellular α2δ and the cytosolic β subunits—yields various Cav2.2 heteromers that exhibit distinct membrane distribution and biophysical properties1,2,7,8. Cav2.2 channels in the primary afferent terminals are involved in pain signalling. Suppression of Cav2.2 activity thus represents a strategy for the development of analgesic agents911.

Cav2.2 can be potently and selectively inhibited by a number of peptide toxins, some of which have been exploited for pharmacological applications3,1214. Among these, the ω-conotoxin MVIIA was developed to produce ziconotide (also known as SNX-111 and Prialt), which has been approved by the US Food and Drug Administration for the treatment of severe pain46. However, its intrathecal application and adverse effects have limited the broad use of ziconotide15,16. High-resolution structures of ziconotide-bound Cav2.2 are required to elucidate the molecular basis for the subtype-specific inhibition and to facilitate drug discovery.

Cryo-electron microscopy (cryo-EM) structures of rabbit Cav1.1 and human Cav3.1 alone and in complex with various modulators have previously been published1721. Despite substantial sequence variations, the structures of Cav1.1 and Cav3.1 are similar in their transmembrane regions, which comprise four homologous repeats (designated I, II, III and IV; throughout, subscript Roman numerals indicate location on the respective homologous repeat) each. The transmembrane segments 5 and 6 (S5 and S6) from the four repeats enclose the central pore domain that is surrounded by the four VSDs, each of which is constituted by S1 to S4 (Supplementary Fig. 1). In contrast to the conserved transmembrane domain, the long linkers between the repeats (the I–II linker and the II–III linker) and the C-terminal segments vary substantially both in length and in the primary sequence. These cytosolic segments remain unresolved in the structures of Cav and closely related Nav channels.

The gating of the pore domain is coupled to the movements of VSDs in response to membrane potential changes, in a mechanism that is known as the electromechanical coupling. Despite minor deviations, all of the previously reported structures of Cav1.1 and Cav3.1 channels exhibit similar ‘inactivated’ states, in which the pore domain is closed and all four VSDs are in the ‘up’ conformations1821. Investigating electromechanical coupling necessitates structural resolution of the channels in distinct conformations.

Cryo-EM analysis of the Cav2.2 complex

Details of protein production and cryo-EM analysis of human Cav2.2 in the absence (apo) and presence of purchased ziconotide (determined at 3.1 and 3.0 Å resolution, respectively (Extended Data Fig. 1)) are provided in the Methods. The 3D electron microscopy reconstructions of apo Cav2.2 and Cav2.2–ziconotide are nearly identical, and show only limited structural deviations. Unless indicated otherwise, the structural analysis discussed below is based on Cav2.2–ziconotide because of its higher resolution.

We could reliably assign side chains to α1, α2δ−1, ziconotide and the α1-interacting guanylate kinase domain of β3 (Extended Data Figs. 24, Extended Data Table 1, Supplementary Video 1). The distal domain of β3 and the C-terminal domain of α1 were docked as rigid bodies. In addition to the protein components, one Ca2+ ion is coordinated by the signature EEEE motif22 (Glu314, Glu663, Glu1365 and Glu1655) at the outer site of the selectivity filter (Extended Data Fig. 3c). Sixteen sugar moieties were assigned to eight glycosylated Asn residues (seven on α2δ−1 and one on α1). Among the resolved cholesterol, phospholipid and detergent molecules that surround α1, a phosphatidylinositol 4,5-bisphosphate (PIP2) molecule binds to the interface of VSDII and the pore domain in the inner leaflet (Fig. 1a, Extended Data Fig. 4).

Fig. 1 |. Specific pore blockade of Cav2.2 by ziconotide.

Fig. 1 |

a, Overall structure of the Cav2.2–ziconotide complex at an averaged resolution of 3.0 Å. CTD, C-terminal domain; Zi, ziconotide. The resolved lipid, cholesterol and cholesterol hemisuccinate molecules are shown as black sticks. Sugar moieties are omitted for visual clarity. All structure figures were prepared in PyMol with the same colour scheme. b, Ziconotide is caged by the ECLs. The sequence of ziconotide is shown above, with the three disulfide bonds indicated. The surface electrostatic potential, shown in semi-transparent presentation, was calculated in PyMol. c, Specific coordination of ziconotide by α1 of Cav2.2. The residues that are not conserved in Cav channels are labelled blue (in all Figures). The bound Ca2+ is shown as green sphere, and the EEEE motif (Glu314, Glu663, Glu1365 and Glu1655) that determines Ca2+ selectivity is shown as thin sticks. Letters in parentheses denote backbone groups, N for amide and O for carbonyl oxygen (in all Figures) d, ECLIII and α2δ−1 concertedly move upward to accommodate ziconotide. When the structures of the Cav2.2–ziconotide (coloured according to domain) and apo Cav2.2 (blue) are superimposed relative to α1, the only deviation (indicated by orange arrows) occurs in ECLIII and α2δ−1. Inset, upward shift of ECLIII to avoid clash with ziconotide. The distances between the indicated residues are in Å.

The overall structures of apo Cav2.2 and rat Cav1.1 can be superimposed with a root mean square deviation of 1.10 Å over 1,728 Cα atoms. The α2δ−1 subunit and its interface with the extracellular segments of α1 remain identical, despite minor structural differences in ECLI and ECLIII (Extended Data Fig. 5). On the cytosolic side, the helix after S6I bends to be parallel with the membrane plane and interacts with β3 through its α1-interacting domain (AID) motif23. We refer to this transverse helix as the AID helix. These common structures—which have been illustrated in detail for Cav1.118—are not discussed further here: here we focus on ziconotide recognition and repeat II, which exhibits several features that are distinct from known structures of eukaryotic Cav and Nav channels, for detailed analysis.

Pore blockade of Cav2.2 by ziconotide

Ziconotide has previously been reported to specifically block Cav2.2 with a half-maximal inhibitory concentration of 0.7–1.8 nM24,25. Our structure reveals that the blocker, which contains three disulfide bonds, is nestled in the electronegative cavity that surrounds the entrance to the selectivity filter, and that specific recognition is mediated mainly by charged and polar residues on the P1 and P2 helices and the ECLs in repeats II, III and IV. Although ECLI is also in the vicinity of ziconotide, none of the repeat I residues is directly engaged in ziconotide coordination (Fig. 1b, c, Extended Data Fig. 6a).

The binding pose of ziconotide is reminiscent of that of μ-conotoxin KIIIA (KIIIA) in its complex structure with Nav1.2, in which Lys7 of KIIIA points to the outer mouth of the selectivity filter26 (Extended Data Fig. 6b). Ziconotide does not possess an equivalent residue to directly seal the entrance to the vestibule of the selectivity filter. Instead, it blocks ion entrance by neutralizing the outer electronegativity and sterically hindering the ion access path to the entrance of the selectivity filter (Fig. 1b). To neutralize the acidic residues, ziconotide engages Arg10 and Tyr13 to bind to Asp664, which marks the beginning of P2II. Ser19 interacts with Glu1659, which is on the first helical turn of P2IV. Thr17 interacts with Asp1345 on ECLIII, and Arg21 and Lys4 interact with Asp1628 and Asp1629, respectively, on ECLIV (Fig. 1c).

Our structure-guided sequence comparison showed that four of the eight ziconotide-coordinating residues in Cav2.2 (Thr643, Asp1345, Lys1372 and Asp1629) are not conserved in other Cav channels (Fig. 1c, Supplementary Fig. 1), which explains the subtype specificity of pore blockade by ziconotide. Consistent with our structure, Y13A or R10A substitutions have previously been shown to substantially reduce pore blocking by ziconotide, and Y13R abolished its activity 24,27,28.

The only structural change upon ziconotide binding is a slight tilt of α2δ−1 and ECLIII of α1 as a rigid body. The upward movement of ECLIII is necessary to avoid a clash with ziconotide (Fig. 1d, Extended Data Fig. 6c, d). The α2δ−1 subunit has previously been shown to reduce ziconotide affinity29, which may result from the energy penalty for lifting ECLIII and α2δ−1 to accommodate ziconotide.

Cytosolic segments unique to Cav2

On the cytosolic side, S6II is extraordinarily long and contains seven additional helical turns (residues 714 to 739), which we designate the cytosolic segment of S6II (S6IIC). The ensuing sequence (residues 740 to 786) forms two additional helices (which we designate cytosolic helix (CH)1II and CH2II) that fold back towards the membrane. CH1II is anti-parallel with S6IIC, whereas CH2II is perpendicular to it (Fig. 2a). S6IIC is not conserved among the three Cav families and the sequence corresponding to CH1II and CH2II is missing in Cav1 and Cav3 families, which affords an explanation for the lack of these structural elements in Cav1.1 and Cav3.118,20 (Supplementary Fig. 1).

Fig. 2 |. Cytosolic segments unique to Cav2 in the II–III linker.

Fig. 2 |

a, An extended cytosolic segment of S6II is followed by CH1 and CH2. The sequence for S6IIC (residues 714–742) is not conserved among the Cav family and the ensuing helices are missing in Cav1 and Cav3 channels. Supplementary Fig. 1 shows the sequence alignment. b, β3 is sandwiched between the AID helix and S6IIC. Inset, detailed interactions between α1 and β3. Polar interactions are indicated by red dashed lines. c, CH2II simultaneously interacts with all four S6 segments below the intracellular gate. A tilted side view of the S6 tetrahelical bundle is shown. d, S6III is three helical turns shorter in Cav2.2 than in Cav1.1. Unwinding the last three turns of S6III yields space to accommodate CH2II. Cav1.1 (Protein Data Bank (PDB) code 5GJW) is coloured in wheat. e, CH2II tightens the closed intracellular gate. Several layers of hydrophobic residues on the S6 tetrahelical bundle intertwine to close the intracellular gate. Trp768, a Cav2-specific residue, interacts with several gating residues.

The U-turn between S6IIC and CH1II mediates weak interaction with β3, in addition to the well-characterized primary interface between the AID and the guanylate kinase domain of β33032 (Fig. 2b). Glu740, which is conserved in Cav2 family only and marks the end of S6II, may be hydrogen-bonded to His348 on the α9 helix of β3. Glu743, which is positioned at the tip of the U-turn, may interact with the amide of Glu111 in the SH3 domain of β3 (Fig. 2b inset). With this additional interface, β3 is sandwiched between AID and S6II.

CH1II has only limited contact with S6IIC, whereas CH2II interacts with all four S6 segments (Fig. 2c). As the sequence corresponding to CH2II is not found in Cav1 and Cav3 channels, this structural feature is probably unique to Cav2 channels (Supplementary Fig. 1). To accommodate CH2II, the last three helical turns in S6III are unwound, resulting in a shorter S6III than that in Cav1.1. The N-terminal residues of CH2II are placed immediately beneath the closed intracellular gate (Fig. 2d, e)

The intracellular gate of Cav2.2 comprises a large number of residues that intertwine to form a hydrophobic structural core. At the bottom of this exceptionally thick hydrophobic gate is Trp768 on CH2II, which is surrounded by Ala360, Ala713, Ile1417 and Phe1711 from the four S6 segments (Fig. 2e insets). By pulling all four S6 segments together and directly engaging Trp768 to secure the gate, CH2II may facilitate the closing of the pore domain.

VSDII in a ‘down’ conformation

The four VSDs of Cav2.2 carry different numbers of gating charge residues: five in VSDI (R1 to R4, and K5; to make the numbering of the gating charge residues consistent, we designate the residue on the first helical turn of the S4 segment as 1), VSDII (R2, R3, R4, K5 and K6) and VSDIV (R2, R3, R4, K5 and R6), and six in VSDIII (K1, R2, R3, R4, K5 and K6) (Fig. 3a). K5 in VSDI, K5 and K6 in VSDIII and K5 and R6 in VSDIV are below, and the other gating charge residues are above, the occluding Phe on S2 (a state that we define as up). By contrast, in VSDII R3, R4, K5 and K6 are all below the occluding Phe525, and only R2 is above it (a state that we define as down) (Fig. 3a). When the four VSDs are superimposed, S4II slides down by about 11–12 Å from the other S4 segments, and the S1, S2 and S3 segments are relatively well superimposed (Fig. 3b).

Fig. 3 |. VSDII in a down state.

Fig. 3 |

a, Structure of the four VSDs. In each VSD, the gating charge residues (gating charge numbering in parentheses) on S4 and the surrounding residues that may facilitate gating charge transfer are shown as sticks. An1, An2, conserved acidic or polar residues on S2. The gating charge residues above and below the occluding Phe on S2 are labelled dark cyan and brown, respectively. b, Distinct conformation of VSDII. When the four VSDs are superimposed, S4II slides down by about 12 Å compared to the S4 segments in other VSDs. c, Conformational changes of the α1 subunit in Cav1.1 and Cav2.2. A cytosolic view of the superimposed α1 structures is shown; the CTD is omitted for clarity. The major structural shifts from Cav1.1 to Cav2.2 are indicated by blue arrows. Right, structural comparison of the diagonal repeats II and IV. d, Coupled shifts of VSDII and S6II. Left, the downward motion of S4II is accompanied by the rotation of VSDII and the displacement of S4–5II and the nearby S6IIC. Right, pronounced shifts of VSDII between Cav1.1 and Cav2.2 in the context of the overall α1 structure. The distances shown on the left and right indicate the linear displacement of the Cα atoms and side chains of the corresponding gating charge residues, respectively. Supplementary Video 2 shows the conformational shifts.

In VSDI, VSDIII and VSDIV, the one or two extracellular helical turns of S4 are in the form of an α-helix, and the remaining segments conform to a 310 helix. In VSDII, the entire S4 is a 310 helix, and R2, R3, R4 and K5 all line the same side. K6 (Lys590), which is at the tip of a sharp and short turn that immediately precedes the S4–5II segment, is on the opposite side to the other four gating charge residues and its side chain projects into the cytosol (Fig. 3a, b). Compared to the deactivated structures of toxin-bound VSDIV in the NavPaS-1.7 chimera33 and rat Nav1.534, S4II in Cav2.2 is lower by one more helical turn (Extended Data Fig. 7a).

Structural comparison of the α1 subunits in Cav2.2 and Cav1.1 shows that the pore domain, VSDI and VSDIV remain nearly identical. Major shifts occur in repeats II and III (Fig. 3c, Extended Data Fig. 7b). VSDIII undergoes a slight clockwise rigid-body rotation from Cav1.1 to Cav2.2 in the intracellular view. The motion of VSDII involves both domain rotation and intradomain rearrangement (Fig. 3c, d, Extended Data Fig. 7c, d, Supplementary Video 2). The S1, S2 and S3 segments in VSDII of Cav1.1 and Cav2.2 remain nearly identical in domain comparison, whereas S4 is dislodged by about 12 Å towards the intracellular side from Cav1.1 to Cav2.2 through a spiral sliding, leading to the transfer of two gating charge residues (R3 and R4) across the occluding Phe on S2 (Extended Data Fig. 7d). The most marked shift occurs to K6, the amine group of which is displaced by 25 Å through both secondary structure transition and side-chain swing (Fig. 3d, Supplementary Video 2).

Accompanying the rotation of VSDII and the marked shift of S4II from Cav1.1 to Cav2.2, the ensuing N terminus of S4–5II is pushed down by about 3 Å. The motion of S4–5II drives the adjacent S6IIC to shift accordingly, establishing a potential transmission path from VSDII motion to pore gating (Fig. 3d, Extended Data Fig. 8, Supplementary Video 2). As only VSDII is in the down conformation and the other VSDs are up, we refrain from assigning a functional state to the present Cav2.2 structure. As, to our knowledge, a down structure of a VSD has not previously been captured in the absence of mutagenesis or modulators, we attempt to identify the structural elements that stabilize the deactivated state of VSDII.

Stabilization of the down VSDII by PIP2

The intracellular segments that are unique to Cav2 contribute to the stabilization of the down conformation of VSDII. S6IIC, which interacts extensively with S4–5II, contacts the AID helix and CH2II through several polar residues (Fig. 4a). The AID helix binds S0, S2 and S3 in VSDII, preventing rotation of these segments. Simultaneous association of S0II, S2–3II, S4–5II and S6IIC with the straight AID helix may restrain VSDII from rotation, therefore stabilizing the present conformation. More than half of the residues in the extensive interaction network are Cav2-specific, which provides a potential explanation for the distinct VSDII conformations in Cav2.2 (Fig. 4a, b).

Fig. 4 |. The down conformation of VSDII is stabilized by several intracellular segments and a bound PIP2.

Fig. 4 |

a, Cytosolic segments that are unique to Cav2 stabilize the down VSDII. Inset, a network of extensive polar interactions among the AID helix, S4–5II, S6II and CH2II. b, The AID helix is an organizing centre for segments within and near VSDII. The straight AID helix, in addition to mediating the channel modulation by the β subunits, may serve as a lever that couples the motion of VSDII to S6I and S6II. c, The bound PIP2 favours a down conformation of VSDII. Left, VSDII and the ensuing S4–5II in the up conformation as in Cav1.1 (wheat) would clash with PIP2 in its current binding pose. Right, a cytosolic view of PIP2 coordination by polar residues in Cav2.2 (coloured according to domain).

The bound PIP2 also appears to favour a down VSDII, as its binding pose is incompatible with the up conformation of Cav1.1 VSDII. The head group of PIP2 wedges into a cytosolic cavity in VSDII through the interface of S3 and S4, and the tails are coordinated by hydrophobic residues on S3 toS6 in repeat II and S5 and S6 in repeat III (Fig. 4c, Extended Data Fig. 9a, b). The PIP2-binding site is reminiscent of that in KCNQ135, but is at a higher position (Extended Data Fig. 9c). The 5-phosphate group of PIP2 is coordinated by R4 (Arg584) and K5 (Lys587) on S4II. The head group is anchored by Ser544 (which marks the beginning of S3II), the amide group of Phe546 and Arg596 on S4–5II (Fig. 4c, Extended Data Fig. 9a). When VSDII moves upward, its interface with the neighbouring S5III is also rearranged (Extended Data Fig. 9d). Consequently, PIP2, the displacement of which may be restrained by the hydrophobic membrane, can no longer retain these polar and hydrophobic interactions, but clashes with the lifted S4–5II (Fig. 4c, Extended Data Fig. 9a). Therefore, the bound PIP2 may stabilize the down conformation of VSDII.

Discussion

The cryo-EM structures of the human Cav2.2 complex reveal the molecular basis for the specific pore blockade of Cav2.2 by ziconotide and provide a mechanistic interpretation for lowered ziconotide affinity in the presence of α2δ−1 (Fig. 1d). The down conformation of VSDII, which is stabilized by previously unresolved cytosolic segments and a PIP2 molecule, provides important insights into the electromechanical coupling in Cav channels.

The AID helix, which follows S6I through a tight turn and interacts with S6II, is sandwiched by VSDII and β3 (Fig. 4b). It may thus serve as a lever to couple the conformational changes of VSDII with S6I, S6II and the intracellular segments. The β subunits may also modulate channel activity in part through the interaction with AID.

PIP2 has previously been shown to both cause a right shift of the activation voltage and reduce the current rundown of Cav2.136. The PIP2-binding site observed in this study may account for the voltage-dependent inhibitory modulation as it stabilizes the down conformation of VSDII. It remains to be investigated whether a separate PIP2-binding site is responsible for the voltage-independent rundown reduction, a mechanism that may involve Gβγ proteins37. Structural elucidation of representative members in the three Cav families lay the foundation for future structural and mechanistic investigation of the electromechanical coupling and the regulation of Cav channels by a variety of modulators3,8.

METHODS

No statistical methods were used to predetermine sample size. The experiments were not randomized, and investigators were not blinded to allocation during experiments and outcome assessment.

Transient expression of the human Cav2.2 complex in HEK293F cells

Codon-optimized cDNAs of CACNA1B for full-length Cav2.2 α1 (2,339 residues, Uniprot Q00975–1), CACNA2D1 for α2δ−1 (1,103 residues, Uniprot P54289–1) and CACNB3 for β3 (484 residues, Uniprot P54284–1) were synthesized (BGI Geneland Scientific) and subcloned into the pCAG vector. An amino-terminal Flag tag and a carboxy-terminal His10 tag were added to each subunit in Cav2.2–ziconotide and untagged α2δ−1 was used for the apo Cav2.2 complex. DNA sequences were examined in SnapGene. HEK293F suspension cells (Thermo Fisher Scientific, R79007) were cultured in Freestyle 293 medium (Thermo Fisher Scientific) at 37 °C, supplied with 5% CO2 under 60% humidity. When cell density reached 1.5–2.0 × 106 cells per ml, a mixture of expression plasmids and polyethylenimine (Polysciences) was added into cell culture to initiate the transient expression of human Cav2.2 following a standard transfection protocol. No further authentication was performed for the commercially available cell line. Mycoplasma contamination was not tested.

Protein purification of human Cav2.2 and complex preparation with ziconotide

Approximately 72 h after transfection, the HEK293F cells were collected by centrifugation at 3,600g for 10 min and resuspended in the lysis buffer containing 25 mM HEPES (pH 7.4), 150 mM NaCl and the protease inhibitor cocktail containing 2.6 μg ml−1 aprotinin and 1.4 μg ml−1 pepstatin. The suspension was supplemented with glyco-diosgenin (GDN) (Anatrace) to a final concentration of 1% (w/v), n-dodecyl-β-d-maltopyranoside (DDM, Anatrace) to 0.2% (w/v), and cholesteryl hemisuccinate Tris salt (CHS) (Anatrace) to 0.04% (w/v). After incubation at 4 °C overnight, the mixture was centrifuged at 35,000g for 30 min, and the supernatant was applied to anti-Flag M2 affinity resin (Sigma). The resin was rinsed with wash buffer (buffer W) containing 25 mM HEPES (pH 7.4), 150 mM NaCl, 2 mM CaCl2 and 0.01% GDN. Eluted with buffer W plus 0.2 mg ml−1 Flag peptide (synthesized by GenScript), the eluent was concentrated using a 100-kDa cut-off Amicon (Millipore) and further purified through size-exclusion chromatography (Superose 6 10/300 GL, GE Healthcare) that was preequilibrated in buffer W. The peak fractions were pooled and concentrated to a concentration of about 20 mg ml−1 with α2δ−1 in excess. For structural determination of Cav2.2–ziconotide, purchased ziconotide (Alomone labs) was added to the concentrated protein solution at a final concentration of 2 mg ml−1. The mixture was incubated at 4 °C for 30 min before making cryo-grids.

Cryo-EM sample preparation and data collection

Aliquots of 3.5 μl concentrated apo Cav2.2 or Cav2.2–ziconotide were loaded onto glow-discharged holey carbon grids (Quantifoil Cu R1.2/1.3, 300 mesh for Cav2.2–ziconotide, Quantifoil Au R1.2/1.3, 300 mesh for the apo channel), which were blotted for 6 s and plunge-frozen in liquid ethane cooled by liquid nitrogen using a Vitrobot Mark IV (Thermo Fisher) at 8 °C with 100% humidity. Grids were transferred to a Titan Krios electron microscope (Thermo Fisher) operating at 300 kV and equipped with a Gatan Gif Quantum energy filter (slit width 20 eV) and spherical aberration (Cs) image corrector. Micrographs were recorded using a K2 Summit counting camera (Gatan) in super-resolution mode with a nominal magnification of 105,000×, resulting in a calibrated pixel size of 0.557 Å. Each stack of 32 frames was exposed for 5.6 s, with an exposure time of 0.175 s per frame. The total dose for each stack was about 50 e per Å2. SerialEM was used for fully automated data collection38. All 32 frames in each stack were aligned, summed and dose-weighted using MotionCorr239 and twofold-binned to a pixel size of 1.114 Å per pixel. The defocus values were set from −1.9 to −2.1 μm and estimated by Gctf40.

Image processing

A total of 3,384 and 1,317 cryo-EM micrographs were collected for Cav2.2–ziconotide and the apo channel, respectively, and 1,909,156 (Cav2.2–ziconotide) and 928,665 (apo Cav2.2) particles were auto-picked by RELION-3.0. Particle picking was performed using 2D classes of rabbit Cav1.1 (EMD-22426) in the side and tilted views as reference. All subsequent 2D and 3D classification and refinement was performed with RELION-3.041. Two rounds of reference-free 2D classification were performed to remove ice spots, contaminants and aggregates, yielding 1,101,746 and 807,595 particles for Cav2.2–ziconotide and apo Cav2.2, respectively. The particles were processed with a global search with K = 1 to determine the initial orientation alignment parameters using bin2 particles. The electron microscopy map of rabbit Cav1.1 (EMD-22426), low-pass-filtered to 20 Å, was used as an initial reference21. The output of the 35–40 iterations was further applied to local angular search 3D classification with four classes. A total of 305,200 and 159,079 particles (for Cav2.2–ziconotide and apo Cav2.2, respectively) were selected by combining the good classes of the local angular search 3D classification. The particles were then re-extracted using a box size of 280 and pixel size of 1.114 Å. These particles yielded reconstructions at 3.2 Å and 3.3 Å (for Cav2.2–ziconotide and apo Cav2.2, respectively) after 3D auto-refinement with an adapted mask. Multi-reference 3D classification using bin1 particles after Bayesian polishing resulted in a reconstruction at 3.0 Å from 170,839 particles for Cav2.2–ziconotide and 3.1 Å from 56,616 particles for the apo channel. Focused refinement was performed using a core mask for β3 and α1 S6II.

All 2D classification, 3D classification and 3D auto-refinement was performed with RELION 3.0. Resolutions were estimated using the gold-standard Fourier shell correlation 0.143 criterion with high-resolution noise substitution42,43.

Model building and refinement

Cav2.2, together with the other two Cav2 subtypes (Cav2.1 and Cav2.3) and the four Cav1 subtypes (Cav1.1 to Cav1.4), are high-voltage-activated, in contrast to the low-voltage-gated Cav3.1, Cav3.2 and Cav3.3 channels44 (Supplementary Fig. 1). Model building of Cav2.2 was thus based on the structure of the Cav1.1 channel complex. The starting model of Cav2.2 α1 subunit and α2δ−1 were built in SWISS-MODEL45 based on the structure of rabbit Cav1.1 (PDB code 5GJW), and that of β3 was based on the crystal structure of rat β3 (PDB code 1VYU). The starting models of Cav2.2 and ω-conotoxin MVIIA (PDB code 1OMG) were then manually docked into the 3.0 Å toxin-bound electron microscopy map in Chimera46. The model was manually adjusted in COOT47, followed by refinement against the corresponding maps by phenix.real_space_refine program in PHENIX48 with secondary structure and geometry restraints. For model building of the apo channel, coordinates for Cav2.2 α1, α2δ−1 and β3 were docked into the 3.1 Å apo Cav2.2 map separately and manually adjusted in COOT.

The excellent map quality supports reliable assignment of 2,622 side chains in both structures. Similar to the structures of other eukaryotic Cav and Nav channels18,20,4956, most of the cytosolic segments in the α1 subunit are invisible, including residues 1–85 in the N terminus, residues 407–463 in the I–II linker, residues 787–1138 in the II–III linker, and residues 1838–2339 in the C terminus.

Statistics of the map reconstruction and model refinement can be found in Extended Data Table 1. All structure figures were prepared in PyMol57.

Reporting summary

Further information on research design is available in the Nature Research Reporting Summary linked to this paper.

Extended Data

Extended Data Fig. 1 |. Cryo-EM analysis of the human Cav2.2 complex alone and in the presence of ziconotide.

Extended Data Fig. 1 |

a, Representative electron micrograph and 30 classes of 2D class averages for Cav2.2–ziconotide. The green circles indicate particles in distinct orientations. The box size for the 2D averages is 312 Å. Scale bar, 50 nm. Left, a half view of one micrograph out of 3,384 in total for Cav2.2–ziconotide. b, Workflow for electron microscopy data processing (Methods). c, The gold-standard Fourier shell correlation (FSC) curves for the 3D reconstructions. The graph was prepared in GraphPad Prism. d, FSC curves of the refined model versus the summed map that it was refined against (black); of the model versus the first half map (red); and of the model versus the second half map (green). Z-complex, Cav2.2–ziconotide.

Extended Data Fig. 2 |. Cryo-EM structure of the human Cav2.2 complex bound with ziconotide.

Extended Data Fig. 2 |

a, Heat map for local resolution of the complex. The resolution map was calculated in Relion 3.0 and prepared in Chimera. Top inset, bound ziconotide (labelled as Zi) is well-resolved. Bottom inset, resolution of the β3 subunit, after a focused refinement, allows for reliable model building using the crystal structure of rat β3 (PDB code 1VYU) as a template. b, Overall structure of the complex at an averaged resolution of 3.0 Å. Left, the complex comprises the α1 core subunit (silver), the extracellular α2δ−1 subunit (light pink for α2 and green for δ) and the cytosolic β3 subunit (wheat). The peptide pore blocker ziconotide is coloured brown. The resolved lipid, cholesterol and CHS molecules are shown as black sticks. The bound PIP2 is shown as black ball-and-sticks. Sugar moieties are shown as thin sticks. Right, surface presentation of the structure. The four repeats are coloured grey, cyan, yellow, and pale green. The III–IV linker and the CTD are coloured orange and pale purple, respectively.

Extended Data Fig. 3 |. Electron microscopy densities for representative segments of Cav2.2–ziconotide.

Extended Data Fig. 3 |

a, Electron microscopy maps for representative segments in α1 and β3. The densities for the β3 segments are from focused refinement, and the others are from the overall map. All the densities shown are contoured at 4σ. b, The electron microscopy map for ziconotide. c, Electron microscopy densities for the bound Ca2+ ion and surrounding residues in the selectivity filter. The maps were prepared in PyMol.

Extended Data Fig. 4 |. Lipids resolved in the structures.

Extended Data Fig. 4 |

a, A PIP2 molecule binds to VSDII in both structures. All the densities shown are contoured at 3σ. b, The densities for the resolved cholesterol (Cho) and CHS molecules in Cav2.2–ziconotide. c, Lipids resolved in the structure of Cav2.2–ziconotide. The α1 subunit are shown in two opposite side views. The numbering for cholesterol and CHS is consistent with that in b. Two phospholipids are also resolved and assigned as phosphatidylethanolamine (PE).

Extended Data Fig. 5 |. Structural comparison of the Cav1.1 and Cav2.2 channel complexes.

Extended Data Fig. 5 |

a, Superimposition of the overall structures of human Cav2.2 (apo) and rabbit Cav1.1 (PDB code 5GJW). For visual clarity, the γ subunit in the endogenous Cav1.1 complex is not shown. The conformational shift of VSDII from Cav1.1 (wheat) to Cav2.2 (blue) is indicated by the blue arrow. b, Identical structures of the α2δ−1 subunit in the two channel complexes. A detailed structural description of the α2δ−1 subunit can be found in a previous publication18. c, Structural differences of the ECLs between Cav1.1 and Cav2.2. An extracellular view of the superimposed α1 subunits in the two channels is shown.

Extended Data Fig. 6 |. Conformational shifts of Cav2.2 upon ziconotide binding.

Extended Data Fig. 6 |

a, ECLI does not participate in ziconotide coordination. An extracellular view perpendicular to that in Fig. 1c is shown. b, Slightly different mode of action of KIIIA for Nav1.226. Lys7 in KIIIA directly blocks the outer mouth of the selectivity filter vestibule of Nav1.2 (PDB code 6J8E), in a manner similar to a cork. Ziconotide lacks an equivalent basic residue. c, Relative shift of α2δ−1 between apo (blue) and ziconotide-bound Cav2.2 (domain-coloured) when the two structures are superimposed relative to the α1 subunit. The rest of the complex remains identical except for ECLIII. d, Concerted motion of α2δ−1 and ECLIII of α1. The α2δ−1 subunit in the two structures can be superimposed with a root mean square deviation of 0.28 Å over 847 Cα atoms, indicating nearly identical conformations. When the two structures are superimposed relative to α2δ−1, the entire α1 undergoes a relative shift—except for ECLIII, which stays as a rigid body with α2δ−1.

Extended Data Fig. 7 |. Conformational changes of VSDII and VSDIII between Cav1.1 and Cav2.2.

Extended Data Fig. 7 |

a, Structural comparison of Cav2.2 VSDII with other VSDs that exhibit down conformations. To make the nomenclature consistent, we define the gating charge residue on the first helical turn of the S4 segment as R1. The PDB accession codes are 6NT4 for VSDIV in the chimeric NavPaS-1.7, 7K18 for toxin-bound VSDIV in rat Nav1.5, 4G80 for the antibody-locked VSD of a voltage-sensitive phosphatase, and 6UQF for the VSD of HCN1 in hyperpolarized conformation. b, Structural comparison of Cav1.1 and Cav2.2 shows a slight rotation of VSDIII around the pore domain. The superimposed structures of the diagonal repeats I and III of Cav1.1 (wheat) and Cav2.2 (domain-coloured) are shown. c, VSDIII remains nearly rigid in these two structures. When the structures of VSDIII in the two channels are individually superimposed, the S4 segment and the gating charge residues align well. d, Marked shift of S4II between Cav1.1 and Cav2.2 when the two structures are compared relative to VSDII. S4II undergoes a combination of spiral sliding and secondary structural transition. S1, S2 and S3 remain nearly unchanged in these two VSDII structures, which suggests a concerted rotation of the other three segments pivoting around S4.

Extended Data Fig. 8 |. A closed pore domain with one small fenestration.

Extended Data Fig. 8 |

a, The pore domain is in a closed conformation. Four perpendicular side views of the pore domain are shown. S4–5II is pushed downward as a result of the sliding of S4II. b, Side walls that involve S6II are sealed without fenestration. Side views of the pore domain surface are shown. There is only one fenestration on the interface of repeats III and IV.

Extended Data Fig. 9 |. A PIP2 molecule may help to stabilize the down conformation of Cav2.2 VSDII.

Extended Data Fig. 9 |

a, The binding pose for PIP2 is incompatible with an up VSDII. Left, coordination of the head group of PIP2 by Cav2.2. Side local view of VSDII is shown. Right, in an up state of VSDII (as in Cav1.1), R4 and K5 can no longer interact with the PIP2 head group, and S4–5II directly clashes with PIP2. Structures of Cav1.1 and Cav2.2 are superimposed relative to the α1 subunit and Cav2.2 is omitted to highlight the relative position of PIP2 to Cav1.1. b, The hydrophobic tails of PIP2 interact extensively with multiple segments in repeats II and III. Hydrophobic residues on segments S3 to S6 in repeat II and S5 and S6 in repeat III contact the two tails of PIP2. c, The PIP2 molecule in the Kv channel KCNQ1 is bound at a similar, but lower, position. The PDB code for the KCNQ1 structure is 6V01. d, Rearrangement of the interface of VSDII and pore domain between Cav2.2 and Cav1.1. Cav1.1 is coloured with the same scheme as for Cav2.2. Alternative sets of hydrophobic residues between the gating charge residues on S4II are used for interacting with S5III in Cav1.1 and Cav2.2 as a result of the rotation of S4II. The sequence numbers for corresponding VSDII residues in these two channels differ by 50, and those for S5III residues differ by 354. As labelled in the parentheses, Val1298 and Phe1292 on the S5III segment of Cav2.2 are at loci corresponding to Cys944 and Leu938 in Cav1.1, respectively.

Extended Data Table 1 |.

Statistics for data collection and structural refinement

Cav2.2-ziconotide
(EMDB-23867)
(PDB 7MIX)
Cav2.2-apo
(EMDB-23868)
(PDB 7MIY)
Data collection and processing
Magnification 105,000 105,000
Voltage (kV) 300 300
Electron dose (e-/Å2) 50 50
Defocus range (μm) −2.1~−1.9 −2.1~−1.9
Pixel size (Å) 1.114 1.114
Symmetry C1 C1
Initial particle images (no.) 1,909,156 928,665
Final particle images (no.) 170,839 56,616
Map resolution (Å) 3.0 3.1
 FSC threshold 0.143 0.143
Map resolution range (Å) 2.8-31.2 2.9-50
Refinement
Initial model used (PDB code) 5GJW, 1VYU, 1OMG Cav2.2-ziconotide
Model resolution (Å) 3.3 3.4
 FSC threshold 0.5 0.5
Map sharpening B factor (Å2) −79 −59
Model composition
 Non-hydrogen atoms 21654 21485
 Protein residues 2623 2598
 Ligands 28 28
B factors (Å2)
 Protein 58.27 78.42
 Ligand 69.59 62.16
R.m.s deviations
 Bond lengths (Å) 0.004 0.003
 Bond angles (°) 0.759 0.727
Validation
 MolProbity score 2.00 1.95
 Clashscore 9.72 9.68
 Poor rotamers (%) 1.47 1.31
Ramachandran plot
 Favored (%) 94.76 94.99
 Allowed (%) 5.12 4.86
 Disallowed (%) 0.12 0.16

Supplementary Material

Supplementary Figure 1
Supplementary Video 1
Download video file (7.4MB, mov)
Supplementary Video 2
Download video file (10.5MB, mov)

Acknowledgements

We thank X. Pan for sharing the expression plasmids for the Cav2.2 complex subunits as a gift; X. Fan for critical discussion on cryo-EM data processing; and the cryo-EM facility at Princeton Imaging and Analysis Center. The work was supported by a grant from the NIH (5R01GM130762).

Footnotes

Online content Any methods, additional references, Nature Research reporting summaries, source data, extended data, supplementary information, acknowledgements, peer review information; details of author contributions and competing interests; and statements of data and code availability are available at https://doi.org/10.1038/s41586-021-03699-6.

Competing interests The authors declare no competing interests.

Supplementary Information The online version contains supplementary material available at https://doi.org/10.1038/s41586-021-03699-6].

Data availability

The atomic coordinates and electron microscopy maps for Cav2.2 in complex with ziconotide and alone have been deposited in the PDB with the accession codes 7MIX (with ziconotide) and 7MIY (without ziconotide) and in the Electron Microscopy Data Bank with the codes EMD-23867 (with ziconotide) and EMD-23868 (without ziconotide), respectively. The atomic coordinates of the proteins for structural comparison in this study can be found in the PDB: rabbit Cav1.1 (5GJW) (https://www.rcsb.org/structure/5GJW), toxin-bound human Nav1.2 (6J8E) (https://www.rcsb.org/structure/6J8E), toxin-bound NavPaS-1.7 chimera (6NT4) (https://www.rcsb.org/structure/6NT4), toxin-bound rat Nav1.5 (7K18) (https://www.rcsb.org/structure/7K18), Ci-VSP (4G80) (https://www.rcsb.org/structure/4G80), HCN1 (6UQF) (https://www.rcsb.org/structure/6UQF) and KCNQ1 (6V01) (https://www.rcsb.org/structure/6V01). Expression plasmids for the Cav2.2 subunits are available from the corresponding author upon reasonable request.

References

  • 1.Dubel SJ et al. Molecular cloning of the alpha-1 subunit of an omega-conotoxin-sensitive calcium channel. Proc. Natl Acad. Sci. USA 89, 5058–5062 (1992). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Komuro H & Rakic P Selective role of N-type calcium channels in neuronal migration. Science 257, 806–809 (1992). [DOI] [PubMed] [Google Scholar]
  • 3.Zamponi GW, Striessnig J, Koschak A & Dolphin AC The physiology, pathology, and pharmacology of voltage-gated calcium channels and their future therapeutic potential. Pharmacol. Rev 67, 821–870 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.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 36, 1651–1658 (1998). [DOI] [PubMed] [Google Scholar]
  • 5.Miljanich GP Ziconotide: neuronal calcium channel blocker for treating severe chronic pain. Curr. Med. Chem 11, 3029–3040 (2004). [DOI] [PubMed] [Google Scholar]
  • 6.Schmidtko A, Lötsch J, Freynhagen R & Geisslinger G Ziconotide for treatment of severe chronic pain. Lancet 375, 1569–1577 (2010). [DOI] [PubMed] [Google Scholar]
  • 7.Lin Y, McDonough SI & Lipscombe D Alternative splicing in the voltage-sensing region of N-type CaV2.2 channels modulates channel kinetics. J. Neurophysiol 92, 2820–2830 (2004). [DOI] [PubMed] [Google Scholar]
  • 8.Dolphin AC Voltage-gated calcium channels and their auxiliary subunits: physiology and pathophysiology and pharmacology. J. Physiol 594, 5369–5390 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Saegusa H et al. Suppression of inflammatory and neuropathic pain symptoms in mice lacking the N-type Ca2+ channel. EMBO J. 20, 2349–2356 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Snutch TP Targeting chronic and neuropathic pain: the N-type calcium channel comes of age. NeuroRx 2, 662–670 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.McGivern JG Targeting N-type and T-type calcium channels for the treatment of pain. Drug Discov. Today 11, 245–253 (2006). [DOI] [PubMed] [Google Scholar]
  • 12.McCleskey EW et al. Omega-conotoxin: direct and persistent blockade of specific types of calcium channels in neurons but not muscle. Proc. Natl Acad. Sci. USA 84, 4327–4331 (1987). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ellinor PT, Zhang JF, Horne WA & Tsien RW Structural determinants of the blockade of N-type calcium channels by a peptide neurotoxin. Nature 372, 272–275 (1994). [DOI] [PubMed] [Google Scholar]
  • 14.Stocker JW, Nadasdi L, Aldrich RW & Tsien RW Preferential interaction of ω-conotoxins with inactivated N-type Ca2+ channels. J. Neurosci 17, 3002–3013 (1997). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Deer T et al. Effectiveness and safety of intrathecal ziconotide: interim analysis of the patient registry of intrathecal ziconotide management (PRIZM). Pain Pract. 18, 230–238 (2018). [DOI] [PubMed] [Google Scholar]
  • 16.Nair AS, Poornachand A & Kodisharapu PK Ziconotide: indications, adverse effects, and limitations in managing refractory chronic pain. Indian J. Palliat. Care 24, 118–119 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wu J et al. Structure of the voltage-gated calcium channel Cav1.1 complex. Science 350, aad2395 (2015). [DOI] [PubMed] [Google Scholar]
  • 18.Wu J et al. Structure of the voltage-gated calcium channel Cav1.1 at 3.6 Å resolution. Nature 537, 191–196 (2016). [DOI] [PubMed] [Google Scholar]
  • 19.Zhao Y et al. Molecular basis for ligand modulation of a mammalian voltage-gated Ca2+ channel. Cell 177, 1495–1506.e12 (2019). [DOI] [PubMed] [Google Scholar]
  • 20.Zhao Y et al. Cryo-EM structures of apo and antagonist-bound human Cav3.1. Nature 576, 492–497 (2019). [DOI] [PubMed] [Google Scholar]
  • 21.Gao S & Yan N Structural basis of the modulation of the voltage-gated calcium ion channel Cav 1.1 by dihydropyridine compounds. Angew. Chem. Int. Ed. Engl 60, 3131–3137 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Yang J, Ellinor PT, Sather WA, Zhang JF & Tsien RW Molecular determinants of Ca2+ selectivity and ion permeation in L-type Ca2+ channels. Nature 366, 158–161 (1993). [DOI] [PubMed] [Google Scholar]
  • 23.Pragnell M et al. Calcium channel β-subunit binds to a conserved motif in the I–II cytoplasmic linker of the α1-subunit. Nature 368, 67–70 (1994). [DOI] [PubMed] [Google Scholar]
  • 24.Kim JI, Takahashi M, Ohtake A, Wakamiya A & Sato K Tyr13 is essential for the activity of ω-conotoxin MVIIA and GVIA, specific N-type calcium channel blockers. Biochem. Biophys. Res. Commun 206, 449–454 (1995). [DOI] [PubMed] [Google Scholar]
  • 25.Benjamin ER et al. Pharmacological characterization of recombinant N-type calcium channel (Cav2.2) mediated calcium mobilization using FLIPR. Biochem. Pharmacol 72, 770–782 (2006). [DOI] [PubMed] [Google Scholar]
  • 26.Pan X et al. Molecular basis for pore blockade of human Na+ channel Nav1.2 by the μ-conotoxin KIIIA. Science 363, 1309–1313 (2019). [DOI] [PubMed] [Google Scholar]
  • 27.Sousa SR et al. Novel analgesic ω-conotoxins from the vermivorous cone snail Conus moncuri provide new insights into the evolution of conopeptides. Sci. Rep 8, 13397 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Nadasdi L et al. Structure-activity analysis of a Conus peptide blocker of N-type neuronal calcium channels. Biochemistry 34, 8076–8081 (1995). [DOI] [PubMed] [Google Scholar]
  • 29.Mould J et al. The α2δ auxiliary subunit reduces affinity of ω-conotoxins for recombinant N-type (Cav2.2) calcium channels. J. Biol. Chem 279, 34705–34714 (2004). [DOI] [PubMed] [Google Scholar]
  • 30.Chen YH et al. Structural basis of the α1–β subunit interaction of voltage-gated Ca2+ channels. Nature 429, 675–680 (2004). [DOI] [PubMed] [Google Scholar]
  • 31.Van Petegem F, Clark KA, Chatelain FC & Minor DL Jr. Structure of a complex between a voltage-gated calcium channel β-subunit and an α-subunit domain. Nature 429, 671–675 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Opatowsky Y, Chen CC, Campbell KP & Hirsch JA Structural analysis of the voltage-dependent calcium channel β subunit functional core and its complex with the α1 interaction domain. Neuron 42, 387–399 (2004). [DOI] [PubMed] [Google Scholar]
  • 33.Clairfeuille T et al. Structural basis of α-scorpion toxin action on Nav channels. Science 363, eaav8573 (2019). [DOI] [PubMed] [Google Scholar]
  • 34.Jiang D et al. Structural basis for voltage-sensor trapping of the cardiac sodium channel by a deathstalker scorpion toxin. Nat. Commun 12, 128 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Sun J & MacKinnon R Structural basis of human KCNQ1 modulation and gating. Cell 180, 340–347.e9 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Wu L, Bauer CS, Zhen XG, Xie C & Yang J Dual regulation of voltage-gated calcium channels by PtdIns(4,5)P2. Nature 419, 947–952 (2002). [DOI] [PubMed] [Google Scholar]
  • 37.Rodríguez-Menchaca AA, Adney SK, Zhou L & Logothetis DE Dual regulation of voltage-sensitive ion channels by PIP2. Front. Pharmacol 3, 170 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Mastronarde DN Automated electron microscope tomography using robust prediction of specimen movements. J. Struct. Biol 152, 36–51 (2005). [DOI] [PubMed] [Google Scholar]
  • 39.Zheng SQ et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods 14, 331–332 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Zhang K Gctf: real-time CTF determination and correction. J. Struct. Biol 193, 1–12 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Zivanov J et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife 7, e42166 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Rosenthal PB & Henderson R Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol 333, 721–745 (2003). [DOI] [PubMed] [Google Scholar]
  • 43.Chen S et al. High-resolution noise substitution to measure overfitting and validate resolution in 3D structure determination by single particle electron cryomicroscopy. Ultramicroscopy 135, 24–35 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Nowycky MC, Fox AP & Tsien RW Three types of neuronal calcium channel with different calcium agonist sensitivity. Nature 316, 440–443 (1985). [DOI] [PubMed] [Google Scholar]
  • 45.Waterhouse A et al. SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Res. 46, W296–W303 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Pettersen EF et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem 25, 1605–1612 (2004). [DOI] [PubMed] [Google Scholar]
  • 47.Emsley P, Lohkamp B, Scott WG & Cowtan K Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Adams PD et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Shen H et al. Structure of a eukaryotic voltage-gated sodium channel at near-atomic resolution. Science 355, eaal4326](2017). [DOI] [PubMed] [Google Scholar]
  • 50.Yan Z et al. Structure of the Nav1.4-β1 complex from electric eel. Cell 170, 470–482.e11 (2017). [DOI] [PubMed] [Google Scholar]
  • 51.Shen H, Liu D, Wu K, Lei J & Yan N Structures of human Nav1.7 channel in complex with auxiliary subunits and animal toxins. Science 363, 1303–1308 (2019). [DOI] [PubMed] [Google Scholar]
  • 52.Pan X et al. Structure of the human voltage-gated sodium channel Nav1.4 in complex with β1. Science 362, eaau2486](2018). [DOI] [PubMed] [Google Scholar]
  • 53.Jiang D et al. Structure of the cardiac sodium channel. Cell 180, 122–134.e10 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Li Z et al. Structural basis for pore blockade of the human cardiac sodium channel Nav1.5 by the antiarrhythmic drug quinidine. Angew. Chem. Int. Edn Engl 60, 11474–11480(2021). [DOI] [PubMed] [Google Scholar]
  • 55.Pan X et al. Comparative structural analysis of human Nav1.1 and Nav1.5 reveals mutational hotspots for sodium channelopathies. Proc. Natl Acad. Sci. USA, 118, e2100066118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Li Z et al. Structure of human Nav1.5 reveals the fast inactivation-related segments as a mutational hotspot for the long QT syndrome. Proc. Natl Acad. Sci. USA 118, e2100069118(2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.DeLano WL The PyMOL Molecular Graphics System, http://www.pymol.org (2002).

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Figure 1
Supplementary Video 1
Download video file (7.4MB, mov)
Supplementary Video 2
Download video file (10.5MB, mov)

Data Availability Statement

The atomic coordinates and electron microscopy maps for Cav2.2 in complex with ziconotide and alone have been deposited in the PDB with the accession codes 7MIX (with ziconotide) and 7MIY (without ziconotide) and in the Electron Microscopy Data Bank with the codes EMD-23867 (with ziconotide) and EMD-23868 (without ziconotide), respectively. The atomic coordinates of the proteins for structural comparison in this study can be found in the PDB: rabbit Cav1.1 (5GJW) (https://www.rcsb.org/structure/5GJW), toxin-bound human Nav1.2 (6J8E) (https://www.rcsb.org/structure/6J8E), toxin-bound NavPaS-1.7 chimera (6NT4) (https://www.rcsb.org/structure/6NT4), toxin-bound rat Nav1.5 (7K18) (https://www.rcsb.org/structure/7K18), Ci-VSP (4G80) (https://www.rcsb.org/structure/4G80), HCN1 (6UQF) (https://www.rcsb.org/structure/6UQF) and KCNQ1 (6V01) (https://www.rcsb.org/structure/6V01). Expression plasmids for the Cav2.2 subunits are available from the corresponding author upon reasonable request.

RESOURCES