ABSTRACT
Ion channels play key roles in human physiology and are important targets in drug discovery. The atomic-scale structures of ion channels provide invaluable insights into a fundamental understanding of the molecular mechanisms of channel gating and modulation. Recent breakthroughs in deep learning-based computational methods, such as AlphaFold, RoseTTAFold, and ESMFold have transformed research in protein structure prediction and design. We review the application of AlphaFold, RoseTTAFold, and ESMFold to structural modeling of ion channels using representative voltage-gated ion channels, including human voltage-gated sodium (NaV) channel - NaV1.8, human voltage-gated calcium (CaV) channel – CaV1.1, and human voltage-gated potassium (KV) channel – KV1.3. We compared AlphaFold, RoseTTAFold, and ESMFold structural models of NaV1.8, CaV1.1, and KV1.3 with corresponding cryo-EM structures to assess details of their similarities and differences. Our findings shed light on the strengths and limitations of the current state-of-the-art deep learning-based computational methods for modeling ion channel structures, offering valuable insights to guide their future applications for ion channel research.
Keywords: Structural modeling, voltage-gated sodium channels, voltage-gated calcium channels, voltage-gated potassium chnanels, AlphaFold, RoseTTAFold, ESMFold
Introduction
Ion channels play key roles in human physiology and have been established as important targets in drug discovery [1,2]. The atomic-scale structures of ion channels provide invaluable insights into a fundamental understanding of the molecular mechanisms of channel gating and modulation. The recent advancements in cryo-electron microscopy (cryo-EM) produced a remarkable increase in the number of high-resolution structures of ion channels [3–7]. Multiple ion channel structures have been resolved in various putative physiological states and in complex with auxiliary subunits, small molecules, and natural peptides, providing crucial insights into the molecular mechanisms underlying their modulation.
In parallel to advancements in cryo-EM, breakthroughs in deep learning-based computational methods, such as AlphaFold [8] from Google’s DeepMind and RosetTTAFold [9,10] from David Baker’s Institute for Protein Design at the University of Washington, have been transforming research in protein structure prediction. These methods utilize deep neural networks trained on co-evolution information from multiple sequence alignments derived from protein sequence database (UniProt) [11] and protein structural data derived from Protein Data Bank (PDB) [12] to predict protein structures. AlphaFold and RoseTTAFold based methods have been applied to protein design and modeling of protein complexes [13–19]. Additionally, large language models of protein sequences, such as Meta AI’s ESMFold [20], trained on millions of protein sequences and using billions of parameters, provide rapid protein structure predictions, although with slightly lower accuracy compared to AlphaFold and RosettaFold. The ability to predict protein structures with high accuracy holds tremendous promise in transforming the field of drug discovery. Notably, the AlphaFold Structural Database currently contains over 200 million protein models predicted by AlphaFold [21] and the ESM Metagenomic Atlas contains more than 700 million protein models predicted by ESMFold [20].
Despite the remarkable achievements in protein structure prediction using deep-learning-based methods, the performance of these methods on challenging targets like ion channels remains to be determined. This is particularly significant when considering the structural heterogeneity of ion channel physiologically relevant states. To examine the structural modeling of ion channels using the deep-learning-based methods, we applied AlphaFold2 [8], RosetTTAFold2 [9], and ESMFold [20] to predict structures of representative voltage-gated ion channels, including human voltage-gated sodium (NaV) channel – NaV1.8, human voltage-gated calcium (CaV) channel – CaV1.1, and human voltage-gated potassium (KV) channel – KV1.3. We compared AlphaFold2, RoseTTAFold2, and ESMFold structural models of NaV1.8, CaV1.1, and KV1.3 with corresponding cryo-EM structures of NaV1.8 [22], CaV1.1 [23], and KV1.3 [24] to assess details of their similarities and differences. Our findings shed light on the strengths and limitations of the current state-of-the-art deep learning-based computational methods for modeling ion channel structures, offering valuable insights to guide their future applications for ion channel research.
Structural modeling of voltage-gated sodium (Nav) channels
Voltage-gated sodium (NaV) channels are responsible for initiating and propagating action potentials, the electrical signals facilitating communication between excitable cells [1,25–27]. There are nine NaV channel subtypes, from NaV1.1 to NaV1.9. The NaV1.1, NaV1.2, and NaV1.6 subtypes are predominantly expressed in the central nervous system [28]. The NaV1.4 and NaV1.5 subtypes are mainly expressed in skeletal and cardiac muscles, respectively [28]. The peripheral nervous system primarily expresses NaV1.7, NaV1.8, and NaV1.9 subtypes [28]. Dysfunctions in these channels can lead to serious health issues, including epilepsy, cardiac arrhythmias, muscle weakness, and chronic pain. The advancement of cryo-EM has facilitated the resolution of mammalian NaV subtypes, ranging from NaV1.1 to NaV1.8, significantly enhancing our understanding of their structure, gating, and modulation [22,29–35].
The voltage-dependent gating, sodium conduction, and modulation by natural peptides and small molecule drugs are performed by the NaV channel α subunit [1,25–27]. Auxiliary NaV channel β subunits (β1-β4) are co-expressed with the α subunit and modulate the channel function [36]. We selected the NaV1.8 channel α subunit, as an example NaV channel, for structure prediction using AlphaFold2 [8] and ColabFold as the computational platform [37]. ColabFold’s AlphaFold2 pipeline employs MMseqs2 multiple sequence alignment method [38,39], which is a more efficient alternative to Jackhmmer multiple sequence alignment method [40] used in DeepMind’s original AlphaFold2 pipeline [8]. The MMseqs2 method has considerably accelerated the AlphaFold2 protein structure prediction pipeline performance while maintaining comparable accuracy [37]. The protein sequence of the SCN10A gene, which encodes the human NaV1.8 (hNaV1.8) α subunit (UniProt ID: Q9Y5Y9), was used as input into ColabFold’s AlphaFold_mmseqs2 notebook for structure prediction. We assessed the quality of predicted AlphaFold2 models of hNaV1.8 using predicted local distance difference test (pLDDT) confidence score. Generally, pLDDT values above 90 represent very high confidence, pLDDT values between 70 and 90 represent good confidence, pLDDT values between 50 and 70 represent low confidence, and pLDDT values below 50 represent very low confidence [8]. We also compared similarities and differences to the resolved hNaV1.8 structures (PDB: 7WE4, 7WEL, 7WFR, and 7WFW) [22] using alpha carbon root mean square deviation (Cα RMSD). We assessed individual transmembrane voltage-sensing domains (VSD-I, VSD-II, VSD-III, and VSD-IV), the pore domain, the extracellular loop (ECL) regions, and the overall model topology.
Our results showed that AlphaFold2 could predict the majority of the hNaV1.8 domains with very high confidence scores (pLDDT >90), ESMFold could predict VSDs, the pore domain, and ECL regions with good confidence (70 < pLDDT < 90), while RoseTTAFold2 predicted most transmembrane regions with low confidence (50 < pLDDT < 70) and predicted the pore domain with good confidence (Figure 1).
Figure 1.

AlphaFold2, RoseTTAFold2, and ESMFold models of hNaV1.8. a) plot of pLDDT confidence score versus hNaV1.8 residue position for AlphaFold2 (AF), RoseTTAFold2 (RF2), and ESMFold (ESM) models. right panel, multiple sequence alignment of hNaV1.8 sequence and its homologs identified by MMseqs2 method [38,39] and used for AlphaFold modeling of hNaV1.8. a total number of homologous sequences identified per hNaV1.8 residue position is shown by a black trace. b) transmembrane (left panel) and extracellular (right panel) views of AlphaFold model of hNaV1.8. c) transmembrane (left panel) and extracellular (right panel) views of RoseTTAFold2 model of hNaV1.8. d) transmembrane (left panel) and extracellular (right panel) views of ESMFold model of hNaV1.8. AlphaFold2, RoseTTAFold2, and ESMFold models are colored by confidence score (pLDDT) from very low confidence (red) to good confidence (yellow) to high confidence (blue).
Relative to published structures, the overall topology of the models closest resembles the apo state (PDB: 7WFW), with AlphaFold2 having the lowest Cα RMSD at 2.0 Å (Figure 2(a)). This is exemplified by the AlphaFold2 model being able to predict VSD-II, VSD-II, VSD-IV (Figure 2(b)), and the domain III-IV intracellular linker (Figure 2(c)) with less than 2.0 Å Cα RMSD to the apo state structure (VSD-I is not resolved in the apo state). Notably, all methods produced pore domain Cα RMSD 2.0 Å or less relative to the apo-state, with AlphaFold2 having the lowest Cα RMSD at 0.72 Å (Figure 2(d)).
Figure 2.

Comparison of AlphaFold2, RoseTTAFold2, and ESMFold models and cryoEM structures of hNaV1.8. a) transmembrane (left panel) and extracellular (right panel) views of AlphaFold2, RoseTTAFold2, and ESMFold models superimposed to cryoEM structures of hNaV1.8 (PDB: 7WE4, 7WEL, 7WFR, and 7WFW) [22]. Intrinsically disordered N- and C-termini predicted by AlphaFold2, RoseTTAFold2, and ESMFold with low confidence were removed for clarity. b) transmembrane view of AlphaFold2, RoseTTAFold2, and ESMFold VSD-I, VSD-II, VSD-III, and VSD-IV models superimposed to apo-state hNaV1.8 structure (PDB: 7WFW). c) extracellular view of AlphaFold2, RoseTTAFold2, and ESMFold domain III-IV intracellular linker models superimposed to apo-state hNaV1.8 structure. d) extracellular view of AlphaFold2, RoseTTAFold2, and ESMFold pore domain (PD) models superimposed to apo-state hNaV1.8 structure.
For all three methods, the N-terminal domain (NTD) residues from M1 to N11, domain I-II intracellular loop region residues from L442 to D602, domain II-III intracellular loop region residues from S935 to L1092, and C-terminal domain (CTD) residues from N1875 to P1956 have low confidence scores (Figure 1). The low prediction confidence scores suggest that these regions are either highly flexible or inherently disordered in the absence of interacting intracellular partners. Notably, AlphaFold2 also predicted with good and high confidence scores parts of the NTD preceding the VSD-I (residues from N12 to S130), domain II-III intracellular loop region residues from N908 to R934, domain III-IV intracellular loop region from K1427 to N1463 that contains the intracellular gate important for NaV channel fast inactivation, and segments of the CTD (residues from T1733 to S1874). The NTD, domain I-II intracellular loop region, domain I-III intracellular loop region, and CTD regions are absent in currently resolved structures of hNaV1.8 (PDB IDs: 7WE4, 7WEL, 7WFR, and 7WFW) [22] but have been resolved in cryo-EM structures of other NaV channel subtypes [22,29–35]. The domain III-IV intracellular linker residues from K1427 to N1463 consistently had the closest match between hNaV1.8 structures and AlphaFold2 models (Cα RMSD = 1.1 Å), compared to ESMFold models (Cα RMSD = 2.6–2.8 Å) and RoseTTAFold2 models (Cα RMSD = 5.5–5.6 Å) (Figure 2(c)).
Focusing on the pore region, the AlphaFold2, ESMFold, and RoseTTAFold2 models of hNaV1.8 align closely with the experimentally resolved structures of hNaV1.8 (Cα RMSD = 0.7–1.8 Å) Notably, the AlphaFold2 model exhibits only minor differences over the selectivity filter, P1-helix, P2-helix, S5, and S6 segments (Figure 2(d)). One noticeable difference is the conformation of domain I S6 segment (DI-S6). AlphaFold2’s model shows a slight deviation in the helical turn near F386 residue in DI-S6, causing its side chain to point toward the DI-IV fenestration and pack together with M1716 in DIV-S6 (Figure 3(a)). In contrast, the cryo-EM structures of hNaV1.8 resolved in complex with a small molecule-based compound (A-803467) and in the apo state show F386 pointing downwards and M1716 oriented away from the fenestration. This difference between the AlphaFold2 model and the cryo-EM structure of hNaV1.8 may arise due to a different conformation of DI-S6 captured in the AlphaFold2 model. Interestingly, AlphaFold2 predicted a conformation of the ECL in domain I (ECL-I) with high confidence, which was partially unresolved in the cryo-EM structures of hNaV1.8. Specifically, residues 279 to 282 and 289 to 297 in ECL-I adopted a helical conformation with high confidence in the AlphaFold2 model of hNaV1.8 but are unstructured in the cryo-EM structure of hNaV1.8 (Figure 3(a)). For the same residues in ECL-I, ESMFold adopts a helical conformation between residues 283 to 291 with very low confidence (pLDDT <50), and RoseTTAFold2 adopts a loop conformation between residues 279 to 297 with very low confidence (Figure 3(a)).
Figure 3.

Comparison of specific regions in AlphaFold2, RoseTTAFold2, and ESMFold models to cryoEM structures of hNaV1.8. a) transmembrane (left and middle panel), and extracellular (right panel) views of F386 relative to M1716 at the domains I-IV (DI-IV) interface fenestration region in AlphaFiold2 model superimposed to apo-state hNaV1.8 structure (PDB: 7WFW). b) comparison of AlphaFold2, RoseTTAFold2, and ESMFold extracellular loop in domain I (ECL-I) model predictions relative to the partially resolved ECL-I in the apo-state hNaV1.8 structure. c) comparison of VSD-I S4 gating charges in AlphaFold2, RoseTTAFold2, and ESMFold models relative to cryoEM structures of hNaV1.8 (PDB: 7WE4, 7WEL, and 7WFR) [22]. d) comparison of VSD-II S4 gating charges in AlphaFold2, RoseTTAFold2, and ESMFold models relative to cryoEM structures of hNaV1.8 (PDB: 7WE4, 7WEL, 7WFR, and 7WFW) [22]. e) comparison of VSD-III S4 gating charges in AlphaFold2, RoseTTAFold2, and ESMFold models relative to cryoEM structures of hNaV1.8 (PDB: 7WE4, 7WEL, 7WFR, and 7WFW) [22]. f) comparison of VSD-IV S4 gating charges in AlphaFold2, RoseTTAFold2, and ESMFold models relative to cryoEM structures of hNaV1.8 (PDB: 7WE4, 7WEL, 7WFR, and 7WFW) [22]. Side chains of gating charge-carrying residues in the S4 segments are shown in stick representation and labeled.
The voltage sensor domains were predicted with high confidence in the AlphaFold2 model, good confidence in the ESMFold model, and low confidence in the RoseTTAFold2 model of hNaV1.8. However, the S3-S4 loop regions in VSD-I and VSD-II were predicted with good confidence in AlphaFold2, low confidence in ESMFold, and very low confidence in RoseTTAFold2, reflecting their flexibility (Figure 1). Despite the high confidence in the VSD predictions, comparison with the experimentally resolved hNaV1.8 structures revealed several key differences. The cryo-EM structures of hNaV1.8 in complex with A-803467 revealed various conformations of VSD-I, represented by class I, II, and III structures (PDB IDs: 7WE4, 7WEL, and 7WFR) [22]. VSD-I was not resolved in the cryo-EM structures of hNaV1.8 in an apo state (PDB: 7WFW) [22]. However, conformations of VSD-I in the AlphaFold2, ESMFold, and RoseTTAFold2 models of hNaV1.8 don’t align with the conformation of VSD-I in any of the cryo-EM structures of hNaV1.8 (Figure 3(c)) (Cα RMSD = 2.5–3.8 Å). With a very high pLDDT confidence score, this discrepancy raises a possibility that AlphaFold2’s model of hNaV1.8 represents another apo state of VSD-I. The AlphaFold2 model of hNaV1.8 VSD-II appears to be in a fully “up” state, with the gating charges R1, R2, and R3 in the S4 segment of VSD-II positioned above the gating charge transfer residue F708 in the S2 segment of VSD-II. In contrast, the class I cryo-EM structure of hNaV1.8 in complex with A-803467 (PDB: 7WE4) shows the gating charge R3 in the S4 segment of VSD-II at the gating charge transfer, considered to be a “half-click” down from the VSD-II state observed in the AlphaFold2 model (see Figure 2(b)). These observations may explain the difference in S3-S4 region conformations in VSD-II between the hNaV1.8 structures and AlphaFold2 models (Figure 3(d)).
In contrast to the models of hNaV1.8 VSD-I and VSD-II, AlphaFold2 conformations of VSD-III and VSD-IV appear to align closely with conformations of corresponding VSD-III and VSD-IV in the cryo-EM structures of hNaV1.8 (VSD-III: Cα RMSD = 0.6–0.8 Å, VSD-IV: Cα RMSD = 1.2–1.3 Å) (Figure 2(b)). ESMFold and RoseTTAFold2 both performed worse in comparison to AlphaFold2 (ESMFold VSD-III: Cα RMSD = 1.5 Å, ESMFold VSD-IV: Cα RMSD = 1.8–1.9 Å, RoseTTAFold2 VSD-III: Cα RMSD = 1.8–1.9 Å, RoseTTAFold2 VSD-IV: Cα RMSD = 2.4 Å). The backbone conformations of VSD-III in the AlphaFold2 model almost identically match those in the resolved structure (PDB: 7WE4), with gating charges consistently at the same position relative to the gating transfer (Figure 3(e)). Similarly, the gating charges in the S4 segment of VSD-IV in the AlphaFold2 models and cryo-EM structures of hNaV1.8 also occupy similar positions (Figure 3(f)). However, the S3-S4 conformations in VSD-IV are significantly different between the AlphaFold2 models and cryo-EM structures of hNaV1.8. As with VSD-I, there is heterogeneity observed in the S3-S4 region of VSD-IV, with multiple conformations resolved in cryo-EM structures of hNaV1.8 (PDB: 7WE4, 7WEL, and 7WFR, and 7WFW) [22]. Similar to the observations with VSD-II, the high confidence prediction of the VSD-IV S3-S4 region conformations by AlphaFold2 suggests a potentially different conformation of this region of hNaV1.8.
Structural modeling of voltage-gated calcium (Cav) channels
Voltage-gated calcium (CaV) channels mediate Ca2+ influx upon depolarization of cell membrane potentials [1,2,25,41,42]. Ten subtypes of CaV channels are divided into three main subfamilies, CaV1, CaV2, and CaV3, each serving distinct and crucial roles in physiological functions [28]. The CaV1 channels are responsible for muscle contraction, hormone secretion, and integrating synaptic inputs. The CaV2 channels play a key role in rapid communication in nerve cells. The CaV3 channels are crucial for the repetitive firing of action potentials in rhythmically firing cells, such as cardiac myocytes and thalamic neurons, contributing to regulating heart rhythm and synchronizing neural activities.
Cav1.1 structure is composed of α1, α2, β, γ, and γ subunits with α1 subunit responsible for voltage-dependent gating, calcium conduction, and modulation by small molecule drugs [23,43,44]. The protein sequence of CACNA1S gene encoded human CaV1.1 (hCaV1.1) α subunit (UniProt ID: Q13698) was used as input into AlphaFold2, RoseTTAFold2, and ESMFold for structure prediction. We assessed the quality of predicted models using the pLDDT confidence score, compared how close they are to the resolved rabbit CaV1.1 structure (rCaV1.1) (PDB: 5GJV) [43], and discussed the structural variations.
Our results showed that most of hCaV1.1 structure was predicted with high confidence by AlphaFold2 (overall pLDDT = 71.4) and ESMFold (overall pLDDT = 70.6) and with low confidence by RoseTTAFold2 (overall pLDDT = 54.1) (Figures 4 and 5). Specifically, hCaV1.1 VSD-I, VSD-II, VSD-III, and VSD-IV, pore domain, and ECL regions are predicted with high confidence (pLDDT >90) by AlphaFold2 and ESMFold and with good confidence by RoseTTAFold2 (60 < pLDDT < 90). However, most of the N-terminal domain formed by residues from M1 to K18, part of domain I-II intracellular loop region residues from L386 to A402, part of domain II-III intracellular loop region residues from L705 to E768, and most of the CTD regions formed by residues from Y1546 to S1780 and from G1835 to L1837 have been predicted with low confidence (pLDDT <50) by AlphaFold2, ESMFold, and RoseTTAFold2 (Figure 4(a)). This low confidence prediction suggests that these regions are either highly flexible or inherently disordered in the absence of interacting intracellular partners (Figure 4(a)). Interestingly, AlphaFold2 also predicted parts of the NTD from P19 to K51, part of domain I-II intracellular loop region residues from R347 to K385, part of domain II-III intracellular loop region residues from G689 to K704, domain III-IV intracellular loop region residues from G689 to K704 from E1073 to P1106, and CTD region residues from N1383 to G1545 with good confidence (70 < pLDDT < 90) and high confidence (pLDDT >90) (Figure 4(a,b)).
Figure 4.

AlphaFold2, RoseTTAFold2, and ESMFold models of hCaV1.1. a) plot of pLDDT confidence score versus hCaV1.1 residue position for AlphaFold2 (AF), RoseTTAFold2 (RF2), and ESMFold (ESM) models. right panel, multiple sequence alignment of hCaV1.1 sequence and its homologs identified by MMseqs2 method [38,39] and used for AlphaFold modeling of hCaV1.1. a total number of homologous sequences identified per hCaV1.1 residue position is shown by a black trace. b) transmembrane (left panel) and extracellular (right panel) views of AlphaFold model of hCaV1.1. c) transmembrane (left panel) and extracellular (right panel) views of RoseTTAFold2 model of hCaV1.1. d) transmembrane (left panel) and extracellular (right panel) views of ESMFold model of hCaV1.1. AlphaFold2, RoseTTAFold2, and ESMFold models are colored by confidence score (pLDDT) from very low confidence (red) to good confidence (yellow) to high confidence (blue).
Figure 5.

Comparison of AlphaFold2, RoseTTAFold2, and ESMFold models and cryoEM structures of hCaV1.1. a) Transmembrane (left panel) and extracellular (right panel) views of AlphaFold2, RoseTTAFold2, and ESMFold models superimposed to cryoEM structure of hCaV1.1 (PDB: 5GJV). b) transmembrane view of AlphaFold2 model of hCaV1.1 superimposed to cryoEM structure of hCaV1.1 (PDB: 5GJV). c) transmembrane view of RoseTTAFold2 model of hCaV1.1 superimposed to cryoEM structure of hCaV1.1 (PDB: 5GJV). d) transmembrane view of ESMFold model of hCaV1.1 superimposed to cryoEM structure of hCaV1.1 (PDB: 5GJV). AlphaFold2, RoseTTAFold2, and ESMFold models are colored by confidence score (pLDDT) from very low confidence (red) to good confidence (yellow) to high confidence (blue). e) transmembrane view of AlphaFold model of hCaV1.1 VSD-I, VSD-II, VSD-III, and VSD-IV colored in blue superimposed with cryoEM structures of hCaV1.1 (PDB: 5GJV) colored in purple. Side chains of gating charge-carrying residues in the S4 segments are shown in stick representation and labeled.
In the AlphaFold2, ESMFold, and RoseTTAFold2 models of hCaV1.1, the extracellular regions of the pore relatively closely match the cryo-EM structure of rCaV1.1 [43], exhibiting only minor differences over the selectivity filter, P1-helix, P2-helix, and ECL regions (Cα RMSD = 1.09 Å) (Figure 5(a-d)). However, conformations of the intracellular half of the S6 segments are captured in a different state in the AlphaFold2, ESMFold, and RoseTTAFold2 model of hCaV1.1 compared to the cryo-EM structure of rCaV1.1 (Cα RMSD = 4.96 Å) (Figure 5(a-d)).
The voltage sensor domains were predicted with high confidence (pLDDT >90) in the AlphaFold2 model of hCaV1.1. However, a significant drop in confidence was observed in the S3-S4 regions, reflecting their higher flexibility (Figure 5(b,e)). Notably, VSD-I state in the AlphaFold2 model of hCaV1.1 is matching closely VSD-I state in the cryo-EM structure of rCaV1.1 (Figure 5(e)). Similarly, the gating charges in the S4 segment of VSD-I in the AlphaFold2 model of hCaV1.1 and cryo-EM structures of rCaV1.1 also occupy similar positions (Figure 5(e)). VSD-II state in the AlphaFold2 model of hCaV1.1 has the S4 segment positioned a “half-click” down compared to VSD-II state in the cryo-EM structure of rCaV1.1 (Figure 5(e)). VSD-III state in AlphaFold2 model of hCaV1.1 has the S4 segment positioned a “half-click” up compared to VSD-III state in the cryo-EM structure of rCaV1.1 (Figure 5(e)). VSD-IV state in AlphaFold2 model of hCaV1.1 matches closely VSD-IV state in the cryo-EM structure of rCaV1.1 (Figure 5(e)). Similarly, the gating charges in the S4 segment of VSD-IV in the AlphaFold2 model of hCaV1.1 and cryo-EM structures of rCaV1.1 also occupy similar positions. The S3-S4 loop regions in VSD-I, VSD-III, and VSD-IV have not been resolved in the cryo-EM structure of rCaV1.1 (PDB: 5GJV) [43]. Notably, the S3-S4 loop regions in VSD-I, VSD-III, and VSD-IV AlphaFold2 models of hCaV1.1 have good confidence (pLDDT >70) (Figure 5(e)).
Structural modeling of voltage-gated potassium (Kv) channels
Voltage-gated potassium (KV) channels mediate K+ efflux upon membrane depolarization and regulate membrane potential [1,2]. There are 40 subtypes of KV channels divided into 12 main subfamilies, from KV1 to KV12 [28]. Differently from NaV and CaV channels, KV channels are tetramers where each VSD-PD pair forms a separate subunit. The KV1.3 channels regulate membrane potential and calcium signaling in lymphocytes and oligodendrocytes [45]. These channels form homotetramers with the VSDs and PDs forming the membrane-spanning region and a cytosolic tetramerization (T1) domain [24]. The protein sequence of KCNA3 gene encoded human KV1.3 (hKV1.3) (UniProt ID: P22001) was used as input into AlphaFold2, RoseTTAFold2, and ESMFold for structure prediction. We assessed the quality of predicted models using the pLDDT confidence score, compared how close they are to the resolved hKV1.3 structures (PDB: 7SSX and 7SSY) [24], and discussed the structural variations.
Our results showed that AlphaFold2 predicted the transmembrane region of hKV1.3 formed by VSDs and pore domain with high confidence (pLDDT >90), as illustrated in Figures 6 and 7. The first part of the N-terminal region formed by residues from M1 to E104 was predicted with low confidence (pLDDT <50) (Figure 6(a)). The second part of the N-terminal region that comprises the T1 domain formed by residues from R105 to S230 was predicted with good confidence (70 < pLDDT < 90) or high confidence (pLDDT >90) and closely matched cryo-EM structures of hKV1.3 [24] N-terminal region formed by the same region (Cα RMSD = 1.4 Å) (Figure 6(a)). All of the pore region formed by residues from M395 to T491 was predicted with good confidence (70 < pLDDT < 90) or high confidence (pLDDT >90) and closely matched cryo-EM structure of hKV1.3 [24] pore region (PDB: 7SSX) (Cα RMSD = 1.4 Å) (Figure 6(a,b)). C-terminal region formed by residues from E492 to V575 has low confidence (pLDDT <50) (Figure 6(a)). This low prediction confidence suggests that this region is either highly flexible or inherently disordered in the absence of interacting partners. Most of the VSD region formed by residues from P232 to S381 was predicted with good confidence (70 < pLDDT < 90) or high confidence (pLDDT >90) and closely matched cryo-EM structure of hKV1.3 [24] VSD region (Cα RMSD = 1.9 Å) (Figures 6(a) and 7(c)). However, the S1-S2 and S3-S4 loop regions were predicted with low confidence (pLDDT <50), reflecting unstructured nature of these regions that have not been resolved in the cryo-EM structures of hKV1.3 [24] (Figure 7(c)). The relative position of the gating charges in the S4 segment revealed a similar state for the modeled VSD as the one observed in the cryo-EM experimental structures (Figure 7(c)).
Figure 6.

AlphaFold2, RoseTTAFold2, and ESMFold models of hKV1.3. a) plot of pLDDT confidence score versus hKV1.3 residue position for AlphaFold2 (AF), RoseTTAFold2 (RF2), and ESMFold (ESM) models. right panel, multiple sequence alignment of hKV1.3 sequence and its homologs identified by MMseqs2 method [38,39] and used for AlphaFold modeling of hKV1.3. a total number of homologous sequences identified per hKV1.3 residue position is shown by a black trace. b) transmembrane (left panel) and extracellular (right panel) views of AlphaFold model of hKV1.3. c) transmembrane (left panel) and extracellular (right panel) views of RoseTTAFold2 model of hKV1.3. d) transmembrane (left panel) and extracellular (right panel) views of ESMFold model of hKV1.3. Unstructured N- and C- terminals are not shown for clarity in AlphaFold, RoseTTAFold2 and ESMFold models. AlphaFold2, RoseTTAFold2, and ESMFold models are colored by confidence score (pLDDT) from very low confidence (red) to good confidence (yellow) to high confidence (blue).
Figure 7.

Comparison of specific regions in AlphaFold2, RoseTTAFold2, and ESMFold models to cryoEM structures of hKV1.3. a) extracellular view of AlphaFold2, RoseTTAFold2, and ESMFold models of hKV1.3 superimposed with cryoEM structures of hKV1.3 (PDBs: 7SSX and 7SSY). b) superimposition of AlphaFold, RoseTTAFold2, and ESMFold models. The yellow arrow indicates the tilt observed in the ESMFold model of the T1 domain. c) transmembrane view of AlphaFold, RoseTTAFold2, and ESMFold models of hKV1.3 VSD. The side chains of the gating charges located in the S4 segment are shown in stick representation and labeled. d) intracellular view of AlphaFold2, RoseTTAFold2, and ESMFold models of hKV1.3 T1 domain after superimposition of the full models with the cryoEM structures of hKV1.3 colored in purple (PDB: 7SSX) and cyan (PDB: 7SSY). e) superimposition of individual T1 domains from AlphaFold2, RoseTTAFold2, and ESMFold models with the T1 domains from cryoEM structures of hKV1.3. f) extracellular view of AlphaFold2, RoseTTAFold2, and ESMFold models of hKV1.3 pore domain superimposed with cryoEM structures of hKV1.3 pore.
RoseTTAFold2 prediction for KV1.3 showed the expected architecture with the transmembrane domains, building up the VSDs and PDs, and the intracellular T1 domain (Figure 6(a)). The confidence of the prediction was good at the well-structured regions (80 < pLDDT < 90), but lower compared to the AF models, while it presented low confidence (pLDDT <50) at the N and C terminal regions, as well as at the S1-S2 and S3-S4 unstructured loops of the VSDs (Figure 6(a,c)). ESMFold prediction resulted in a similar model that also had good confidence prediction (80 < pLDDT < 90) at the well-structured transmembrane domains and T1 domain, and low (pLDDT <50) at the N and C terminal regions and VSD extracellular loops (Figures 6(a,d)). The predicted state of the VSDs was similar to the one in the AF models (Figure 7(c)).
Notably, the ESMFold model of KV1.3 presented a different arrangement of the T1 domains, in which the domain independently has the expected fold when compared to the available cryo-EM structures (Figure 7(e)) (RMSD <1 Å), but the domain of each subunit shifts outwards (Figure 7(b), yellow arrow) breaking the interactions among the four subunit T1 domains that are observed in AlphaFold2 and RoseTTAFold2 models and solved structures (Figure 7(d)). The predicted structure for the KV1.3 pore in AlphaFold2, RoseTTAFold2, and ESMFold models closely matched the one observed in experimental structures (RMSD = 1.5 Å) (Figure 7(f)).
Conclusions
Deep learning-based methods, such as AlphaFold [8], RosetTTAFold [9,10], and ESMFold [20] are useful for predicting structures of transmembrane regions of ion channels, including the voltage-sensing and pore domains, with high confidence. The extracellular and intracellular loop regions and intracellular N- and C-termini regions can be potentially predicted with high confidence if they are formed by α-helical or β-sheet secondary structure. Deep learning-based methods may predict alternative conformations of ion channels compared to known structures of identical or homologous ion channels. However, the accuracy of alternative ion channel conformations is only determined once confirmed by structural and experimental data. Modeling unstructured extracellular and intracellular loop regions and intracellular N- and C-termini regions remains challenging in the absence of potential protein partners to stabilize specific conformations of these regions. Structure prediction of ion channels using deep learning-based methods might be useful for designing therapeutics and molecular probes targeting specific ion channel subtypes. Finally, structural modeling of ion channels in complex with other proteins deep learning-based methods might reveal molecular mechanisms of ion channel modulation by extracellular, transmembrane, and intracellular proteins.
Acknowledgement
We dedicate this review to the memory of Dr. William A. Catterall – a creative scientist, supportive mentor, and inspiring collaborator.
Correction Statement
This article has been republished with minor changes. These changes do not impact the academic content of the article.
Funding Statement
This work was supported by National Institute of Neurological Disorders and Stroke grant1R61NS127285-01 to V.Y.-Y., National Heart, Lung, and Blood Institute grants 1R01HL159304-01 to Dr. Rose Dixon (V.Y-Y., Co-I) and R01HL128537 to Dr. Colleen Clancy (V.Y-Y., Co-I).
Disclosure statement
V. Yarov-Yarovoy reported personal fees from Gerson Lehrman Group, Grunenthal, Novo Ventures, and Praxis Precision Medicines outside the submitted work; in addition, V. Yarov-Yarovoy had a patent to “Peptides targeting sodium channels to treat pain” pending. No other disclosures were reported.
Data availability statement
AlphaFold2, ESMFold, and RoseTTAFold2 models of hNaV1.8, hCaV1.1, and hKV1.3 are available for download through the DRYAD database (https://doi.org/10.5061/dryad.z08kprrn0).
References
- [1].Hille B. Ion channels of excitable membranes. Sunderland, MA): Sinauer Associates; 2001. [Google Scholar]
- [2].Zheng J, Trudeau MC. Textbook of ion channels. Boca Raton: CRC Press; 2023. [Google Scholar]
- [3].Cao E, Liao M, Cheng Y, et al. TRPV1 structures in distinct conformations reveal activation mechanisms. Nature. 2013;504(7478):113–15. doi: 10.1038/nature12823 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Cheng Y. Single-particle cryo-EM at crystallographic resolution. Cell. 2015;161(3):450–7. doi: 10.1016/j.cell.2015.03.049 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [5].Armache JP, Cheng Y. Single-particle cryo-EM: beyond the resolution. Natl Sci Rev. 2019;6(5):864–6. doi: 10.1093/nsr/nwz127 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Cheng Y. Single-particle cryo-EM—How did it get here and where will it go. Science. 2018;361(6405):876–880. doi: 10.1126/science.aat4346 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Yip KM, Fischer N, Paknia E, et al. Atomic-resolution protein structure determination by cryo-EM. Nature. 2020;587(7832):157–61. doi: 10.1038/s41586-020-2833-4 [DOI] [PubMed] [Google Scholar]
- [8].Jumper J, Evans R, Pritzel A, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–9. doi: 10.1038/s41586-021-03819-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [9].Baek M, Anishchenko I, Humphreys IR, et al. Efficient and accurate prediction of protein structure using RoseTTAFold2. bioRxiv 2023:2023.05.24.542179.
- [10].Baek M, DiMaio F, Anishchenko I, et al. Accurate prediction of protein structures and interactions using a three-track neural network. Science. 2021;373(6557):871–6. doi: 10.1126/science.abj8754 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [11].UniProt C. UniProt: the universal protein knowledgebase in 2023. Nucleic Acids Res. 2023;51(D1):D523–D531. doi: 10.1093/nar/gkac1052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [12].Berman HM, Westbrook J, Feng Z, et al. The protein data bank. Nucleic Acids Res. 2000;28(1):235–242. doi: 10.1093/nar/28.1.235 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Humphreys IR, Pei J, Baek M, et al. Computed structures of core eukaryotic protein complexes. Science. 2021;374(6573):eabm4805. doi: 10.1126/science.abm4805 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Evans R, O’Neill M, Pritzel A, et al. Protein complex prediction with AlphaFold-Multimer. bioRxiv 2022:2021.10.04.463034.
- [15].Tunyasuvunakool K, Adler J, Wu Z, et al. Highly accurate protein structure prediction for the human proteome. Nature. 2021;596(7873):590–6. doi: 10.1038/s41586-021-03828-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [16].Watson JL, Juergens D, Bennett NR, et al. De Novo design of protein structure and function with RFdiffusion. Nature. 2023;620(7976):1089–1100. doi: 10.1038/s41586-023-06415-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Watson JL, Juergens D, Bennett NR, et al. Broadly applicable and accurate protein design by integrating structure prediction networks and diffusion generative models. bioRxiv 2022:2022.12.09.519842.
- [18].Wang J, Lisanza S, Juergens D, et al. Scaffolding protein functional sites using deep learning. Science. 2022;377(6604):387–94. doi: 10.1126/science.abn2100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [19].Rettie SA, Campbell KV, Bera AK, et al. Cyclic peptide structure prediction and design using AlphaFold. bioRxiv 2023:2023.02.25.529956.
- [20].Lin Z, Akin H, Rao R, et al. Evolutionary-scale prediction of atomic-level protein structure with a language model. Science. 2023;379(6637):1123–30. doi: 10.1126/science.ade2574 [DOI] [PubMed] [Google Scholar]
- [21].Varadi M, Anyango S, Deshpande M, et al. AlphaFold protein structure database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022;50(D1):D439–D44. doi: 10.1093/nar/gkab1061 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [22].Huang X, Jin X, Huang G, et al. Structural basis for high-voltage activation and subtype-specific inhibition of human Na v 1.8. Proc Natl Acad Sci U S A. 2022;119(30):e2208211119. doi: 10.1073/pnas.2208211119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [23].Wu J, Yan Z, Li Z, et al. Structure of the voltage-gated calcium channel Cav1.1 complex. Science. 2015;350(6267):aad2395. doi: 10.1126/science.aad2395 [DOI] [PubMed] [Google Scholar]
- [24].Selvakumar P, Fernandez-Marino AI, Khanra N, et al. Structures of the T cell potassium channel Kv1.3 with immunoglobulin modulators. Nat Commun. 2022;13(1):3854. doi: 10.1038/s41467-022-31285-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Catterall WA, Lenaeus MJ, Gamal El-Din TM. Structure and pharmacology of voltage-gated sodium and calcium channels. Annu Rev Pharmacol Toxicol. 2020;60(1):133–54. doi: 10.1146/annurev-pharmtox-010818-021757 [DOI] [PubMed] [Google Scholar]
- [26].Catterall WA. Forty years of sodium channels: structure, function, pharmacology, and epilepsy. Neurochem Res. 2017;42(9):2495–504. doi: 10.1007/s11064-017-2314-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Ahern CA, Payandeh J, Bosmans F, et al. The hitchhiker’s guide to the voltage-gated sodium channel galaxy. J Gen Physiol. 2016;147(1):1–24. doi: 10.1085/jgp.201511492 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Alexander SP, Mathie A, Peters JA, et al. The concise guide to pharmacology 2021/22: ion channels. Br J Pharmacol. 2021;178 Suppl 1:S157–S245. doi: 10.1111/bph.15539 [DOI] [PubMed] [Google Scholar]
- [29].Pan X, Li Z, Jin X, et al. Comparative structural analysis of human Na v 1.1 and Na v 1.5 reveals mutational hotspots for sodium channelopathies. Proc Natl Acad Sci U S A. 2021;118(11):118. doi: 10.1073/pnas.2100066118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Pan X, Li Z, Huang X, et al. Molecular basis for pore blockade of human Na+ channel Na v 1.2 by the μ-conotoxin KIIIA. Science. 2019;363(6433):1309–1313. doi: 10.1126/science.aaw2999 [DOI] [PubMed] [Google Scholar]
- [31].Li X, Xu F, Xu H, et al. Structural basis for modulation of human Na(V)1.3 by clinical drug and selective antagonist. Nat Commun. 2022;13(1):1286. doi: 10.1038/s41467-022-28808-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Pan X, Li Z, Zhou Q, et al. Structure of the human voltage-gated sodium channel Na v 1.4 in complex with β1. Science. 2018;362(6412):eaau2486. doi: 10.1126/science.aau2486 [DOI] [PubMed] [Google Scholar]
- [33].Jiang D, Shi H, Tonggu L, et al. Structure of the cardiac sodium channel. Cell. 2020;180(1):122–34 e10. doi: 10.1016/j.cell.2019.11.041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Fan X, Huang J, Jin X, et al. Cryo-EM structure of human voltage-gated sodium channel Na v 1.6. Proc Natl Acad Sci U S A. 2023;120(5):e2220578120. doi: 10.1073/pnas.2220578120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Shen H, Liu D, Wu K, et al. Structures of human Nav1.7 channel in complex with auxiliary subunits and animal toxins. Science. 2019;363(6433):1303–8. doi: 10.1126/science.aaw2493 [DOI] [PubMed] [Google Scholar]
- [36].Bouza AA, Isom LL. Voltage-Gated Sodium Channel beta Subunits and Their Related Diseases. Handb Exp Pharmacol. 2018;246:423–450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Mirdita M, Schutze K, Moriwaki Y, et al. ColabFold: making protein folding accessible to all. Nat Methods. 2022;19(6):679–82. doi: 10.1038/s41592-022-01488-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Mirdita M, Steinegger M, Soding J, et al. MMseqs2 desktop and local web server app for fast, interactive sequence searches. Bioinformatics. 2019;35(16):2856–8. doi: 10.1093/bioinformatics/bty1057 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Steinegger M, Soding J. MMseqs2 enables sensitive protein sequence searching for the analysis of massive data sets. Nat Biotechnol. 2017;35(11):1026–8. doi: 10.1038/nbt.3988 [DOI] [PubMed] [Google Scholar]
- [40].Johnson LS, Eddy SR, Portugaly E. Hidden Markov model speed heuristic and iterative HMM search procedure. BMC Bioinf. 2010;11(1):431. doi: 10.1186/1471-2105-11-431 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Nanou E, Catterall WA. Calcium channels, synaptic plasticity, and neuropsychiatric disease. Neuron. 2018;98(3):466–81. doi: 10.1016/j.neuron.2018.03.017 [DOI] [PubMed] [Google Scholar]
- [42].Catterall WA. Voltage-gated calcium channels. Cold Spring harbor perspectives in biology. Cold Spring Harbor Perspect Biol. 2011;3(8):a003947–a003947. doi: 10.1101/cshperspect.a003947 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Wu J, Yan Z, Li Z, et al. Structure of the voltage-gated calcium channel Cav1.1 at 3.6 Å resolution. Nature. 2016;537(7619):191–196. doi: 10.1038/nature19321 [DOI] [PubMed] [Google Scholar]
- [44].Zhao Y, Huang G, Wu J, et al. Molecular basis for ligand modulation of a Mammalian Voltage-Gated Ca(2+) channel. Cell. 2019;177(6):1495–506 e12. doi: 10.1016/j.cell.2019.04.043 [DOI] [PubMed] [Google Scholar]
- [45].Grissmer S, Nguyen AN, Aiyar J, et al. Pharmacological characterization of five cloned voltage-gated K+ channels, types Kv1.1, 1.2, 1.3, 1.5, and 3.1, stably expressed in mammalian cell lines. Mol Pharmacol. 1994;45:1227–1234. [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
AlphaFold2, ESMFold, and RoseTTAFold2 models of hNaV1.8, hCaV1.1, and hKV1.3 are available for download through the DRYAD database (https://doi.org/10.5061/dryad.z08kprrn0).
