Next Article in Journal
Discovery of the Genus Anapleus Horn, 1873 from Cretaceous Kachin Amber (Coleoptera: Histeridae)
Next Article in Special Issue
Sensitivity Differences and Biochemical Characteristics of Laodelphax striatellus (Fallén) to Seven Insecticides in Different Areas of Shandong, China
Previous Article in Journal
Where Does Honey Bee (Apis mellifera L.) Pollen Come from? A Study of Pollen Collected from Colonies at Ornamental Plant Nurseries
Previous Article in Special Issue
Large-Scale Monitoring of the Frequency of Ryanodine Receptor Target-Site Mutations Conferring Diamide Resistance in Brazilian Field Populations of Fall Armyworm, Spodoptera frugiperda (Lepidoptera: Noctuidae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pyrethroids in an AlphaFold2 Model of the Insect Sodium Channel

1
Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, ON L8S 4K1, Canada
2
Sechenov Institute of Evolutionary Physiology & Biochemistry, Russian Academy of Sciences, Saint Petersburg 194223, Russia
3
Almazov National Medical Research Centre, Saint Petersburg 197341, Russia
4
Department of Biology, Duke University, Durham, NC 27708, USA
*
Author to whom correspondence should be addressed.
Insects 2022, 13(8), 745; https://doi.org/10.3390/insects13080745
Submission received: 17 July 2022 / Revised: 11 August 2022 / Accepted: 15 August 2022 / Published: 18 August 2022
(This article belongs to the Special Issue The Evolution of Pesticide Resistance)

Abstract

:

Simple Summary

Pyrethroids are a major class of insecticides for controlling arthropod pests and human disease transmitting vectors. Resistance to pyrethroid insecticides presents a serious obstacle to effective pest control. One major mechanism of pyrethroid resistance is caused by mutations in the voltage-gated sodium channel, the target of pyrethroid toxic action. Previous mutational analyses, coupled with homology modeling, predicted two pyrethroid receptor sites on insect sodium channels. Mutations that are located in or close to these receptor sites likely confer resistance by directly reducing pyrethroid binding affinity. However, many mutations appear to be far from the receptor sites, and how these mutations confer pyrethroid resistance remains unknown. In this study, we employed the AlphaFold2 neural network to generate new models of mosquito and cockroach sodium channels. We then used computational methods to dock pyrethroids in the open and closed states of sodium channels to understand the atomic mechanisms of interactions between pyrethroids and sodium channels. Our analysis suggests that mutations residing beyond the receptor sites alter the action of pyrethroids by allosterically affecting the geometry of the receptor sites. The information gained from this study could be valuable for a structure-based design of novel pyrethroids.

Abstract

Pyrethroid insecticides stabilize the open state of insect sodium channels. Previous mutational, electrophysiological, and computational analyses led to the development of homology models predicting two pyrethroid receptor sites, PyR1 and PyR2. Many of the naturally occurring sodium channel mutations, which confer knockdown resistance (kdr) to pyrethroids, are located within or close to these receptor sites, indicating that these mutations impair pyrethroid binding. However, the mechanism of the state-dependent action of pyrethroids and the mechanisms by which kdr mutations beyond the receptor sites confer resistance remain unclear. Recent advances in protein structure prediction using the AlphaFold2 (AF2) neural network allowed us to generate a new model of the mosquito sodium channel AaNav1-1, with the activated voltage-sensing domains (VSMs) and the presumably inactivated pore domain (PM). We further employed Monte Carlo energy minimizations to open PM and deactivate VSM-I and VSM-II to generate additional models. The docking of a Type II pyrethroid deltamethrin in the models predicted its interactions with many known pyrethroid-sensing residues in the PyR1 and PyR2 sites and revealed ligand-channel interactions that stabilized the open PM and activated VSMs. Our study confirms the predicted two pyrethroid receptor sites, explains the state-dependent action of pyrethroids, and proposes the mechanisms of the allosteric effects of various kdr mutations on pyrethroid action. The AF2-based models may assist in the structure-based design of new insecticides.

1. Introduction

Pyrethroid insecticides are used extensively to control arthropod pests and human disease-transmitting vectors (WHO, 2007). The toxic effect of pyrethroids is due to their prolonged activation of voltage-gated sodium channels, which are critical for electrical signaling in the nervous system [1,2,3,4,5]. Development of pyrethroid resistance is a major threat to their sustained use in pest and vector control. Knockdown resistance (kdr) to pyrethroids is a well-known mechanism of pyrethroid resistance caused by naturally occurring mutations in sodium channels [5,6,7,8]. In the absence of cryo-EM structures of insect sodium channels within insecticides, the computational docking of pyrethroids aids in the understanding of the atomic mechanisms of their action and resistance.
The pore-forming α-subunit of eukaryotic channels folds from a single polypeptide chain of four homologous domains, I-IV. Each domain contains six transmembrane helical segments (S1–S6). Segments S1 to S4 constitute a voltage-sensing module (VSM). Four VSMs in domains I-IV are connected through linker-helices (S4–S5) to the pore-forming module (PM). The latter contains four outer helices S5, four pore-lining inner helices S6, and four membrane re-entrant P-loops. In the resting channel, the negative membrane potential forces positively charged S4 segments “down” in the cytoplasmic direction. Upon membrane depolarization, the S4 segments move outward which, through linker-helices S4–S5, initiates the opening of the activation gate at the cytoplasmic ends of S6s.
The first model of the pyrethroid-bound sodium channel [9] was elaborated for the house fly open sodium channel, using the crystal structure of a voltage-gated potassium channel Kv1.2 as a template [10]. In this model, pyrethroids bind to the lipid-exposed interface formed by the linker-helix IIS4-S5, the outer helix IIS5, and the inner helix IIIS6. We referred to the first pyrethroid receptor site in the II/III domain interface as PyR1. Later, we elaborated a model of the second pyrethroid receptor site in the I/II domain interface, PyR2, which contains ligand-binding residues in helices IL45, IS5, IS6, and IIS6 [11]. We further proposed that the simultaneous binding of pyrethroids to both PyR1 and PyR2 sites is necessary for the effective prolonging of the sodium channel opening [12].
Molecular modeling was further employed to understand the interactions of pyrethroid with the sodium channels in different insect species. For example, a hypothesis was proposed to explain the preferential binding of acaricidal pyrethroids to the sodium channels of ticks/mites [13]. The docking of permethrin and fenvalerate to the honeybee sodium channel revealed that the PyR1 site is similar to that in other insects [14]. Selective resistance of the bumble bee sodium channel to tau-fluvalinate was explained [15].
Identification of residues that are critical for the binding and action of pyrethroids is valuable not only for understanding the interactions of pyrethroids with sodium channels, but also for resistance monitoring. For example, systematic mutational analysis, coupled with computational modeling, uncovered a V253A mutation in the S4–S5 linker in domain I that reduced mosquito sodium channel sensitivity to pyrethroids [11]. Interestingly, a different substitution, V253F, at the same position was recently detected and concluded to be associated with pyrethroid resistance in an Ae. aegypti mosquito population in Brazil [16]. The role of this mutation has recently been functionally confirmed in the Xenopus oocyte system [17].
Due to a limited precision of previous homology models, which are based on the crystal and cryo-EM structures of potassium channels and prokaryotic sodium channels, the mechanisms of the state and dependent action of pyrethroids within and beyond the PyR1 and PyR2 sites are not well understood. Here, we employed the AlphaFold2 (AF2) neural network software to model the mosquito sodium channel AaNav1-1. The AF2 neural network is trained on crystal and cryo-EM structures of transmembrane proteins, including ion channels, in the absence of membrane potential. In the experimental structures of ion channels, VSMs are activated (S4 segments are in the “up” position), and PM is presumably inactivated. As a consequence, the AF2 model of AaNav1-1 predicted the channel with activated VSMs and inactivated PM.
We employed Monte Carlo energy minimizations (MCM) to impose the open PM conformation of the AaNav1-1 channel, as seen in the cryo-EM structure of the open rat Nav1.5 channel [18], docked a Type I pyrethroid permethrin (PMT) and a Type II pyrethroid deltamethrin (DMT) to this model, and found that the ligands are in direct contact with all residues whose kdr mutations within the PyR1 and PyR2 sites are reported. We further inactivated PM and deactivated VSMs in silico with bound DMT and monitored the DMT-bound contacts to understand why pyrethroids stabilize the activated channels. We further analyzed state-dependent contacts of those residues beyond the PyR1 and PyR2 sites, which are subject to kdr mutations, and found that these contacts stabilize the mutual disposition of the channel segments that contribute to the PyR1 and PyR2 sites. We propose that respective kdr mutations would allosterically affect the geometry of the PyR1 and/or PyR2 sites and hence, the affinity of the ligands. Our study integrates available experimental data and explains the state-dependent action of pyrethroids that bind within or beyond the PyR1 and PyR2 sites. The proposed structural models suggest mechanisms by which kdr mutations affect the action of pyrethroids and by which they may assist in the structure-based design of new pyrethroid insecticides.

2. Materials and Methods

2.1. AlphaFold2 Models

The AlphaFold2 (AF2) neural network software [19] is a cutting-edge computational technology, which in the latest CASP 2020 competition (Critical Assessment of Techniques for Protein Structure Prediction) greatly outperformed other methods in terms of the number of correct predictions and their accuracy [20]. We employed the AF2 software installed at the Digital Research Alliance of Canada (alliancecan.ca) to build full-fledged models of the mosquito sodium channel AaNav1-1 (Aedes aegypti, GenBank: AGG53064.1) and the cockroach sodium channel BgNav1-1a (Blattella germanica, GenBank: U73583.1). Cytoplasmic N- and C-ends and inter-domain linkers I/II and II/III of sodium channels are known to interact with cytoplasmic proteins, but details of these interactions are unknown or poorly understood. Since the AF2 models ignore the impact of cytoplasmic proteins, the respective cytoplasmic segments are disordered. We removed these segments from the models.

2.2. Modeling Channel States with the Open PM and Resting VSMs

We used the cryo-EM structure of rat Nav1.5 channel in the open state [18] to transform the PM of the insect channels in the open state using the methodology of large-scale in silico transformations of PM in full-fledged channel models, as described by the authors of [21]. Voltage-sensing domains were transformed from the activated to resting states, as described in the study in [22]. The designation of different models of the channels is given in Supplementary Figure S1B.

2.3. Residue Designations and 3D Alignment of Channel Structures

We used a residue-labeling scheme, which is universal for P-loop channels [23], as shown in Figure 1 and Supplemental Figure S1. This labeling scheme highlights symmetric locations of residues in different channel domains. We also used residue numbers in the housefly voltage-sensitive sodium channel (Musca domestica, GenBank: AAB47604).
Different models and experimental structures were 3D-aligned by minimizing the root mean square deviations of Cα atoms in P1 helices, the most 3D-consereved segments of P-loop channels, by matching atoms in the Kv1.2-Kv2.1 channel (PDB code: 2R9R).

2.4. Energy Optimizations, Ligand Docking, and Visualization

Calculations were performed with the ZMM program (http://zmmsoft.ca) that optimizes molecular structures in the space of generalized (internal) coordinates [24,25] using Monte Carlo energy minimizations [26]. Docking of pyrethroids from multiple starting poses into the PyR1 and PyR2 sites was performed as described elsewhere [12,27]. Molecular images were created using the PyMol Molecular Graphics System, Version 0.99rc6 (Schrödinger, LLC, New York, NY, USA).

3. Results and Discussions

3.1. The AF2 Model of AaNav1-1

Supplemental Figure S2 shows the best-ranking full-fledged AlphaFold2 model of the mosquito sodium channel AaNav1-1 superimposed with the cryo-EM structure of the non-functional cockroach sodium channel NavPaS [28]. The folding of the AF2 model in the transmembrane and extracellular regions is similar to that seen in the cryo-EM structures of sodium channels, which contain the activated voltage-sensing modules (VSMs) and the pore module (PM) in presumably inactivated state [29,30,31]. In addition, the AF2 model predicted several structured segments in cytoplasmic parts of the channel, which are not resolved in the cryo-EM structures of the sodium channels. We further refer to this model iAaNav1-1, where the superscripted prefix ‘i’ indicates the inactivated PM.
We used the ZMM program to calculate the energy of the AF2 model. Since both the AF2 software and the ZMM program employ the AMBER force field (although different versions), the ZMM energy of the iAaNav1-1 model had favorable energy and lacked any clashes. MC-minimization of the iAaNav1-1 model with fixed backbones and elastic bond angles in proline residues further decreased the energy.
We also generated AF2 models of the cockroach sodium channel iBgNav1-1a. The best-ranking AF2 models of the mosquito and cockroach sodium channels have very similar transmembrane parts, but some differences in the C-terminal domain and within a large extracellular loop (Figure S3A,B). These segments are far from the PyR1 and PyR2 sites. Importantly, the PyR1 and PyR2 sites are practically identical in the AF2 models of both channels (Figure S3C,D). Therefore, we further used only the mosquito sodium channel models. The 3D structures of the sodium channels of other insect organisms may deviate from that of AaNav1-1, especially in the extracellular and cytoplasmic loops. However, the fact that many different insect organisms are sensitive to the representative Type I and Type II pyrethroids DMT and PMT suggests that the geometry of the PyR1 and PyR2 sites in these organisms is generally similar. Peculiarities of the PyR1 and PyR2 sites in different organisms may underlie the species-selective action of some pyrethroids [15]. Analysis of these peculiarities would require the AF2 modeling of respective channels and the docking of species-specific pyrethroids. Such analysis is beyond the goals of this study.
Table 1 shows kdr mutations for which respective wild-type residues are within 4 Å from the pyrethroids that we docked into the PyR1 and PyR2 sites, as described below (Section 3.2 and Section 3.3). Table 2 shows engineered mutations within the PyR1/PyR2 sites that were generated and confirmed earlier in functional studies.

3.2. Docking of PMT in the PyR1 Site in Model iAaNav1-1

We generated 2000 random starting poses of PMT at the PyR1 site of iAaNav1-1 (Figure 1b), MC-minimized the energy for each starting pose, and collected the ligand-binding poses for which the energy of ligand–channel interactions was within 5 kcal/mol above the apparent global energy minimum. In the collected ensemble, we have chosen the PMT binding pose in which the ligand interacts with the maximal number of experimentally known pyrethroid-sending residues (Figure 1c). These include 12 residues whose kdr mutations are listed in Table 1 and two residues whose engineered substitutions affect the action of pyrethroids in functional studies (Table 2). This binding pose resembles the orientation of bifenthrin and PMT in the Kv1.2-based model of the housefly sodium channel [9,68] and the complex of τ-fluvalinate with the NavMs-based homology model of the open bumble bee sodium channel [15].

3.3. Docking of DMT in the PyR1 and PyR2 Sites in Model oAaNav1-1

The sampling strategy illustrated in Figure 1 was used to dock DMT in the PyR1 and PyR2 sites. The iAaNav1-1 model generated above using the AF2 neural network predicts the AaNav1-1 channel structure in the energetically most preferable state, with activated VSMs and presumably inactivated PM. However, pyrethroids are known to stabilize sodium channels in the activated state [5]. Therefore, we transformed iAaNav1-1 into the state with the open PM by stepwise moving Cα atoms of transmembrane helices in PM towards positions in the cryo-EM structure of the rat Nav1.5 channel, whose PM is captured in the open state [18]. Superposition of the obtained open-PM model oAaNav1-1 with the starting model iAaNav1-1 is shown in Supplemental Figure S4a,b. The in silico opening of PM in the iAaNav1-1 model caused rather small changes at the PyR1 site (Figure S4c). Therefore, the binding pose of PMT in the PyR1 site of the oAaNav1-1 model is likely similar to that in the iAaNav1-1 model (Figure 1c). However, the PM activation caused significant changes in the PyR2 site, especially in the IIS6 helix with pyrethroid sensing residues, including L2i16 that moved closer to the linker-helix IS4-S5 (Figure S4d).
Pyrethroids, particularly Type II pyrethroids, such as DMT, preferably bind to the activated state of insect sodium channels and exhibit strong state-dependent action [70,71]. Therefore, model oAaNav1-1 with the activated VSMs and in silico opened PM was then used to dock DMT in sites PyR1 and PyR2. We performed unbiased docking of DMT in the PyR1 and PyR2 sites from multiple starting poses as described for PMT in the iAaNav1-1a model (Figure 1b). Representative low-energy structures from the collected ensembles of binding poses are shown in Figure 2. In both the PyR1 and PyR2 sites, we found DMT binding poses in two opposite orientations. In one orientation, the dimethylcyclopropane ring fit the corner between the S4–S5 and S5 helices, whereas the terminal aromatic ring deeply penetrated into the fenestrations, as previously predicted [11,12]. In another ensemble, the tricyclic ring penetrated into the fenestrations and the terminal aromatic ring bound between helices S4–S5 and S6, as in models of bifenthrin and PMT in the Kv1.2-based models of the housefly sodium channel [9,68]. The latter orientation is also seen in our model of τ-fluvalinate in the bumble bee channel [15] and in the current AF2-based model of iAaNav1-1 with PMT in the PyR1 (Figure 1c).
In both ensembles, the nitrile group of DMT formed favorable contacts with polar residues in the S5 helices. In the PyR1 site, the bromine atoms approached the sulfur atom either in C2i14 (Figure 2a) or in M2k11 (Figure 2b), forming favorable bromine–sulfur halogen bonds [72]. In the binding pose shown in Figure 2c, DMT approached L3i5. Mutation F3i5L substantially decreased binding of several pyrethroids in the bumble bee sodium channel [15]. In our model oAaNav1-1, replacement L3i15F would enhance the binding of DMT due to π-stacking with the terminal aromatic ring of the ligand. The major difference between models of PMT in iAaNav1.1 (Figure 1c) and DMT in oAaNav1-1 (Figure 2a) is a favorable interaction of the DMT nitrile group with T2o10.
In both orientations of DMT in the PyR2 site, the ligand was with 4 Å from seven residues whose kdr mutations are reported (Table 1). These include two residues, V1k11 and T1o13, whose kdr mutations were recently functionally confirmed [17]. The energy of ligand–channel interactions in the opposed DMT-binding poses was comparable. We are not aware of experimental data that would enable us to discriminate between these binding poses of DMT and favor one of them.
As compared with the model of iAaNav1-1 with PMT in PyR1 (Figure 1c), the model of oAaNav1.1 with DMT in PyR1 (Figure 2a) has two peculiarities. Firstly, the DMT nitrile group H-bonded to T2o10. Secondly, the 2,2-dibromoetheny group in DMT, which is larger and more polarized than the 2,2-dichloroethenyl group in PMT [30], would establish stronger contacts with the channel residues.

3.4. In Silico Inactivating oAaNav1-1a with Two DMT Molecules in the PyR1 and PyR2 Sites

To understand how pyrethroids inhibit the channel transition from the open to the inactivated state, we in silico transformed model oAaNav1-1, with two bound DMT molecules (Figure 2a,b), back to the inactivated state by stepwise moving Cα atoms in helices S4–S5, S5, and S6 to their positions in the original AF2-based model iAaNav1-1, which we name o-iAaNav1-1.
The folding of the in silico inactivated o-iAaNav1-1 model with DMT molecules in the PyR1 and PyR2 sites is very similar to that of the original model iAaNav1-1, with inactivated PM. The linker helix IIS4-S5, which contains M2k11, shifted in the cytoplasmic direction and closer to VSM-II (Figure 3a). DMT in the PyR1 site shifted to preserve a contact between its terminal aromatic ring and the linker-helix. The 2,2-dibromoetheny moiety moved farther from the fenestration, which apparently shrunk, and lost its favorable halogen bonds with the sulfur atom of C2o14 and the oxygen atom of T2o10 (cf. Figure 3b,c). Overall, the favorable DMT contacts with the PyR1 site in oAaNav1-1 would resist the channel transition into the inactivated state, where DMT contacts with o-iAaNav1-1 are weaker. This could be one cause of the preferable binding of pyrethroids to the activated channels.
Figure 1. (A) Transmembrane topology of voltage-gated insect sodium channels indicating residues within PyR1 (circles) or PyR2 (triangles), or beyond PyR1 and PyR2 (squares), for which kdr mutations have been functionally confirmed in Xenopus oocytes. Residues are labeled according to the scheme, which is universal for P-loop channels [11,23]. A residue label includes the amino acid one-letter code, the domain number (I–IV), segment type (k, the linker-helix S4–S5; i, the inner helix S6; p, the P-loop; and o, the outer helix S5), and relative number of the residue in the segment. (B) Ensemble of PMT starting poses in the PyR1 site of the mosquito sodium channel model with inactivated pore domain, iAaNav1-1. (C) A low-energy binding pose of PMT in PyR1. Side chains of residues within 4 Å from PMT are shown as sticks. Labels of residues for which kdr mutations are reported (Table 1) are bold-underlined. Labels of residues for which kdr mutations are unknown, but engineered mutations are shown to affect the action of pyrethroids (Table 2), are bold-typed.
Figure 1. (A) Transmembrane topology of voltage-gated insect sodium channels indicating residues within PyR1 (circles) or PyR2 (triangles), or beyond PyR1 and PyR2 (squares), for which kdr mutations have been functionally confirmed in Xenopus oocytes. Residues are labeled according to the scheme, which is universal for P-loop channels [11,23]. A residue label includes the amino acid one-letter code, the domain number (I–IV), segment type (k, the linker-helix S4–S5; i, the inner helix S6; p, the P-loop; and o, the outer helix S5), and relative number of the residue in the segment. (B) Ensemble of PMT starting poses in the PyR1 site of the mosquito sodium channel model with inactivated pore domain, iAaNav1-1. (C) A low-energy binding pose of PMT in PyR1. Side chains of residues within 4 Å from PMT are shown as sticks. Labels of residues for which kdr mutations are reported (Table 1) are bold-underlined. Labels of residues for which kdr mutations are unknown, but engineered mutations are shown to affect the action of pyrethroids (Table 2), are bold-typed.
Insects 13 00745 g001
Figure 2. Alternative binding modes of deltamethrin in the PyR1 (A,B) and PyR2 (C,D) sites of the mosquito sodium channel with an in silico opened pore domain, oAaNav1-1. Side chains of residues within 4 Å from DMT are shown. Labels of residues for which kdr mutations are reported (Table 1) are bold-underlined. Labels of residues for which kdr mutations are unknown, but the engineered mutations that are shown to affect the action of pyrethroids (Table 2) are bold-typed.
Figure 2. Alternative binding modes of deltamethrin in the PyR1 (A,B) and PyR2 (C,D) sites of the mosquito sodium channel with an in silico opened pore domain, oAaNav1-1. Side chains of residues within 4 Å from DMT are shown. Labels of residues for which kdr mutations are reported (Table 1) are bold-underlined. Labels of residues for which kdr mutations are unknown, but the engineered mutations that are shown to affect the action of pyrethroids (Table 2) are bold-typed.
Insects 13 00745 g002
The major change in the PyR2 site upon the channel inactivation from the open state involves a large shift of L2i16 (Figure 3d), which lost its favorable contacts with the aromatic ring of DMT (cf. Figure 3e,f). These transformations may be considered as another cause of preferable binding of pyrethroids to the activated channels. Thus, favorable DMT interactions with the PyR1 and PyR2 sites in the open, but not inactivated, PM contribute to the state-dependent action of pyrethroids.

3.5. In Silico Deactivating VSM-I and VSM-II in oAaNav1-1 with DMT in the PyR1 and PyR2 Sites

Linker helices IS4-S5 and IIS4-S5, which contain pyrethroid-sensing residues, are immediately connected to the voltage-sensing helices in VSM-I and VSM-II, respectively. To explore the impact of deactivation of VSM-I and VSM-II on the binding of pyrethroids in PM, we in silico moved helices IS4 and IIS4 from the activated “up” positions to the resting “down” positions. The models with the deactivated VSMs are designated o-dAaNav1-1. The computations were performed with two DMT ligands bound to the PyR1 and PyR2 sites, with the aromatic rings in the interfaces between helices S4–S5 and S6, and the dimethylcyclopropane rings in the fenestrations (Figure 4a,c). The voltage-sensing arginines were step-wise shifted in the cytoplasmic direction, and the structure was MC-minimized at each step, as previously described [22]. As expected, the N-ends of the linker helices IS4-S5 and IIS4-S5 shifted down with the S4 helices.
DMT-sensing valine V1k11 in the linker helix IS4-S5 of PyR2 [12] and DMT-sensing methionine M2k11 [73] in the linker helix IIS4-S5 receded from the DMT molecules. DMT interactions with F3i17 in PyR1 and L2i16 in PyR2 weakened, while most of other ligand–channel contacts changed insignificantly. Interaction of DMT and PMT with V1k11 was predicted by molecular modeling and confirmed in functional studies [12]. Mutation V267/1k11F was recently found in pyrethroid-resistant Ae. aegypti mosquitoes [16], and the role of this mutation, V267/1k11F, in pyrethroid resistance has been functionally confirmed in Xenopus oocytes [17]. The weaker contacts of DMT with pyrethroid-sensing residues in the linker helices IS4-S5 and IIS4-S5 could be the third cause of the preferable binding of pyrethroids to the activated channels.
Some kdr mutations within PyR1/PyR2 sites may have dual effect on the action of pyrethroids. The mutations may affect the channel gating and decrease the open-state probability of the channel, as previously reported [70,74,75]. This would directly or indirectly enhance the action of kdr mutations on the ligand binding. For example, several kdr mutations involving V410/1i19 are reported: V410/1i19M/A/G in Helicoverpa zea [32], V410/1i19M in Heliothis virescens [74], and V410/1i19L in Cimex luctularis [33]. In our Kv1.2-based model, V410/1i19 was rather far from the PyR2 site [12]. In the AF2-based open-state model oAaNav1-1, V410/1i19 directly interacts with the PyR2-bound DMT in both binding poses (Figure 2b,d). Smaller substitutions V410/1i19A/G would decrease these interactions, while larger substitutions V410/1i19M/L would repel DMT from its binding site, supporting direct interactions of DMT with V4101i19. On the other hand, mutations V1i19M/L are known to affect the channel gating [5], likely due to repulsion of the large substitutions from the nearby M1p47 (Figure S5c). The repulsion would affect the position of the cytoplasmic half of IS6, which contributes to the activation gate at the S6 bundle crossing. IS6 also has L1i22, which contributes to the PyR2 site [12].

3.6. Allosteric Effects on Pyrethroid Action of kdr Mutations beyond the PyR1 and PyR2 Sites

Besides the kdr mutations within the PyR1 and PyR2 sites, multiple kdr mutations are located beyond the PyR1 and PyR2 sites and thus likely allosterically affect the action of pyrethroids (Table 3). The AF2-based models of AaNav1-1 allowed us to explore the contacts of the majority of respective WT residues and suggest mechanisms by which kdr mutations of respective residues allosterically affect the action of pyrethroids.

3.6.1. M827I in the VSM-II Linker IIS1-S2

D802 in the BgNav1-1a cockroach sodium channel is critical for the channel gating and action of pyrethroid insecticides [82]. In our recent study, D802 is proposed to form a salt bridge with the voltage-sensing arginine IIR1, the outermost arginine in IIS4 (Figure 5a), thus stabilizing the activated conformation of IIS4 [22]. Interestingly, as shown in Figure 5a, M827 approaches D823, which corresponds to D802 in BgNav1-1a. The hydrophobic methyl group of M827 would strengthen the salt bridge D823:IIR1, while much shorter isoleucine M827I would not stabilize the salt bridge. As a result, the “up” conformation of IIS4, which is necessary for favorable interactions of M2k11 with DMT in the PyR1 site (Figure 4a), would be destabilized.

3.6.2. G943A and Q945R in IIS6

Glycine G943/2o24 is involved in the knob-into-hole contact with F2p33 and F2i8 (Figure 5b), thus stabilizing the mutual disposition of helices IIS5 and IIS6. Mutation G2o24A would push away the extracellular ends of helices IIS5 and IIS6, which contribute, respectively, to the PyR1 and PyR2 sites.
Glutamine Q945/2o26 interacts with several residues in the extracellular loops of VSM-I (Figure 5b). Arginine Q2o26R would repel the voltage-sensing arginine IR1 and thus destabilize the “up” conformation of IS4 and linker-helix IS4-S5. The interaction of pyrethroids with the pyrethroid-sensing V1k11 [12,16] in the linker-helix IS4-S5 of the PyR2 site (Figure 4b) could be indirectly affected.

3.6.3. S989/2p55P and D1763Y in P-Loops

Mutation S989/2p55P at the C-end of helix IIP2 is found only in combination with V1016/2i18G [83]. In the AF2 model, V1016/2i18 is involved in inter-segment contacts with L2o9, L2o13, and L2p47. These contacts stabilize the mutual disposition of helices IIS5 and IIS6, which contribute to PyR1, as well as to the geometry of loop IIP1-P2, with the selectivity-filter glutamate E2p50 (Figure 5c). Mutation V1016/2i18G eliminates these contacts. The destabilization of loop IIP1-P2 would affect the ion selectivity of the sodium channel [84]. Mutation S989/2p55P eliminates H-bond S989/2p55---R2p45, whereas released R2p45 would find another H-bonding partner. This may compensate for the impact of mutation V1016/2i18G on the ion selectivity (Figure 5c).
Mutation D1763Y in the extracellular loop IVS5-P1 is also found in combination with V1016/2i18G [61]. In the AF2 model, D1763 is salt-bridged with K1774 at the C-end of loop IVS5-P1. Elimination of the salt bridge in mutant D1763Y would affect the stability of the IVP1-P2 region, which contributes to the selectivity filter. The fact that two mutations, S989P and D1763Y, which are far beyond PyR1 or PyR2, are found in combination with V1016/2i18G, suggests that the double mutations may compensate for possible alterations in channel functional properties associated with individual mutations, which could cause some fitness issues in mosquitoes carrying these individual mutations.

3.6.4. F1538/3i17L, D1549/3i28V, and E1553/3i32G in IIIS6

Phenylalanine F3i7 is involved in multiple hydrophobic contacts with P-loop residues in helices IIIP1 and IIIP2, thus stabilizing the P-loop domain (Figure 6d). Double mutation F1538/3i7L + M4i3I would destabilize this domain and affect the position of helix IIIS6, which contributes to the PyR1 site.
Glutamate E1553/3i32 forms an H-bond with S2i36 (Figure 6e). One helical turn above, D1549/3i28 is close to S2i34. Although an H-bond D1549/3i28---S2i34 is not seen in the iAaNav1-1 model, its formation is possible upon rotation of the serine side chain, which lacks any H-bonding partners. The double mutation D1549/3i28V + E1553/3i32G, which has been detected in pyrethroid-resistant natural populations of Helicoverpa armigera, would destroy the inter-domain contact at the cytoplasmic end of PM, increasing the flexibility of helices IIS6 and IIIS6, which contribute, respectively, to the PyR2 and PyR1 sites.
Thus, the AF2 model suggests mechanisms by which most of the kdr mutations listed in Table 3 affect the geometry of the PyR1 or/and PyR2 sites. However, for several kdr mutations located in the cytoplasmic segments, which are unresolved in the AF2 model, the allosteric mechanisms of action on the PyR1 or PyR2 sites remain unclear.

3.7. Engineered Substitutions Not Reported as kdr Mutations

The effects on the channel gating and sensitivity to pyrethroids of many residue substitutions from the systematic mutational analysis of insect sodium channels have been explored in functional studies (Table 2 and Table 4). While these mutations have not been observed as kdr mutations, functional studies have shed additional light on the mechanisms of the channel gating and sensitivity to pyrethroids. The AF2-based models also allowed us to further rationalize observed effects of some engineered mutations within the PyR1 and PyR2 sites (Table 2), as well as beyond the two sites (Table 4).
The Kv1.2-based model of the open AaNav1-1 channel suggested that I1k7 contributes to the PyR2 site; mutation I1k7A facilitated activation and inactivation of the AaNav1-1 channel and strongly decreased its sensitivity to DMT and PMT [11]. In the AF2-based model oAaNav1-1, I1k7makes tight hydrophobic inter-domain contacts with four residues (Figure 5d). DMT in the PyR2 site approaches I1k7, but it falls short of making direct contact with it. Mutation I1k7A would affect mutual disposition of helices IS4-S5, IIS5, and IIS6, which contribute to PyR1 and PyR2. Mutation I1k7A would also release contacts of the linker helix IS4-S5 with IIS5 and affect the channel gating. Facilitated inactivation would decrease the channel sensitivity to pyrethroids, as observed in previous experiments [11].
In the rotationally symmetric position of the open channel model, L2k7 is in tight hydrophobic contact with V3o12 and F3o16 in IIIS5 and F3i17 in IIIS6 (Figure 6f). Mutations L2k7F/I strongly facilitate activation and inactivation of the sodium channel from D. melanogaster [68]. These mutations would affect mutual disposition of helices IIS4-S5, IIS5, and IIIS6, which contribute to PyR1. This may explain why these mutations decrease the channel sensitivity to DMT [68].
Serine S1i29 is rather far from the PyR1-bound DMT, but mutation S1i29A strongly affects the channel sensitivity to PMT and DMT [12]. In model oAaNav1-1, S1i29 forms an intersegment H-bond with N1o5 (Figure 5e). Mutation S1i29A eliminates this H-bond and would affect mutual disposition of helices IS6 and IS5, both of which contribute to the PyR1 site.
Substitutions D823G/A/K, but not D823E, would eliminate the salt bridge of D823 with R1 in VSM-II, which stabilizes the “up” state of the voltage-sensing helix IIS4 [22]. This would destabilize the “up” conformation of IIS4 and the IIS4-S5 linker in which M2k11 interacts with PMT (Figure 1c) and DMT (Figure 2a,c). Destabilization of the IIS4-activated conformation would separate M2k11 from the pyrethroids (Figure 4a) and thus decrease their action, as demonstrated experimentally [82].
An H-bond between the side chain of N2o8 and the backbone carbonyl of G1k3 (Figure 5f) stabilizes the mutual disposition of helices IIS4-S5 and IIS5, which contribute to the PyR1 site. This may explain why engineered mutation N2o8I decreased the action of DMT. It should be noted that N2o8I, as well as Lk7F/I, which do not form direct contacts with pyrethroids, strongly decreased the action of DMT, but not PMT [68]. In view of our models, DMT is more sensitive than PMT to the mutations-associated deformations in the PyR1 site because the nitrile group of DMT makes an additional contact with T2o10 in the PyR1 site (Figure 2a,b). The fact that C2i14 makes specific contacts with DMT in both of its binding orientations (Figure 2a,c), but not with PMT (Figure 1c), explains why mutation C2i14A decreases the channel sensitivity to DMT, but not to PMT [68].

3.8. Cryo-EM Structure of NavPaS vs. AF2 Model of AaNav1-1

Earlier, the cryo-EM structure [28] was used to visualize the location of some pyrethroid-sensing residues in the PyR1 site of the mosquito sodium channels [87]. The overall folding of the NavPaS channel and the AF2 model iAaNav1-1 is similar, but not identical (Figure S2). While the NavPaS structure greatly advanced our understanding of the sodium channel structure, the AF2-based models of insect sodium channels appear more appropriate for analyzing pyrethroid action because (i) the NavPaS is a non-functional channel due to unknown cause(s), (ii) The AF2 models predicted many specific inter-residue contacts beyond the PyR1 and PyR2 sites (Figure 5 and Figure 6), which explain the allosteric effects of respective mutations, and (iii) the AF2 models show structures of some segments that are unresolved in the NavPaS structure. While these segments do not contribute to the PyR1 or PyR2 sites, they seem important for predicting the channel states with the open PM and deactivated VSMs. The in silico transformations of the AF2 model in this study allowed us to understand the structural causes of the state-dependent action of pyrethroids.

3.9. Limitations of Our Modeling Approach

Molecular dynamics simulations and molecular modeling with Monte Carlo energy minimizations, which employ high-resolution crystal or cryo-EM structures of the sodium channels, are two major approaches in theoretical studies aimed to understand atomic details of interactions of these large transmembrane proteins with drugs and toxins, e.g., the work described in [88,89,90,91]. AF2-based models and in silico transformed models are not precise enough to unambiguously predict the lowest-energy ligand-binding poses. Furthermore, our computations ignore entropy, do not use explicit lipid and water molecules, and employ a rather simple treatment of electrostatic interactions. Among multiple low-energy binding poses found in the PyR1 and PyR2 sites, we presented those where pyrethroids form the maximal number of contacts with experimentally known pyrethroid-sensing residues. Energy characteristics of ligand–channel complexes are just approximate estimates of the enthalpy contribution to the free energy of ligand binding in a specific pose. Nevertheless, our results, which are obtained without any biasing of the ligand contacts with specific residues, are consistent with multiple sets of experimental data. The AF2-based models advance our understanding of how the binding of pyrethroids within and beyond the PyR1 and PyR2 sites stabilizes the activated channel conformation.

4. Conclusions

Our computations, which employed the AF2 neural network and in silico transformation of the channel structures, demonstrate that the majority of kdr and engineered mutations are located within the PyR1 and PyR2 sites. The models also suggest mechanisms by which mutations beyond the PyR1 and PyR2 sites affect the action of pyrethroids. Our study resolved the long-standing question regarding why pyrethroids preferably bind to activated channels and stabilize the channels in the activated state. Our models show that pyrethroids lose or weaken contacts with some pyrethroid-sensing residues in PyR1 and PyR2 upon the pore domain transition from the open to inactivated state and upon the transition of voltage-sensing helices in VSM-I and VSM-II from the activated (up) to deactivated (down) positions. The bound pyrethroids resist these transitions, thus stabilizing the channel in the activated state. We believe that the proposed models will be useful for the structure-based design of novel pyrethroids.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/insects13080745/s1. Figure S1: Sequence alignment of domains I-IV of the AaNav1-1 channel and list of models used in this work. Figure S2: Intra-membrane and extracellular views at the cryo-EM structure of cockroach non-functional sodium channel NavPaS and the AF2 model of the mosquito sodium channel AaNav1-1. Figure S3: AF2 models of the AaNav1-1 and BgNav1-1a channels. Figure S4: The AF2 model of the channel with inactivated PM overlaid over the channel model with in silico opened pore domain. Figure S5: Intersegment contacts of residues in the PyR1 site whose kdr mutations may exert dual impact on the action of pyrethroids.

Author Contributions

Conceptualization, B.S.Z. and K.D.; methodology, B.S.Z.; software, B.S.Z.; investigation, B.S.Z. and K.D.; resources B.S.Z. and K.D.; writing—original draft preparation, B.S.Z.; writing—review and editing, B.S.Z. and K.D.; funding acquisition, K.D. and B.S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant to K.D. from NIH (GM057440) and grants to B.S.Z. from the Natural Sciences and Engineering Research Council of Canada (RGPIN-2020-07100) and the Russian Science Foundation (22-15-00186).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No experimental data were generated in this study.

Acknowledgments

We thank Funmilayo Egunjobi for reviewing the final version of the manuscript. Computations were performed using facilities provided by Compute Ontario (https://www.computeontario.ca/) and the Digital Research Alliance of Canada (https://www.alliancecan.ca).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AF2: AphaFold2; DMT, deltamethrin; kdr, knockdown resistance; MCM, Monte Carlo minimization; PM, pore module; PMT, permethrin; PyR1, pyrethroid receptor site 1; PyR2, pyrethroid receptor site 2; VSM, voltage-sensing module; WHO, World Health Organization.

References

  1. Narahashi, T. Mechanisms of action of pyretrhoids on sodium and calcium channel gating. Neuropharmacol. Pestic. Action 1986, 36–60. [Google Scholar]
  2. Narahashi, T. Neuronal ion channels as the target sites of insecticides. Pharmacol. Toxicol. 1996, 79, 1–14. [Google Scholar] [CrossRef] [PubMed]
  3. Bloomquist, J.R. Ion channels as targets for insecticides. Ann. Rev. Entomol. 1996, 41, 163–190. [Google Scholar] [CrossRef] [PubMed]
  4. Voth, A.R.; Ho, P.S. The role of halogen bonding in inhibitor recognition and binding by protein kinases. Curr. Top. Med. Chem. 2007, 7, 1336–1348. [Google Scholar] [CrossRef] [PubMed]
  5. Dong, K.; Du, Y.; Rinkevich, F.; Nomura, Y.; Xu, P.; Wang, L.; Silver, K.; Zhorov, B.S. Molecular biology of insect sodium channels and pyrethroid resistance. Insect Biochem. Mol. Biol. 2014, 50, 1–17. [Google Scholar] [CrossRef]
  6. Soderlund, D. Sodium channels. In Comprehensive Molecular Insect Science; Gilbert, L.I., Latrou, K., Gill, S.S., Eds.; Elsevier: New York, NY, USA, 2005; Volume 5, pp. 1–24. [Google Scholar]
  7. Davies, T.G.; Field, L.M.; Usherwood, P.N.; Williamson, M.S. DDT, pyrethrins, pyrethroids and insect sodium channels. IUBMB Life 2007, 59, 151–162. [Google Scholar] [CrossRef]
  8. Rinkevich, F.D.; Du, Y.; Dong, K. Diversity and Convergence of Sodium Channel Mutations Involved in Resistance to Pyrethroids. Pestic. Biochem. Physiol. 2013, 106, 93–100. [Google Scholar] [CrossRef]
  9. O’Reilly, A.O.; Khambay, B.P.; Williamson, M.S.; Field, L.M.; Wallace, B.A.; Davies, T.G. Modelling insecticide-binding sites in the voltage-gated sodium channel. Biochem. J. 2006, 396, 255–263. [Google Scholar] [CrossRef]
  10. Long, S.B.; Campbell, E.B.; Mackinnon, R. Crystal structure of a mammalian voltage-dependent Shaker family K+ channel. Science 2005, 309, 897–903. [Google Scholar] [CrossRef]
  11. Du, Y.; Nomura, Y.; Satar, G.; Hu, Z.; Nauen, R.; He, S.Y.; Zhorov, B.S.; Dong, K. Molecular evidence for dual pyrethroid-receptor sites on a mosquito sodium channel. Proc. Natl. Acad. Sci. USA 2013, 110, 11785–11790. [Google Scholar] [CrossRef]
  12. Du, Y.; Nomura, Y.; Zhorov, B.S.; Dong, K. Rotational Symmetry of Two Pyrethroid Receptor Sites in the Mosquito Sodium Channel. Mol. Pharmacol. 2015, 88, 273–280. [Google Scholar] [CrossRef]
  13. O’Reilly, A.O.; Williamson, M.S.; Gonzalez-Cabrera, J.; Turberg, A.; Field, L.M.; Wallace, B.A.; Davies, T.G. Predictive 3D modelling of the interactions of pyrethroids with the voltage-gated sodium channels of ticks and mites. Pest Manag. Sci. 2014, 70, 369–377. [Google Scholar] [CrossRef]
  14. Gosselin-Badaroudine, P.; Moreau, A.; Delemotte, L.; Cens, T.; Collet, C.; Rousset, M.; Charnet, P.; Klein, M.L.; Chahine, M. Characterization of the honeybee AmNaV1 channel and tools to assess the toxicity of insecticides. Sci. Rep. 2015, 5, 12475. [Google Scholar] [CrossRef]
  15. Wu, S.; Nomura, Y.; Du, Y.; Zhorov, B.S.; Dong, K. Molecular basis of selective resistance of the bumblebee BiNav1 sodium channel to tau-fluvalinate. Proc. Natl. Acad. Sci. USA 2017, 114, 12922–12927. [Google Scholar] [CrossRef]
  16. Itokawa, K.; Furutani, S.; Takaoka, A.; Maekawa, Y.; Sawabe, K.; Komagata, O.; Tomita, T.; de Lima Filho, J.L.; Alves, L.C.; Kasai, S. A first, naturally occurring substitution at the second pyrethroid receptor of voltage-gated sodium channel of Aedes aegypti. Pest Manag. Sci. 2021, 77, 2887–2893. [Google Scholar] [CrossRef]
  17. Sun, H.; Nomura, Y.; Du, Y.; Liu, Z.; Zhorov, B.S.; Dong, K. Two kdr mutations at the predicted second pyrethroid receptor site in sodium channels of Aedes aegypti and Nilaparvata lugens. Insect Biochem. Mol. Biol. 2022; in press. [Google Scholar] [CrossRef]
  18. Jiang, D.; Banh, R.; Gamal El-Din, T.M.; Tonggu, L.; Lenaeus, M.J.; Pomes, R.; Zheng, N.; Catterall, W.A. Open-state structure and pore gating mechanism of the cardiac sodium channel. Cell 2021, 184, 5151–5162.e11. [Google Scholar] [CrossRef]
  19. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Zidek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  20. Callaway, E. ‘It will change everything’: DeepMind’s AI makes gigantic leap in solving protein structures. Nature 2020, 588, 203–204. [Google Scholar] [CrossRef]
  21. Korkosh, V.S.; Kiselev, A.M.; Mikhaylov, E.N.; Kostareva, A.A.; Zhorov, B.S. Atomic Mechanisms of Timothy Syndrome-Associated Mutations in Calcium Channel Cav1.2. Front. Physiol. 2019, 10, 335. [Google Scholar] [CrossRef]
  22. Zhorov, B.S.; Du, Y.; Song, W.; Luo, N.; Gordon, D.; Gurevitz, M.; Dong, K. Mapping the interaction surface of scorpion beta-toxins with an insect sodium channel. Biochem. J. 2021, 478, 2843–2869. [Google Scholar] [CrossRef]
  23. Zhorov, B.S.; Tikhonov, D.B. Potassium, sodium, calcium and glutamate-gated channels: Pore architecture and ligand action. J. Neurochem. 2004, 88, 782–799. [Google Scholar] [CrossRef]
  24. Zhorov, B.S. Vector method for calculating derivatives of energy of atom-atom interactions of complex molecules according to generalized coordiantes. J. Struct. Chem. 1981, 22, 4–8. [Google Scholar] [CrossRef]
  25. Garden, D.P.; Zhorov, B.S. Docking flexible ligands in proteins with a solvent exposure- and distance-dependent dielectric function. J. Comput. Aided Mol. Des. 2010, 24, 91–105. [Google Scholar] [CrossRef]
  26. Li, Z.; Scheraga, H.A. Monte Carlo-minimization approach to the multiple-minima problem in protein folding. Proc. Natl. Acad. Sci. USA 1987, 84, 6611–6615. [Google Scholar] [CrossRef]
  27. Tikhonov, D.B.; Zhorov, B.S. Mechanism of sodium channel block by local anesthetics, antiarrhythmics, and anticonvulsants. J. Gen. Physiol. 2017, 149, 465–481. [Google Scholar] [CrossRef]
  28. Shen, H.; Zhou, Q.; Pan, X.; Li, Z.; Wu, J.; Yan, N. Structure of a eukaryotic voltage-gated sodium channel at near-atomic resolution. Science 2017, 355. [Google Scholar] [CrossRef]
  29. Jiang, D.; Shi, H.; Tonggu, L.; Gamal El-Din, T.M.; Lenaeus, M.J.; Zhao, Y.; Yoshioka, C.; Zheng, N.; Catterall, W.A. Structure of the Cardiac Sodium Channel. Cell 2020, 180, 122–134.e10. [Google Scholar] [CrossRef]
  30. Auffinger, P.; Hays, F.A.; Westhof, E.; Ho, P.S. Halogen bonds in biological molecules. Proc. Natl. Acad. Sci. USA 2004, 101, 16789–16794. [Google Scholar] [CrossRef]
  31. 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 2019, 363, 1303–1308. [Google Scholar] [CrossRef]
  32. Hopkins, B.W.; Pietrantonio, P.V. The Helicoverpa zea (Boddie) (Lepidoptera: Noctuidae) voltage-gated sodium channel and mutations associated with pyrethroid resistance in field-collected adult males. Insect Biochem. Mol. Biol. 2010, 40, 385–393. [Google Scholar] [CrossRef]
  33. Yoon, K.S.; Kwon, D.H.; Strycharz, J.P.; Hollingsworth, C.S.; Lee, S.H.; Clark, J.M. Biochemical and molecular analysis of deltamethrin resistance in the common bed bug (Hemiptera: Cimicidae). J. Med. Entomol. 2008, 45, 1092–1101. [Google Scholar] [CrossRef]
  34. Singh, O.P.; Dykes, C.L.; Das, M.K.; Pradhan, S.; Bhatt, R.M.; Agrawal, O.P.; Adak, T. Presence of two alternative kdr-like mutations, L1014F and L1014S, and a novel mutation, V1010L, in the voltage gated Na+ channel of Anopheles culicifacies from Orissa, India. Malar. J. 2010, 9, 146. [Google Scholar] [CrossRef] [PubMed]
  35. Brengues, C.; Hawkes, N.J.; Chandre, F.; McCarroll, L.; Duchon, S.; Guillet, P.; Manguin, S.; Morgan, J.C.; Hemingway, J. Pyrethroid and DDT cross-resistance in Aedes aegypti is correlated with novel mutations in the voltage-gated sodium channel gene. Med. Vet. Entomol. 2003, 17, 87–94. [Google Scholar] [CrossRef] [PubMed]
  36. Saavedra-Rodriguez, K.; Urdaneta-Marquez, L.; Rajatileka, S.; Moulton, M.; Flores, A.E.; Fernandez-Salas, I.; Bisset, J.; Rodriguez, M.; McCall, P.J.; Donnelly, M.J.; et al. A mutation in the voltage-gated sodium channel gene associated with pyrethroid resistance in Latin American Aedes aegypti. Insect Mol. Biol. 2007, 16, 785–798. [Google Scholar] [CrossRef] [PubMed]
  37. Tan, W.L.; Li, C.X.; Wang, Z.M.; Liu, M.D.; Dong, Y.D.; Feng, X.Y.; Wu, Z.M.; Guo, X.X.; Xing, D.; Zhang, Y.M.; et al. First detection of multiple knockdown resistance (kdr)-like mutations in voltage-gated sodium channel using three new genotyping methods in Anopheles sinensis from Guangxi Province, China. J. Med. Entomol. 2012, 49, 1012–1020. [Google Scholar] [CrossRef] [PubMed]
  38. Martinez-Torres, D.; Chandre, F.; Williamson, M.S.; Darriet, F.; Berge, J.B.; Devonshire, A.L.; Guillet, P.; Pasteur, N.; Pauron, D. Molecular characterization of pyrethroid knockdown resistance (kdr) in the major malaria vector Anopheles gambiae s.s. Insect Mol. Biol. 1998, 7, 179–184. [Google Scholar] [CrossRef] [PubMed]
  39. Stump, A.D.; Atieli, F.K.; Vulule, J.M.; Besansky, N.J. Dynamics of the pyrethroid knockdown resistance allele in western Kenyan populations of Anopheles gambiae in response to insecticide-treated bed net trials. Am. J. Trop. Med. Hyg. 2004, 70, 591–596. [Google Scholar] [CrossRef]
  40. Kim, H.; Baek, J.H.; Lee, W.-J.; Lee, S.-H. Frequency detection of pyrethroid resistance allele in Anopheles sinensis populations by real-time PCR amplification of specific allele (rtPASA). Pestic. Biochem. Physiol. 2007, 87, 54–61. [Google Scholar] [CrossRef]
  41. Cassanelli, S.; Cerchiari, B.; Giannini, S.; Bizzaro, D.; Mazzoni, E.; Manicardi, G.C. Use of the RFLP-PCR diagnostic test for characterizing MACE and kdr insecticide resistance in the peach potato aphid Myzus persicae. Pest Manag. Sci. 2005, 61, 91–96. [Google Scholar] [CrossRef]
  42. Jones, C.M.; Liyanapathirana, M.; Agossa, F.R.; Weetman, D.; Ranson, H.; Donnelly, M.J.; Wilding, C.S. Footprints of positive selection associated with a mutation (N1575Y) in the voltage-gated sodium channel of Anopheles gambiae. Proc. Natl. Acad. Sci. USA 2012, 109, 6614–6619. [Google Scholar] [CrossRef]
  43. Liu, Z.; Valles, S.M.; Dong, K. Novel point mutations in the German cockroach para sodium channel gene are associated with knockdown resistance (kdr) to pyrethroid insecticides. Insect Biochem. Mol. Biol. 2000, 30, 991–997. [Google Scholar] [CrossRef]
  44. Guerrero, F.D.; Jamroz, R.C.; Kammlah, D.; Kunz, S.E. Toxicological and molecular characterization of pyrethroid-resistant horn flies, Haematobia irritans: Identification of kdr and super-kdr point mutations. Insect Biochem. Mol. Biol. 1997, 27, 745–755. [Google Scholar] [CrossRef]
  45. Wu, M.; Gotoh, H.; Waters, T.; Walsh, D.B.; Lavine, L.C. Identification of an alternative knockdown resistance (kdr)-like mutation, M918L, and a novel mutation, V1010A, in the Thrips tabaci voltage-gated sodium channel gene. Pest Manag. Sci. 2013. [Google Scholar] [CrossRef]
  46. Sonoda, S.; Tsukahara, Y.; Ashfaq, M.; Tsumuki, H. Genomic organization of the para-sodium channel alpha-subunit genes from the pyrethroid-resistant and -susceptible strains of the diamondback moth. Arch. Insect Biochem. Physiol. 2008, 69, 1–12. [Google Scholar] [CrossRef]
  47. Forcioli, D.; Frey, B.; Frey, J.E. High nucleotide diversity in the para-like voltage-sensitive sodium channel gene sequence in the western flower thrips (Thysanoptera: Thripidae). J. Econ. Entomol. 2002, 95, 838–848. [Google Scholar] [CrossRef]
  48. Sonoda, S. Molecular analysis of pyrethroid resistance conferred by target insensitivity and increased metabolic detoxification in Plutella xylostella. Pest Manag. Sci. 2010, 66, 572–575. [Google Scholar] [CrossRef]
  49. Bass, C.; Schroeder, I.; Turberg, A.; Field, L.M.; Williamson, M.S. Identification of mutations associated with pyrethroid resistance in the para-type sodium channel of the cat flea, Ctenocephalides felis. Insect Biochem. Mol. Biol. 2004, 34, 1305–1313. [Google Scholar] [CrossRef]
  50. Rinkevich, F.D.; Su, C.; Lazo, T.; Hawthorne, D.; Tingey, W.; Naimov, S.; Scott, J.G. Multiple evolutionary origins of knockdown resistance (kdr) in pyrethroid-resistant Colorado potato beetle, Leptinotarsa decemlineata. Pestic. Biochem. Physiol. 2012, 104, 192–200. [Google Scholar] [CrossRef]
  51. Marshall, K.L.; Moran, C.; Chen, Y.; Herron, G.A. Detection of kdr pyrethroid resistance in the cotton aphid, Aphis gossypii (Hemiptera: Aphididae), using a PCR-RFLP assay. J. Pestic. Sci. 2012, 37, 169–172. [Google Scholar] [CrossRef]
  52. Carletto, J.; Martin, T.; Vanlerberghe-Masutti, F.; Brevault, T. Insecticide resistance traits differ among and within host races in Aphis gossypii. Pest Manag. Sci. 2009, 66, 301–307. [Google Scholar] [CrossRef]
  53. Karatolos, N.; Gorman, K.; Williamson, M.S.; Denholm, I. Mutations in the sodium channel associated with pyrethroid resistance in the greenhouse whitefly, Trialeurodes vaporariorum. Pest Manag. Sci. 2012, 68, 834–838. [Google Scholar] [CrossRef]
  54. Morin, S.; Williamson, M.S.; Goodson, S.J.; Brown, J.K.; Tabashnik, B.E.; Dennehy, T.J. Mutations in the Bemisia tabaci para sodium channel gene associated with resistance to a pyrethroid plus organophosphate mixture. Insect Biochem. Mol. Biol. 2002, 32, 1781–1791. [Google Scholar] [CrossRef]
  55. Gonzalez-Cabrera, J.; Davies, T.G.; Field, L.M.; Kennedy, P.J.; Williamson, M.S. An Amino Acid Substitution (L925V) Associated with Resistance to Pyrethroids in Varroa destructor. PLoS ONE 2013, 8, e82941. [Google Scholar] [CrossRef]
  56. Toda, S.; Morishita, M. Identification of three point mutations on the sodium channel gene in pyrethroid-resistant Thrips tabaci (Thysanoptera: Thripidae). J. Econ. Entomol. 2009, 102, 2296–2300. [Google Scholar] [CrossRef]
  57. Lee, S.H.; Smith, T.J.; Ingles, P.J.; Soderlund, D.M. Cloning and functional characterization of a putative sodium channel auxiliary subunit gene from the house fly (Musca domestica). Insect Biochem. Mol. Biol. 2000, 30, 479–487. [Google Scholar] [CrossRef]
  58. Hodgdon, H.E.; Yoon, K.S.; Previte, D.J.; Kim, H.J.; Aboelghar, G.E.; Lee, S.H.; Clark, J.M. Determination of knockdown resistance allele frequencies in global human head louse populations using the serial invasive signal amplification reaction. Pest Manag. Sci. 2010, 66, 1031–1040. [Google Scholar] [CrossRef]
  59. Roditakis, E.; Tsagkarakou, A.; Vontas, J. Identification of mutations in the para sodium channel of Bemisia tabaci from Crete, associated with resistance to pyrethroids. Pestic. Biochem. Physiol. 2006, 85, 161–166. [Google Scholar] [CrossRef]
  60. Srisawat, R.; Komalamisra, N.; Eshita, Y.; Zheng, M.; Ono, K.; Itoh, T.Q.; Matsumoto, A.; Petmitr, S.; Rongsriyam, Y. Point mutations in domain II of the voltage-gated sodium channel gene in deltamethrin-resistant Aedes aegypti (Diptera: Culicidae). Appl. Entomol. Zool. 2010, 45, 275–282. [Google Scholar] [CrossRef]
  61. Chang, C.; Shen, W.K.; Wang, T.T.; Lin, Y.H.; Hsu, E.L.; Dai, S.M. A novel amino acid substitution in a voltage-gated sodium channel is associated with knockdown resistance to permethrin in Aedes aegypti. Insect Biochem. Mol. Biol. 2009, 39, 272–278. [Google Scholar] [CrossRef]
  62. Pridgeon, J.W.; Appel, A.G.; Moar, W.J.; Liu, N. Variability of resistance mechanisms in pyrethroid resistant German cockroaches (Dictyoptera: Blattellidae). Pestic. Biochem. Physiol. 2002, 73, 149–156. [Google Scholar] [CrossRef]
  63. Kwon, D.H.; Clark, J.M.; Lee, S.H. Cloning of a sodium channel gene and identification of mutations putatively associated with fenpropathrin resistance in Tetranychus urticae. Pestic. Biochem. Physiol. 2010, 97, 93–100. [Google Scholar] [CrossRef]
  64. Kawada, H.; Higa, Y.; Komagata, O.; Kasai, S.; Tomita, T.; Thi Yen, N.; Loan, L.L.; Sanchez, R.A.; Takagi, M. Widespread distribution of a newly found point mutation in voltage-gated sodium channel in pyrethroid-resistant Aedes aegypti populations in Vietnam. PLoS Negl. Trop. Dis. 2009, 3, e527. [Google Scholar] [CrossRef] [PubMed]
  65. Kushwah, R.B.S.; Kaur, T.; Dykes, C.L.; Ravi Kumar, H.; Kapoor, N.; Singh, O.P. A new knockdown resistance (kdr) mutation, F1534L, in the voltage-gated sodium channel of Aedes aegypti, co-occurring with F1534C, S989P and V1016G. Parasit. Vectors 2020, 13, 327. [Google Scholar] [CrossRef] [PubMed]
  66. Katsavou, E.; Vlogiannitis, S.; Karp-Tatham, E.; Blake, D.P.; Ilias, A.; Strube, C.; Kioulos, I.; Dermauw, W.; Van Leeuwen, T.; Vontas, J. Identification and geographical distribution of pyrethroid resistance mutations in the poultry red mite Dermanyssus gallinae. Pest Manag. Sci. 2020, 76, 125–133. [Google Scholar] [CrossRef] [PubMed]
  67. He, H.; Chen, A.C.; Davey, R.B.; Ivie, G.W.; George, J.E. Identification of a point mutation in the para-type sodium channel gene from a pyrethroid-resistant cattle tick. Biochem. Biophys. Res. Commun. 1999, 261, 558–561. [Google Scholar] [CrossRef]
  68. Usherwood, P.N.; Davies, T.G.; Mellor, I.R.; O’Reilly, A.O.; Peng, F.; Vais, H.; Khambay, B.P.; Field, L.M.; Williamson, M.S. Mutations in DIIS5 and the DIIS4-S5 linker of Drosophila melanogaster sodium channel define binding domains for pyrethroids and DDT. FEBS Lett. 2007, 581, 5485–5492. [Google Scholar] [CrossRef]
  69. Du, Y.; Lee, J.E.; Nomura, Y.; Zhang, T.; Zhorov, B.S.; Dong, K. Identification of a cluster of residues in transmembrane segment 6 of domain III of the cockroach sodium channel essential for the action of pyrethroid insecticides. Biochem. J. 2009, 419, 377–385. [Google Scholar] [CrossRef]
  70. Soderlund, D.M. State-Dependent Modification of Voltage-Gated Sodium Channels by Pyrethroids. Pestic. Biochem. Physiol. 2010, 97, 78–86. [Google Scholar] [CrossRef]
  71. Vais, H.; Williamson, M.S.; Goodson, S.J.; Devonshire, A.L.; Warmke, J.W.; Usherwood, P.N.; Cohen, C.J. Activation of Drosophila sodium channels promotes modification by deltamethrin. Reductions in affinity caused by knock-down resistance mutations. J. Gen. Physiol. 2000, 115, 305–318. [Google Scholar] [CrossRef]
  72. Ford, M.C.; Saxton, M.; Ho, P.S. Sulfur as an Acceptor to Bromine in Biomolecular Halogen Bonds. J. Phys. Chem. Lett. 2017, 8, 4246–4252. [Google Scholar] [CrossRef]
  73. Lee, S.H.; Smith, T.J.; Knipple, D.C.; Soderlund, D.M. Mutations in the house fly Vssc1 sodium channel gene associated with super-kdr resistance abolish the pyrethroid sensitivity of Vssc1/tipE sodium channels expressed in Xenopus oocytes. Insect Biochem. Mol. Biol. 1999, 29, 185–194. [Google Scholar] [CrossRef]
  74. Park, Y.; Taylor, M.F.; Feyereisen, R. A valine421 to methionine mutation in IS6 of the hscp voltage-gated sodium channel associated with pyrethroid resistance in Heliothis virescens F. Biochem. Biophys. Res. Commun. 1997, 239, 688–691. [Google Scholar] [CrossRef]
  75. Xu, Q.; Zhang, L.; Li, T.; Zhang, L.; He, L.; Dong, K.; Liu, N. Evolutionary adaptation of the amino acid and codon usage of the mosquito sodium channel following insecticide selection in the field mosquitoes. PLoS ONE 2012, 7, e47609. [Google Scholar] [CrossRef]
  76. Pittendrigh, B.; Reenan, R.; ffrench-Constant, R.H.; Ganetzky, B. Point mutations in the Drosophila sodium channel gene para associated with resistance to DDT and pyrethroid insecticides. Mol. Gen. Genet. 1997, 256, 602–610. [Google Scholar] [CrossRef]
  77. Kristensen, M. Identification of sodium channel mutations in human head louse (Anoplura: Pediculidae) from Denmark. J. Med. Entomol. 2005, 42, 826–829. [Google Scholar] [CrossRef]
  78. Fallang, A.; Denholm, I.; Horsberg, T.E.; Williamson, M.S. Novel point mutation in the sodium channel gene of pyrethroid-resistant sea lice Lepeophtheirus salmonis (Crustacea: Copepoda). Dis. Aquat. Organ. 2005, 65, 129–136. [Google Scholar] [CrossRef]
  79. Wang, R.; Liu, Z.; Dong, K.; Elzen, P.J.; Pettis, J.; Huang, Z. Association of novel mutations in a sodium channel gene with fluvinate resistance in the mite, Varroa destructor. J. Apic. Res. 2002, 41, 17–25. [Google Scholar] [CrossRef]
  80. Head, D.J.; McCaffery, A.R.; Callaghan, A. Novel mutations in the para-homologous sodium channel gene associated with phenotypic expression of nerve insensitivity resistance to pyrethroids in Heliothine lepidoptera. Insect Mol. Biol. 1998, 7, 191–196. [Google Scholar] [CrossRef]
  81. Tsagkarakou, A.; Van Leeuwen, T.; Khajehali, J.; Ilias, A.; Grispou, M.; Williamson, M.S.; Tirry, L.; Vontas, J. Identification of pyrethroid resistance associated mutations in the para sodium channel of the two-spotted spider mite Tetranychus urticae (Acari: Tetranychidae). Insect Mol. Biol. 2009, 18, 583–593. [Google Scholar] [CrossRef]
  82. Du, Y.; Song, W.; Groome, J.R.; Nomura, Y.; Luo, N.; Dong, K. A negative charge in transmembrane segment 1 of domain II of the cockroach sodium channel is critical for channel gating and action of pyrethroid insecticides. Toxicol. Appl. Pharmacol. 2010, 247, 53–59. [Google Scholar] [CrossRef]
  83. Du, Y.; Nomura, Y.; Zhorov, B.S.; Dong, K. Sodium Channel Mutations and Pyrethroid Resistance in Aedes aegypti. Insects 2016, 7, 60. [Google Scholar] [CrossRef]
  84. Zhorov, B.S. Possible Mechanism of Ion Selectivity in Eukaryotic Voltage-Gated Sodium Channels. J. Phys. Chem. B 2021, 125, 2074–2088. [Google Scholar] [CrossRef]
  85. Chen, M.; Du, Y.; Nomura, Y.; Zhu, G.; Zhorov, B.S.; Dong, K. Alanine to valine substitutions in the pore helix IIIP1 and linker-helix IIIL45 confer cockroach sodium channel resistance to DDT and pyrethroids. Neurotoxicology 2017, 60, 197–206. [Google Scholar] [CrossRef]
  86. Chen, M.; Du, Y.; Nomura, Y.; Zhu, G.; Zhorov, B.S.; Dong, K. Mutations of two acidic residues at the cytoplasmic end of segment IIIS6 of an insect sodium channel have distinct effects on pyrethroid resistance. Insect Biochem. Mol. Biol. 2017, 82, 1–10. [Google Scholar] [CrossRef] [PubMed]
  87. Chen, M.; Du, Y.; Wu, S.; Nomura, Y.; Zhu, G.; Zhorov, B.S.; Dong, K. Molecular evidence of sequential evolution of DDT- and pyrethroid-resistant sodium channel in Aedes aegypti. PLoS Negl. Trop. Dis. 2019, 13, e0007432. [Google Scholar] [CrossRef]
  88. Nguyen, P.T.; DeMarco, K.R.; Vorobyov, I.; Clancy, C.E.; Yarov-Yarovoy, V. Structural basis for antiarrhythmic drug interactions with the human cardiac sodium channel. Proc. Natl. Acad. Sci. USA 2019, 116, 2945–2954. [Google Scholar] [CrossRef]
  89. Jiang, D.; Tonggu, L.; Gamal El-Din, T.M.; Banh, R.; Pomes, R.; Zheng, N.; Catterall, W.A. Structural basis for voltage-sensor trapping of the cardiac sodium channel by a deathstalker scorpion toxin. Nat. Commun. 2021, 12, 128. [Google Scholar] [CrossRef] [PubMed]
  90. Buyan, A.; Sun, D.; Corry, B. Protonation state of inhibitors determines interaction sites within voltage-gated sodium channels. Proc. Natl. Acad. Sci. USA 2018, 115, E3135–E3144. [Google Scholar] [CrossRef] [PubMed]
  91. Mazola, Y.; Marquez Montesinos, J.C.E.; Ramirez, D.; Zuniga, L.; Decher, N.; Ravens, U.; Yarov-Yarovoy, V.; Gonzalez, W. Common Structural Pattern for Flecainide Binding in Atrial-Selective Kv1.5 and Nav1.5 Channels: A Computational Approach. Pharmaceutics 2022, 14, 1356. [Google Scholar] [CrossRef] [PubMed]
Figure 3. In silico inactivating DMT-bound model oAaNav1-1. (A) Superposition of models oAaNav1-1 (green) and o-iAaNav1-1 (pink) with DMT in PyR1. (B,C) Surface representation of models oAaNav1-1 (B) and o-iAaNav1-1 (C) with DMT in site PyR1. Note a significant shift of the CBr2 group that lost halogen bonds with T2o10 and C2o14. (D) Superposition of models oAaNav1-1 and o-iAaNav1-1 with DMT in site PyR2. (E,F) Surface representation of models oAaNav1-1. (E) and o-iAaNav1-1 (F) with DMT in PyR2. Note a significant shift of L2i16 that lost favorable contacts with the DMT aromatic ring.
Figure 3. In silico inactivating DMT-bound model oAaNav1-1. (A) Superposition of models oAaNav1-1 (green) and o-iAaNav1-1 (pink) with DMT in PyR1. (B,C) Surface representation of models oAaNav1-1 (B) and o-iAaNav1-1 (C) with DMT in site PyR1. Note a significant shift of the CBr2 group that lost halogen bonds with T2o10 and C2o14. (D) Superposition of models oAaNav1-1 and o-iAaNav1-1 with DMT in site PyR2. (E,F) Surface representation of models oAaNav1-1. (E) and o-iAaNav1-1 (F) with DMT in PyR2. Note a significant shift of L2i16 that lost favorable contacts with the DMT aromatic ring.
Insects 13 00745 g003
Figure 4. In-silico deactivating VSMs in DMT-bound model oAaNav1-1. (A) In site PyR1, the side chain of M2k11 moves significantly from DMT. (B) In site PyR2, the side chains of I2i13 and L2i16 move from DMT.
Figure 4. In-silico deactivating VSMs in DMT-bound model oAaNav1-1. (A) In site PyR1, the side chain of M2k11 moves significantly from DMT. (B) In site PyR2, the side chains of I2i13 and L2i16 move from DMT.
Insects 13 00745 g004
Figure 5. Residues beyond the PyR1 and PyR2 sites form intersegment contacts whose destruction would allosterically affect the action of pyrethroids. Labels of residues that are subject to kdr mutations are bold-underlined (see Section 3.6 for more detail). Panels (AF) show, respectively, contacts of residues M827, G2o24, L2p47/S2.p55, I1k7, S1i29 and G1k3.
Figure 5. Residues beyond the PyR1 and PyR2 sites form intersegment contacts whose destruction would allosterically affect the action of pyrethroids. Labels of residues that are subject to kdr mutations are bold-underlined (see Section 3.6 for more detail). Panels (AF) show, respectively, contacts of residues M827, G2o24, L2p47/S2.p55, I1k7, S1i29 and G1k3.
Insects 13 00745 g005
Figure 6. Residues beyond the PyR1 and PyR2 sites form intersegment contacts whose destruction would allosterically affect the action of pyrethroids. Labels of residues that are subject to kdr mutations are bold-underlined (see Section 3.6 for more detail). Labels of residues, which are not subject of kdr mutations, but to their engineered substitutions, are explored in functional studies and are bold-typed. (see Section 3.8 for more detail). Panels (AF) show, respectively, contacts of residues A3k10, A3p4, M3i3, F3i7, D3i28/E3i32 and L2k7.
Figure 6. Residues beyond the PyR1 and PyR2 sites form intersegment contacts whose destruction would allosterically affect the action of pyrethroids. Labels of residues that are subject to kdr mutations are bold-underlined (see Section 3.6 for more detail). Labels of residues, which are not subject of kdr mutations, but to their engineered substitutions, are explored in functional studies and are bold-typed. (see Section 3.8 for more detail). Panels (AF) show, respectively, contacts of residues A3k10, A3p4, M3i3, F3i7, D3i28/E3i32 and L2k7.
Insects 13 00745 g006
Table 1. Kdr mutations in the PyR1 and PyR2 sites of insect sodium channels.
Table 1. Kdr mutations in the PyR1 and PyR2 sites of insect sodium channels.
Mutation 1LocationSpeciesReferenceFigure
V253/1k11FPyR2Aedes aegypti[16]2b,d
T267/1o13APyR2Plant hopper[17]2b,d
V410/1i19MPyR2Helicoverpa zea[32]2b,d
V410/1i19LPyR2Cimex luctularis[33]2b,d
V410/1i19G/APyR2Helicoverpa zea[32]2b,d
V1010/2i12L + L1014/2i16SPyR2 + PyR2Anopheles culicifacies[34]2b,d
I1011/2i13MPyR2Aedes aegypti[35]2b,d
I1011/2i13VPyR2Aedes aegypti[36]2b,d
N1013/2i15SPyR2Anopheles sinensis[37]2b,d
L1014/2i16FPyR2Anopheles gambiae[38]2b,d
L1014/2i16SPyR2Anopheles arabiensis[39]2b,d
L1014/2i16HPyR2Helicoverpa zea[32]2b,d
L1014/2i16CPyR2Anopheles sinensis[40]2b,d
L1014/2i16WPyR2Anopheles sinensis[37]2b,d
L1014/2i16F + F979SPyR2 + IIP1Myzus persicae[41]2b,d
L1014/2i16F + N1575YPyR2 + III-IVAnopheles gambiae[42]2b,d
L1014/2i16F+E435K+C785RPyR2Blattella germanica[43]2b,d
M918/2k11T + L1014/2i16FPyR1 + PyR2Haematobia i. irritans[44]1c, 4a,b,d
M918/2k11L + V1010/2i12APyR1 + PyR2Thrips tabaci[45]1c, 2a
M918/2k11I + L1014/2i16FPyR1 + PyR2Plutella xylostella[46]1c, 2a,b,d
T929/2o10I + L1014/2i16FPyR1 + PyR2Frankliniella occidentalis[47]1c, 2a,b,d
T929/2o10I + L1014/2i16FPyR1 + PyR2Plutella xylostella[48]1c, 2a,b,d
T929/2o10C + L1014/2i16FPyR1 + PyR2Frankliniella occidentalis[47]1c, 2a
T929/2o10V + L1014/2i16FPyR1 + PyR2Ctenocephalides felis[49]1c, 2a
T929/2o10N + L1014/2i16FPyR1 + PyR2L. decemlineata[50]1c, 2a,b,d
F979/2p44S + L1014/2i16FPyR1 + PyR2Myzus persicae[41]1c, 2a,b,d
M918/2k11TPyR1Aphis gossypii[51]1c, 2a
M918/2k11LPyR1Aphis gossypii[52]1c, 2a
M918/2k11L + L925/2o6IPyR1 + PyR1Trialeurodesvaporariorum[53]1c, 2a
M918/2k11VPyR1Bemisia tabaci[54]1c, 2a
L925/2o6IPyR1Bemisia tabaci[54]1c, 2a
L925/2o6VPyR1Varroa destructor[55]1c, 2a
T929/2o10IPyR1Thrips tabaci[56]1c, 2a
T929/2o10I + L932/2o13FPyR1 + PyR1P. humanus capitis[57]1c, 2a
T929/2o10I + M827IPyR1 + VSM2P. humanus capitis[58]1c, 2a
T929/2o10CPyR1Frankliniella occidentalis[47]1c, 2a
T929/2o10VPyR1Bemisia tabaci[59]1c, 2a
L932/2o13F + M827IPyR1 + VSM2P. humanus capitis[58]1c, 2a
I936/2o17VPyR1Helicoverpa zea[32]1c, 2a
L982/2p47WPyR1Aedes aegypti[35]1c
V1016/2i18GPyR1Aedes aegypti[35]2a,c
V1016/2i18G + S989/2p55PPyR1Aedes aegypti[60]2a,c
V1016/2i18G + D1763YPyR1Aedes aegypti[61]2a,c
V1016/2i18IPyR1Aedes aegypti[36]2a,c
F1020/2i22SPyR1Blattella germanica[62]2a
L1024/2i26VPyR1Tetranychus urticae[63]1c, S5b
F1534/3i13CPyR1Aedes aegypti[64]1c, 2a
F1534/3i13LPyR1Aedes aegypti[65]1c, 2a
F1537/3i16LPyR1Dermanyssus gallinae[66]1c
F1538/3i17IPyR1Rhipicephalus microplus[67]1c, 2a
1 Each mutation is labeled with the position of the residue in the house fly sodium channel (GenBank accession number: X96668), followed by another label using the universal labeling system for P-loop channels, as explained in the Figure 1 legend.
Table 2. Engineered mutations within PyR2 and PyR1 that affect the action of pyrethroids.
Table 2. Engineered mutations within PyR2 and PyR1 that affect the action of pyrethroids.
Mutation 1LocationChannelReferenceFigure
L260/1o6APyR2AaNav1-1[12]2b,d
I264/1o10CPyR2AaNav1-1[11]2b,d
I413/1i22APyR2AaNav1-1[12]2b,d
M918/2k11TPyR1DmNav[68]4c,2a,c
V922/2o3I aPyR1BiNav1-1[15]6c
L925/2o6IPyR1DmNav[68]1c
T929/2o10 IPyR1DmNav[68]1c,2a
L932/2o13FPyR1DmNav[68]1c,2a
C933/2o14APyR1DmNav[68]1c,2a
I936/2o17VPyR1DmNav[68]1c,2a
N1013/2i15SPyR2AaNav1-1[12]2b,d
F1020/2i22SPyR1AaNav1-1[12]1c,2a
L1023/2i25APyR1AaNav1-1[12]2a
L1024/2i26APyR1AaNav1-1[12]1c
F1526/3i5LPyR1 (τFVL)BiNav1-1[15]6c
V1529/3i8APyR1 (τFVL)BiNav1-1[15]1c,2a
I1533/3i12APyR1BgNav1-1a[69]1c,2a
F1537/3i16APyR1BgNav1-1a[69]2a
a M2o3 in AaNav1-1a. 1 Each mutation is labeled with the position of the residue in the house fly sodium channel (GenBank accession number: X96668), followed by another label using the universal labeling system for P-loop channels, as explained in the Figure 1 legend.
Table 3. Kdr Mutations beyond the PyR1 and PyR2 sites.
Table 3. Kdr Mutations beyond the PyR1 and PyR2 sites.
Mutation 1SpeciesReferenceImpactFigure
D59GBlattella germanica[43]
A99SCulex quinquefasciatus[75]
I254/1k12NDrosophila melanogaster[76]
E435/1i45KBlattella germanica[43]
C785RBlattella germanica[43]
M827/IIS1-S2IP. humanus capitis[58]PyR15a
G943/2o24AP. humanus capitis[77]PyR1, PyR25b
Q945/2o26RL. salmonis[78]PyR25b
S989/2p55P + V1016/2i18GAedes aegypti[60]IIP1-P2 + PyR15c, 2a,c
A1101T + P1879/CTDSPlutella xylostella[46]II/III + III/IV
A1410/3k10VDrosophila melanogaster[76]PyR16a
A1494/3p47VDrosophila melanogaster[76]PyR16b
M1524/3i3IDrosophila melanogaster[76]PyR16c
F1528/3i7L + M1823/4i3IVarroa destructor[79]PyR1 + IVS66d
D1549/3i28V + E1553/3i32GHelicoverpa armigera[80]PyR1, PyR26e
A1215D + F1538/3i17ITetranychus urticae[81]II/III + PyR1
W1594RCulex quinquefasciatus[75]N-end of IVS0
1 Each mutation is labeled with the position of the residue in the house fly sodium channel (GenBank accession number: X96668), followed by another label using the universal labeling system for P-loop channels, as explained in the Figure 1 legend.
Table 4. Engineered mutations beyond PyR1 and PyR2 that affect the action of pyrethroids.
Table 4. Engineered mutations beyond PyR1 and PyR2 that affect the action of pyrethroids.
Mutation aSpeciesReferenceImpactFigure
I249/1k7AAedes aegypti[11]PyR2, PyR15d
S420/1i29AAedes aegypti[12]PyR15e
D823G/A/KBlattella germanica[82]PyR15a
L914/2k7F/IDrosophila melanogaster[68]PyR16f
N927/2o8IDrosophila melanogaster[68]PyR1, PyR25f
A1410/3k10 VBlattella germanica[85]PyR16a
A1494/3p47VBlattella germanica[85]PyR16b
G1535/3i14A bBlattella germanica[69]PyR1
N1541/3i20ABlattella germanica[69]PyR1
D1549/3i28VBlattella germanica[86] 6e
a Each mutation is labeled with the position of the residue in the house fly sodium channel (GenBank accession number: X96668), followed by another label using the universal labelling system for P-loop channels, as explained in the Figure 1 legend. b Gating hinge in IIIS6.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhorov, B.S.; Dong, K. Pyrethroids in an AlphaFold2 Model of the Insect Sodium Channel. Insects 2022, 13, 745. https://doi.org/10.3390/insects13080745

AMA Style

Zhorov BS, Dong K. Pyrethroids in an AlphaFold2 Model of the Insect Sodium Channel. Insects. 2022; 13(8):745. https://doi.org/10.3390/insects13080745

Chicago/Turabian Style

Zhorov, Boris S., and Ke Dong. 2022. "Pyrethroids in an AlphaFold2 Model of the Insect Sodium Channel" Insects 13, no. 8: 745. https://doi.org/10.3390/insects13080745

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop