Next Article in Journal
Histopathological Changes in the Liver, Heart and Kidneys Following Malayan Pit Viper (Calloselasma rhodostoma) Envenoming and the Neutralising Effects of Hemato Polyvalent Snake Antivenom
Next Article in Special Issue
Cationic PLGA Nanoparticle Formulations as Biocompatible Immunoadjuvant for Serum Production and Immune Response against Bothrops jararaca Venom
Previous Article in Journal
In Silico Evaluation of Sesquiterpenes and Benzoxazinoids Phytotoxins against Mpro, RNA Replicase and Spike Protein of SARS-CoV-2 by Molecular Dynamics. Inspired by Nature
Previous Article in Special Issue
Crotoxin B: Heterologous Expression, Protein Folding, Immunogenic Properties, and Irregular Presence in Crotalid Venoms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

µ-Conotoxins Targeting the Human Voltage-Gated Sodium Channel Subtype NaV1.7

by
Kirsten L. McMahon
1,
Hue N. T. Tran
1,
Jennifer R. Deuis
1,
David J. Craik
1,
Irina Vetter
1,2,* and
Christina I. Schroeder
1,3,*,†
1
Institute for Molecular Bioscience, The University of Queensland, Brisbane, QLD 4072, Australia
2
The School of Pharmacy, The University of Queensland, Woolloongabba, QLD 4102, Australia
3
Center for Cancer Research, National Cancer Institute, National Institutes of Health, Frederick, MD 21702, USA
*
Authors to whom correspondence should be addressed.
Current address: Department of Peptide Therapeutics, Genentech, 1 DNA Way, South San Francisco, CA 94080, USA.
Submission received: 15 July 2022 / Revised: 25 August 2022 / Accepted: 26 August 2022 / Published: 30 August 2022
(This article belongs to the Special Issue Biotechnological Potential of Animal Venom and Toxins)

Abstract

:
µ-Conotoxins are small, potent, peptide voltage-gated sodium (NaV) channel inhibitors characterised by a conserved cysteine framework. Despite promising in vivo studies indicating analgesic potential of these compounds, selectivity towards the therapeutically relevant subtype NaV1.7 has so far been limited. We recently identified a novel µ-conotoxin, SxIIIC, which potently inhibits human NaV1.7 (hNaV1.7). SxIIIC has high sequence homology with other µ-conotoxins, including SmIIIA and KIIIA, yet shows different NaV channel selectivity for mammalian subtypes. Here, we evaluated and compared the inhibitory potency of µ-conotoxins SxIIIC, SmIIIA and KIIIA at hNaV channels by whole-cell patch-clamp electrophysiology and discovered that these three closely related µ-conotoxins display unique selectivity profiles with significant variations in inhibitory potency at hNaV1.7. Analysis of other µ-conotoxins at hNaV1.7 shows that only a limited number are capable of inhibition at this subtype and that differences between the number of residues in loop 3 appear to influence the ability of µ-conotoxins to inhibit hNaV1.7. Through mutagenesis studies, we confirmed that charged residues in this region also affect the selectivity for hNaV1.4. Comparison of µ-conotoxin NMR solution structures identified differences that may contribute to the variance in hNaV1.7 inhibition and validated the role of the loop 1 extension in SxIIIC for improving potency at hNaV1.7, when compared to KIIIA. This work could assist in designing µ-conotoxin derivatives specific for hNaV1.7.
Key Contribution: K.L.M., I.V. and C.I.S. conceptualized the project, K.L.M. carried out peptide synthesis, N.M.R.; and electrophysiology with assistance from H.N.T. (peptide chemistry), and J.R.D., and I.V. (electrophysiology). K.L.M. wrote the manuscript which was edited and reviewed by I.V. and C.I.S. All authors read and approved the manuscript prior to submission.

1. Introduction

Marine cone snail venom comprises a diverse mixture of disulfide-rich ion channel modulators called conotoxins [1]. µ-Conotoxins, characterised by a conserved cysteine framework (i.e., CC–C–C–CC), potently inhibit voltage-gated sodium (NaV) channels through a pore blockage mechanism [2]. Although humans express nine different NaV channel subtypes, µ-conotoxins have a preference for TTX-sensitive (TTX-s) subtypes, NaV1.1–1.4, 1.6 and NaV1.7, over TTX-resistant (TTX-r) subtypes NaV1.5, 1.8 and 1.9 [3,4,5]. Of particular interest is their inhibitory potency at NaV1.7, as this subtype has therapeutic potential for the treatment of pain [6,7]. However, the translation from hit to drug lead is hindered by the lack of specificity of µ-conotoxins for NaV channel subtypes. Given the essential role of NaV channels in the generation and propagation of action potentials, off-target inhibition of NaV channels could lead to serious unwanted side-effects. Therefore, identifying µ-conotoxin residues contributing to potent and selective inhibition of human (h)NaV1.7 channels would aid in the development of µ-conotoxin-derived therapeutics for diseases such as chronic pain.
Of the 22 µ-conotoxins characterised to date, only a few display potent inhibition of hNaV1.7 [4,8]. We recently reported the discovery of SxIIIC, a novel µ-conotoxin that inhibits hNaV1.7 with nanomolar potency, though non-selectively [9]. SxIIIC has high sequence homology with two other µ-conotoxins, SmIIIA and CnIIIC [10,11], both of which have been evaluated against mammalian NaV channels, but only CnIIIC was characterised at human isoforms (hNaV1.7 IC50 485 ± 94 nM) [4]. KIIIA is the only other µ-conotoxin reported to potently inhibit hNaV1.7 (IC50 97–363 nM), although also in a non-selective fashion [8]. Several studies have used KIIIA to investigate residues essential for NaV inhibition and identified several within the C-terminal α-helix of the peptide [8,12,13]. Those studies resulted in approaches for µ-conotoxin optimisation, including isolating the helical residues and downsizing KIIIA [14,15,16,17,18,19]. For example, Khoo and colleagues stabilised the KIIIA pharmacophore with a lactam helix, and while the selectivity profile across NaV channel subtypes remained unaffected, the stabilised helix only yielded micromolar inhibitory potency [14]. In another example, a peptidomimetic based on the KIIIA pharmacophore residues was designed by Brady et al., but only showed micromolar inhibition of NaV1.7 (species not reported) [15]. Other researchers have opted for a combination approach and designed ‘mini’ and ‘midi’ variants by combining two µ-conotoxins [16,17], or combining µ-conotoxins and spider toxins [18,19]; however, these too showed limited, non-selective inhibitory potency at hNaV1.7.
In the present study, we used electrophysiology assays to characterise the hNaV channel selectivity profiles of µ-conotoxins closely related to SxIIIC, including SmIIIA and KIIIA for NaV1.7 (neuronal and pain related) over NaV1.4 (muscle) sodium channel. Through structure-activity relationship analysis of hNaV1.7-inhibiting µ-conotoxins, we discovered point mutations that affect potency and selectivity at hNav1.7. These findings could lead to the development of subtype-selective µ-conotoxin derivatives with potential for therapeutic development.

2. Results

2.1. Oxidation of Native µ-Conotoxins and Analogues

SxIIIC analogues including [Δ1,2]SxIIIC, [R16A]SxIIIC, [R16H]SxIIIC, [R16Q]SxIIIC, [D17A]SxIIIC, [R20A]SxIIIC, [R20E]SxIIIC, [R20H]SxIIIC, [R20Q]SxIIIC, [R20W]SxIIIC were produced using thermodynamic folding as the majority of these peptides produced one major isomer during folding as observed for native SxIIIC, and this major isomer was used for our electrophysiology studies (Figure S2). These peptides were not analysed further using NMR or high-resolution MS (HR-MS) for disulfide bond connectivity confirmation due to the similarities in RP-HPLC chromatogram compared to native SxIIIC suggesting native disulfide connectivity. SmIIIA, the KIIIA analogue [loop1R]KIIIA and SxIIIC analogues ([Δ7,8]SxIIIC, [Δ6–9]SxIIIC and [G8R]SxIIIC) were produced using regioselective folding as they did not form one major isomer during thermodynamic folding conditions (Figure S3). SxIIIC analogues [Δ1,2:R16H]SxIIIC, [Δ1,2:R20A]SxIIIC, [R16H:R20A]SxIIIC and [Δ1,2:R16H:R20A]SxIIIC) produced two main peaks during RP-HPLC analysis following thermodynamic folding (Figure S4) and both peaks were evaluated using 1D 1H NMR (Figure S5) and compared to the NMR spectrum of native SxIIIC. When comparing the NMR spectrum of the SxIIIC peak 1 analogues (blue arrow in Figure S4), several additional or missing peaks were noted in amide region compared to the NMR spectrum of native SxIIIC. The 1D 1H NMR of the second peak more closely resembled native SxIIIC suggesting native disulfide bond conformation (Figure S5), leading us to use the second peak (red arrow in Figure S4) for electrophysiology inhibitory potency evaluation.

2.2. Selectivity of µ-Conotoxins for hNaV Channel Subtypes

µ-Conotoxin SmIIIA has been reported to only weakly inhibit mammalian NaV1.7 [20,21], despite sharing ~90% sequence homology with the recently described non-selective hNaV1.7 inhibitor SxIIIC [9]. KIIIA has also been shown to inhibit hNaV1.7, albeit at various potencies probably due to the use of different species of NaV1.7 and different recording systems [8,22]. To gain a comprehensive understanding of potency and selectivity of SmIIIA and KIIIA across human NaV channel subtypes, we assessed the peptides by automated whole-cell patch-clamp electrophysiology in HEK293 cells expressing hNaV1.1–1.7/β1 and CHO cells expressing hNaV1.8/β3 and compared inhibitory potency with our results for SxIIIC in the same cell lines. Concentration-response curves (Figure 1A) were used to determine IC50 values for the µ-conotoxins at each channel subtype (Table 1). Comparison of IC50 values revealed differences in subtype selectivity between the three µ-conotoxins (Figure 1B). All three µ-conotoxins most potently inhibited hNaV1.4, with SxIIIC (IC50 15 ± 11 nM) and SmIIIA (IC50 14 ± 2 nM) being significantly more potent than KIIIA (IC50 67 ± 15 nM). A similar pattern of inhibitory potency was observed for hNaV1.3 and hNaV1.6, while no differences in potency were seen at hNaV1.2. At NaV1.1, SxIIIC and KIIIA showed similar inhibitory potency, while SmIIIA was the least potent. KIIIA, SmIIIA and SxIIIC showed minimal to no inhibition of TTX-r hNaV1.5 and SmIIIA and SxIIIC showed no inhibition of TTX-r hNaV1.8 up to 1 µM (Table 1, Figure S6). Intriguingly, µ-conotoxin inhibitory potency at hNaV1.7 was most varied, with a 9-fold difference between SmIIIA (IC50 41 ± 4 nM) and KIIIA (IC50 379 ± 43 nM) and a 4-fold difference between SmIIIA and SxIIIC (IC50 152 ± 22 nM). This makes SmIIIA, in our hands, the most potent µ-conotoxin inhibitor of hNaV1.7 reported to date.

2.3. A Limited Number of µ-Conotoxins Display Inhibitory Potency at hNaV1.7

One of the challenges in comparing µ-conotoxin inhibitory potency between different studies is the differences in methods and species (human, mouse, and rat) used. While several µ-conotoxins have been evaluated at hNaV1.7 channels [4,23,24,25], only a few studies have directly compared µ-conotoxin inhibitory potency at hNaV1.7 in the same cell background. Therefore, we expanded our screen to analyse the hNaV1.7 inhibition of other conotoxins by automated whole-cell patch-clamp electrophysiology. CnIIIC inhibited hNaV1.7 with similar potency to KIIIA (IC50 546 ± 42 nM), whereas TIIIA, GIIIA, GIIIB and SIIIA showed no inhibitory potency at hNaV1.7 up to concentrations of 1 µM (Figure 2). These results are consistent with previous studies where TIIIA and GIIIB (3 µM), assayed using two-electrode voltage-clamp electrophysiology, were inactive at hNaV1.7 [25], and SIIIA (10 µM) only produced ~75% block at hNaV1.7 when evaluated using whole-cell patch-clamp methods [24].

2.4. Charged Residues in Loop 3 Affect Selectivity for hNaV1.7

Interestingly, except for SIIIA, all µ-conotoxins that potently inhibit hNaV1.7 contain five residues within loop 3 (the region between the fourth and fifth cysteine residues), whereas those that were inactive only contained four residues (Table 2). We were therefore interested in investigating the role of the charged residues within this loop in hNaV1.7 inhibition and used SxIIIC as a scaffold to design analogues with either charge-conserving or charge-neutralising substitutions (Table S1). We used SxIIIC in preference over the more active SmIIIA µ-conotoxin due to the ease of synthesis of SxIIIC compared to SmIIIA, which has proved challenging to fold. Whole-cell patch-clamp electrophysiology experiments comparing inhibitory potency of the mutants at NaV1.7 and NaV1.4 (Figure 3) showed that replacement of Arg16 by Ala, [R16A]SxIIIC, resulted in a two-fold loss of inhibition (IC50 334.1 ± 25.5 nM), whereas replacement of Arg16 with a neutral sidechain in [R16Q]SxIIIC did not significantly affect the potency of SxIIIC (Table 3). Interestingly, substituting Arg 16 with His [R16H]SxIIIC (His carries a neutral charge at physiological pH due to its low pKa value, 6.0) we observed a three-fold improvement in inhibition of hNaV1.7 (IC50 66.5 ± 2.5 nM) compared to SxIIIC. We also assessed the neighbouring residue Asp17 and observed no significant change in inhibition with the Ala-replacement analogue [D17A]SxIIIC (IC50 123.5 ± 18.1 nM). Surprisingly, we also observed no difference for the Arg20 analogue [R20A]SxIIIC (IC50 142.3 ± 12.9 nM), despite the analogous residue in KIIIA (Arg 14—KIIIA numbering) having previously been reported to shift subtype selectivity in favour of hNaV1.7 [8].
We next evaluated the inhibitory potency of these analogues, along with truncated SxIIIC, at hNaV1.4 to observe differences in subtype selectivity. Compared to SxIIIC (IC50 12.8 ± 13.0 nM), the R16H and R16Q mutations significantly reduced potency at hNaV1.4 (IC50 80.3 ± 24.7 nM and 99.5 ± 12.9 nM, respectively). Interestingly, this reduction gave IC50 values equipotent to hNaV1.7 (Figure 3). The same trend was observed for truncated [~1,2]SxIIIC and the [R20A]SxIIIC analogue, with both displaying reduced potency at hNaV1.4 (IC50 96.3 ± 13.0 nM and 64.8 ± 15.4 nM, respectively) but retaining inhibitory potency at hNaV1.7. Although [R16A]SxIIIC slightly lost potency at hNaV1.4 (IC50 64.8 ± 15.4 nM; Table 3), this did not affect subtype selectivity (Figure 3). In contrast, the D17A mutation in SxIIIC did not affect inhibition of either hNaV1.7 or hNaV1.4 (IC50 14.6 ± 7.7 nM). Together these results reveal that several SxIIIC residues, including Arg1, Gly2, Arg16 and Arg20, may contribute to selectivity between hNaV1.4 and hNaV1.7.

2.5. Double and Triple SxIIIC Mutants

Given the minimal disruption to hNaV1.7 potency caused by some SxIIIC mutations, we designed and synthesised double and triple mutant SxIIIC analogues (Table S1). Compared to native SxIIIC, all four analogues lost significant inhibitory potency at both subtypes (Figure 4). At hNaV1.7, the double mutants [Δ1,2; R16H], [Δ1,2; R20A] and [R16H; R20A]SxIIIC lost 4-, 3- and 6-fold potency, respectively (Table 3), whereas at hNaV1.4, the difference was much larger, with all three analogues decreasing potency by an average of 18-fold (Table 3). Notably, the triple mutant [Δ1,2; R16H; R20A] only displayed ~50% inhibition at hNaV1.7 at 1 µM and was not active at hNaV1.4 at 1 µM, the highest concentration tested (Figure 4 and Figure S7).

2.6. Key Structural Differences between hNaV1.7 Inhibiting µ-Conotoxins

The main sequence differences of SxIIIC and SmIIIA compared to KIIIA are, a two-residue N-terminal extension and a four-residue loop 1 extension (Figure 5A). To explore the significance of these extensions, we compared the NMR solution structures (PDB: SxIIIC 6X8R, SmIIIA 1Q2J and KIIIA 2LXG). Notably, while the cysteine framework of ‘native’ KIIIA [27] differed from SxIIIC and SmIIIA, this does not appear to affect the overall structure of these µ-conotoxins. Comparison of secondary Hα chemical shift deviations from random coil values showed similar shifts between residues Arg13–Ala20 (Figure 5A), indicating conserved structural features across this region. The N-terminal half of the µ-conotoxins displayed the most variation. Backbone alignment of 3D NMR structures between Cys10 and Cys22 (SxIIIC numbering; Figure 5B) showed that the extra residues in loop 1 of SxIIIC and SmIIIA extend the top half of the structure. We propose that this section contributes to the increased potency of these two µ-conotoxins at hNaV1.7. Furthermore, a substitution at Ala19 in SxIIIC to Ser19 in SmIIIA and KIIIA did not appear to affect the shifts for this residue. However, the neighbouring residue Arg20 (SxIIIC numbering) did show differences between the three µ-conotoxins and suggesting a different side-chain orientation of this residue.

2.7. N-Terminal Deletion Does Not Significantly Affect the Potency of SxIIIC

To assess the effects of N-terminal extensions observed in SxIIIC and SmIIIA compared to KIIIA, we designed the truncated SxIIIC analogue [Δ1,2]SxIIIC (Table S1). Due to difficulties synthesising native and modified SmIIIA, only SxIIIC and KIIIA analogues were pursued. In whole-cell patch-clamp electrophysiology experiments, we found that deletion of the N-terminal residues of SxIIIC did not significantly affect potency at hNaV1.7 (IC50 106.2 ± 19.4 nM) compared to native SxIIIC (IC50 152.2 ± 26.8 nM) (p > 0.05) (Figure 6).

2.8. Loop 1 Truncation Reduces Potency of SxIIIC at hNaV1.7

Interestingly, when we deleted residues within loop 1 of SxIIIC ([Δ7,8]SxIIIC), we observed reduced potency (IC50 313.6 ± 36.6 nM; Table 3) compared to native SxIIIC. As these IC50 values were similar to KIIIA (IC50 383.5 ± 15.2 nM; Figure 6), which has a similarly truncated loop 1, these results suggest that the additional loop 1 residues in SxIIIC are required for increased inhibition of hNaV1.7 compared to KIIIA. Furthermore, when we introduced an extra charged residue into loop 1 of SxIIIC ([G8R]SxIIIC) to mimic SmIIIA, we did not observe a significant change in potency (IC50 155.7 ± 18.9 nM; Table 3) compared to SxIIIC. This observation suggests that the additional charged residue in loop 1 of SmIIIA does not cause the 4.4-fold difference in inhibition of hNaV1.7 observed between SxIIIC and SmIIIA. Unfortunately, previously mentioned difficulties in producing SmIIIA limited our ability to confirm this using SmIIIA analogues. Finally, when modifying KIIIA to include additional residues within loop 1 ([loop1R]KIIIA), we expected this analogue to mimic the inhibitory potency of SxIIIC. However, we observed no significant change in potency (IC50 305.1 ± 23.9 nM; Table 3) compared to native KIIIA, suggesting that either the disulfide connectivity of KIIIA is unfavourable for loop 1 extensions, or that this peptide binds in a different orientation.

3. Discussion

µ-Conotoxins are potent inhibitors of disease-relevant NaV channels, yet their promiscuous activity across different subtypes has hindered their potential for therapeutic development, as off-target activity may lead to undesirable side effects. Due to an apparent lack of inhibitory potency of many µ-conotoxins at hNaV1.7, only a few studies have analysed structure–activity relationships at this subtype. Here, we identified key features of µ-conotoxins of SxIIIC, SmIIIA and KIIIA that contribute to inhibition of hNaV1.7, including loop length, and single point mutations that affect selectivity towards hNaV1.4 without affecting potency at hNaV1.7.
We initially assessed two µ-conotoxins, SmIIIA and KIIIA, by whole-cell patch-clamp methods, to evaluate their selectivity profiles for hNaV channel subtypes. Compared to SxIIIC [9], these highly homologous µ-conotoxins displayed distinct selectivity profiles (Table 4). Interestingly, our results are not in agreement with mammalian selectivity profiles reported previously for SmIIIA [20] or KIIIA [12,30]. While these inconsistencies may be attributed to the use of different techniques (two-electrode voltage- or patch-clamp methods on oocytes vs. mammalian cells; manual vs. automated techniques), or the presence of different auxiliary β-subunits, the most plausible explanation is species differences (human, rat or mouse). The earlier studies used rat isoforms, as opposed to the human subtypes used in this study. Despite the high sequence homology between species, it is not uncommon for µ-conotoxins to display different inhibitory potency across species. Such interspecies variability highlights the need to study human isoforms when evaluating the therapeutic potential of µ-conotoxins. We showed that SmIIIA inhibits hNaV1.7 more potently than both SxIIIC (~4-fold) and KIIIA (~10-fold). In our hands, SmIIIA is the most potent µ-conotoxin inhibitor of hNaV1.7 (IC50 41 ± 4 nM) reported to date.
As only a limited number of µ-conotoxins have been assessed at human NaV isoforms, we explored the ability of µ-conotoxins GIIIA, GIIIB, TIIIA, CnIIIC and SIIIA to inhibit hNaV1.7 by whole-cell patch-clamp methods. CnIIIC was found to be equipotent with KIIIA, whereas a lack of inhibitory potency was observed for the other four µ-conotoxins at the highest concentration tested (1 µM). The results for GIIIA, GIIIB and TIIIA were consistent with the homologous µ-conotoxin GIIIC, previously shown to be inactive at hNaV1.7 [23]. While the results of SIIIA were expected [24], it is interesting that SIIIA (at 10 nmol, corresponding to 0.7 mg/kg) can produce an analgesic effect in a murine model of inflammatory pain [31]. Leipold et al. [3] introduced the Domain II pore loop of hNaV1.7 into rNaV1.4 and identified Asn889 (hNaV1.7) as responsible for the reduced sensitivity of hNaV1.7 to SIIIA. The corresponding residue in mNaV1.7 is conserved, suggesting that this residue alone may not be responsible for species differences. Another possible explanation for the lack of activity of SIIIA at hNaV1.7 may be the apparent low net charge (+1.7) compared to hNaV1.7 inhibiting µ-conotoxins, which all have a net charge of +2.8 to +4.7 (Figure 7). While µ-conotoxins share key similarities such as a conserved disulfide network, the backbone sequences between these cysteines, referred to as loops 1–3, are varied. For example, GIIIA, GIIIB, GIIIC, PIIIA and TIIIA all have four residues in loop 3 (between Cys4–Cys6) and are classed in the M4-branch µ-conotoxins [32]. In contrast, SIIIA, KIIIA, CnIIIC, SmIIIA and SxIIIC have five residues in loop 3 and are classified as M5-branch µ-conotoxins [32]. M4-branch µ-conotoxins have an increased overall net charge (≥+4.8) compared to M5-branch µ-conotoxins (Figure 7), suggesting that overall net change may contribute to µ-conotoxins’ ability to inhibit hNaV1.7.
To explore the role of residues in loop 3, we designed a series of µ-conotoxin analogues, opting to base our designs on SxIIIC, as opposed to the more potent SmIIIA, due to difficulties in folding the latter µ-conotoxin. Our experience in synthesising SmIIIA is consistent with literature that reports synthetic production of SmIIIA often yields multiple isomers when folded under thermodynamic conditions [33]. In the current study we used a regioselective approach for SmIIIA synthesis to ensure desired disulfide connectivity; however, despite successful direction of the disulfide bonds during regioselective SmIIIA synthesis, the yield was low, and purification proved surprisingly challenging considering the high sequence homology with SxIIIC. In contrast, SxIIIC could easily be oxidised thermodynamically, resulting in a single major isomer, or using regioselective oxidation. In addition to targeted Ala-mutations in SxIIIC, in lieu of an Ala-walk (replacing all non-Cys and Gly residues with Ala), which has been conducted for other µ-conotoxins [12,25,34,35], we were interested in exploring the consequences of different chemical properties at specific residues and thus employed rational design for our SxIIIC analogues.
Replacement of Arg16 of SxIIIC with an Ala ([R16A]SxIIIC) reduced potency two-fold at hNaV1.7. Our results are in contrast to those of McArthur and colleagues, who found Ala-replacement of the equivalent residue in KIIIA ([R10A]KIIIA) increased inhibition at hNaV1.7 [8]. That study did not report the disulfide connectivity of KIIIA, and activity studies lacked β-subunit expression, which has since been shown to affect the affinity of µ-conotoxins to NaV channels [21]. Interestingly, we observed improved potency at hNaV1.7 and decreased inhibition at hNaV1.4 with a SxIIIC His-substitution at position 16, [R16H]SxIIIC. Whilst His theoretically carries a neutral charge at physiological pH, the charge state of this amino acid can be highly dependent on the environment created by surrounding amino acids and may carry a positive charge, which together with the additional bulk of the amino acid compared to the native Arg could explain the improved potency at NaV1.7. Another analogue, [R20A]SxIIIC, also did not affect hNaV1.7 inhibition and lost potency at hNaV1.4; however, the peptide did not display altered subtype selectivity. These results are also in contrast to Arg20 KIIIA mutations previously reported [8] that shifted selectivity towards hNaV1.7 from rNaV1.2 and rNaV1.4.
Using the three SxIIIC analogues that did not affect inhibition at hNaV1.7 but reduced potency at hNaV1.4, we designed and synthesised a series of double and triple mutants. None of the multiple SxIIIC mutant analogues improved potency at either channel tested (hNaV1.4 and hNaV1.7) or shifted selectivity in favour of hNaV1.7. While these results were disappointing, they were not completely unexpected. µ-Conotoxins have naturally evolved to potently and selectively target NaV channels, and the manipulation of such an evolutionarily refined natural product can be challenging. However, one explanation for the loss in potency may have been in choosing the incorrect isomer following thermodynamic oxidation (Figures S4 and S5). To assess if the folding contributed to the loss in inhibitory potency, we additionally tested the first eluting peak for each of the analogues against hNaV1.4 and found them to be less active than the second eluting peak (Figure S7). In summary, none of the multiple mutant analogues improved the potency or selectivity of SxIIIC.
Given the potential clinical applications of selective NaV1.7 inhibition, we compared the NMR structures of three µ-conotoxins with potent inhibition of hNaV1.7, i.e., SxIIIC, SmIIIA and KIIIA. We identified two major differences, in the form of extensions in the N-terminal and loop 1 residues of SxIIIC and SmIIIA, both absent in KIIIA. As many residues within the highly conserved C-terminal region of the µ-conotoxins have been shown to be critical for µ-conotoxin inhibition [8,12,13,36], we proposed that these structural differences observed in the N-terminal region influenced the differences in the inhibitory potency at hNaV1.7. To evaluate this, we designed an additional series of µ-conotoxins analogues using SxIIIC.
By creating a truncated N-terminal analogue, we found that the N-terminal residues of SxIIIC are not essential for inhibition of hNaV1.7. However, truncated SxIIIC, which maintained inhibitory potency at hNaV1.7, compared to native SxIIIC, displayed decreased potency at hNaV1.4. The N-terminal extension of SxIIIC, and potentially SmIIIA, is likely to be essential for inhibition at other subtypes, as seen with µ-conotoxin KIIIB (identical to KIIIA but with a two-residue N-terminal extension) which inhibits rNaV1.2 with 5-fold greater affinity than KIIIA [27]. These results are consistent with truncated SIIIA, which showed that these N-terminal residues played a non-essential role in rNaV1.2 inhibition [34]. It could be suggested that these N-terminal residues play a role in guiding the µ-conotoxin to the binding site in different NaV channel subtypes. In contrast to the N-terminal extension, we discovered that residues within loop 1 of SxIIIC influenced inhibition of hNaV1.7. When loop 1 residues were removed to mimic KIIIA, the potency of SxIIIC reverted to IC50 values similar to KIIIA. As many studies interested in optimising KIIIA for NaV1.7 inhibition have focused on downsizing or reducing the µ-conotoxins to only retain residues contributing to the pharmacophore [14,15,16], our findings identify for the first time the importance of the residues within loop 1 of SxIIIC for potent inhibition of hNaV1.7. Interestingly, when we reintroduced extra loop 1 residues into KIIIA, we did not observe the expected improved potency. By revisiting the NMR structure, we observed that although KIIIA had a different disulfide connectivity to SxIIIC, the overall conformation of the disulfide bonds did not change. This suggests that KIIIA with I–V, II–IV, III–VI connectivity might not be optimal for the addition of residues in loop 1 and this could be explored in future studies.
In addition to improved selectivity for NaV1.7, which is essential to avoid severe systemic side effects, species selectivity of the µ-conotoxins currently limits preclinical efficacy test in relevant rodent models of pain. Accordingly, future work, such as modelling or structural studies using relevant channel states and subtypes, will be required to facilitate drug development efforts, although oral delivery of peptides remains a challenge that would likely limit µ-conotoxin-based therapeutics to parenteral administration.
In conclusion, we have identified that the number of residues within loops 1 and 3 of SxIIIC, SmIIIA and KIIIA contributes to potency at hNaV1.7. Overall, the findings of this study provide valuable information for future design of hNaV1.7 specific µ-conotoxin derived inhibitors.

4. Methods

4.1. Peptide Synthesis

KIIIA, SIIIA, SxIIIC and analogues including [Δ1,2]SxIIIC, [R16A]SxIIIC, [R16H]SxIIIC, [R16Q]SxIIIC, [D17A]SxIIIC, [R20A]SxIIIC, [R20E]SxIIIC, [R20H]SxIIIC, [R20Q]SxIIIC, [R20W]SxIIIC, [Δ1,2:R16H]SxIIIC [Δ1,2:R20A]SxIIIC, [R16H:R20A]SxIIIC and [Δ1,2:R16H:R20A]SxIIIC) were synthesised using methods described earlier [9]. Protecting groups trityl (Trt), acetamidomethyl (Acm) and 4,4′-dimethylsulfinylbenzhydryl (Msbh) were used for regioselective protection of SmIIIA and SxIIIC analogues ([Δ7,8]SxIIIC, [Δ6–9]SxIIIC and [G8R]SxIIIC) with connectivity Cys3–Cys15(Msbh)/Cys4–Cys21(Acm)/Cys10–Cys22(Trt). The KIIIA analogue [loop1R]KIIIA was produced with connectivity Cys1–Cys15(Msbh)/Cys2–Cys9(Acm)/Cys4–Cys16(Trt). All peptides were assembled on a CEM Liberty Prime automatic synthesiser (CEM, Matthews, NC, USA) using rink amide-AM resin. Peptides were cleaved from solid support and side chains simultaneously deprotected in 92.5% trifluoroacetic acid (TFA)/2.5% triisopropylsilane/2.5% water/2.5% 2,2-(ethylenedioxy)diethanethiol for 2 h at room temperature. Excess TFA acid was evaporated by N2 flow, followed by peptide precipitation in ice-cold diethyl ether and centrifugation. Peptides were redissolved in solvent B (50% acetonitrile (ACN)/0.01% TFA) and lyophilised. TIIIA, GIIIA and GIIIB were synthesised as previously described [25,34]. A list of peptide sequences can be found in Supplementary Materials (Table S1).

4.2. Oxidation

Most SxIIIC analogues were oxidised thermodynamically to form disulfide bonds as described earlier [9]. However, due to difficulties obtaining the desired connectivity using thermodynamic oxidation, some peptides were oxidised using orthogonal protection of pairs of Cys residues. This regioselective oxidation was done as described previously [26,37], whereby linear peptides with Trt group removed were dissolved in acetic acid (2 mg/mL) and stirred at room temperature. To oxidise the first Cys pair, one equivalent of I2 (dissolved in MeOH) was added dropwise until the reaction turned pale-yellow and the colour persisted. Five equivalents of free Trp were added to minimise side-reactions with iodine. Following 15 min of stirring, the reaction was diluted with water and hydrochloric acid to a final volume of 50% acetic acid and 1% hydrochloric acid. To remove the Acm group and form the second disulfide bond, eight equivalents of I2 were added to the reaction and stirred for a further 30 min. The reaction was quenched with aqueous ascorbic acid until the solution became colourless. The intermediate product was isolated by RP-HPLC on a Shimadzu LC-20AT system equipped with an SPD-20A Prominence UV/VIS detector, and a FRC-10A fraction collector. Peptides were diluted to <2% acetic acid and loaded onto a Gemini, 5 µm C18 110 Å, 250 × 10 mm column (Phenomenex, Torrance, CA, USA). Peptides were purified using linear gradient between 5–25% solvent B over 40 min at 3 mL/min (solvent A: 0.05% TFA in H2O; solvent B: 90% ACN/0.05% TFA in H2O). Fractions containing the desired product were identified by electrospray ionisation-mass spectrometry (ESI-MS) and lyophilised. To remove the Msbh protecting groups and form the final disulfide bond, peptides were dissolved in TFA (1 mg/mL) and cooled on ice. Five equivalents of free Trp, dimethyl sulfate (1% v/v) and 20 equivalents of NaI were added and stirred for 15 min. The reaction was quenched by diluting the reaction with aqueous ascorbic acid (10 nM; 15 times the initial volume of TFA). The final product was isolated by RP-HPLC using a Gemini, 5 µm C18 110 Å, 250 × 5 mm column and a linear gradient 0–20% solvent B at 1 mL/min over 40 min. Fractions containing the desired product were identified by ESI-MS (Shimadzu LCMS-2020), lyophilised and stored at −20 °C. For analytical RP-HPLC and ESI-MS of the final products, see Figures S1–S4.

4.3. Cell Culture

HEK293 cells heterologously expressing hNaV1.1–1.7/β1 (SB Drug Discovery, Glasgow, UK) or CHO cells heterologously expressing hNaV1.8/β3 in a tetracycline-inducible system (ChanTest, Cleveland, OH, USA) were maintained in Minimum Essential Medium Eagle (M5650) supplemented with 10% v/v foetal bovine serum, 2 mM L-glutamine, and the selection antibiotics as recommended by the manufacturer. Cells were incubated at 37 °C with 5% CO2 and split every 3–4 days when reaching 70–80% confluency using the dissociation reagent TrypLE Express (Thermo Fisher Scientific, Scoresby, VIC, Australia). Expression of hNaV1.8 was induced by the addition of tetracycline (1 μg/mL) for 48 h prior to assays.

4.4. Whole-Cell Patch-Clamp Electrophysiology Assays

Automated whole-cell patch-clamp recordings were performed with a QPatch-16 automated electrophysiology platform (Sophion Bioscience, Ballerup, Denmark) as previously described [9].
The extracellular solution consisted of (in mM) 145 NaCl, 4 KCl, 2 CaCl2, 1 MgCl2, 10 HEPES, and 10 glucose, pH to 7.4 with NaOH (adjusted to 305 mOsm/L with sucrose). The intracellular solution consisted of (in mM) 140 CsF, 1 EGTA, 5 CsOH, 10 HEPES, and 10 NaCl, pH to 7.3 with CsOH (adjusted to 320 mOsm/L with sucrose). Peptides were diluted in extracellular solution with 0.1% bovine serum albumin and each concentration was incubated for 5 min. hNaV1.1–1.7 currents were elicited by a 50 ms test pulse to −20 mV from a holding potential of −90 mV (repetition interval 20 s). hNaV1.8 currents were elicited by a 50 ms test pulse to +10 mV from a holding potential of −90 mV (repetition interval 20 s) in the presence of TTX (1 µM) to inhibit endogenous TTX-sensitive current in CHO cells. Recordings were taken at ambient room temperature (22 °C). Peak current post-peptide addition (I) was normalised to peak current of buffer control (I0). IC50s were determined by plotting difference in peak current (I/I0) and log peptide concentration. Concentration-response curves were fitted using the log (inhibitor) vs. response-variable slope (four parameters) Equation (1) to obtain pIC50 values for each biological replicate.
y = 1/(1 + 10^((LogIC50 − x) × HillSlope))
The corresponding IC50s were used to compute mean ± standard error of the mean from at least n = 3 biological replicates.

5. Data Analysis

Data were analysed using GraphPad Prism 9.3.1. Statistical significance, defined as p < 0.05, was determined using One-way ANOVA with Tukey’s multiple comparison of IC50 or pIC50 values fitted for individual biological replicates as described above.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/toxins14090600/s1. Figure S1: Analytical RP-HPLC traces with corresponding ESI-MS spectra of native μ-conotoxins used in this study; Figure S2: Analytical RP-HPLC traces with corresponding ESI-MS spectra of single mutants and N-terminus truncated SxIIIC analogues used in this study; Figure S3: Crude and analytical RP-HPLC traces with corresponding ESI-MS spectra following regioselective oxidation of SxIIIC analogues used in this study; Figure S4: RP-HPLC traces repre-senting the crude oxidation of double and triple mutant SxIIIC analogues highlighting the two main folding isomers (peak 1—blue arrow, and peak 2—red arrow), and analytical RPHPLC traces of isolated fractions (peak 2—red arrow), with corresponding ESI-MS spectra of peak 2 (red arrow); Figure S5: 1D 1H NMR spectra of the amide region of the first and second peaks from the double and triple mutant SxIIIC analogues; Figure S6: Concentration response curves of A KIIIA, SmIIIA, SxIIIC at TTX-r hNaV1.5/b1 overexpressed in HEK293 cells, and B SmIIIA and SxIIIC at TTX-r hNaV1.8/b3 overexpressed in CHO cells using automated whole cell patch clamp electrophysiol-ogy; Figure S7: Concentration response curves for the two major peaks (First peak—A, blue arrow in Figure S4; Second peak—B, red arrow in Figure S4) of the double and triple mutants against hNaV1.4 as assessed by whole-cell patch-clamp electrophysiology; Table S1: Sequence table of µ-conotoxins and analogues used in this study.

Author Contributions

K.L.M., C.I.S. and I.V. conceived the study. K.L.M. carried out chemical synthesis, peptide folding, NMR structure calculations, cell culture, and electrophysiology studies. H.N.T.T. assisted with chemical synthesis and peptide folding. J.R.D. and I.V. assisted with electrophysiology analysis. D.J.C. assisted with supervision. K.L.M. wrote the manuscript, which was reviewed and edited by J.R.D., I.V. and C.I.S. All authors assisted with the final editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Australian National Health and Medical Research Council (NMHRC) through Project and Ideas Grants (APP1080405, 2002860), a Career Development Fellowship (APP1162503) awarded to IV, and an Early Career Fellowship (APP1139961) awarded to JRD. CIS was an Australian Research Council (ARC) Future Fellow (FT160100055). Work in DJC’s laboratory is supported by the ARC Centre of Excellence for Innovations in Peptide and Protein Science (CE200100012) and by an NHMRC Investigator Grant (GNT2009564). Australian Government Research Training Program Scholarships supported KLM and HNTT through the University of Queensland. This research was supported [in part] by the Intramural Research Program of the National Cancer Institute, Center for Cancer Research (Grant ZIA BC 012003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Acknowledgments

We thank Richard J. Lewis from the Institute for Molecular Biosciences at The University of Queensland for kindly gifting µ-conotoxins TIIIA, GIIIA and GIIIB.

Conflicts of Interest

K.L.M., H.N.T.T., J.R.D., D.J.C. and I.V. declare that they have no conflicts of interest with the contents of this article. C.I.S. is an employee of Genentech, Inc. and shareholder of Roche.

References

  1. Olivera, B.M.; Seger, J.; Horvath, M.P.; Fedosov, A.E. Prey-Capture Strategies of Fish-Hunting Cone Snails: Behavior, Neurobiology and Evolution. Brain Behav. Evol. 2015, 86, 58–74. [Google Scholar] [CrossRef]
  2. Hui, K.; Lipkind, G.; Fozzard, H.A.; French, R.J. Electrostatic and Steric Contributions to Block of the Skeletal Muscle Sodium Channel by μ-Conotoxin. J. Gen. Physiol. 2002, 119, 45–54. [Google Scholar] [PubMed]
  3. Leipold, E.; Markgraf, R.; Miloslavina, A.; Kijas, M.; Schirmeyer, J.; Imhof, D.; Heinermann, S.H. Molecular determinants for the subtype specificity of μ-conotoxin SIIIA targeting neuronal voltage-gated sodium channels. Neuropharmacology 2011, 61, 105–111. [Google Scholar] [CrossRef] [PubMed]
  4. Markgraf, R.; Leipold, E.; Schirmeyer, J.; Paolini-Bertrand, M.; Hartley, O.; Heinemann, S.H. Mechanism and molecular basis for the sodium channel subtype specificity of µ-conopeptide CnIIIC. Br. J. Pharmacol. 2012, 167, 576–586. [Google Scholar] [CrossRef]
  5. Li, R.A.; Ennis, I.L.; Xue, T.; Nguyen, H.M.; Tomaselli, G.F.; Godin, A.L.; Marban, E. Molecular Basis of Isoform-specific μ-Conotoxin Block of Cardiac, Skeletal Muscle, and Brain Na+ Channels. J. Biol. Chem. 2003, 278, 8717–8724. [Google Scholar] [PubMed]
  6. Vetter, I.; Deuis, J.R.; Mueller, A.; Israel, M.R.; Starobova, H.; Zhang, A.; Rash, L.D.; Mobli, M. NaV1.7 as a pain target—From gene to pharmacology. Pharmacol. Ther. 2017, 172, 73–100. [Google Scholar] [PubMed]
  7. Dib-Hajj, S.D.; Waxman, S.G. Sodium Channels in Human Pain Disorders: Genetics and Pharmacogenomics. Annu. Rev. Neurosci. 2019, 42, 87–106. [Google Scholar]
  8. McArthur, J.R.; Singh, G.; McMaster, D.; Winkfein, R.; Tieleman, D.P.; French, R.J. Interactions of Key Charged Residues Contributing to Selective Block of Neuronal Sodium Channels by μ-Conotoxin KIIIA. Mol. Pharmacol. 2011, 80, 573–584. [Google Scholar] [CrossRef]
  9. McMahon, K.L.; Tran, H.N.T.; Deuis, J.R.; Lewis, R.J.; Vetter, I.; Schroeder, C.I. Discovery, Pharmacological Characterisation and NMR Structure of the Novel µ-Conotoxin SxIIIC, a Potent and Irreversible NaV Channel Inhibitor. Biomedicines 2020, 8, 391. [Google Scholar] [CrossRef]
  10. West, P.J.; Bulaj, G.; Garrett, J.E.; Olivera, B.M.; Yoshikami, D. μ-Conotoxin SmIIIA, a potent inhibitor of tetrodotoxin-resistant sodium channels in amphibian sympathetic and sensory neurons. Biochemistry 2002, 41, 15388–15393. [Google Scholar]
  11. Favreau, P.; Benoit, E.; Hocking, H.G.; Carlier, L.; D’hoedt, D.; Leipold, E.; Markgraf, R.; Schlumberger, S.; Cordova, M.A.; Gaertner, H.; et al. A novel µ-conopeptide, CnIIIC, exerts potent and preferential inhibition of NaV1.2/1.4 channels and blocks neuronal nicotinic acetylcholine receptors. Br. J. Pharmacol. 2012, 166, 1654–1668. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, M.M.; Green, B.R.; Catlin, P.; Fiedler, B.; Azam, L.; Chadwick, A.; Terlau, H.; McArthur, J.R.; French, R.J.; Gulyas, J.; et al. Structure/function characterization of μ-conotoxin KIIIA, an analgesic, nearly irreversible blocker of mammalian neuronal sodium channels. J. Biol. Chem. 2007, 282, 30699–30706. [Google Scholar] [PubMed] [Green Version]
  13. Van der Haegen, A.; Peigneur, S.; Tytgat, J. Importance of position 8 in μ-conotoxin KIIIA for voltage-gated sodium channel selectivity. FEBS J. 2011, 278, 3408–3418. [Google Scholar] [CrossRef]
  14. Khoo, K.K.; Wilson, M.J.; Smith, B.J.; Zhang, M.M.; Gulyas, J.; Yoshikami, D.; River, J.E.; Bulaj, G.; Norton, R.S. Lactam-stabilized helical analogues of the analgesic μ-conotoxin KIIIA. J. Med. Chem. 2011, 54, 7558–7566. [Google Scholar] [CrossRef] [PubMed]
  15. Brady, R.M.; Zhang, M.; Gable, R.; Norton, R.S.; Baell, J.B. De novo design and synthesis of a μ-conotoxin KIIIA peptidomimetic. Bioorg. Med. Chem. Lett. 2013, 23, 4892–4895. [Google Scholar] [PubMed]
  16. Stevens, M.; Peigneur, S.; Dyubankova, N.; Lescrinier, E.; Herdewijn, P.; Tytgat, J.J. Design of bioactive peptides from naturally occurring μ-conotoxin structures. Biol. Chem. 2012, 287, 31382–31392. [Google Scholar]
  17. Lebbe, E.K.; Peigneur, S.; Brullot, W.; Verbiest, T.; Tytgat, J. Ala-7, His-10 and Arg-12 are crucial amino acids for activity of a synthetically engineered μ-conotoxin. Peptides 2014, 53, 300–306. [Google Scholar] [CrossRef] [PubMed]
  18. Peigneur, S.; Cheneval, O.; Maiti, M.; Leipold, E.; Heinemann, S.H.; Herdewijn, P.; de Lima, M.E.; Craik, D.J.; Schroeder, C.I. Where cone snails and spiders meet: Design of small cyclic sodium-channel inhibitors. FASEB J. 2019, 33, 3693–3703. [Google Scholar]
  19. Peigneur, S.; da Costa Oliveira, C.; de Sousa Fonseca, F.C.; McMahon, K.L.; Mueller, A.; Cheneval, O.; Nogueira Freitas, A.C.; Starobova, H.; Gama Duarte, I.D.; Craik, D.J.; et al. Small cyclic sodium channel inhibitors. Biochem. Pharmacol. 2020, 183, 114291. [Google Scholar] [CrossRef]
  20. Schroeder, C.I.; Ekberg, J.; Nielsen, K.J.; Adams, D.; Loughnan, M.L.; Thomas, L.; Adams, D.J.; Alewood, P.F.; Lewis, R.J. Neuronally selective μ-conotoxins from Conus striatus utilize an α-helical motif to target mammalian sodium channels. J. Biol. Chem. 2008, 283, 21621–21628. [Google Scholar] [PubMed]
  21. Lewis, R.J.; Schroeder, C.I.; Ekberg, J.; Nielsen, K.J.; Loughnan, M.; Thomas, L.; Adams, D.A.; Drinkwater, R.; Adams, D.J.; Alewood, P.F. Isolation and structure-activity of μ-conotoxin TIIIA, a potent inhibitor of tetrodotoxin-sensitive voltage-gated sodium channels. Mol. Pharmacol. 2007, 71, 676–685. [Google Scholar] [PubMed]
  22. Dekan, Z.; Mobli, M.; Pennington, M.W.; Fung, E.; Nemeth, E.; Alewood, P.F. Total Synthesis of Human Hepcidin through Regioselective Disulfide-Bond Formation by using the Safety-Catch Cysteine Protecting Group 4,4′-Dimethylsulfinylbenzhydryl. Angew. Chem. Int. Ed. Engl. 2014, 53, 2931–2944. [Google Scholar] [PubMed]
  23. Tran, H.N.T.; McMahon, K.L.; Deuis, J.R.; Vetter, I.; Schroeder, C.I. Structural and functional insights into the inhibition of human voltage-gated sodium channels by μ-conotoxin KIIIA disulfide isomers. J. Biol. Chem. 2022, 298, 101728. [Google Scholar] [PubMed]
  24. Wilson, M.J.; Yoshikami, D.; Azam, L.; Gajewiak, J.; Olivera, B.M.; Bulaj, G.; Zhang, M.-M. μ-Conotoxins that differentially block sodium channels NaV1.1 through 1.8 identify those responsible for action potentials in sciatic nerve. Proc. Natl. Acad. Sci. USA 2011, 108, 10302–10307. [Google Scholar]
  25. Zhang, M.M.; Wilson, M.J.; Azam, L.; Gajewiak, J.; Rivier, J.E.; Bulaj, G.; Olivera, B.M.; Yoshikami, D. Co-expression of NaVβ subunits alters the kinetics of inhibition of voltage-gated sodium channels by pore-blocking μ-conotoxins. Br. J. Pharmacol. 2013, 168, 1597–1610. [Google Scholar]
  26. Tran, H.N.T.; Tran, P.; Deuis, J.R.; Agwa, A.J.; Zhang, A.H.; Vetter, I.; Schroeder, C.I. Enzymatic Ligation of a Pore Blocker Toxin and a Gating Modifier Toxin: Creating Double-Knotted Peptides with Improved Sodium Channel NaV1.7 Inhibition. Bioconjug. Chem. 2020, 31, 64–73. [Google Scholar]
  27. Harvey, P.J.; Kurniawan, N.D.; Finol-Urdaneta, R.K.; McArthur, J.R.; van Lysebetten, D.; Dash, T.S.; Hill, J.M.; Adams, D.J.; Durek, T.; Craik, D.J. NMR Structure of μ-Conotoxin GIIIC: Leucine 18 Induces Local Repacking of the N-Terminus Resulting in Reduced NaV Channel Potency. Molecules 2018, 23, 2715. [Google Scholar] [CrossRef] [Green Version]
  28. Leipold, E.; Ullrich, F.; Thiele, M.; Tietze, A.A.; Terlau, H.; Imhof, D.; Heinermann, S.H. Subtype-specific block of voltage-gated K+ channels by μ-conopeptides. Biochem. Biophys. Res. Commun. 2017, 482, 1135–1140. [Google Scholar]
  29. Khoo, K.K.; Gupta, K.; Green, B.R.; Zhang, M.M.; Watkins, M.; Olivera, B.M.; Balaram, P.; Yoshikami, D.; Bulaj, G.; Norton, R.S. Distinct disulfide isomers of μ-conotoxins KIIIA and KIIIB block voltage-gated sodium channels. Biochemistry 2012, 51, 9826–9835. [Google Scholar]
  30. Wishart, D.S.; Bigam, C.G.; Holm, A.; Hodges, R.S.; Sykes, B.D. 1H, 13C and 15N random coil NMR chemical shifts of the common amino acids. I. Investigations of nearest-neighbor effects. J. Biomol. NMR 1995, 5, 67–81. [Google Scholar]
  31. Keizer, D.W.; West, P.J.; Lee, E.F.; Yoshikami, D.; Olivera, B.M.; Bulaj, G.; Norton, R.S. Structural basis for tetrodotoxin-resistant sodium channel binding by μ-conotoxin SmIIIA. J. Biol. Chem. 2003, 278, 46805–46813. [Google Scholar] [PubMed]
  32. Khoo, K.K.; Feng, Z.P.; Smith, B.J.; Zhang, M.M.; Yoshikami, D.; Olivera, B.M.; Bulaj, G.; Norton, R.S. Structure of the analgesic μ-conotoxin KIIIA and effects on the structure and function of disulfide deletion. Biochemistry 2009, 48, 1210–1219. [Google Scholar]
  33. Green, B.R.; Catlin, P.; Zhang, M.M.; Fiedler, B.; Bayudan, W.; Morrison, A.; Norton, R.S.; Smith, B.J.; Yoshikami, D.; Olivera, B.M.; et al. Conotoxins Containing Nonnatural Backbone Spacers: Cladistic-Based Design, Chemical Synthesis, and Improved Analgesic Activity. Chem. Biol. 2007, 14, 399–407. [Google Scholar] [PubMed]
  34. Jacob, R.B.; McDougal, O.M. The M-superfamily of conotoxins: A review. Cell. Mol. Life Sci. 2010, 67, 17–27. [Google Scholar]
  35. Paul George, A.A.; Heimer, P.; Maass, A.; Hamaekers, J.; Hofmann-Apitius, M.; Biswas, A.; Imhof, D. Insights into the folding of disulfide-rich μ-conotoxins. ACS Omega 2018, 3, 12330–12340. [Google Scholar] [PubMed]
  36. Sato, K.; Ishida, Y.; Wakamatsu, K.; Kato, R.; Honda, H.; Ohizumi, Y.; Nakamura, H.; Ohya, M.; Lancelin, J.M.; Khoda, D. Active site of mu-conotoxin GIIIA, a peptide blocker of muscle sodium channels. J. Biol. Chem. 1991, 266, 16989–16991. [Google Scholar] [PubMed]
  37. Pan, X.; Li, Z.; Huang, X.; Huang, G.; Gao, S.; Shen, H.; Liu, L.; Lei, J.; Yan, N. Molecular basis for pore blockade of human Na+ channel Nav1.2 by the μ-conotoxin KIIIA. Science 2019, 363, 1309–1313. [Google Scholar] [PubMed]
Figure 1. Comparison of pharmacological inhibitory potency of µ-conotoxins SxIIIC, SmIIIA and KIIIA as assessed by whole-cell patch-clamp electrophysiology across hNaV1.1–1.7/β1 overexpressed in HEK cells. (A) Concentration-response curves and (B) comparison of µ-conotoxin potency across NaV channel subtypes. SxIIIC data from [9]. Data are presented as mean ± SEM with individual data points (n = 3–10). Statistical significance was calculated by ordinary one-way ANOVA analysis and Tukey’s multiple comparisons test (* p-value < 0.05).
Figure 1. Comparison of pharmacological inhibitory potency of µ-conotoxins SxIIIC, SmIIIA and KIIIA as assessed by whole-cell patch-clamp electrophysiology across hNaV1.1–1.7/β1 overexpressed in HEK cells. (A) Concentration-response curves and (B) comparison of µ-conotoxin potency across NaV channel subtypes. SxIIIC data from [9]. Data are presented as mean ± SEM with individual data points (n = 3–10). Statistical significance was calculated by ordinary one-way ANOVA analysis and Tukey’s multiple comparisons test (* p-value < 0.05).
Toxins 14 00600 g001
Figure 2. Extended comparison of µ-conotoxins inhibiting hNaV1.7 assessed by whole-cell patch-clamp electrophysiology across hNaV1.7/β1 overexpressed in HEK293 cells. Inhibition of hNaV1.7 by five additional µ-conotoxins, CnIIIC, SIIIA, TIIIA, GIIIA and GIIIB, was evaluated by automated whole-cell patch-clamp methods as described above. CnIIIC was the only additional µ-conotoxin to inhibit hNaV1.7 current. SIIIA, TIIIA, GIIIA and GIIIB displayed no activity up to the highest concentration tested (1 µM). GIIIC was previously found to be inactive at hNaV1.7 at 1 µM [26] * denotes data from references). Data are presented as mean ± SEM with individual data points (n = 3–4).
Figure 2. Extended comparison of µ-conotoxins inhibiting hNaV1.7 assessed by whole-cell patch-clamp electrophysiology across hNaV1.7/β1 overexpressed in HEK293 cells. Inhibition of hNaV1.7 by five additional µ-conotoxins, CnIIIC, SIIIA, TIIIA, GIIIA and GIIIB, was evaluated by automated whole-cell patch-clamp methods as described above. CnIIIC was the only additional µ-conotoxin to inhibit hNaV1.7 current. SIIIA, TIIIA, GIIIA and GIIIB displayed no activity up to the highest concentration tested (1 µM). GIIIC was previously found to be inactive at hNaV1.7 at 1 µM [26] * denotes data from references). Data are presented as mean ± SEM with individual data points (n = 3–4).
Toxins 14 00600 g002
Figure 3. Effect of single SxIIIC mutations on potency and selectivity at hNaV1.7. pIC50 values ± SEM are displayed for hNaV1.4 (maroon bars) and hNaV1.7 (blue bars). [R16A]SxIIIC, [D17A]SxIIIC and [R20A]SxIIIC maintained selectivity for hNaV1.4 over hNaV1.7. In contrast, the mutations ([Δ1,2], [R16H], [R16Q]) did not affect inhibitory potency at hNaV1.7 but reduced potency at hNaV1.4, resulting in no significant selectivity between the subtypes. Data are presented as mean ± SEM with individual data points (n = 3–5). Statistical significance was determined by unpaired t-test (* p-value < 0.05).
Figure 3. Effect of single SxIIIC mutations on potency and selectivity at hNaV1.7. pIC50 values ± SEM are displayed for hNaV1.4 (maroon bars) and hNaV1.7 (blue bars). [R16A]SxIIIC, [D17A]SxIIIC and [R20A]SxIIIC maintained selectivity for hNaV1.4 over hNaV1.7. In contrast, the mutations ([Δ1,2], [R16H], [R16Q]) did not affect inhibitory potency at hNaV1.7 but reduced potency at hNaV1.4, resulting in no significant selectivity between the subtypes. Data are presented as mean ± SEM with individual data points (n = 3–5). Statistical significance was determined by unpaired t-test (* p-value < 0.05).
Toxins 14 00600 g003
Figure 4. Effect of double and triple SxIIIC mutations on potency and subtype-selectivity at hNaV1.4 and hNaV1.7. Comparative potency of analogues displayed for hNaV1.4 (maroon bars) and hNaV1.7 (blue bars). [Δ1,2; R16H] and [R16H:R20A] maintained selectivity for hNaV1.4 over hNaV1.7. The triple mutant [Δ1,2; R16H; R20A] lost inhibitory potency. Data are presented as mean ± SEM with individual data points (n = 3–4). Statistical significance was determined by unpaired t-test (* p-value < 0.05).
Figure 4. Effect of double and triple SxIIIC mutations on potency and subtype-selectivity at hNaV1.4 and hNaV1.7. Comparative potency of analogues displayed for hNaV1.4 (maroon bars) and hNaV1.7 (blue bars). [Δ1,2; R16H] and [R16H:R20A] maintained selectivity for hNaV1.4 over hNaV1.7. The triple mutant [Δ1,2; R16H; R20A] lost inhibitory potency. Data are presented as mean ± SEM with individual data points (n = 3–4). Statistical significance was determined by unpaired t-test (* p-value < 0.05).
Toxins 14 00600 g004
Figure 5. Structural comparison of µ-conotoxins SxIIIC, SmIIIA and KIIIA. (A) Comparison of secondary Hα-chemical shifts compared to random coil values [28] for three µ-conotoxins SxIIIC, SmIIIA and KIIIA [29]. Sequence alignment shows high sequence homology between µ-conotoxins with disulfide connectivity (yellow lines). (B) Representative 3D NMR structures (PDB: SxIIIC 6X8R, SmIIIA 1Q2J and KIIIA 2LXG) of µ-conotoxins superimposed across residues Cys10–Cys22 (SxIIIC numbering; RMSD = 1.146 Å) shows how the conserved C-terminal residues contribute to similar backbone fold, whereas the N-terminal and loop 1 extensions (inserts) extend the structures of SxIIIC and SmIIIA.
Figure 5. Structural comparison of µ-conotoxins SxIIIC, SmIIIA and KIIIA. (A) Comparison of secondary Hα-chemical shifts compared to random coil values [28] for three µ-conotoxins SxIIIC, SmIIIA and KIIIA [29]. Sequence alignment shows high sequence homology between µ-conotoxins with disulfide connectivity (yellow lines). (B) Representative 3D NMR structures (PDB: SxIIIC 6X8R, SmIIIA 1Q2J and KIIIA 2LXG) of µ-conotoxins superimposed across residues Cys10–Cys22 (SxIIIC numbering; RMSD = 1.146 Å) shows how the conserved C-terminal residues contribute to similar backbone fold, whereas the N-terminal and loop 1 extensions (inserts) extend the structures of SxIIIC and SmIIIA.
Toxins 14 00600 g005
Figure 6. Effects of N-terminal and loop 1 mutations on µ-conotoxin potency at hNaV1.7. Comparative potency of analogues at hNaV1.7. N-terminal deletions of SxIIIC ([Δ1,2]SxIIIC) and the addition of loop 1 residues ([loop1R]KIIIA) into KIIIA did not significantly affect potency compared to native µ-conotoxins. Loop 1 deletions of SxIIIC ([Δ7,8] significantly reduced potency of SxIIIC, similar to that of KIIIA. Introduction of an additional charged residue into SxIIIC loop 1 ([G8R]SxIIIC) did not alter potency compared to native SxIIIC. Data are presented as mean ± SEM with individual data points (n = 3–4). Statistical significance for SxIIIC analogues was determined by ordinary one-way ANOVA with Dunnett’s multiple comparisons test compared to native SxIIIC (* p-value < 0.05). Statistical significance for KIIIA and loop1 RKIIIA was determined by unpaired t-test (* p-value < 0.05).
Figure 6. Effects of N-terminal and loop 1 mutations on µ-conotoxin potency at hNaV1.7. Comparative potency of analogues at hNaV1.7. N-terminal deletions of SxIIIC ([Δ1,2]SxIIIC) and the addition of loop 1 residues ([loop1R]KIIIA) into KIIIA did not significantly affect potency compared to native µ-conotoxins. Loop 1 deletions of SxIIIC ([Δ7,8] significantly reduced potency of SxIIIC, similar to that of KIIIA. Introduction of an additional charged residue into SxIIIC loop 1 ([G8R]SxIIIC) did not alter potency compared to native SxIIIC. Data are presented as mean ± SEM with individual data points (n = 3–4). Statistical significance for SxIIIC analogues was determined by ordinary one-way ANOVA with Dunnett’s multiple comparisons test compared to native SxIIIC (* p-value < 0.05). Statistical significance for KIIIA and loop1 RKIIIA was determined by unpaired t-test (* p-value < 0.05).
Toxins 14 00600 g006
Figure 7. Representation of the surface profile of µ-conotoxins and corresponding net charge. Colours represent charged residues (basic–blue; acidic–red; neutral–grey). µ-Conotoxins active against hNaV1.7 (yellow shaded box) appear on the M5-branch and have an overall net charge of between +2.8–4.7.
Figure 7. Representation of the surface profile of µ-conotoxins and corresponding net charge. Colours represent charged residues (basic–blue; acidic–red; neutral–grey). µ-Conotoxins active against hNaV1.7 (yellow shaded box) appear on the M5-branch and have an overall net charge of between +2.8–4.7.
Toxins 14 00600 g007
Table 1. Summary of µ-conotoxin IC50 values. IC50 values displayed in nM ± SEM.
Table 1. Summary of µ-conotoxin IC50 values. IC50 values displayed in nM ± SEM.
SubtypeSxIIIC aSmIIIAKIIIA
NaV1.1132 ± 12235 ± 15.0136 ± 18
NaV1.2364 ± 54172 ± 49186 ± 32
NaV1.389 ± 1195 ± 9>1000
NaV1.415 ± 1114 ± 267 ± 15
NaV1.5>1000>1000>1000
NaV1.6125 ± 11106 ± 37762 ± 77
NaV1.7152 ± 2241 ± 4379 ± 43
NaV1.8>1000>1000>1000
n = 3–10 cells per data point. a Results from [9].
Table 2. Comparison of µ-conotoxin sequences.
Table 2. Comparison of µ-conotoxin sequences.
µ-ConotoxinSequenceInhibitory Potency at hNaV1.7
GIIIARDCCTOOKKCKDRQCKOQ-RCCA *No
GIIIBRDCCTOORKCKDRRCKOM-KCCA *No
GIIICRDCCTOOKKCKDRRCKOL-KCCA *No
TIIIARHGCCKGOKGCSSRECROQ-HCC *No
CnIIICGCCNGPKGCSSKWCRDHARCC *Yes
KIIIACCN—CSSKWCRDHSRCC *Yes
SIIIAZNCCNG—GCSSKWCRDHARCC *No
SmIIIAZRCCNGRRGCSSRWCRDHSRCC *Yes
SxIIICRGCCNGRGGCSSRWCRDHARCC *Yes
* C-terminal amidation, Z-pyroglutamic acid, O-hydroxyproline.
Table 3. Summary of IC50 values for SxIIIC analogues at hNaV1.7 and hNaV1.4. IC50 displayed in nM ± SEM.
Table 3. Summary of IC50 values for SxIIIC analogues at hNaV1.7 and hNaV1.4. IC50 displayed in nM ± SEM.
AnaloguehNaV1.7Fold Difference to SxIIIChNav1.4Fold Difference to SxIIIC
SxIIIC152.2 ± 26.812.8 ± 13.0
[Δ1,2]SxIIIC106.2 ± 19.4ns96.3 ± 13.0↓7-fold
[Δ7,8]SxIIIC313.6 ± 36.6↓2-fold
[Δ6–9]SxIIIC317.9 ± 100.0↓2-fold
[G8R]SxIIIC155.7 ± 18.9ns
[Loop1R]KIIIA305.1 ± 23.9↓2-fold *
[R16A]SxIIIC334.1 ± 25.5↓2-fold38.0 ± 13.0↓3-fold
[R16H]SxIIIC66.5 ± 2.5↑2-fold80.3 ± 24.7↓6-fold
[R16Q]SxIIIC151.3 ± 24.0ns99.5 ± 12.9↓7-fold
[D17A]SxIIIC123.5 ± 18.1ns14.6 ± 7.7ns
[R20A]SxIIIC142.3 ± 12.9ns64.8 ± 15.4↓5-fold
[Δ1,2; R16H]666.6 ± 211.6↓4-fold221.4 ± 26.9↓17-fold
[Δ1,2; R20A]481.8 ± 646.9↓3-fold222.3 ± 70.4↓17-fold
[R16H; R20A]888.5 ± 916.4↓6-fold247.8 ± 57.7↓19-fold
[Δ1,2; R16H; R20A]>1000>1000
n = 3–4 cells per data point. ns—no significant difference. * compared to KIIIA. ↓ fold-decrease, and ↑ fold-increase in potency compared to SxIIIC.
Table 4. Comparison of µ-conotoxin selectivity profiles of hNaV channels.
Table 4. Comparison of µ-conotoxin selectivity profiles of hNaV channels.
µ-ConotoxinSelectivity Profile
SxIIIC1.4 > 1.3 > 1.6 ≈ 1.1 ≈ 1.7 > 1.2 >> 1.5 ≈ 1.8
SmIIIA1.4 > 1.7 > 1.3 ≈ 1.6 > 1.2 > 1.1 >> 1.5 ≈ 1.8
KIIIA1.4 > 1.1 > 1.2 > 1.7 > 1.6 > 1.3 >> 1.5 ≈ 1.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

McMahon, K.L.; Tran, H.N.T.; Deuis, J.R.; Craik, D.J.; Vetter, I.; Schroeder, C.I. µ-Conotoxins Targeting the Human Voltage-Gated Sodium Channel Subtype NaV1.7. Toxins 2022, 14, 600. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins14090600

AMA Style

McMahon KL, Tran HNT, Deuis JR, Craik DJ, Vetter I, Schroeder CI. µ-Conotoxins Targeting the Human Voltage-Gated Sodium Channel Subtype NaV1.7. Toxins. 2022; 14(9):600. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins14090600

Chicago/Turabian Style

McMahon, Kirsten L., Hue N. T. Tran, Jennifer R. Deuis, David J. Craik, Irina Vetter, and Christina I. Schroeder. 2022. "µ-Conotoxins Targeting the Human Voltage-Gated Sodium Channel Subtype NaV1.7" Toxins 14, no. 9: 600. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins14090600

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