Next Article in Journal
Preparation and Neutralization Efficacy of Novel Jellyfish Antivenoms against Cyanea nozakii Toxins
Next Article in Special Issue
Animal Venoms and Their Components: Molecular Mechanisms of Action
Previous Article in Journal
Combination of Extrusion and Fermentation with Lactobacillus plantarum and L. uvarum Strains for Improving the Safety Characteristics of Wheat Bran
Previous Article in Special Issue
ERK and mTORC1 Inhibitors Enhance the Anti-Cancer Capacity of the Octpep-1 Venom-Derived Peptide in Melanoma BRAF(V600E) Mutations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Three-Finger Neurotoxins from Naja melanoleuca Cobra Venom Interact with GABAA and Nicotinic Acetylcholine Receptors

1
Shemyakin-Ovchinnikov Institute of Bioorganic Chemistry, ul. Miklukho-Maklaya 16/10, 117997 Moscow, Russia
2
Moscow Institute of Physics and Technology, 141700 Dolgoprudny, Russia
3
Institute of Molecular Medicine, Sechenov First Moscow State Medical University, ul. Trubetskaya 8, bld. 2, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Submission received: 9 December 2020 / Revised: 16 February 2021 / Accepted: 17 February 2021 / Published: 20 February 2021
(This article belongs to the Special Issue Animal Venoms and Their Components: Molecular Mechanisms of Action)

Abstract

:
Cobra venoms contain three-finger toxins (TFT) including α-neurotoxins efficiently binding nicotinic acetylcholine receptors (nAChRs). As shown recently, several TFTs block GABAA receptors (GABAARs) with different efficacy, an important role of the TFTs central loop in binding to these receptors being demonstrated. We supposed that the positive charge (Arg36) in this loop of α-cobratoxin may explain its high affinity to GABAAR and here studied α-neurotoxins from African cobra N. melanoleuca venom for their ability to interact with GABAARs and nAChRs. Three α-neurotoxins, close homologues of the known N. melanoleuca long neurotoxins 1 and 2, were isolated and sequenced. Their analysis on Torpedo californica and α7 nAChRs, as well as on acetylcholine binding proteins and on several subtypes of GABAARs, showed that all toxins interacted with the GABAAR much weaker than with the nAChR: one neurotoxin was almost as active as α-cobratoxin, while others manifested lower activity. The earlier hypothesis about the essential role of Arg36 as the determinant of high affinity to GABAAR was not confirmed, but the results obtained suggest that the toxin loop III may contribute to the efficient interaction of some long-chain neurotoxins with GABAAR. One of isolated toxins manifested different affinity to two binding sites on Torpedo nAChR.
Key Contribution: The long chain α-neurotoxins isolated from Naja melanoleuca cobra venom inhibited GABAA receptor; one of them being almost as efficient as α-cobratoxin. Another toxin distinguished two binding sites in Torpedo nicotinic acetylcholine receptor and showed the highest affinity to acetylcholine binding proteins.

Graphical Abstract

1. Introduction

Ligand-gated ion channels are classified into several families of membrane-bound receptors. The Cys-loop family of pentameric receptors is represented widely in the muscle and nervous systems as well as in immune and other cells and plays prominent roles. The Cys-loop family of vertebrates includes nicotinic acetylcholine receptor (nAChR), as well as serotonin type 3, γ-aminobutyric acid (GABAAR), and glycine receptors. The receptors of this family are characterized by a conserved sequence of 13 amino acid residues confined by two cysteines forming a disulfide bond (Cys loop) in the N-terminal extracellular domain of each subunit [1]. Among these receptors, excitatory nAChRs are the targets of numerous natural compounds including toxins from snake venom [2,3]. Snake venom neurotoxins of the three-finger toxin (TFT) family efficiently inhibit some nAChR subtypes. For example, α-bungarotoxin (α-Bgt) from krait Bungarus multicinctus venom binds muscle-type as well as neuronal α7 and α9 nAChRs with nanomolar affinities and blocks ion current through the ion channel of these receptors. α-Cobratoxin (α-Ctx) from cobra Naja kaouthia venom exhibits similar effects. Based on location and function, nAChRs can be divided into muscle and neuronal ones [4,5,6]. Neuronal nAChRs are localized in the central and peripheral nervous system and are involved in the transmission of fast nerve impulses. Muscle-type nAChRs are located on the postsynaptic membranes of the neuromuscular junction and transmit signals for muscle contraction. Each muscle-type nAChR has two copies of α1 and one each of the β1, δ, and γ subunits (in the embryonic receptor) or ε subunit (in the adult form). In this receptor type, two binding sites for agonists and competitive antagonists are located in the extracellular domain at the interfaces of the α1–δ and α1–γ (or α1–ε) subunits [5]. Two binding sites have different affinity to acetylcholine, the α–γ site possessing a ∼35–40-fold higher affinity for acetylcholine than the α–δ and α–ε sites [7]. Some natural toxins, including several α-conotoxins, waglerins, and few TFTs, bind with different affinities to these two binding sites as well. Thus, a short α-neurotoxin from cobra N. mossambica mossambica (NmmI) was shown to bind with high affinity to α–γ and α–δ subunit interfaces but had a markedly reduced affinity to the α–ε interface [8]. Furthermore, the mutation of Lys27 to Glu27 in NmmI affected binding at the α–γ site more than the α–δ site [9] making this mutant selective for α–δ site. The similar results were obtained for two mutants of α-Ctx K23E and K49E, which demonstrated the different affinities to two toxin-binding sites on Torpedo nAChR with higher and lower affinities at the α–δ and α–γ sites, respectively [10]. The data for nonconventional TFT, candoxin from B. candidus also suggest its differential affinity for the α–γ or α–δ sites at the muscle nAChR [11]. We have found recently that so called αδ-bungarotoxins from B. candidus venom show different affinity to two binding sites in muscle-type nAChRs, manifesting higher activity at the interface of α–δ subunits [12].
Until recently α-Bgt and α-Ctx were considered as very specific markers of nAChRs. However, the works of our and other groups [13,14,15] have shown that these toxins inhibit also GABAAR. Similarly to nAChRs, each GABAAR is composed of five subunits, including most frequently two α subunits, two β subunits, and one from γ, δ, ε, θ, or π subunit. However, contrary to nAChRs which are cation channels, GABAAR are chloride channels. In mammals, 19 different subunits (six α, three β, three γ, δ, ε, θ, π, and three ρ) are known; they form a wide variety of GABAA receptor subtypes with distinct subunit composition and unique pharmacological properties [16]. In the brain, the most abundant GABAAR isoforms are αβγ and αβδ [17]. GABAA αβγ receptors are widely distributed in the brain, while αβδ receptors constitute only a small proportion [18]. Among α subunits, the most abundant is α1 which is often colocalized with highly expressed β2 and γ2 subunits [18]. The α2 and α3 subunits are less abundant. Among the β subunits, β2 is most abundant, β3 is reasonably highly expressed, and β1 is least common. Up to 80% of GABAARs contain the γ2 subunit [16]. In the brain, α1β2γ2 is the most common isoform. At present, several clinically used compounds target GABAAR and still there is a need in finding new molecules for potential design of novel drugs.
Acetylcholine binding proteins (AChBPs), soluble proteins mostly from mollusks, are remarkable structural homologues of the ligand-binding domains of all Cys-loop receptors [19]. While transmembrane and intracellular motifs are absent in these proteins, the important elements necessary for ligand binding including the C- and F-loops are structurally conserved. This fact made AChBPs and their mutants perfect tools for structural studies on pentameric ligand-gated ion channels. The fact that many studies have revealed some inconsistency in the activity profiles on nAChR and AChBP for a number of ligands was compensated by high-resolution structures of complexes of these ligands with AChBPs. Based on these data, more reliable models of the complexes of the corresponding ligands with the full-length nAChRs were built in order to identify the key amino acid residues and molecular mechanisms that determine the receptor-ligand interactions. In particular, the molecular basis for the high selectivity of α-conotoxin LvIA for α3β2 nAChR [20] or the difference in affinity of two other α-conotoxins’ analogues to human and rat α7 nAChR [21] were explained. At present, AChBPs from the mollusks Aplysia californica and Lymnaea stagnalis [22] are widely used, the former is closer pharmacologically to homooligimeric nAChRs while the latter is closer to heterooligomeric receptors.
α-Bgt, being a very efficient blocker of nAChRs, showed fairly weak activity on GABAAR inhibiting α1β3γ2 receptors by only 19% at 10 µM [15]. In contrast, α-Ctx inhibited GABAAR quite effectively manifesting half-maximal inhibitory concentration (IC50) of 236 nM at α1β3γ receptor [15]. Some other snake neurotoxins inhibited GABAAR as well, although not so effectively as α-Ctx. It was suggested that Arg36, present in α-Ctx and being valine in this position of α-Bgt, might be responsible for efficient interaction of α-Ctx with GABAAR [6]. In order to test this hypothesis, α-Ctx analogues with or without Arg36 need to be used. Although the most straightforward way would be testing the α-Ctx analogue(s) obtained through site-directed mutagenesis at position occupied by Arg36, there are some problems that should be solved on this way. First one is that α-Ctx contains five disulfide bridges the correct formation of which in mutant should be proved. This may require either determination of spatial structure (e.g., by X-ray or NMR) or combined used of selective disulfide modification and peptide mass fingerprinting. Moreover, the cleavage of some chimeric proteins often used to obtain the protein of interest or direct expression of the protein may add extra amino acid residues at N-terminus, which may influence the biological activity. We believed that use of natural α-Ctx analogues containing either Arg36 or other residue at this position may be the easiest way to test our hypothesis. If natural α-Ctx analogue with Arg36 is not as active as α-Ctx itself, then this arginine residue is not the only determinant of high α-Ctx affinity to GABAAR. In this respect, our attention was attracted to the cobra N. melanoleuca venom in which the presence of two highly homologous long chain neurotoxins 1 and 2 was shown [23,24]: neurotoxin 1 contained valine and neurotoxin 2 contained arginine residue at the position corresponding to Arg36 in α-Ctx. They seemed a good pair to clarify the role the respective residue.
In the present work from N. melanoleuca venom we isolated three α-neurotoxins (two with arginine and one with valine at the position corresponding to Arg36 in α-Ctx) and one muscarinic toxin-like protein. Only one neurotoxin (designated here as TX-NM4) was identical to the earlier described N. melanoleuca neurotoxin 2, another (TX-NM3-1) was its analogue with 5 substitutions, while instead of neurotoxin 1 we isolated its analogue with 3 substitutions. Analyzing binding of the isolated toxins to GABAARs and inhibition of the ion currents elicited by GABA, we did not find any evidence for the role of Arg36. However, molecular modeling showed that α-Ctx loop III had contact with α1 subunit of GABAAR and this contact might contribute to the efficient binding of the toxin to the receptor. Thus, our work expands the range of TFTs capable of inhibiting the GABAAR subtypes. In addition, we found that one N. melanoleuca toxin (Tx-NM2) showed unequal affinities for the two ligand-binding sites in the Torpedo californica nAChR and thus widened the range of naturally occurring TFTs that differently interact with the two binding sites in muscle-type nAChRs. Keeping in mind the future modeling and structural studies of the interaction of N. melanoleuca toxins with nAChRs and GABAARs, we investigated the interaction of these toxins with AChBPs from A. californica and L. stagnalis as plausible models. Tx-NM2 was found to be the most potent in binding with both AChPBs.

2. Results

2.1. Isolation and Characterization of Naja melanoleuca Toxins

To isolate toxins, a three-step chromatographic procedure was used. Gel filtration on Sephadex G50 was the first step (Figure 1a). The analysis of obtained fractions by mass spectrometry showed that fraction 6 contained proteins with molecular masses in the range of 6–8 kDa. In cobra venom, these masses are characteristic of three-finger toxins. As mentioned in introduction, we were interested in long neurotoxins 1 and 2 possessing the molecular masses of 8040.76 and 7756.59 Da, respectively. Therefore, the fraction 6 containing three-finger toxins was further separated by ion exchange chromatography on HEMA BIO 1000CM column (Figure 1b) and fractions obtained were analyzed by mass spectrometry. Fractions 2, 3, and 4 containing toxins with molecular masses of about 8 kDa were further purified by reversed phase HPLC (Figure S1). As a result, one toxin was obtained from fraction 2 (Tx-NM2), two toxins from fraction 3 (Tx-NM3-1, Tx-NM3-2) and one toxin from fraction 4 (Tx-NM4). Finally, four toxins with molecular masses of 8030.68 (Tx-NM2), 7787.577 (Tx-NM3-1), 7441.497 (Tx-NM3-2), and 7756.581 (Tx-NM4) Da were purified (Table 1, Figure S2).
The amino acid sequences of isolated toxins were determined by tandem mass spectrometry. For this purpose, the toxins were reduced, carbamidomethylated and digested with trypsin or chymotrypsin. The peptides obtained were analyzed by LC-MS with simultaneous de novo mass spectrometry sequencing. As a result, the complete amino acid sequences of Tx-NM2, Tx-NM3-1, and Tx-NM4 as well as partial amino acid sequence of Tx-NM3-2 were determined (Figure 2).
It was found that the amino acid sequence of Tx-NM4 exactly corresponds to that of long type α-neurotoxin 2 from N. melanoleuca (Uniprot accession number P01388). The sequence of Tx-NM2 differs from that of N. melanoleuca long type α-neurotoxin 1 (P01383) in three positions: at position 50, Lys residue is changed to Thr, Gln56 to Glu, and Met72 is oxidized. Tx-NM3-1 is very similar to long type α-neurotoxin 2 (P01388), but its sequences differ in 5 positions (Figure 2). We were not able to determine the amino acid sequence of Tx-NM3-2 at the short C-terminal fragment, however this toxin is very similar to muscarinic toxin-like protein 1 (P82462) from N. kaouthia cobra venom (Figure 2).

2.2. Biological Activity of N. melanoleuca Toxins

2.2.1. Interaction with GABAAR

The biological activity of isolated toxins was studied on rat GABAARs obtained by heterologous expression of different combination of α, β and γ subunits. The most common combinations of GABAAR subunits were used. They included α1β2γ2, α1β3γ2, and α3β2γ2 receptors. As concerns the inclusion of γ2 subunit, it has been shown earlier that increase in the ratio of mRNA/cDNA relative to α1 and β2 subunits (up to 10:1:1) resulted in more homogenous α1β2γ2 receptors in oocytes [25]. In our work, we used exactly this high γ2 subunit ratio relative α1 and β2 (10:1:1). Moreover, the presence of γ subunit in the receptors was proved in electrophysiology experiments as shown below.
  • Competition of N. melanoleuca toxins with fluorescently labeled α-Ctx
In this work, we used the most prevalent types of GABAARs. We have shown previously that α-Ctx labelled with Alexa Fluor 546 fluorophore efficiently stained Neuro 2a cells heterologously expressing α1β3γ2 GABAAR and this staining was inhibited by native α-Ctx and some other toxins [15]. α-Ctx was shown to be the most active on this receptor type [15], and we used it first to test the activity of N. melanoleuca toxins (Figure 3). The activity of N. melanoleuca toxins on the α1β3γ2 receptor was tested with the mentioned fluorescence assay. It was found that at 10 μM Tx-NM2 and Tx-NM3-1 inhibited Alexa Fluor 546 α-Ctx binding to GABAAR as efficiently as native α-Ctx, while Tx-NM4 was less active (Figure 3).
  • Electrophysiology measurements
Functional activities of the isolated toxins were studied by two-electrode voltage clamp on mouse GABAARs heterlogously expressed in Xenopus laevis oocytes (Figure 4 and Figure 5). Three combinations of receptor subunits were used in this study: α1β2γ2, α3β2γ2, and α1β3γ2. The incorporation of γ subunit was confirmed in separate experiments which are shown for the α1β3γ2 receptor (Figure 4c). GABA-induced currents at the α1β3γ2 receptor were not inhibited by Zn2+ ions, while inhibition was observed at α1β3 receptors (not shown). The incorporation of γ subunit in α1β3γ2 receptors was further confirmed by their sensitivity to diazepam (Figure 4c).
First, we checked the activity of toxins on the common receptor types α1β2γ2 and α3β2γ2, α1β2γ2 being the most common isoform (Figure 4). Finally, the quantitative characteristics for binding (IC50) were determined at α1β3γ2 type (Figure 5). On α1β2γ2 and α3β2γ2 receptors at concentration of 10 μM, Tx-NM3-1 almost completely inhibited GABA-evoked currents (Figure 4b). Two other toxins Tx-NM4 and Tx-NM2 were less effective and Tx-NM3-2 showed extremely low activity (Figure 4b). Interestingly, the efficiency of interactions of Tx-NM3-1, Tx-NM3-2, and Tx-NM4 with two GABAAR subtypes (α1β2γ2 and α3β2γ2) was different and the largest difference was observed for Tx-NM4 (Figure 4b).
More detailed studies of activity were performed at α1β3γ2 GABAAR. The dependences of GABA-elicited currents on toxin concentrations were investigated and IC50 values were determined. The inhibition curves for the most active toxins Tx-NM2 and Tx-NM3-1 are shown in Figure 5.
The data for all four toxins studied are given in the Table 2. On this receptor subtype, Tx-NM3-1 was the most active, followed by Tx-NM2. Tx-NM4 was one order of magnitude less active than the last two toxins, and Tx-NM3-2 manifested only very weak activity.
Summing up all the data concerning the action of the N. melanoleuca TFTs on GABAARs, we can conclude that Tx-NM3-1 was the most active, Tx-NM2 was a little bit less active, Tx-NM4 manifested very weak activity, and Tx-NM3-2 was practically inactive.

2.2.2. Interaction with nAChRs

The binding of isolated toxins to nAChRs was studied by competitive radioligand method using radioactive α-Bgt (125I-αBgt) as a ligand. Interaction with nAChRs of muscle type from Torpedo electric organ and of human α7 type heterologously expressed in GH4C1 cells was investigated. It was found that Tx-NM2, Tx-NM3-1, and Tx-NM4 efficiently inhibited binding to both receptor subtypes (Figure 6). IC50 values for studied toxins were in the nanomolar range (Table 3).
Interestingly, at Torpedo nAChR, the experimental data for Tx-NM2 differed greatly from those for Tx-NM3-1 and Tx-NM4 and fit two-site binding model with fairly smooth inhibition curve (Figure 6a). The difference in affinity of Tx-NM2 to these binding sites was about one order of magnitude (1 nM versus 8.66 nM) (Table 3). The affinities of Tx-NM3-1 and Tx-NM4 to Torpedo nAChR were almost similar, Tx-NM3-1 being a slightly more active. At the α7 nAChR, the most active was Tx-NM3-1 with IC50 value of 4.84 nM, while Tx-Nm2 and Tx-NM4 were less active manifesting IC50 values of 13.02 and 26.89 nM, respectively (Figure 6b). At a concentration of 1 µM, toxin Tx-NM3-2 inhibited 125I-αBgt binding by only about 40% and was not analyzed further.

2.2.3. Interaction with Acetylcholine Binding Proteins from Lymnaea stagnalis and Aplysia californica

Competitive radioligand assay with 125I-αBgt was applied also for testing the interaction of N. melanoleuca toxins with acetylcholine binding proteins (AChBPs) from L. stagnalis and A. californica. The results obtained showed that Tx-NM2 was the most active (Figure 7). The affinity of this toxin to both proteins exceeded that of α-Bgt and α-Ctx (Table 4). However, the affinity of Tx-NM3-1 and Tx-NM4 to both proteins was the lowest among the toxins studied (Figure 7, Table 4).

2.3. Molecular Modelling

The efficiency of Tx-NM3-1 and Tx-NM4 interactions with α1β3γ2 GABAAR differs more than an order of magnitude (Table 2). However, the amino acid sequences of these toxins differ only in five positions—three located at the tips of loops II and two at the tip of loop III. This may suggest the involvement of the loop III in interaction with GABAAR, in addition to loop II the importance of which for binding was shown earlier [15]. Interestingly, the amino acid sequences of Tx-NM3-1 and α-Ctx are practically identical in this loop III region: positions 49–58 in Figure 8. To find structural elements which may explain the differences in the interaction of N. melanoleuca toxins with GABAAR, we performed the molecular modeling of toxin spatial structures and structure of the complex between GABAAR and toxin.
The models of N. melanoleuca toxin structures were constructed by homology modeling in Swissmodel service using α-Ctx structure (PDB 1YI5) as a template. The loop III structures for N. melanoleuca toxins are quite similar (Figure 9a).
Resulting homology models were submitted to protein docking Tox Dock instrument on the Rosetta server [26,27] to generate plausible starting structures of the toxins-GABAAR complexes. Flexibility of toxin molecules and receptors were taken into account and model structures were subsequently subjected to short molecular dynamics using GROMACS 5.0 package [28]. The resulting structures were visualized and inspected in UCSF Chimera [29] (Figure 10). All models showed similar positions with the TFT’s loop II buried under the C-loop of beta3-subunit of GABAAR (Figure 10a–e for the side view and Figure 10f–j for the top view from the extracellular side). Modeled complexes were compared to published X-ray structures of α-Bgt in complex with α7 nAChR/AChBP chimera [30] (Figure 10l) and of α-Ctx in complex with AChBP [31] (Figure 10m) as well as cryo-EM structures of α-Bgt in complex with muscle nAChR [32] (Figure 10k). Structures of principal subunits (α1 and α7 nAChRs, AChBP(+) and β3 GABAAR) of all mentioned above complexes were aligned in UCSF Chimera and TFT positions relative to the receptor were inspected. Interestingly, all putative complexes with GABAAR showed TFT position tilted towards the complementary subunit as compared to nAChR complexes (Figure 10n). In such tilted position the loop III of TFT molecule can form contacts with the complementary subunit in the GABAAR binding site.
The positions toxin loops II placed to the orthosteric GABAAR ligand binding site show a remarkable correlation with the functional inhibitory activity of the modeled toxins. As follows from our measurements, all tested TFTs can be ranked according to their inhibitory activity toward GABAAR in the following way: α-Ctx = Tx-NM3-1 > Tx-NM2 > Tx-NM4 > α-Bgt. On the other hand, distances between β-carbons of Thr 202 located at the tip of loop C of β3 subunit and Arg 65 located on the complementary side of the orthosteric ligand binding site of the receptor α1 subunit are as follows: for α-Ctx—17 Å, Tx-NM3-1—15 Å, Tx-NM2—14 Å, Tx-NM4—13 Å, and α-Bgt—12 Å. The distance between these atoms may be used as a simple metric of loop C closure upon ligand binding and may be a possible predictor of the inhibitory activity of the TFT. The stronger toxin is bound to receptor and the deeper inserted in the binding pocket, the longer is distance between the indicated amino acid residues. However, this may not be true for low molecular ligands. Thus, cryo-EM structure of GABAAR with non-peptidic antagonist bicuculline (PDB 6HUK) is characterized by the even shorter distance between β-carbons of the above mentioned Thr and Arg residues, that is only 10 Å.
According to molecular modeling of the α-Ctx-GABAAR complex performed earlier [15], loop III contacts GABAAR through Asp56 salt bridges with Arg65 and Arg171 of the α1 subunit and Thr53 forming hydrogen bond with Val177 backbone of the α1 subunit. Asp56 is present in the amino acid sequence of Tx-NM3-1 but is replaced by glutamic acid in α-Bgt and Tx-NM2 and by asparagine in Tx-NM4 (Figure 9b,c). Thr53 present in α-Ctx and Tx-NM3-1 is replaced by proline in α-Bgt, Tx-Nm2, and Tx-Nm4. These replacements might substantially weaken the interaction of toxins with GABAAR. However, further more extensive molecular modeling and structure–function relationship studies are needed to investigate details of TFT-GABAAR pharmacophore in details.

3. Discussion

Earlier it was shown that several snake venom toxins blocked GABAAR [13,14,15]. We have found that the snake venom toxins manifested a different efficiency of the interaction with the receptor, α-Ctx being the most active [15]. As mentioned in the introduction, we suggested that the Arg36 present in α-Ctx (Arg39 in the alignment given in Figure 2) might determine the highest activity of this toxin against GABAAR. To check if this is true, we decided to compare the activity of toxins which differ at this position and for this purpose isolated toxins from N. melanoleuca cobra venom. Two long type α-neurotoxins were isolated earlier from this venom: long neurotoxin 1 (P01383) [24] containing valine residue at the position 38 (Figure 2), and long neurotoxin 2 (P01388) [23] with Arg38 in the sequence (Figure 2). The only information about biological activity available for these toxins was LD50 of 1.5 mg/kg for neurotoxin 1 by intraperitoneal injection [24]. We wanted to get these particular toxins, but in our work we isolated 4 toxins with molecular masses in the range of 7–8 kDa, characteristic for long type α-neurotoxins (Table 1). Their analysis by high resolution mass spectrometry showed that one of the isolated toxins, designated by us as Tx-NM4, was identical to long neurotoxin 2, while toxin Tx-NM3-1 was long neurotoxin 2 analogue with substitutions in five positions (Figure 2). Tx-NM2 represented an analogue of N. melanoleuca long neurotoxin 1 (P01383) and Tx-NM3-2 was homologous to muscarinic toxin-like protein 1 (MTLP-1, P82462) from N. kaouthia cobra venom. The differences between the published amino acid sequences and those determined in the present work may be explained by different geographical origin of snakes from which the venoms were obtained.
The biological activity of the isolated toxins was assessed against several molecular targets. In the studies on GABAAR, competition experiments with fluorescently labeled α-Ctx and electrophysiological assays were carried out using several combinations of receptor subunits. Both methods gave the consistent results. Toxin Tx-NM3-1 was the most active on all GABAAR subtypes studied and manifested activity similar to that of α-Ctx. However, at α1β3γ2 receptor subtype Tx-NM3-1 was less active than α-Ctx displaying IC50 of 680 nM in contrast to 236 nM for α-Ctx [15]. Tx-NM2 was slightly less active than Tx-NM3-1 showing IC50 of 1.25 µM at α1β3γ2 GABAAR. Tx-NM4 was a weaker antagonist than Tx-NM2 and Tx-NM3-1 at all subunit combinations, while Tx-NM3-2 was practically inactive. At the inhibition of α1β3γ2 GABAAR, the difference in affinity between the most active Tx-NM3-1 and the least active Tx-NM4 was more than one order of magnitude. Tx-NM4 at 10 µM more potently inhibited α1β2γ2 than α3β2γ2 GABAAR. Thus, the earlier suggestion about essential role of Arg36 (Arg39 in Figure 2 and Figure 8) in α-Ctx binding to GABAAR was not confirmed, as both Tx-NM3-1 and Tx-NM4 contain arginine residue at this position. The analysis of toxin amino acid sequences revealed the identity of loop III sequence in α-Ctx and Tx-NM3-1. Molecular modeling of α-Ctx-GABAAR complex revealed the contacts between loop III of α-Ctx and α1 subunit of GABAAR. It was found that Asp56 of α-Ctx loop III forms salt bridges with Arg65 and Arg171 of the α1 subunit, and Thr53 forms a hydrogen bond with Val177 backbone of the α1 subunit (Figure 9b). The aspartic acid and threonine residues are present at these positions in Tx-NM3-1 as well. The interaction of these residues with α1 subunit of GABAAR may contribute to the stronger binding of α-Ctx and Tx-NM3-1 to the receptor.
Tx-NM3-2 is the muscarinic toxin-like protein and its amino acid sequence differs greatly from those of the long type neurotoxins to which other studied toxins belong. The muscarinic toxin-like protein was tested on GABAAR for the first time and it showed some activity, albeit very weak, against α3β2γ2 GABAAR (Figure 4). This is the first indication of the activity of muscarinic toxin-like protein against GABAAR.
The activity of all isolated toxins was studied in competition experiments for binding to Torpedo and α7 nAChRs as well as to two AChBPs. The neurotoxins tested showed a relatively high affinity to both receptor subtypes. It was found that at Torpedo nAChR, Tx-NM2 distinguished two binding sites, the affinity differing by about an order of magnitude (Table 3). Earlier, it was shown that the two binding sites in Torpedo and muscle types nAChRs were distinguished by αδ-bungarotoxins from the krait B. candidus venom [12], the difference being 17-fold for αδ-bungarotoxin-1. Comparison of the amino acid sequences of N. melanoleuca toxins with those of bungarotoxins and α-Ctx showed that the sequence of Tx-NM2 has a higher identity to that of αδ-bungarotoxin-1 than to α-Btx and α-Ctx (69% versus 61 and 55%, respectively, Figure 8). At the same time Tx-NM3-1 has 81% identical residues with α-Ctx and only 50% with αδ-bungarotoxin-1. Moreover, the sequence characteristics that were supposed to be responsible for difference in activity between α-Btx and αδ-bungarotoxin-1, i.e., the shortened loop I, the change of Phe residue in position 32 to Trp as well as Arg25 to Thr, are present in Tx-NM2. In addition, only Tx-NM2 and αδ-bungarotoxin-1 contain Tyr residue in position 4 and a positively charged residue in position 5. This similarity in structural features and activity between Tx-NM2 and αδ-bungarotoxin-1 supports the earlier ideas [12] on the capacity of long chain neurotoxin to distinguish two binding sites in the muscle-type nAChRs. It was found that Tx-NM3-1 was the most active against α7 nAChR (IC50 4.84 nM), followed by Tx-NM2 and Tx-NM4 with IC50 of 13.02 and 26.89 nM, respectively. Although the affinities of these toxins for nAChR are much higher than for GABAAR, their rank coincides with that observed for GABAAR suggesting that both receptors might have similar structural elements involved in toxin binding to them.
AChBPs are considered as models of the extracellular domains of Cys-loop receptors, in particular of nAChRs. Interestingly, AChBPs has only 20–24% sequence identity with nAChRs, however their pharmacological properties are similar to those of nAChRs. Studies of the binding of N. melanoleuca toxins to AChBPs from L. stagnalis and A. californica showed that toxin Tx-NM2 possessed the highest affinity to both proteins, while toxins Tx-NM3-1 and Tx-NM4 were several orders of magnitude less active. On A. californica AChBP, Tx-NM3-1 was the least active, which is very different from its activity against nAChR and GABAAR, where it was most active. This once again highlights the differences between AChBPs and the extracellular domains of nAChR and the GABAAR.
Considering the above data for nAChRs and AChBPs, one can conclude that for these targets, N. melanoleuca toxins manifest affinities similar to those of α-Btx and α-Ctx, and Tx-NM2 on AChBP from L. stagnalis showed the affinity higher than these two neurotoxins.
To obtain some information about possible molecular mechanisms determining the specificity of TFT interaction with GABAAR, we performed molecular modeling of the complexes formed by N. melanoleuca toxins, α-Bgt and α-Ctx with GABAAR. The modeling showed that the orientation of toxin molecules at GABAAR and nAChRs were different and toxin molecules in complexes with GABAAR tilted towards the complementary subunit. This tilt can lead to close contact of toxin loop III with the α1 subunit of the receptor. More detailed consideration revealed the amino acid residues which can form salt bridges (Asp56 in α-Ctx) and hydrogen bond (Thr53 in α-Ctx) with α1 subunit. These amino acid residues are present in the amino acid sequence of Tx-NM3-1, interacting with GABAAR with affinity similar to that of α-Ctx, but are replaced by other residues in less active toxins. Such substitutions may lead to a significant decrease in the efficiency of toxin interactions with GABAAR.

4. Conclusions

In our previous paper [15] it was found that among several snake toxins studied only α-Ctx manifested high activity against GABAAR and we put forward the hypothesis about essential role of Arg36 as the determinant of high affinity to GABAAR. To check this hypothesis, in this work we additionally studied several snake toxins for their ability to interact with GABAAR. For this purpose, four toxins were isolated from African cobra N. melanoleuca venom and their amino acid sequences were established by mass spectrometry. The amino acid sequence of one toxin was identical to that of previously known N. melanoleuca long neurotoxin 2. The second toxin sequenced differed from that of neurotoxin 2 in five positions. The third one was homologous to N. melanoleuca long neurotoxin 1; its sequence differed from that of neurotoxin 1 in three positions. One more toxin was homologous to muscarinic toxin-like protein from N. kaouthia venom and this is the first muscarinic toxin-like protein isolated from African cobra venom. The interaction of N. melanoleuca toxins with Torpedo and α7 nAChRs as well as with AChBPs and several subtypes of GABAARs was studied. One of isolated toxins, being the most active on AChBPs, manifested different affinity to two binding sites on Torpedo nAChR. Together with earlier data [12] this may indicate that there is a group of long type α-neurotoxins capable to bind with different affinity to two binding sites in the muscle type nAChR. All N. melanoleuca toxins interacted with the GABAAR much weaker than with the nAChR: one neurotoxin was almost as active as α-Ctx, while others manifested lower activity. The earlier hypothesis about essential role of Arg36 as the sole determinant of high toxin affinity to GABAAR was not confirmed, but the results of molecular modeling suggest that the loop III may contribute to the efficient interaction of some long-chain neurotoxins with GABAAR. Experimental proof of the modeling data will be the task of our future work.

5. Materials and Methods

5.1. Materials

All salts obtained from local suppliers were of analytical grade or higher. Venom of cobra N. melanoleuca was from Latoxan (Valence, France). Acetonitrile was purchased from Catrosa Reaktiv LLC (Moscow, Russia), and trifluoroacetic acid from Merck KGaA (Darmstadt, Germany). GH4C1 cells transfected with hα7 nAChR cDNA were a gift of the Eli-Lilly Co. (London, UK). Muscle-type nAChR-enriched membranes from the electric organs of Torpedo californica were kindly provided by Prof. F. Hucho (Free University of Berlin, Germany). Acetylcholine binding proteins from L. stagnalis and A. californica were from Prof. A.B. Smit (Faculty of Earth and Life Sciences, Vrije Universiteit, Amsterdam).

5.2. Isolation of Neurotoxins

A 600 mg sample of dried N. melanoleuca venom was dissolved in 0.1 M ammonium acetate buffer, pH 6.2, and applied to a Sephadex G50s column (4.5 × 150 cm) equilibrated in the same buffer. The column was eluted at flow rate 32 mL/min. The fractions obtained were pooled as shown in Figure 1a. Fraction 6 was further separated on a HEMA BIO 1000CM column (4 × 250 mm) (Tessek, Prague, Czech Republic) in an ammonium acetate gradient from 5 to 500 mM (pH 7.5) in 100 min at flow rate 1.0 mL/min (Figure 1b). Fractions 2, 3 and 4 were freeze-dried and further purified by reversed phase chromatography on Jupiter C18 column (10 × 250 mm, Phenomenex, Torrance, CA, USA) in in a gradient of acetonitrile 20–35% in 60 min in the presence of 0.1% trifluoroacetic acid, at a flow rate of 2.0 mL/min (Figure S1). After freeze-drying, the obtained proteins were used for further studies.

5.3. Mass Spectrometry Analysis

For mass spectrometry measurements, the carbamidomethylated toxins were digested with trypsin and chymotrypsin at a 1:50 (w/w) ratio overnight at 37 °C. Desalting of peptides was carried out using SDB-RPS StageTips that were prepared as described earlier [33]. After overnight digestion, peptide solution was acidified by equal volume of 2% (v/v) TFA and peptides were loaded on SDB-RPS StageTip by centrifugation at 200× g. StageTip was washed by 50 µL 0.2% (v/v) TFA and peptides were eluted by 50 μL 50% (v/v) acetonitrile, 5% (v/v) ammonia, lyophilized, and stored at −80 °C. Before analyses, peptides were dissolved in 20 µL of 2% (v/v) acetonitrile, 0.1% (v/v) TFA, and sonicated for 2 min. Samples were loaded to a home-made trap column 20 × 0.1 mm, packed with Inertsil ODS3 3 μm sorbent (GLSciences, Tokyo, Japan ), in the loading buffer (2% ACN, 98% H2O, 0.1% TFA) at 10 μL/min flow and separated at RT in a home-packed [34] fused-silica column 300 × 0.1 mm packed with Reprosil PUR C18AQ 1.9 (Dr. Maisch, Ammerbuch-Entringen, Germany) into an emitter prepared with P2000 Laser Puller (Sutter Instrument, Novato, CA, USA). Reverse-phase chromatography was performed with an Ultimate 3000 Nano LC System (Thermo Fisher Scientific, Waltham, MA, USA), which was coupled to a Q Exactive Plus benchtop Orbitrap mass spectrometer (Thermo Fisher Scientific) via a nanoelectrospray source (Thermo Fisher Scientific). Peptide samples were eluted with a linear gradient of 80% ACN, 19.9% H2O, 0.1% FA (buffer B) in 99.9% H2O, 0.1% FA (solvent A) from 4 to 36% of solvent B in 60 min at 0.5 μL/min flow, intact toxins were separated by linear gradient from 10 to 60% of solvent B in 18 min at 0.5 μL/min flow.
MS raw files were analyzed by PEAKS Studio 8.5 (Waterloo, ON, Canada) [35] and peak lists were searched against Serpentes Uniprot-Tremble FASTA (canonical and isoform) database version of May 2018 (144954 entries) with cysteine carbamidomethylation as a fixed modification and methionine oxidation and asparagine and glutamine deamidation as variable modifications. Enzyme specificity in the database search was set to trypsin with semi-specific digest mode. False discovery rate was set to 0.01 for peptide-spectrum matches and was determined by searching a reverse database. Peptide identification was performed with an allowed initial precursor mass deviation up to 10 p.p.m. and an allowed fragment mass deviation 0.05 Da.

5.4. Expression of GABAAR in Xenopus Oocytes

Xenopus laevis oocytes were prepared as described [36]. The work with oocytes was approved by the Shemyakin-Ovchinnikov Institute of Bioorganic Chemistry RAS with the approval number IACUC 251/2018 26.02.18. Plasmid DNAs encoding rat GABAA receptor subunits in pCI mammalian expression vector (Promega, Madison, WI, USA) were kindly provided by Dr. M. Ernst from the Medical University of Vienna. Next day after harvesting oocytes were injected by means of Nanoject II (Drummond Scientific, Broomall, PA, USA) with 2 ng mixture containing vector DNAs encoding receptor α1 or 3, β2 or 3, and γ2 subunits at 1:1:10 mass ratio. Injected oocytes were incubated at 18 °C for 2–3 days in ND96 solution (5 mM HEPES/NaOH pH 7.4, 96 mM NaCl, 2 mM KCl, 1.8 mM CaCl2, 2 mM MgCl2) supplemented with gentamycin at 40 μg/mL.

5.5. Two-Electrode Voltage Clamp Electrophysiology Assay

Oocyte was placed in the flow through chamber combined with a nylon grid holding the bath of ND96 solution. Membrane potential of oocyte was clamped at −60mV by TURBO TEC-03X (npi electronic GmbH, Tamm, Germany). Electrodes were pulled from borosilicate capillary (Warner Instruments, Holliston, MA, USA) and filled with 3M KCl solution. Currents were recorded and digitized by WinWCP (University of Strathclyde, Glasgow, UK) software. After stable amplitude of 10 μM GABA-evoked currents was obtained oocyte was pre-incubated with toxin ND96-based solution for three minutes followed by co-application with 10 μM GABA. After steady baseline potential was achieved usually in a five-minute washout session oocyte was perfused with GABA again for current amplitude stability control purposes and pre-incubated with next sample. All solutions were applied manually with automatic pipette in a volume of 200 μL (Eppendorf, Hamburg, Germany). To confirm γ subunit incorporation, GABA response in the presence of 50 μM Zn2⁺ and 1 μM diazepam was tested under the same conditions. Data for toxins are presented as the amplitude ratio of the currents elicited by 10 μM GABA in the presence of toxin to the average of the currents in control elicited by 10 μM GABA before co-application with toxin (100% control current amplitude) in the same oocyte. Data were collected from at least three oocytes from three different batches. Representative current traces, bar graphs and toxin-dependent dose response curves were plotted by Origin 8.1 (OriginLab, Northampton, MA, USA).

5.6. Mammalian Cell Culture

Mouse neuroblastoma Neuro2a cells (Russian collection of cell cultures, Institute of Cytology, Saint Petersburg, Russia) were routinely cultured in incubator (Sanyo, Osaka, Japan) at 37 °C and 5% CO2 in tissue culture treated T25 flasks (SPL, Pocheon, Korea) containing 5 mL Dulbecco’s modified Eagle’s medium (DMEM, Paneco, Moscow, Russia) supplemented with 10% fetal bovine serum (FBS, PAA Laboratories, Pasching, Austria), 50 units/mL streptomycin and 50 µg/mL penicillin. Cells were splitted 1:10 at confluency twice a week non-enzymatically using Versene solution (Paneco, Moscow, Russia).

5.7. Fluorescent Ligand Competition Assay

Neuro2a cells sub-cultured 1:5 24 h before transfection were growing on clear 96-well plate (Corning, Corning, NY, USA) in a complete DMEM (Paneco, Moscow, Russia). Lipofectamine (Invitrogen, Waltham, MA, USA) mediated transfection was performed with equal amounts (0.14 µg/mL) of pCI plasmid expression vectors encoding rat α1, β3, γ2 GABAAR subunits. Transfected Neuro2a cells were grown at 37 °C in 5% CO2 incubator for 72 h, at the day of experiment medium was substituted for extracellular solution containing (in mM) 140 NaCl, 2 CaCl2, 2.8 KCl, 4 MgCl2, 20 HEPES, 10 glucose at pH 7.4. Cells were pre-incubated with 10 µM of toxins for 15 min at room temperature followed by 20 min of incubation with 50 nM Alexa Fluor 546 α-Ctx conjugate in the final volume of 100 µL. Afterwards cells were washed 3 times with two-fold excess of extracellular solution. To control the level of non-specific fluorescence experiment with 10 µM α-Ctx was run under the same conditions. By means of epifluorescent microscope IX71 (Olympus, Tokyo, Japan) equipped with CCD camera pictures of 3 fields chosen on each plate well in bright-field illumination were taken. Fluorescence was counted using CellX and ImageJ open-source software. Intensity of the fluorescence was normalized on integral intensity of the plate well incubated in presence of 50 nM Alexa Fluor 546 α-Ctx conjugate. Each experimental point is an average of integral intensity independently measured on 6 plate wells from three separate passages ± SEM.

5.8. nAChR Competition Radioligand Assay

The competition binding assays with radio-iodinated α-bungarotoxin (125I-α-Bgt) were performed as in [37]. Briefly, suspension of GH4C1 cells stably transfected with human α7 nAChR (0.4 nM αBgt binding sites) were incubated in 50 μL binding buffer (20 mM Tris-HCl, pH 8.0, containing 1 mg/mL bovine serum albumin) for 90 min with various amounts of toxins. Thereafter, 0.1–0.2 nM 125I-α-Bgt (500 Ci/mmol) was added, and after an additional 5 min incubation, cell suspensions were applied to GF/C glass filters (Cytiva, Marlborough, MA, USA) pretreated with 0.3% polyethyleneimine. The samples were then washed (3 × 4 mL) with 20 mM cold Tris-HCl buffer, pH 8.0, containing 0.1 mg/mL bovine serum albumin and bound radioactivity was measured with a Wallac 1470 Wizard Gamma Counter (PerkinElmer, Waltham, MA, USA). Nonspecific 125I-αBgt binding was determined in the presence of 200-fold excess of α-Ctx. The competition binding assays with AChBPs were performed as described in [38].

5.9. Molecular Modeling

The molecular models of N. melanoleuca toxin structures were constructed using free bioinformatic tool Swissmodel (https://swissmodel.expasy.org/ (accessed on 12 February 2021)). α-Ctx structure (PDB 1YI5) was used as template. The molecular model of α-Ctx-GABAAR complex was taken from [15]. Briefly, the model of orthosteric binding site was constructed basing on extracellular domain from X-ray structure of β3 GABAAR subunit [39]. Swissmoldel service was used to build the extracellular domain of α1 GABAAR subunit. RMSD between modelled α1 extracellular domain and recently published cryo-EM structure was calculated by UCSF Chimera Match maker tool and did not exceed 1.2 Å. Homology models were submitted to protein docking “Tox Dock” instrument on the Rosetta server [26,27] to generate structures of the toxin-GABAAR complexes. Flexibility of toxin molecules and receptors were taken into account and model structures were subsequently subjected to short molecular dynamics using GROMACS 5.0 package. Briefly, putative structures of TFT-GABAAR complexes obtained via “Tox Dock” were energy minimized using steepest descent minimization algorithm to a maximum force <1000 kJ/mol/nm. Short-range electrostatic and Van der Waals cut-offs were set to 1 Å, periodic boundary conditions were constructed with rhombododecaedron simulation box having 1.2 Å from protein image to the closest boundary. After energy minimization two consequent constrained 100 ps molecular dynamics were performed to equilibrate systems in NVT (constant number of particles, volume, and temperature) and NPT (constant number of particles, pressure, and temperature) conditions. Equilibration was followed by 100 ps of unconstrained molecular dynamics [28]. The resulting structures were visualized and inspected in UCSF Chimera [29]. To compare TFT positions “Match” and “Match-align” functions of UCSF Chimera were used.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2072-6651/13/2/164/s1, Figure S1: Separation of fractions 2 (a), 3 (b), and 4 (c) after ion exchange chromatography (Figure 1b) by reversed phase HPLC on Jupiter C18 column (10 × 250 mm, Phenomenex, Torrance, CA, USA) in a gradient of acetonitrile 20–35% in 60 min in the presence of 0.1% trifluoroacetic acid, at a flow rate of 2.0 mL/min. Figure S2: Extracted-ion chromatogram (XIC) from an LC-MS analysis and high-resolution spectra at z = 7 for toxins Tx-NM2, XIC at m/z = 1148.5–1149.5 (a); Tx-MN3-1, XIC at m/z = 1113.6–1114.6 (b); Tx-Nm3-2, XIC at m/z = 1064.3–1065.3 (c); and Tx-NM4, XIC at m/z = 1109.3–1110.3 (d).

Author Contributions

Conceptualization, D.K., I.K., V.T. and Y.U.; funding acquisition, D.K., I.K., V.T. and Y.U.; investigation, L.S., E.K., R.Z., T.A., D.K. and I.K.; methodology, R.Z.; resources, T.A.; supervision, V.T. and Y.U.; writing—original draft, L.S., E.K. and D.K.; writing—review and editing, I.K., V.T. and Y.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Foundation for Basic Research according to the research project No 20-54-00033.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki, and approved by the Ethics Committee of the Shemyakin-Ovchinnikov Institute of Bioorganic Chemistry RAS (the approval number IACUC 251/2018 26.02.18).

Informed Consent Statement

Not applicable.

Data Availability Statement

All data is contained within this article and supplementary materials.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Nys, M.; Kesters, D.; Ulens, C. Structural insights into Cys-loop receptor function and ligand recognition. Biochem. Pharmacol. 2013, 86, 1042–1053. [Google Scholar] [CrossRef]
  2. Dutertre, S.; Nicke, A.; Tsetlin, V.I. Nicotinic acetylcholine receptor inhibitors derived from snake and snail venoms. Neuropharmacology 2017, 127, 196–223. [Google Scholar] [CrossRef]
  3. Tsetlin, V.I.; Kasheverov, I.E.; Utkin, Y.N. Three-finger proteins from snakes and humans acting on nicotinic receptors: Old and new. J. Neurochem. 2020. [Google Scholar] [CrossRef]
  4. Zoli, M.; Pucci, S.; Vilella, A.; Gotti, C. Neuronal and Extraneuronal Nicotinic Acetylcholine Receptors. Curr. Neuropharmacol. 2018, 16, 338–349. [Google Scholar] [CrossRef]
  5. Cetin, H.; Beeson, D.; Vincent, A.; Webster, R. The Structure, Function, and Physiology of the Fetal and Adult Acetylcholine Receptor in Muscle. Front. Mol. Neurosci. 2020, 13, 581097. [Google Scholar] [CrossRef] [PubMed]
  6. Papke, R.L.; Lindstrom, J.M. Nicotinic acetylcholine receptors: Conventional and unconventional ligands and signaling. Neuropharmacology 2020, 168, 108021. [Google Scholar] [CrossRef]
  7. Nayak, T.K.; Bruhova, I.; Chakraborty, S.; Gupta, S.; Zheng, W.; Auerbach, A. Functional differences between neurotrans-mitter binding sites of muscle acetylcholine receptors. Proc. Natl. Acad. Sci. USA 2014, 111, 17660–17665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Osaka, H.; Malany, S.; Molles, B.E.; Sine, S.M.; Taylor, P. Pairwise electrostatic interactions between alpha-neurotoxins and gamma, delta, and epsilon subunits of the nicotinic acetylcholine receptor. J. Biol. Chem. 2000, 275, 5478–5484. [Google Scholar] [CrossRef] [Green Version]
  9. Ackermann, E.J.; Taylor, P. Nonidentity of the α-Neurotoxin Binding Sites on the Nicotinic Acetylcholine Receptor Revealed by Modification in α-Neurotoxin and Receptor Structures. Biochemistry 1997, 36, 12836–12844. [Google Scholar] [CrossRef] [PubMed]
  10. Antil, S.; Servent, D.; Ménez, A. Variability among the Sites by Which Curaremimetic Toxins Bind to Torpedo Acetylcholine Receptor, as Revealed by Identification of the Functional Residues of α-Cobratoxin. J. Biol. Chem. 1999, 274, 34851–34858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Nirthanan, S.; Charpantier, E.; Gopalakrishnakone, P.; Gwee, M.C.; Khoo, H.E.; Cheah, L.S.; Kini, R.M.; Bertrand, D. Neuromuscular effects of candoxin, a novel toxin from the venom of the Malayan krait (Bungarus candidus). Br. J. Pharmacol. 2003, 139, 832–844. [Google Scholar] [CrossRef] [Green Version]
  12. Utkin, Y.N.; Kuch, U.; Kasheverov, I.E.; Lebedev, D.S.; Cederlund, E.; Molles, B.E.; Polyak, I.L.; Ivanov, I.A.; Prokopev, N.A.; Ziganshin, R.H.; et al. Novel long-chain neurotoxins from Bungarus candidus distinguish the two binding sites in muscle-type nicotinic acetylcholine receptors. Biochem. J. 2019, 476, 1285–1302. [Google Scholar] [CrossRef]
  13. McCann, C.M.; Bracamontes, J.; Steinbach, J.H.; Sanes, J.R. The cholinergic antagonist α-bungarotoxin also binds and blocks a subset of GABA receptors. Proc. Natl. Acad. Sci. USA 2006, 103, 5149–5154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Hannan, S.; Mortensen, M.; Smart, T.G. Snake neurotoxin α-bungarotoxin is an antagonist at native GABA(A) receptors. Neuropharmacology 2015, 93, 28–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kudryavtsev, D.S.; Shelukhina, I.V.; Son, L.V.; Ojomoko, L.O.; Kryukova, E.V.; Lyukmanova, E.N.; Zhmak, M.N.; Dolgikh, D.A.; Ivanov, I.A.; Kasheverov, I.E.; et al. Neurotoxins from Snake Venoms and α-Conotoxin ImI Inhibit Functionally Active Ionotropic γ-Aminobutyric Acid (GABA) Receptors. J. Biol. Chem. 2015, 290, 22747–22758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Olsen, R.W.; Sieghart, W. International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: Classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacol. Rev. 2008, 60, 243–260. [Google Scholar] [CrossRef] [Green Version]
  17. Feng, H.J.; Forman, S.A. Comparison of αβδ and αβγ GABAA receptors: Allosteric modulation and identification of subunit arrangement by site-selective general anesthetics. Pharmacol. Res. 2018, 133, 289–300. [Google Scholar] [CrossRef] [PubMed]
  18. Sieghart, W.; Sperk, G. Subunit composition, distribution and function of GABA(A) receptor subtypes. Curr. Top Med. Chem. 2002, 2, 795–816. [Google Scholar] [CrossRef]
  19. Camacho-Hernandez, G.A.; Taylor, P. Lessons from nature: Structural studies and drug design driven by a homologous sur-rogate from invertebrates, AChBP. Neuropharmacology 2020, 179, 108108. [Google Scholar] [CrossRef] [PubMed]
  20. Yu, J.; Zhu, X.; Zhang, L.; Kudryavtsev, D.; Kasheverov, I.; Lei, Y.; Zhangsun, D.; Tsetlin, V.; Luo, S. Species specificity of rat and human α7 nicotinic acetylcholine receptors towards different classes of peptide and protein antagonists. Neuropharmacology 2018, 139, 226–237. [Google Scholar] [CrossRef] [PubMed]
  21. Zhu, X.; Pan, S.; Xu, M.; Zhang, L.; Yu, J.; Yu, J.; Wu, Y.; Fan, Y.; Li, H.; Kasheverov, I.E.; et al. High Selectivity of an α-Conotoxin LvIA Analogue for α3β2 Nicotinic Acetylcholine Receptors Is Mediated by β2 Functionally Important Residues. J. Med. Chem. 2020, 63, 13656–13668. [Google Scholar] [CrossRef] [PubMed]
  22. Rucktooa, P.; Smit, A.B.; Sixma, T.K. Insight in nAChR subtype selectivity from AChBP crystal structures. Biochem. Pharmacol. 2009, 78, 777–787. [Google Scholar] [CrossRef] [Green Version]
  23. Botes, D.P. Snake venom toxins. The amino acid sequences of toxins b and d from Naja melanoleuca venom. J. Biol. Chem. 1972, 247, 2866–2871. [Google Scholar] [CrossRef]
  24. Shipolini, R.A.; Bailey, G.S.; Banks, B.E.C. The Separation of a Neurotoxin from the Venom of Naja melanoleuca and the Primary Sequence Determination. Eur. J. Biochem. 1974, 42, 203–211. [Google Scholar] [CrossRef]
  25. Boileau, A.J.; Baur, R.; Sharkey, L.M.; Sigel, E.; Czajkowski, C. The relative amount of cRNA coding for gamma2 subunits affects stimulation by benzodiazepines in GABA(A) receptors expressed in Xenopus oocytes. Neuropharmacology 2002, 43, 695–700. [Google Scholar] [CrossRef]
  26. Leffler, A.E.; Kuryatov, A.; Zebroski, H.A.; Powell, S.R.; Filipenko, P.; Hussein, A.K.; Gorson, J.; Heizmann, A.; Lyskov, S.; Tsien, R.W.; et al. Discovery of peptide ligands through docking and virtual screening at nicotinic acetylcholine receptor homology models. Proc. Natl. Acad. Sci. USA 2017, 114, E8100–E8109. [Google Scholar] [CrossRef] [Green Version]
  27. Lyskov, S.; Chou, F.C.; Conchúir, S.Ó.; Der, B.S.; Drew, K.; Kuroda, D.; Xu, J.; Weitzner, B.D.; Renfrew, P.D.; Sripakdeevong, P.; et al. Serverification of molecular model-ing applications: The Rosetta Online Server that Includes Everyone (ROSIE). PLoS ONE 2013, 8, e63906. [Google Scholar] [CrossRef] [Green Version]
  28. Berendsen, H.J.C.; van der Spoel, D.; van Drunen, R. GROMACS: A message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 1995, 91, 43–56. [Google Scholar] [CrossRef]
  29. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, S.; Li, S.X.; Bren, N.; Cheng, K.; Gomoto, R.; Chen, L.; Sine, S.M. Complex between α-bungarotoxin and an α7 nicotinic receptor ligand-binding domain chimaera. Biochem. J. 2013, 454, 303–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Bourne, Y.; Talley, T.T.; Hansen, S.B.; Taylor, P.; Marchot, P. Crystal structure of a Cbtx-AChBP complex reveals essential interactions between snake alpha-neurotoxins and nicotinic receptors. EMBO J. 2005, 24, 1512–1522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Rahman, M.M.; Teng, J.; Worrell, B.T.; Noviello, C.M.; Lee, M.; Karlin, A.; Stowell, M.H.B.; Hibbs, R.E. Structure of the Native Muscle-type Nicotinic Receptor and Inhibition by Snake Venom Toxins. Neuron 2020, 106, 952–962. [Google Scholar] [CrossRef] [PubMed]
  33. Rappsilber, J.; Mann, M.; Ishihama, Y. Protocol for micro-purification, enrichment, pre-fractionation and storage of peptides for proteomics using StageTips. Nat. Protoc. 2007, 2, 1896–1906. [Google Scholar] [CrossRef] [PubMed]
  34. Kovalchuk, S.I.; Jensen, O.N.; Rogowska-Wrzesinska, A. FlashPack: Fast and Simple Preparation of Ultrahigh-performance Capillary Columns for LC-MS. Mol. Cell. Proteom. 2019, 18, 383–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ma, B.; Zhang, K.; Hendrie, C.; Liang, C.; Li, M.; Doherty-Kirby, A.; Lajoie, G. PEAKS: Powerful software for peptidede novo sequencing by tandem mass spectrometry. Rapid Commun. Mass Spectrom. 2003, 17, 2337–2342. [Google Scholar] [CrossRef] [PubMed]
  36. Vulfius, C.A.; Lebedev, D.S.; Kryukova, E.V.; Kudryavtsev, D.S.; Kolbaev, S.N.; Utkin, Y.N.; Tsetlin, V.I. PNU-120596, a positive allosteric modulator of mammalian α7 nicotinic acetylcholine receptor, is a negative modulator of ligandgated chlorideselective channels of the gastropod Lymnaea stagnalis. J. Neurochem. 2020, 155, 274–284. [Google Scholar] [CrossRef] [PubMed]
  37. Lebedev, D.S.; Kryukova, E.V.; Ivanov, I.A.; Egorova, N.S.; Timofeev, N.D.; Spirova, E.N.; Tufanova, E.Y.; Siniavin, A.E.; Kudryavtsev, D.S.; Kasheverov, I.E.; et al. Oligoarginine Peptides, a New Family of Nicotinic Acetylcholine Receptor Inhibitors. Mol. Pharmacol. 2019, 96, 664–673. [Google Scholar] [CrossRef]
  38. Celie, P.H.N.; Kasheverov, I.E.; Mordvintsev, D.Y.; Hogg, R.C.; Van Nierop, P.; Van Elk, R.; Van Rossum-Fikkert, S.E.; Zhmak, M.N.; Bertrand, D.; Tsetlin, V.; et al. Crystal structure of nicotinic acetylcholine receptor homolog AChBP in complex with an α-conotoxin PnIA variant. Nat. Struct. Mol. Biol. 2005, 12, 582–588. [Google Scholar] [CrossRef] [PubMed]
  39. Miller, P.S.; Aricescu, A.R. Crystal structure of a human GABAA receptor. Nat. Cell Biol. 2014, 512, 270–275. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Isolation of N. melanoleuca neurotoxins. (a) Separation of the crude N. melanoleuca venom by gel-filtration on Sephadex G50 column. (b) Separation of fraction 6 by ion exchange chromatography on HEMA BIO 1000CM column.
Figure 1. Isolation of N. melanoleuca neurotoxins. (a) Separation of the crude N. melanoleuca venom by gel-filtration on Sephadex G50 column. (b) Separation of fraction 6 by ion exchange chromatography on HEMA BIO 1000CM column.
Toxins 13 00164 g001
Figure 2. Alignment of amino acid sequences. Long 1 is long type α-neurotoxin 1 from N. melanoleuca P01383 (3L21_NAJME), Long 2 is long type α-neurotoxin 2 from N. melanoleuca P01388 (3L22_NAJME), toxin P82462 (3SUC1_NAJKA) is muscarinic toxin like protein from N. kaouthia venom. All toxins have three loops stabilized by five disulfide bonds. The residues different in α-neurotoxin 1 and Tx-NM2 as well as α-neurotoxin 2 and Tx-NM3-1 are underlined. Arginine and valine residues corresponding to position 36 in α-Ctx sequence are shown in lilac and green, respectively.
Figure 2. Alignment of amino acid sequences. Long 1 is long type α-neurotoxin 1 from N. melanoleuca P01383 (3L21_NAJME), Long 2 is long type α-neurotoxin 2 from N. melanoleuca P01388 (3L22_NAJME), toxin P82462 (3SUC1_NAJKA) is muscarinic toxin like protein from N. kaouthia venom. All toxins have three loops stabilized by five disulfide bonds. The residues different in α-neurotoxin 1 and Tx-NM2 as well as α-neurotoxin 2 and Tx-NM3-1 are underlined. Arginine and valine residues corresponding to position 36 in α-Ctx sequence are shown in lilac and green, respectively.
Toxins 13 00164 g002
Figure 3. Inhibition of Alexa Fluor 546 α-Ctx staining of Neuro 2a cells expressing α1β3γ2 GABAAR by N. melanoleuca toxins at concentrations of 10 μM. Data are shown as percent of control staining without toxins and presented as Mean ± SEM (n = 3).
Figure 3. Inhibition of Alexa Fluor 546 α-Ctx staining of Neuro 2a cells expressing α1β3γ2 GABAAR by N. melanoleuca toxins at concentrations of 10 μM. Data are shown as percent of control staining without toxins and presented as Mean ± SEM (n = 3).
Toxins 13 00164 g003
Figure 4. Activity testing of N. melanoleuca toxins by two-electrode voltage clamp (TEVC) on Xenopus oocytes expressing GABAA α1β2γ2 and α3β2γ2 receptor subtypes. (a) Representative current traces illustrating inhibition of 10 μM GABA-evoked currents by N. melanoleuca toxins at 10 μM concentration. (b) Comparison of the activity of N. melanoleuca toxins at GABAAR. The response to GABA without toxins is taken as 100%. Data are shown for α1β2γ2 and α3β2γ2 GABAAR and presented as Mean ± SEM (n = 3). For toxins Tx-NM3-1, Tx-NM3-2, and Tx-NM4, the difference in interaction with two receptor subtypes is statistically significant. p < 0.05 according to Student’s t test. (c) Example current traces at α1β2γ2 GABAAR illustrating gamma subunit incorporation. GABA potency was not affected by 50 μM of Zn2⁺ but was enhanced by 1 μM of Diazepam (Dz).
Figure 4. Activity testing of N. melanoleuca toxins by two-electrode voltage clamp (TEVC) on Xenopus oocytes expressing GABAA α1β2γ2 and α3β2γ2 receptor subtypes. (a) Representative current traces illustrating inhibition of 10 μM GABA-evoked currents by N. melanoleuca toxins at 10 μM concentration. (b) Comparison of the activity of N. melanoleuca toxins at GABAAR. The response to GABA without toxins is taken as 100%. Data are shown for α1β2γ2 and α3β2γ2 GABAAR and presented as Mean ± SEM (n = 3). For toxins Tx-NM3-1, Tx-NM3-2, and Tx-NM4, the difference in interaction with two receptor subtypes is statistically significant. p < 0.05 according to Student’s t test. (c) Example current traces at α1β2γ2 GABAAR illustrating gamma subunit incorporation. GABA potency was not affected by 50 μM of Zn2⁺ but was enhanced by 1 μM of Diazepam (Dz).
Toxins 13 00164 g004
Figure 5. Activity testing of Tx-NM3-1 and Tx-NM2 at α1β3γ2 GABAAR expressed in Xenopus oocytes by TEVC. (a) Representative current traces elicited by 10 μM GABA in presence of Tx-NM3-1 and Tx-NM2 in 0–10 μM concentration range. (b) Dose–response curves showing the dependence of current elicited by 10 μM GABA on the concentration of toxins Tx-NM3-1 (squares) and Tx-NM2 (circles). The data are presented as Mean ± SEM (n = 3–5).
Figure 5. Activity testing of Tx-NM3-1 and Tx-NM2 at α1β3γ2 GABAAR expressed in Xenopus oocytes by TEVC. (a) Representative current traces elicited by 10 μM GABA in presence of Tx-NM3-1 and Tx-NM2 in 0–10 μM concentration range. (b) Dose–response curves showing the dependence of current elicited by 10 μM GABA on the concentration of toxins Tx-NM3-1 (squares) and Tx-NM2 (circles). The data are presented as Mean ± SEM (n = 3–5).
Toxins 13 00164 g005
Figure 6. Competition of N. melanoleuca toxins with 125I-αBgt for binding to Torpedo (a) and human α7 nAChR expressed in the GH4C1 cell line (b). Each data point is presented as the mean of three independent experiments ± S.E.
Figure 6. Competition of N. melanoleuca toxins with 125I-αBgt for binding to Torpedo (a) and human α7 nAChR expressed in the GH4C1 cell line (b). Each data point is presented as the mean of three independent experiments ± S.E.
Toxins 13 00164 g006
Figure 7. Competition of TFT neurotoxins with 125I-αBgt for binding to acetylcholine binding proteins (AChBPs) from L. stagnalis (a) and A. californica (b). The results of three (on L. stagnalis AChBP) or four (on A. californica AChBP) independent experiments are shown. The exception is Tx-MN4, for which two independent experiments on L. stagnalis AChBP were carried out due to insufficient amount of material for the three experiments. Each data point is presented as the mean ± S.E. NTI—long type neurotoxin I from cobra N. oxiana.
Figure 7. Competition of TFT neurotoxins with 125I-αBgt for binding to acetylcholine binding proteins (AChBPs) from L. stagnalis (a) and A. californica (b). The results of three (on L. stagnalis AChBP) or four (on A. californica AChBP) independent experiments are shown. The exception is Tx-MN4, for which two independent experiments on L. stagnalis AChBP were carried out due to insufficient amount of material for the three experiments. Each data point is presented as the mean ± S.E. NTI—long type neurotoxin I from cobra N. oxiana.
Toxins 13 00164 g007
Figure 8. Comparison of Tx-NM2 amino acid sequence with those of other long chain neurotoxins. αδ-Bgt-1 is αδ-bungarotoxin-1 from B. candidus venom (accession number A1IVR8 in UniProtKB). The amino acid residues identical to those in Tx-NM2 are underlined. The loop III fragments identical in Tx-NM3-1 and α-Ctx are marked in yellow.
Figure 8. Comparison of Tx-NM2 amino acid sequence with those of other long chain neurotoxins. αδ-Bgt-1 is αδ-bungarotoxin-1 from B. candidus venom (accession number A1IVR8 in UniProtKB). The amino acid residues identical to those in Tx-NM2 are underlined. The loop III fragments identical in Tx-NM3-1 and α-Ctx are marked in yellow.
Toxins 13 00164 g008
Figure 9. Homology modeling of N. melanoleuca neurotoxins. (a) Superposition of polypeptide chains of neurotoxins and a zoomed view of the molecular models of loop III of α-Ctx, α-Bgt, and N. melanoleuca neurotoxins. Amino acid residues differing in these TFTs are highlighted. Color code of the respective toxins is shown on the panel (c) in the headers of the respective sequences. (b) A view of the molecular model of complex formed by loop III of α-Ctx with the orthosteric site of the GABAAR. Asp56 from the α-Ctx loop III forms hydrogen bonds (shown as light blue sticks) with Arg65 and Arg171 of the α1 subunit. (c) Sequence alignment of N. melanoleuca neurotoxins with α-Ctx and α-Btx, based on alignments of spatial structures in UCSF Chimera “match-align” tool.
Figure 9. Homology modeling of N. melanoleuca neurotoxins. (a) Superposition of polypeptide chains of neurotoxins and a zoomed view of the molecular models of loop III of α-Ctx, α-Bgt, and N. melanoleuca neurotoxins. Amino acid residues differing in these TFTs are highlighted. Color code of the respective toxins is shown on the panel (c) in the headers of the respective sequences. (b) A view of the molecular model of complex formed by loop III of α-Ctx with the orthosteric site of the GABAAR. Asp56 from the α-Ctx loop III forms hydrogen bonds (shown as light blue sticks) with Arg65 and Arg171 of the α1 subunit. (c) Sequence alignment of N. melanoleuca neurotoxins with α-Ctx and α-Btx, based on alignments of spatial structures in UCSF Chimera “match-align” tool.
Toxins 13 00164 g009
Figure 10. Molecular modeling of TFTs in complexes with GABAAR. (ae) Side views of different TFT complexes with GABAAR extracellular domains; (fj) views of different TFT complexes with GABAAR from the extracellular space; (a,f) show α-Bgt complexes; (b,g) α-Ctx; (c,h) Tx-NM2; (d,i) Tx-NM3-1; (e,j) Tx-NM4. (km) Views of TFT complexes with nAChRs and AChBP from the extracellular space: (k) α-Bgt in complex with muscle nAChR (PDB 6UWZ); (l) α-Bgt in complex with α7 nAChR extracellular domain (PDB 4HQP); (m) α-Ctx in complex with AChBP (PBD 1YI5). (n) Superposition of TFT structures in complexes with nAChRs (black) and in modeled complexes with GABAAR (red). Toxins in complexes with GABAAR show tilt toward complementary subunit. (o) Comparison of loop II positions in different models (color code is the same as on the Figure 9). Position of loop C in the cryo-EM structure of the GABAAR with the non-peptide antagonist bicuculline (PDB 6HUK) is shown in black.
Figure 10. Molecular modeling of TFTs in complexes with GABAAR. (ae) Side views of different TFT complexes with GABAAR extracellular domains; (fj) views of different TFT complexes with GABAAR from the extracellular space; (a,f) show α-Bgt complexes; (b,g) α-Ctx; (c,h) Tx-NM2; (d,i) Tx-NM3-1; (e,j) Tx-NM4. (km) Views of TFT complexes with nAChRs and AChBP from the extracellular space: (k) α-Bgt in complex with muscle nAChR (PDB 6UWZ); (l) α-Bgt in complex with α7 nAChR extracellular domain (PDB 4HQP); (m) α-Ctx in complex with AChBP (PBD 1YI5). (n) Superposition of TFT structures in complexes with nAChRs (black) and in modeled complexes with GABAAR (red). Toxins in complexes with GABAAR show tilt toward complementary subunit. (o) Comparison of loop II positions in different models (color code is the same as on the Figure 9). Position of loop C in the cryo-EM structure of the GABAAR with the non-peptide antagonist bicuculline (PDB 6HUK) is shown in black.
Toxins 13 00164 g010
Table 1. Molecular masses of N. melanoleuca toxins.
Table 1. Molecular masses of N. melanoleuca toxins.
ToxinDetermined Molecular Mass, DaCalculated Molecular Mass, Da
Tx-NM28030.688030.69
Tx-NM3-17787.5777787.602
Tx-NM3-27441.497Not calculated
Tx-NM47756.5817756.602
Table 2. IC50 values for inhibition of currents induced by GABA in α1β3γ2 GABAAR by N. melanoleuca toxins.
Table 2. IC50 values for inhibition of currents induced by GABA in α1β3γ2 GABAAR by N. melanoleuca toxins.
ToxinIC50, µM
Tx-NM3-10.68 ± 0.14
Tx-NM21.25 ± 0.07
Tx-NM4≈10
Tx-NM3-2>>10
Table 3. IC50 values and Hill coefficients (nH) for inhibition of 125I-αBgt binding to Torpedo and human α7 nAChR by N. melanoleuca toxins.
Table 3. IC50 values and Hill coefficients (nH) for inhibition of 125I-αBgt binding to Torpedo and human α7 nAChR by N. melanoleuca toxins.
ToxinIC50, nM (CI 95% 1); nH ± S.E.
Torpedo nAChRα7 nAChR
Tx-NM21.01 (0.89–1.13); 1.92 ± 0.23 and 8.66 (8.31–9.04); 0.67 ± 0.05 29.47 (8.56–10.49); 2.08 ± 0.31
Tx-NM3-11.61 (1.45–1.79); 2.88 ± 0.224.84 (4.36–5.36); 1.81 ± 0.12
Tx-NM42.17 (1.73–2.74); 2.5 ± 0.3326.9 (25.84–27.98); 0.96 ± 0.01
1 CI 95%—95% confidence interval, 2 Two-site binding model.
Table 4. IC50 values for inhibition of 125I-αBgt binding to AChBPs by N. melanoleuca and other known TFT toxins.
Table 4. IC50 values for inhibition of 125I-αBgt binding to AChBPs by N. melanoleuca and other known TFT toxins.
ToxinIC50, nM
AChBP L. stagnalisAChBP A. californica
Tx-NM214.1 ± 0.168.6 ± 3.6
Tx-NM3-1203 ± 103200 ± 600
Tx-NM41160 ± 402350 ± 450
α-Bgt17.2 ± 0.7155 ± 35
α-Ctx25.4 ± 0.8385 ± 45
NT-I 173.4 ± 1.31200 ± 300
1 Long type neurotoxin I from N. oxiana.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Son, L.; Kryukova, E.; Ziganshin, R.; Andreeva, T.; Kudryavtsev, D.; Kasheverov, I.; Tsetlin, V.; Utkin, Y. Novel Three-Finger Neurotoxins from Naja melanoleuca Cobra Venom Interact with GABAA and Nicotinic Acetylcholine Receptors. Toxins 2021, 13, 164. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins13020164

AMA Style

Son L, Kryukova E, Ziganshin R, Andreeva T, Kudryavtsev D, Kasheverov I, Tsetlin V, Utkin Y. Novel Three-Finger Neurotoxins from Naja melanoleuca Cobra Venom Interact with GABAA and Nicotinic Acetylcholine Receptors. Toxins. 2021; 13(2):164. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins13020164

Chicago/Turabian Style

Son, Lina, Elena Kryukova, Rustam Ziganshin, Tatyana Andreeva, Denis Kudryavtsev, Igor Kasheverov, Victor Tsetlin, and Yuri Utkin. 2021. "Novel Three-Finger Neurotoxins from Naja melanoleuca Cobra Venom Interact with GABAA and Nicotinic Acetylcholine Receptors" Toxins 13, no. 2: 164. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins13020164

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