Next Article in Journal
A Comprehensive Genome-Wide Association Study of Carotenoid and Capsaicinoid Contents in Capsicum chinense Germplasm
Previous Article in Journal
Genomic and Phenotypic Characterization of Mastitis-Causing Staphylococci and Probiotic Lactic Acid Bacteria Isolated from Raw Sheep’s Milk
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fifty Years of Animal Toxin Research at the Shemyakin–Ovchinnikov Institute of Bioorganic Chemistry RAS

1
Department of Molecular Neuroimmune Signaling, Shemyakin-Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences, 16/10 Miklukho-Maklay Str., 117997 Moscow, Russia
2
Department of Molecular Neurobiology, Shemyakin-Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences, 16/10 Miklukho-Maklay Str., 117997 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(18), 13884; https://doi.org/10.3390/ijms241813884
Submission received: 18 August 2023 / Revised: 4 September 2023 / Accepted: 6 September 2023 / Published: 9 September 2023

Abstract

:
This review covers briefly the work carried out at our institute (IBCh), in many cases in collaboration with other Russian and foreign laboratories, for the last 50 years. It discusses the discoveries and studies of various animal toxins, including protein and peptide neurotoxins acting on the nicotinic acetylcholine receptors (nAChRs) and on other ion channels. Among the achievements are the determination of the primary structures of the α-bungarotoxin-like three-finger toxins (TFTs), covalently bound dimeric TFTs, glycosylated cytotoxin, inhibitory cystine knot toxins (ICK), modular ICKs, and such giant molecules as latrotoxins and peptide neurotoxins from the snake, as well as from other animal venoms. For a number of toxins, spatial structures were determined, mostly by 1H-NMR spectroscopy. Using this method in combination with molecular modeling, the molecular mechanisms of the interactions of several toxins with lipid membranes were established. In more detail are presented the results of recent years, among which are the discovery of α-bungarotoxin analogs distinguishing the two binding sites in the muscle-type nAChR, long-chain α-neurotoxins interacting with α9α10 nAChRs and with GABA-A receptors, and the strong antiviral effects of dimeric phospholipases A2. A summary of the toxins obtained from arthropod venoms includes only highly cited works describing the molecules’ success story, which is associated with IBCh. In marine animals, versatile toxins in terms of structure and molecular targets were discovered, and careful work on α-conotoxins differing in specificity for individual nAChR subtypes gave information about their binding sites.

1. Introduction

Animal venoms, over a long period of evolution, have become a powerful means of attack and defense. They consist mostly of proteins and peptides, usually called toxins. The best-known terrestrial venomous animals are snakes, scorpions, and spiders. Some invertebrate marine animals like Conus snails, jellyfishes, and sea anemones are also very successful in toxin production. For many centuries, the deadly bites and stings of these animals have been attracting the attention of naturalists and natural medicine specialists. However, it was only in the last century that the study of venoms at the molecular level began. The individual toxins were isolated from venoms and their structures and functions were investigated. Due to their high affinity and selectivity towards various receptor targets, animal toxins are widely used as versatile tools for the study of different physiological processes at the molecular level. Among many others, especially potent was α-bungarotoxin (αBgt), a neurotoxin isolated 60 years ago from the venom of krait Bungarus multicinctus [1], which was used as a first tool in the studies of nicotinic acetylcholine receptor (nAChR) [2]. In the current jubilee Special Issue, we shall consider the research on toxins and receptors performed at the Shemyakin–Ovchinnikov Institute of bioorganic chemistry of the Russian Academy of sciences (IBCh) in Moscow, which started approximately 10 years after the discovery of αBgt—which explains the title of our review. We will discuss various proteins from natural venoms, including those that, like αBgt, share a three-finger spatial structure and interact with nAChRs. The studies of protein and peptide toxins from venomous animals carried out at IBCh will be accompanied by those of their synthetic analogs.

2. Structural Studies of Animal Toxins

2.1. Primary Structure Determination

The first results of the work on the protein toxins at IBCh were the determination of the primary structures of several toxins from the Middle Asian cobra Naja oxiana. Among them were α-neurotoxins, namely short-type neurotoxin II and long-type neurotoxin I [3,4], as well as cytotoxins [5,6] (Figure 1). It should be noted that, in those days, there was no such reliable assistant as modern mass-spectrometry to check the purity and to determine the structure of the obtained peptide or protein: the isolation of an individual toxin required a combination of various chromatography techniques, while Edman degradation was in fact the sole method for primary structure determination. The use of mass-spectrometry greatly facilitated this task and, for example, allowed the detection of the first tryptophan-containing “weak” three-finger toxin (TFT) WTX in Naja kaouthia venom [7] (Figure 1). Unusual forms of snake venom toxins were found at our institute considerably later, and among them were muscarinic toxin-like proteins [8] (Figure 1) and glycosylated cytotoxin [9], the first representative of glycosylated TFTs. All the above-mentioned toxins are from the TFT family. However, snake toxins from other families were discovered as well. They included heterodimeric neurotoxic phospholipases A2 (PLA2) from the venom of viper Vipera nikolskii [10], as well as azemiopsin, a peptide neurotoxin of the new type [11], and original bradykinin-potentiating peptides from viper Azemiops feae [12] (Figure 1).
At the same time, intensive experiments were carried out to study toxic molecules in arthropod venoms. Due to the limitations of the biological tests available, the search for active molecules has mainly been based on toxicity to mammals and/or insect species. Fifteen different toxins from Buthus eupeus scorpion venom, among them insectotoxins I1, I3, I4, and I5, belonging to yet unknown structural type [13], and three polypeptides, M10, M14, and M9, toxic to mammals, were isolated [14] (Figure 1). The complete amino acid sequence of the 66-amino-acid-residue toxins M9 and M14 was established by detecting overlapping peptide fragments using Edman degradation [14]. Another four polypeptide neurotoxins, possessing paralytic activity towards mice, were isolated from the Orthochirus scrobiculosus scorpion’s venom, and, for the major compound Os-3, after tryptic and chymotryptic cleavage, the complete amino acid sequence was established [15].
The development of molecular biology methods (cloning and sequencing of DNA) and mass-spectrometry made protein sequence determination more technologically advanced and productive. With these approaches, many more toxin sequences were determined, including those of phospholipases A2 from snake venoms [10,16], as well as other classes of toxins from scorpion and spider venoms. In particular, the isolation of the toxin named OsI-1, with a molecular weight of 6994 Da, responsible for the O. scrobiculosus scorpion venom’s toxicity to insects, and its partial N-terminal sequence have served as the starting point for gene cloning. The complete primary structure of OsI-1 (Figure 1) was deduced from the cDNA sequence obtained by the rapid amplification of cDNA ends (RACE) method. The mass spectroscopy data indicated the post-translational modification of the precursor protein at maturation by three C-terminal amino acids’ truncation and the amidation of the C-terminus [17].
Other examples are shorter polypeptide toxins BeKm-1 and OSK1 from scorpion venoms (Figure 1); using the same approach, their structures were derived from the nucleotide sequence of the mRNA encoding the complete precursor protein. Further, the examination of BeKm-1’s effects on different Kv channels revealed its selective inhibitory effect on hERG1 ion channels with an IC50 of 3.3 nM, which made it the first selective inhibitor of this channel [18]. In contrast, the OSK1 inhibitory activity had a lack of selectivity: determined initially as small-conductance Ca-activated K+-channels inhibitor [19], this toxin blocked the conductance of voltage-gated potassium channels, the intermediate conductance of calcium-activated potassium channels, and even the mouse muscle nAChR in the micromolar concentration range [20].
This approach also made it possible to determine the complete primary structures of several new toxic compounds found in the venom of sea anemones. Among the most recent was the identification of a series of peptides (Ms11a-1/4) from Metridium senile capable of inhibiting nAChR (Figure 1), three of which demonstrated sub-micromolar affinity for the muscle-type receptor, and one (Ms11a-3) showed maximum efficacy of interaction with α9α10 nAChR [21].
The transition of primary structure determination for toxins from direct sequencing by Edman degradation to mRNA data decoding revealed the heterogeneity of the mRNA in arachnids long before the discovery of gene multiplicity in their genomes. The study of the structural variability in the Os3 neurotoxin from O. scrobiculosus revealed the family of cDNAs that encode the highly homologous Os3-like polypeptides [22]. The further development of techniques and cost reductions for nucleic acid sequencing made it possible to analyze the mRNAs’ multiplicity in the venom glands and the creation of the so-called expressed sequence tag (EST) library.
A pipeline for the processing of large data collected from EST was essential in obtaining the structural information. In IBCh, such an approach for the primary structure of mature toxin search was suggested for spiders’ toxins and further expanded to other venomous species [23,24]. Thus, a comprehensive transcriptome analysis of the EST database was performed for spiders Agelena orientalis and Dolomedes fimbriatus [23,25]. For A. orientalis, by the proteomic analysis of venoms milked individually from a single spider, a lower number of polypeptides was found in each sample than was expected from the transcriptome analysis. However, it was found that the composition of the venom obtained by summing the proteomic data for 20 individual animals was similar in composition to that from the transcriptomics data, but each spider individually produced a limited pool of toxins from the general assortment [26].
Special mention can be made of the discovery of the largest animal toxins named latrotoxins. Dangerous to humans, the venom of the black widow spider Latrodectus mactans tredecimguttatus contains high-molecular-weight toxins (above 150 kDa). The construction of a cDNA library for spider venom glands led to the determination of the primary structure of the 1401-amino-acid α-latrotoxin precursor, from which the first toxin sequence was derived [27]. Further, α-latroinsectotoxin, a presynaptic neurotoxin selective only for insects and composed of 1411 amino acids, was sequenced. It was shown that the mature α-latroinsectotoxin with a molecular mass of approximately 130 kDa was produced by processing in the N-terminal and C-terminal regions of the precursor [28]. A few years later, another insect-specific toxin, δ-latroinsectotoxin, that contained 1214 amino acid residues, was sequenced [29] (Figure 1). The domain organization of all latrotoxins was found to be similar, suggesting that the toxins are a family of related proteins. α-Latrotoxin has made an invaluable contribution to the development of knowledge about the mechanism of synaptic vesicle exocytosis in presynaptic nerve terminals. It is capable of stimulating exocytosis, via binding to two distinct families of neuronal cell surface receptors, neurexins and latrophilins, which share an important function in synaptic cell adhesion [30].

2.2. Toxin Spatial Structure Studies

In the early 1970s, there was no information about the spatial structure of snake venom neurotoxins. At that time, at IBCh, the laboratory of nuclear magnetic resonance was led by well-known biophysicist Prof. V. Bystrov; the dependence of the vicinal coupling constant in peptide 1H-NMR on the dihedral angle is known as the Karplus–Bystrov relationship. It was hoped that some information about the spatial structure of α-neurotoxins could be obtained by the 1H-NMR method, while the chemically modified derivatives of α-neurotoxins would assist the signal assignment in the spectra. It was established that the central loop II and the C-terminal loop III of the short α-neurotoxin NT-II from the N. oxiana venom were in proximity [31] (Figure 2A), but, at that time, 1H-NMR could not solve the total structure even in relatively small proteins. However, this conclusion was in accordance with the complete structure of erabutoxins, the short-chain α-neurotoxins, which was established at that time in the USA by X-ray analysis [32,33].
A positive circumstance that contributed to determining a large number of NMR structures of various TFTs at IBCh was the development in the department led by Prof. M. Kirpichnikov of efficient methods for the expression of naturally occurring neurotoxins and their mutants. Among such proteins were short- and long-chain α-neurotoxins and non-conventional toxins [34,35]. Improvements in NMR methods, as well as the possibility of obtaining recombinant toxins and their mutants, led to the determination of a number of new structures at the department led by Prof. A. Arseniev—for example, the complete structures of naturally occurring neurotoxin II and the non-conventional WTX mutant [36,37] (Figure 2A,B).
In collaboration with Dutch colleagues, the X-ray structure of an unusual dimeric form of α-cobratoxin, where two long-type TFT monomers were connected by two intermolecular disulfides, was determined [38,39] (Figure 2C).
The development of NMR instrumentation allowed the determination of high-resolution structures of peptide and protein neurotoxins. Thus, the first 3D structure of neuronal α-conotoxin ImI in solution was determined [40], and recently the spatial structure of peptide neurotoxin apamin was confirmed [41] (see Figure 2).
For a number of peptide toxins from scorpion venoms, the spatial structures were also resolved by the 1H-NMR technique at IBCh. These include OSK1 from the O. scrobiculosus venom, mentioned above as a nonselective blocker of small-conductance Ca-activated K+ channels [42] (Figure 2D); mammalian neurotoxin M9 from the B. eupeus venom, which was shown to have two slowly exchangeable conformations at acidic pH [43]; the insectotoxin I5A from the B. eupeus containing the α-helical and anti-parallel β-structure [44]; and potassium channel inhibitor BeKm-1 from B. eupeus, which also comprised a short α-helix and a triple-stranded antiparallel β-sheet [45]. The obtained BeKm-1 structure allowed the initiation of the design of point-mutated analogs, followed by the identification of residues important for the binding of BeKm-1 to the human ERG (HERG) potassium channel. Their unusual location in the α-helix and following loop was established, contrary to the β-sheet region in “traditional” scorpion toxins [45].
Several 3D structures of arachnid toxins were determined by NMR as well. Among them, the structure of a modular arachnid toxin, purotoxin-2 (PT2), from the wolf spider Alopecosa marikovskyi (Lycosidae) should be mentioned [46] (Figure 2F). This toxin, which comprises an N-terminal inhibitor cystine knot (ICK) β-structural domain and a C-terminal linear cationic domain, is the founding member of a large family of polypeptides with similar structural motifs.
A non-standard variant of ICK folding was detected by an NMR study of the nAChR-targeting peptides Ms11-2 and Ms11a-3, identified in M. senile sea anemone venom (see above). A structural feature of both compounds was the presence of a prolonged loop between the fifth and sixth cysteines [21] (Figure 2E).

3. Three-Finger Proteins from Snake Venoms in Research on nAChRs

3.1. Overview of α-Neurotoxin–nAChR Relationships

nAChR was the first ion channel receptor to be characterized in detail, and a huge role in this was played by the discovery of αBgt [1], a sophisticated tool for the detection and further detailed investigation of this receptor. Soon, it became clear that αBgt is a member of a large family of diverse protein neurotoxins present in the venoms of different snakes, which may have higher or lower selectivity to different nAChR subtypes. Originally, αBgt was used to block the activity of the muscle-type nAChR from the fish electric organs [2] and, later, radioactive and fluorescent derivatives of αBgt played a role as excellent tools in detecting and measuring the levels of the nAChRs. In particular, with the aid of radioiodinated αBgt, the presence of nAChR, later identified as neuronal α7 nAChR, was shown in the brain [47]. Chromatography on an affinity column with α-cobratoxin (belonging, similarly to αBgt, to the group of the so-called long-chain α-neurotoxins) was one of the methods of nAChR purification [48]. Current knowledge of the nAChR structure and function, including the mechanisms of the toxin–receptor interactions, is presented in textbooks and recent reviews [49,50,51]. Elucidation of the nAChR three-dimensional structure took a long way: from the nAChR low-resolution X-ray structure [52] to the high-resolution X-ray structure of the acetylcholine-binding proteins (AChBPs) [53], an excellent structural model of the ligand-binding domains not only nAChRs but of all Cys-loop ligand-gated ion channels. This important step was followed by the first refined cryo-electron microscopy structure of the Torpedo marmorata ray nAChR [54] and then by high-resolution X-ray structures of microbial proteins GLIC and ELIC [55,56]. The next successes were the high-resolution structures of a couple of neuronal heteromeric nAChRs (on which α-neurotoxins are not acting or acting with low affinity) in complexes with low-molecular agonists and antagonists (see, for example, [57]). Regarding α-neurotoxins, until recently, their complexes with either the AChBPs (built from five identical subunits) [58], pentameric chimeric ligand-binding domain (LBD) of the α7 nAChR subunit [59], or monomeric LBDs of the α1 [60] and α9 [61] nAChR subunits were known. Only in the last two years, cryo-electron microscopy provided the high-resolution structures of αBgt in complex with the muscle-type Torpedo receptor [62] and with the human neuronal α7 nAChR [63], as well as a complex of a short-chain α-neurotoxin with the Torpedo nAChR [64]. This briefly presented information reflects the current state of the α-neurotoxin–nAChR relationships.

3.2. Interaction of α-Neurotoxins Bearing Fluorescent, Spin, or Photoactivatable Labels with the nAChRs

The relatively large amounts of isolated snake venom proteins available at IBCh made possible the preparation of their chemically modified derivatives to obtain information about their spatial structures and for the analysis of the mechanisms of action. Starting from neurotoxin II, a series of singly labeled derivatives containing a fluorescent or spin label at the identified amino acid residues were obtained and used to map the topography of the neurotoxin binding to the T. marmorata nAChR. With the aid of fluorescence and EPR spectroscopy, several labels forming contacts with the receptor were identified, including those on Lys27 in the central loop II and on Lys47 in the C-terminal loop III [65]. In most of the mentioned X-ray and cryo-electron microscopy structures, the toxin was a long-chain α-neurotoxin (α-cobratoxin or α-bungarotoxin), and the major role in contacts with the receptor or its models was played by the toxin central loop II.
Later, in collaboration with Prof. F. Hucho (Free University, Berlin), a series of photoactivatable derivatives of short- and long-chain α-neurotoxins was synthesized and their contacts with various nAChR subunits of the Torpedo californica were characterized [66,67,68]. The results obtained in these experiments also revealed the predominant participation of the toxins’ central loop II in the interaction with nAChR, and, in the case of short-chain neurotoxins, also loop III, which was fully confirmed after publishing the cryo-electron structures of complexes of the Torpedo receptor with both long-chain [62,63] and short-chain [64] neurotoxins.

3.3. Recent Studies on Snake Venom α-Neurotoxins

More recent work on snake venom neurotoxins involved the demonstration of α-cobratoxin’s capacity to block some subtypes of the ionotropic GABA-A receptors [69]. Using, in competition with α-cobratoxin, its mutant and synthetic fragments provided evidence for the involvement of the α-cobratoxin central loop II in binding to the GABA-A receptor. This conclusion is in agreement with the recently solved cryo-electron microscopy structure of the GABA-A receptor complex with α-cobratoxin [70]. The search for toxins acting both on the nAChRs and GABA-A receptor continued and several such novel proteins were isolated from the venom of cobra Naja melanoleuca [71]. Concerning α-bungarotoxin, whose anniversary is celebrated in the present Special Issue, a collaboration with German colleagues should be mentioned, where several new variants of α-bungarotoxin have been isolated, which, contrary to α-bungarotoxin itself, have a different affinity for the two orthosteric binding sites in the T. californica nAChR [72].
At present, a great role among the fluorescent labels is played by green fluorescent protein (GFP) and related proteins. At IBCh, GFP was recombinantly fused with α-cobratoxin for research on the nAChR. However, the results were unsatisfactory and a much higher resolution in the detection of α7 and muscle-type nAChRs in the cell lines was achieved when α-cobratoxin was chemically modified not with full-size GFP but only using its synthesized chromophore [73].

4. Other Snake Venom Toxins

4.1. Cobra Venom Cytotoxins: Spatial Structure and Biological Activity

Cytotoxins (cardiotoxins) belong to the TFT family and are found exclusively in cobra venoms. These toxins nonspecifically kill cells by attacking and destroying the cell membrane. A few cytotoxins from cobra N. oxiana venom were purified and sequenced at IBCh in the early seventies of the last century [5,6] (Figure 1). Further studies were directed to the determination of the detailed spatial structures of the cytotoxins and the molecular mechanisms of their biological effects. Using different variants of NMR spectroscopy, the spatial structures of cytotoxins 1 and 2 in solution and in the membrane environment were determined [74,75,76]. The differences in the mode of interaction with the membrane between different cytotoxin types were revealed and the important role of polypeptide loops I and II was recently demonstrated [77]. The first, and so far unique, among three-fingered toxins, glycosylated cytotoxin was isolated from cobra N. kaouthia venom, where the glycosylation of Asn29 in loop II resulted in a decrease in its biological activity [9]. It was found that the cytotoxicity depended on the structure of loops I and II, the toxins with two adjacent proline residues in loop I being the least active [78,79]. Using the confocal spectral imaging technique and fluorescently labeled cytotoxins, it was shown that cytotoxins penetrated readily into cells and accumulated markedly in the lysosomes, and this accumulation correlated well with the cytotoxic effects [79].
The biological activity of cytotoxins is quite diverse. Thus, they have been shown to manifest antiprotozoal activity [80]. The cardiotoxicity of cytotoxins is well known. However, recent works at IBCh, in collaboration with other institutes of RAS, showed that this activity is also dependent on the structural types of cardiotoxins [81,82]. These differences were observed in the whole heart [82], isolated muscle and aorta [81], as well as in cardiomyocytes [83]. It was shown that the effect of cytotoxins on the cardiovascular system was mainly associated with the disruption of the transport systems responsible for Ca2+ influx [81].

4.2. Non-Conventional Three-Finger Neurotoxin

So-called “weak” snake venom toxins were known long ago and obtained this name because they were virtually nontoxic and their targets were not known. Contrary to long-chain α-neurotoxins, these proteins have an additional disulfide not in loop II, but in loop I. One of these toxins—weak toxin (WTX)—was isolated at IBCh [7] (Figure 1 and Figure 2B), and, in collaboration with Prof. D. Bertrand (Switzerland) and Dr. C. Methfessel (Germany), for the first time, it was demonstrated that WTX at micromolar concentrations inhibited both the muscle-type and neuronal α7 nAChRs [84]. The toxins of this type are now called non-conventional toxins. Further studies demonstrated that among WTX targets can also be some subtypes of muscarinic acetylcholine receptors [37,85]. Because WTX had also effects on blood pressure [86] and was almost nontoxic, it can be considered as a basis for a potential drug.

4.3. Three-Finger Mammalian Proteins

Three-finger proteins are also present in diverse organisms, including mammals, forming the so-called Ly6/uPAR family (see reviews [87,88]). They are mentioned here, close to WTX, because, structurally, they are most similar to nonconventional toxins having the additional fifth disulfide in the N-terminal loop I, but not in the central loop II, like the long-chain α-neurotoxins (see Figure 2B,C). Most of the Ly6/uPAR proteins, like Lynx1, have the glycosylphosphatidylinositol (GPI) anchor at their C-terminus, by which they attach to the membrane, but secreted proteins also exist, like SLURP1, lacking this anchor. Although these proteins do not belong to the animal toxins considered in the present review, we will present brief information about those that act on nAChRs. The reason for their inclusion is that the work at IBCh on the 1H-NMR spatial structures of Lynx1 and SLURP1 analogs, expressed in E. coli, was based on previous work both on the naturally occurring TFTs (see Section 2.2) and those expressed in E. coli. The experience in the production of TFTs in E. coli and their 1H-NMR analysis in the departments of Prof. A. Arseniev and Prof. M. Kirpichnikov allowed, for the first time, the establishment of the structures of SLURP-1 and wsLynx1 (water-soluble Lynx1 analog lacking the GPI tail) [89,90]. In the present review, focusing on toxins from animal venoms, we do not consider in detail the interactions of these proteins with the muscle-type and neuronal nAChRs, but we should mention that an analysis of their activity was performed using the two-electrode voltage-clamp method and by competition with radioiodinated α-bungarotoxin, which allowed to distinguish their interactions with the orthosteric and allosteric binding sites in these receptors [89,91,92].

4.4. Phospholipases A2

The first phospholipase A2 (PLA2) was isolated from the cobra N. oxiana’s venom at IBCh in 1977 [93]; later, its complete amino acid sequence was determined [94]. The determination of the amino acid sequences for PLA2s from other snake venoms was mentioned in Section 2.1 [7,8] (see Figure 1). The studies of PLA2s’ biological activity discovered new properties. It has been shown that the noncytotoxic PLA2 from the N. haje cobra venom is the first thrombin inhibitor of the PLA2 family [95]. PLA2s were found to inhibit nAChRs and this capability did not depend on their enzymatic activity [96,97]. Phospholipase A2 from krait Bungarus fasciatus venom manifested cytotoxicity against human cancer cells in vitro [98]. Recently, the antiviral activity of several snake venom PLA2s was studied. It was found that dimeric PLA2 from the viper Vipera nikolskii and its subunits demonstrated potent virucidal effects, which were related to their phospholipolytic activity and interfered with the binding both of an antibody against angiotensin-converting enzyme 2 (ACE2) and of the receptor-binding domain of the SARS-CoV-2 virus’ glycoprotein S to 293T/ACE2 cells [99]. Dimeric PLA2s had also pronounced virucidal and anti-HIV activity. They inhibited syncytium formation between chronically HIV-infected cells and healthy CD4-positive cells and blocked HIV binding to cells [100], this action being dependent on their catalytic activity. Thus, snake venom PLA2s might be considered as candidates for lead molecules in antiviral drug development.

4.5. Linear Peptides

Finishing this section dealing with snake toxins, it is worth mentioning azemiopsin—a peptide that was isolated at IBCh from the Azemiops feae venom [11] (Figure 1). It does not contain disulfides but quite selectively inhibits the muscle-type nAChRs [11] and, according to preclinical studies, is an efficient myorelaxant [101]. Two new bradykinin-potentiating peptides possessing unique amino acid sequences were isolated from the venom of A. feae [12] (see Figure 1). They have no consensus C-terminal sequence PPIPP but efficiently potentiate the effect of bradykinin.
It should be mentioned that linear toxins (lacking disulfide bonds) were also characterized in the venoms of some spider species. Particularly large content of such peptides having large structural variability was found in the venom of the Lachesana tarabaevi spider and these toxins were named latarcins [102]. Further studies of the spider superfamily Lycosoidea revealed other numerous linear toxins in the venoms with predominantly membrane-destroying activity and cytotoxicity to bacterial and partially to mammalian cells. The mechanisms of interaction with cellular membranes and selectivity to the composition of model lipid membranes were later established for these molecules and described in detail [103]. From the latarcin pool, latarcin 2a (see Figure 1) was the most active among others and had higher cytotoxicity to bacteria as well as to red blood cells.

5. Marine Toxins Acting on the Ligand- or Voltage-Gated Ion Channels

5.1. α-Conotoxins in Research on nAChRs

α-Conotoxins, small neurotoxic peptides with two disulfide bridges found in venoms of the predatory marine mollusks of the Conus genera, emerged in research on nAChRs more than 40 years ago [104,105], and their advantage is their capacity not only to distinguish muscle-type nAChRs from neuronal ones but also to be markers of individual neuronal nAChRs (see, for example, reviews [106,107,108]). Earlier work at IBCh, starting in the 1990s, on the synthesis and analysis of the spatial structures of α-conotoxins [40,109,110] (Figure 2H) was performed in collaboration with Dr. C. Methfessel (Germany). Later, using photoactivatable derivatives of α-conotoxins, their contacts with various subunits of the T. californica nAChR were demonstrated [111,112,113], resulting in the mapping of the orthosteric binding sites of this receptor [113]. The targeted design of the analogs of α-conotoxins has shown that the introduction of additional positive charges in many cases increases the affinity of α-conotoxins for both the muscle-type and neuronal nAChRs [114,115,116]. Naturally occurring α-conotoxins or their designed analogs were used to identify the respective nAChR subtypes in different biological objects—for example, on the neurons of Lymnaea stagnalis [117] or Turkish snail [118]—as well as to characterize the functional role of distinct receptors in these objects. The most extensive research here has been carried out on various immune cells [119,120,121,122] and cancer cell lines [123,124,125].
In collaboration with Profs. A. Smit and T. Sixma from the Netherlands, the first crystal structures of α-conotoxins were determined in complex with the AChBP from Aplysia californica [126,127] (Figure 3A,B). An interesting finding was that α-conotoxins, antagonists of nAChRs, induce an outward shift of AChBP loop C in the orthosteric binding site, while agonists shift it to the AChBP center. At present, this line of work is being continued with Chinese colleagues (Prof. S. Luo) by analyzing the AChBP structures in complex with conotoxins selective towards different subtypes of neuronal nAChRs [128,129,130]. In collaboration with colleagues from Greece, for the first time, the X-ray structure has been determined for the α-conotoxin RgIA, selective for the α9α10 nAChR, in complex with α9 LBD [131] (Figure 3C,D).
The α-conotoxin RgIA analogs inhibiting the α9α10 nAChR are considered as potential drug candidates against the neuropathic pain [132,133]. In view of the important role of Arg residues in RgIA and some other “analgesic” conotoxins, at IBCh, we synthesized a series of oligoarginines (earlier used for the intracellular delivery of various attached compounds) and they were demonstrated to be a new class of inhibitors of distinct nAChR subtypes (including α9α10) [134]. In particular, octa-oligoarginine in a mouse model was shown to be as effective as α-conotoxin RgIA against oxaliplatin-induced neuropathy [135].

5.2. Sea Anemone-Derived Toxins

Several toxins belonging to the APETX-like family with ASIC and/or nAChR inhibitory/potentiating activity have been isolated from H. crispa sea anemones collected in the natural environment. Particular interest was shown in two homologous peptides, Hmg 1b-2 and Hmg 1b-4: both molecules have analgesic activity, similarly to diclofenac in the model of acid-induced muscle pain and acute local inflammation. However, Hmg 1b-4’s analgesic effect was more pronounced and statistically significant; in addition, this compound showed an anxiolytic effect on mouse mice behavior in the open field test [136]. The Hmg 1b-2 and Hmg 1b-4 (originally named Hcr 1b-4) peptides differed in their selectivity to ion channels: the second one has been found to be a potentiator of ASIC3 and inhibitor of ASIC1a at similar concentrations—EC50 1.53 ± 0.07 μM and IC50 1.25 ± 0.04 μM [137]. Hmg 1b-2 (originally named Hcr 1b-2) was an inhibitor of both ASIC3 with IC50 15.9 μM and ASIC1a with IC50 4.8 ± 0.3 μM, but at a concentration of 1 µM, it potentiated ACh-elicited currents of both α7 and α1β1δε nAChRs [137,138,139]. Then, a few novel (see Section 2) nAChR-targeting peptides from the Metridium senile sea anemone were discovered at IBCh and structurally and functionally characterized in collaboration with Chinese and Norwegian colleagues [21] (Figure 1 and Figure 2E).
When studying sea anemone toxins, much attention was paid to different types of receptors. Thus, inhibitors of such important ion channels as TRPV1 [140,141], TRPA1 [142,143], and ASICs [137,138,144] were discovered at IBCh. Some of the found peptides potently inhibit the mammalian nociceptive sensors, produce analgesia in animal models [136,145,146], and may pave the way to novel pain killers.

6. Prominent Molecules from Arthropod Venoms

Among venom-producing terrestrial animals, arthropods are unambiguous leaders both in the variety of structural folds and in the number of unique structures. First of all, there are spiders, whose venom can contain more than a hundred components. We should note that many potent and well-known molecules (in addition to the latrotoxins mentioned above) have been discovered for basic research within the walls of IBCh.
Argiopin, a non-peptide toxin from the spider Argiope lobata’s venom, opened the specific structural class of molecules producing a blocking effect on glutamate receptors participating in synaptic transmission [147]. Further, the number of the active molecules was extended by other compounds from A. lobata—argiopine, argiopinines, and pseudoargiopinines—which could block the same receptors, but their effects on the receptor channels were different from those of argiopin [148]. The latter (often called argiotoxin AgTx-636) was used with the electrophysiology and structural biology techniques to elucidate the mechanisms by which small-molecule blockers selectively inhibit ion channel conductance in calcium-permeable AMPA receptors [149] (Figure 3E). The other well-known bee venom toxin apamin, which is often cited as selective to small-conductance Ca-activated potassium channels (KCa2), was tested against 42 ion channels, KCa, KV, NaV, nAChR, ASIC, and others. As a result, its unique selectivity to KCa2 channels was confirmed [41].
In the venom of the Central Asian spider Geolycosa sp., the first molecule exerting powerful and selective inhibitory action on P2X3 receptors was discovered and named purotoxin-1 (PT1). Since P2X3 purinoreceptors expressed in mammalian sensory neurons play a key role in several processes, including pain perception, the peptide PT1 started a marathon of P2X inhibitor development with the aim of drug design. Animals treated by such peptides demonstrated anti-nociceptive behaviors in models of inflammatory pain [150]. In general, toxin use for drug development is an ongoing trend in the toxinology community. In this regard, interesting data have been obtained for the inhibition by toxins of the skeletal muscle voltage-gated sodium channel NaV1.4 subtype, mutation in which is related to the congenital disease of hypokalemic periodic paralysis. Gating modifier toxins from the Heriaeus melloteei spider can target the mutated channel by modifying the function of the voltage sensing domain, which results in the reduction of the pathogenic gating pore currents through NaV1.4 [151].
From the venom of scorpion O. scrobiculosus, a toxin named OSK2 has been purified and shown to be a potent and selective blocker of Kv1.2 channels (Kd 5.97 nM) [152]. Another ligand, specific to Kv1.2 channels, was purified from B. eupeus venom and named MeKTx11-1 (α-KTx 1.16). It had the highest affinity to KV1.2 (IC50 ∼0.2 nM), while its activity against Kv1.1, Kv1.3, and Kv1.6 was 10,000-, 330-, and 45,000-fold lower, respectively, as measured using the voltage-clamp technique on mammalian channels expressed in Xenopus oocytes [153]. Two other scorpion toxins were successfully used to obtain fluorescent tools to study potassium channels. In particular, the chimeric molecules for the toxin construction approach and full-size fluorescent protein resulted in obtaining the potassium channel blockers eGFP-OSK1 and RFP-AgTx2, which show inhibitory activity and the possibility of visualization as well [154].
For potential-activated sodium channels, a group of toxins with a principally new mechanism of action was found in the venom of spider A. orientalis. These toxins, consisting of 36–38 amino acid residues, were named β/δ-agatoxins and could modulate the insect NaV channel (DmNaV1/tipE) in a unique combination. They shifted the voltage dependence of the channels’ activation towards more hyperpolarized potentials (like site 4 toxins) and induced a non-inactivating persistent Na current (site 3-like action) [155].
An interesting finding was the isolation of an insect-specific toxin from Tibellus oblongus spider venom acting on the insect’s voltage-activated Ca channel. This 41-residue-long toxin, called ω-Tbo-IT1, produced a toxic effect with an LD50 of 19 μg/g on house fly Musca domestica larvae and with an LD50 of 20 μg/g on juvenile Gromphadorhina portentosa cockroaches. In Periplaneta americana cockroach neurons, one type of voltage-gated Ca2+ current was inhibited by ω-Tbo-IT1; thus, the toxin apparently acted as an inhibitor of presynaptic insect Ca2+ channels. Spatial structure analysis by NMR spectroscopy in aqueous solution revealed that the toxin comprised the modified ICK fold with an extended β-hairpin loop and short β-hairpin loop, which can make ”scissor-like” mutual motions [156].
To summarize the data about animal toxins obtained at the IBCh, we present them in the form of two tables. Table 1 gives the list of naturally occurring animal toxins discovered at IBCh, and Table 2 shows the derivatives of a number of toxins used as research tools for the study of their molecular targets.

7. Conclusions

In this review, we aimed to illustrate the research on peptide and protein toxins from animal venoms that started at our institute (IBCh) approximately 50 years ago. It was stimulated in part by the discovery of α-bungarotoxin and by its tremendous role both in the first characterization of nAChRs and in further research. Briefly summarized in this review is the work on animal toxins carried out at our institute in earlier years; in more detail, we present the recent achievements. The reviewed publications illustrate a very long journey from the first primary structures to the present-day role of peptide and protein neurotoxins in research on nAChRs and on different types of ion channels. Apart from neurotoxins, diverse toxins acting on various biological targets have been studied as well. We also hope that the ongoing research is presented in sufficient detail. It should be emphasized that many excellent labs in the world were and are working on protein and peptide neurotoxins, and the fruitful collaboration of IBCh with them is mentioned in the text and is reflected in the author names in the References.

Author Contributions

All authors participated in writing—original draft preparation, V.T., I.S., S.K. and I.K; writing—review and editing, V.T., I.S., S.K. and I.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant 22-24-00769.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chang, C.C.; Lee, C.Y. Isolation of Neurotoxins from the Venom of Bungarus Multicinctus and Their Modes of Neuromuscular Blocking Action. Arch. Int. Pharmacodyn. Ther. 1963, 144, 241–257. [Google Scholar] [PubMed]
  2. Changeux, J.P.; Kasai, M.; Lee, C.Y. Use of a Snake Venom Toxin to Characterize the Cholinergic Receptor Protein. Proc. Natl. Acad. Sci. USA 1970, 67, 1241–1247. [Google Scholar] [CrossRef] [PubMed]
  3. Grishin, E.V.; Sukhikh, A.P.; Lukyanchuk, N.N.; Slobodyan, L.N.; Lipkin, V.M.; Ovchinnikov, Y.A.; Sorokin, V.M. Amino Acid Sequence of Neurotoxin II from Naja naja oxiana Venom. FEBS Lett. 1973, 36, 77–78. [Google Scholar] [CrossRef] [PubMed]
  4. Grishin, E.V.; Sukhikh, A.P.; Slobodyan, L.N.; Ovchinnikov, Y.A.; Sorokin, V.M. Amino Acid Sequence of Neurotoxin I from Naja naja oxiana Venom. FEBS Lett. 1974, 45, 118–121. [Google Scholar] [CrossRef]
  5. Grishin, E.V.; Sukhikh, A.P.; Adamovich, T.B.; Ovchinnikov, Y.A. The Isolation and Sequence Determination of a Cytotoxin from the Venom of the Middle-Asian Cobra Naja naja oxiana. FEBS Lett. 1974, 48, 179–183. [Google Scholar] [CrossRef]
  6. Grishin, E.V.; Sukhikh, A.P.; Adamovich, T.B.; Ovchinnikov, Y.A. Isolation, properties and sequence determination of the two cytotoxins from the venom of the Middle-Asian cobra Naja naja oxiana. Bioorg. Khim. 1976, 8, 1018–1034. [Google Scholar]
  7. Utkin, Y.N.; Kukhtina, V.V.; Maslennikov, I.V.; Eletsky, A.V.; Starkov, V.G.; Weise, C.; Franke, P.; Hucho, F.; Tsetlin, V.I. First Tryptophan-Containing Weak Neurotoxin from Cobra Venom. Toxicon 2001, 39, 921–927. [Google Scholar] [CrossRef]
  8. Kukhtina, V.V.; Weise, C.; Muranova, T.A.; Starkov, V.G.; Franke, P.; Hucho, F.; Wnendt, S.; Gillen, C.; Tsetlin, V.I.; Utkin, Y.N. Muscarinic Toxin-like Proteins from Cobra Venom. Eur. J. Biochem. 2000, 267, 6784–6789. [Google Scholar] [CrossRef]
  9. Osipov, A.V.; Astapova, M.V.; Tsetlin, V.I.; Utkin, Y.N. The First Representative of Glycosylated Three-Fingered Toxins. Cytotoxin from the Naja Kaouthia Cobra Venom. Eur. J. Biochem. 2004, 271, 2018–2027. [Google Scholar] [CrossRef]
  10. Ramazanova, A.S.; Zavada, L.L.; Starkov, V.G.; Kovyazina, I.V.; Subbotina, T.F.; Kostyukhina, E.E.; Dementieva, I.N.; Ovchinnikova, T.V.; Utkin, Y.N. Heterodimeric Neurotoxic Phospholipases A2—The First Proteins from Venom of Recently Established Species Vipera Nikolskii: Implication of Venom Composition in Viper Systematics. Toxicon 2008, 51, 524–537. [Google Scholar] [CrossRef]
  11. Utkin, Y.N.; Weise, C.; Kasheverov, I.E.; Andreeva, T.V.; Kryukova, E.V.; Zhmak, M.N.; Starkov, V.G.; Hoang, N.A.; Bertrand, D.; Ramerstorfer, J.; et al. Azemiopsin from Azemiops Feae Viper Venom, a Novel Polypeptide Ligand of Nicotinic Acetylcholine Receptor. J. Biol. Chem. 2012, 287, 27079–27086. [Google Scholar] [CrossRef]
  12. Babenko, V.V.; Ziganshin, R.H.; Weise, C.; Dyachenko, I.; Shaykhutdinova, E.; Murashev, A.N.; Zhmak, M.; Starkov, V.; Hoang, A.N.; Tsetlin, V.; et al. Novel Bradykinin-Potentiating Peptides and Three-Finger Toxins from Viper Venom: Combined NGS Venom Gland Transcriptomics and Quantitative Venom Proteomics of the Azemiops Feae Viper. Biomedicines 2020, 8, 249. [Google Scholar] [CrossRef]
  13. Grishin, E.V. Structure and Function of Buthus Eupeus Scorpion Neurotoxins. Int. J. Quantum Chem. 1981, 20, 291–298. [Google Scholar] [CrossRef]
  14. Volkova, T.M.; Garsia, A.F.; Telezhinskaia, I.N.; Potapenko, N.A.; Grishin, E.V. [Amino acid sequence of 2 neurotoxins from the scorpion Buthus eupeus venom]. Bioorg. Khim. 1984, 10, 979–982. [Google Scholar] [PubMed]
  15. Volkova, T.M.; Dulubova, I.E.; Telezhinskaia, I.N.; Grishin, E.V. [Toxic components of the venom of the Central Asian scorpion, Orthochirus scrobiculosus]. Bioorg. Khim. 1984, 10, 1100–1108. [Google Scholar] [PubMed]
  16. Tsai, I.-H.; Wang, Y.-M.; Cheng, A.C.; Starkov, V.; Osipov, A.; Nikitin, I.; Makarova, Y.; Ziganshin, R.; Utkin, Y. CDNA Cloning, Structural, and Functional Analyses of Venom Phospholipases A₂ and a Kunitz-Type Protease Inhibitor from Steppe Viper Vipera Ursinii Renardi. Toxicon 2011, 57, 332–341. [Google Scholar] [CrossRef]
  17. Kozlov, S.; Lipkin, A.; Nosyreva, E.; Blake, A.; Windass, J.D.; Grishin, E. Purification and CDNA Cloning of an Insecticidal Protein from the Venom of the Scorpion Orthochirus Scrobiculosus. Toxicon 2000, 38, 361–371. [Google Scholar] [CrossRef]
  18. Korolkova, Y.V.; Kozlov, S.A.; Lipkin, A.V.; Pluzhnikov, K.A.; Hadley, J.K.; Filippov, A.K.; Brown, D.A.; Angelo, K.; Strøbaek, D.; Jespersen, T.; et al. An ERG Channel Inhibitor from the Scorpion Buthus Eupeus. J. Biol. Chem. 2001, 276, 9868–9876. [Google Scholar] [CrossRef]
  19. Grishin, E.; Korolkova, Y.; Kozlov, S.; Lipkin, A.; Nosyreva, E.; Pluzhnikov, K.; Sukhanov, S.; Volkova, T. Structure and Function of the Potassium Channel Inhibitor from Black Scorpion Venom. Pure Appl. Chem. 1996, 68, 2105–2109. [Google Scholar] [CrossRef]
  20. Kasheverov, I.E.; Oparin, P.B.; Zhmak, M.N.; Egorova, N.S.; Ivanov, I.A.; Gigolaev, A.M.; Nekrasova, O.V.; Serebryakova, M.V.; Kudryavtsev, D.S.; Prokopev, N.A.; et al. Scorpion Toxins Interact with Nicotinic Acetylcholine Receptors. FEBS Lett. 2019, 593, 2779–2789. [Google Scholar] [CrossRef] [PubMed]
  21. Kasheverov, I.E.; Logashina, Y.A.; Kornilov, F.D.; Lushpa, V.A.; Maleeva, E.E.; Korolkova, Y.V.; Yu, J.; Zhu, X.; Zhangsun, D.; Luo, S.; et al. Peptides from the Sea Anemone Metridium Senile with Modified Inhibitor Cystine Knot (ICK) Fold Inhibit Nicotinic Acetylcholine Receptors. Toxins 2022, 15, 28. [Google Scholar] [CrossRef]
  22. Lipkin, A.V.; Grishin, E.V. [Variability of the structure of neurotoxins from the scorpion Orthochirus scrobiculosus from various natural habitats]. Bioorg. Khim. 1999, 25, 341–347. [Google Scholar]
  23. Kozlov, S.; Malyavka, A.; McCutchen, B.; Lu, A.; Schepers, E.; Herrmann, R.; Grishin, E. A Novel Strategy for the Identification of Toxinlike Structures in Spider Venom. Proteins 2005, 59, 131–140. [Google Scholar] [CrossRef] [PubMed]
  24. Kozlov, S.A.; Grishin, E.V. The Universal Algorithm of Maturation for Secretory and Excretory Protein Precursors. Toxicon 2007, 49, 721–726. [Google Scholar] [CrossRef]
  25. Kozlov, S.A.; Lazarev, V.N.; Kostryukova, E.S.; Selezneva, O.V.; Ospanova, E.A.; Alexeev, D.G.; Govorun, V.M.; Grishin, E.V. Comprehensive Analysis of the Venom Gland Transcriptome of the Spider Dolomedes Fimbriatus. Sci. Data 2014, 1, 140023. [Google Scholar] [CrossRef] [PubMed]
  26. Shliapnikov, I.M.; Kozlov, S.A.; Fedorov, A.A.; Grishin, E.V. [Comparison of polypeptide compositions from individual Agelena orientalis spider venoms]. Bioorg. Khim. 2010, 36, 81–88. [Google Scholar] [PubMed]
  27. Kiyatkin, N.I.; Dulubova, I.E.; Chekhovskaya, I.A.; Grishin, E.V. Cloning and Structure of CDNA Encoding α-Latrotoxin from Black Widow Spider Venom. FEBS Lett. 1990, 270, 127–131. [Google Scholar] [CrossRef]
  28. Kiyatkin, N.; Dulubova, I.; Grishin, E. Cloning and Structural Analysis of α-Latroinsectotoxin CDNA. Eur. J. Biochem. 1993, 213, 121–127. [Google Scholar] [CrossRef]
  29. Dulubova, I.E.; Krasnoperov, V.G.; Khvotchev, M.V.; Pluzhnikov, K.A.; Volkova, T.M.; Grishin, E.V.; Vais, H.; Bell, D.R.; Usherwood, P.N.R. Cloning and Structure of δ-Latroinsectotoxin, a Novel Insect-Specific Member of the Latrotoxin Family. J. Biol. Chem. 1996, 271, 7535–7543. [Google Scholar] [CrossRef]
  30. Südhof, T.C. Alpha-Latrotoxin and Its Receptors: Neurexins and CIRL/Latrophilins. Annu. Rev. Neurosci. 2001, 24, 933–962. [Google Scholar] [CrossRef]
  31. Arseniev, A.S.; Balashova, T.A.; Utkin, Y.N.; Tsetlin, V.I.; Bystrov, V.F.; Ivanov, V.T.; Ovchinnikov, Y.A. Proton-Nuclear-Magnetic-Resonance Study of the Conformation of Neurotoxin II from Middle-Asian Cobra (Naja naja oxiana) Venom. Eur. J. Biochem. 1976, 71, 595–606. [Google Scholar] [CrossRef]
  32. Low, B.W.; Preston, H.S.; Sato, A.; Rosen, L.S.; Searl, J.E.; Rudko, A.D.; Richardson, J.S. Three Dimensional Structure of Erabutoxin b Neurotoxic Protein: Inhibitor of Acetylcholine Receptor. Proc. Natl. Acad. Sci. USA 1976, 73, 2991–2994. [Google Scholar] [CrossRef]
  33. Tsernoglou, D.; Petsko, G.A. Three-Dimensional Structure of Neurotoxin a from Venom of the Philippines Sea Snake. Proc. Natl. Acad. Sci. USA 1977, 74, 971–974. [Google Scholar] [CrossRef]
  34. Lyukmanova, E.N.; Shenkarev, Z.O.; Schulga, A.A.; Ermolyuk, Y.S.; Mordvintsev, D.Y.; Utkin, Y.N.; Shoulepko, M.A.; Hogg, R.C.; Bertrand, D.; Dolgikh, D.A.; et al. Bacterial Expression, NMR, and Electrophysiology Analysis of Chimeric Short/Long-Chain Alpha-Neurotoxins Acting on Neuronal Nicotinic Receptors. J. Biol. Chem. 2007, 282, 24784–24791. [Google Scholar] [CrossRef] [PubMed]
  35. Lyukmanova, E.N.; Shulepko, M.A.; Tikhonov, R.V.; Shenkarev, Z.O.; Paramonov, A.S.; Wulfson, A.N.; Kasheverov, I.E.; Ustich, T.L.; Utkin, Y.N.; Arseniev, A.S.; et al. Bacterial Production and Refolding from Inclusion Bodies of a “Weak” Toxin, a Disulfide Rich Protein. Biochemistry 2009, 74, 1142–1149. [Google Scholar] [CrossRef] [PubMed]
  36. Golovanov, A.P.; Lomize, A.L.; Arseniev, A.S.; Utkin, Y.N.; Tsetlin, V.I. Two-Dimensional 1H-NMR Study of the Spatial Structure of Neurotoxin II from Naja naja oxiana. Eur. J. Biochem. 1993, 213, 1213–1223. [Google Scholar] [CrossRef]
  37. Lyukmanova, E.N.; Shenkarev, Z.O.; Shulepko, M.A.; Paramonov, A.S.; Chugunov, A.O.; Janickova, H.; Dolejsi, E.; Dolezal, V.; Utkin, Y.N.; Tsetlin, V.I.; et al. Structural Insight into Specificity of Interactions between Nonconventional Three-Finger Weak Toxin from Naja Kaouthia (WTX) and Muscarinic Acetylcholine Receptors. J. Biol. Chem. 2015, 290, 23616–23630. [Google Scholar] [CrossRef] [PubMed]
  38. Osipov, A.V.; Kasheverov, I.E.; Makarova, Y.V.; Starkov, V.G.; Vorontsova, O.V.; Ziganshin, R.K.; Andreeva, T.V.; Serebryakova, M.V.; Benoit, A.; Hogg, R.C.; et al. Naturally Occurring Disulfide-Bound Dimers of Three-Fingered Toxins: A Paradigm for Biological Activity Diversification. J. Biol. Chem. 2008, 283, 14571–14580. [Google Scholar] [CrossRef] [PubMed]
  39. Osipov, A.V.; Rucktooa, P.; Kasheverov, I.E.; Filkin, S.Y.; Starkov, V.G.; Andreeva, T.V.; Sixma, T.K.; Bertrand, D.; Utkin, Y.N.; Tsetlin, V.I. Dimeric α-Cobratoxin X-Ray Structure: Localization of Intermolecular Disulfides and Possible Mode of Binding to Nicotinic Acetylcholine Receptors. J. Biol. Chem. 2012, 287, 6725–6734. [Google Scholar] [CrossRef]
  40. Maslennikov, I.V.; Shenkarev, Z.O.; Zhmak, M.N.; Ivanov, V.T.; Methfessel, C.; Tsetlin, V.I.; Arseniev, A.S. NMR Spatial Structure of Alpha-Conotoxin ImI Reveals a Common Scaffold in Snail and Snake Toxins Recognizing Neuronal Nicotinic Acetylcholine Receptors. FEBS Lett. 1999, 444, 275–280. [Google Scholar] [CrossRef]
  41. Kuzmenkov, A.I.; Peigneur, S.; Nasburg, J.A.; Mineev, K.S.; Nikolaev, M.V.; Pinheiro-Junior, E.L.; Arseniev, A.S.; Wulff, H.; Tytgat, J.; Vassilevski, A.A. Apamin Structure and Pharmacology Revisited. Front. Pharmacol. 2022, 13, 977440. [Google Scholar] [CrossRef] [PubMed]
  42. Jaravine, V.A.; Nolde, D.E.; Reibarkh, M.J.; Korolkova, Y.V.; Kozlov, S.A.; Pluzhnikov, K.A.; Grishin, E.V.; Arseniev, A.S. Three-Dimensional Structure of Toxin OSK1 from Orthochirus Scrobiculosus Scorpion Venom. Biochemistry 1997, 36, 1223–1232. [Google Scholar] [CrossRef] [PubMed]
  43. Pashkov, V.S.; Maiorov, V.N.; Bystrov, V.F.; Hoang, A.N.; Volkova, T.M.; Grishin, E.V. Solution Spatial Structure of ‘Long’ Neurotoxin M9 from the Scorpion Buthus Eupeus by 1H-NMR Spectroscopy. Biophys. Chem. 1988, 31, 121–131. [Google Scholar] [CrossRef]
  44. Arsen’ev, A.S.; Kondakov, V.I.; Maĭorov, V.N.; Volkova, T.M.; Grishin, E.V. [Secondary structure and assignment of signals in two-dimensional 1H-NMR spectra of the Buthus eupeus neurotoxin I5A]. Bioorg. Khim. 1983, 9, 768–793. [Google Scholar] [PubMed]
  45. Korolkova, Y.V.; Bocharov, E.V.; Angelo, K.; Maslennikov, I.V.; Grinenko, O.V.; Lipkin, A.V.; Nosyreva, E.D.; Pluzhnikov, K.A.; Olesen, S.-P.; Arseniev, A.S.; et al. New Binding Site on Common Molecular Scaffold Provides HERG Channel Specificity of Scorpion Toxin BeKm-1. J. Biol. Chem. 2002, 277, 43104–43109. [Google Scholar] [CrossRef] [PubMed]
  46. Oparin, P.B.; Nadezhdin, K.D.; Berkut, A.A.; Arseniev, A.S.; Grishin, E.V.; Vassilevski, A.A. Structure of Purotoxin-2 from Wolf Spider: Modular Design and Membrane-Assisted Mode of Action in Arachnid Toxins. Biochem. J. 2016, 473, 3113–3126. [Google Scholar] [CrossRef]
  47. Polz-Tejera, G.; Schmidt, J.; Karten, H.J. Autoradiographic Localisation of Alpha-Bungarotoxin-Binding Sites in the Central Nervous System. Nature 1975, 258, 349–351. [Google Scholar] [CrossRef] [PubMed]
  48. Karlsson, E.; Heilbronn, E.; Widlund, L. Isolation of the Nicotinic Acetylcholine Receptor by Biospecific Chromatography on Insolubilized Naja Naja Neurotoxin. FEBS Lett. 1972, 28, 107–111. [Google Scholar] [CrossRef]
  49. 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. 2021, 158, 1223–1235. [Google Scholar] [CrossRef]
  50. Changeux, J.-P. Discovery of the First Neurotransmitter Receptor: The Acetylcholine Nicotinic Receptor. Biomolecules 2020, 10, 547. [Google Scholar] [CrossRef]
  51. Nirthanan, S. Snake Three-Finger α-Neurotoxins and Nicotinic Acetylcholine Receptors: Molecules, Mechanisms and Medicine. Biochem. Pharmacol. 2020, 181, 114168. [Google Scholar] [CrossRef]
  52. Mitra, A.K.; McCarthy, M.P.; Stroud, R.M. Three-Dimensional Structure of the Nicotinic Acetylcholine Receptor and Location of the Major Associated 43-KD Cytoskeletal Protein, Determined at 22 A by Low Dose Electron Microscopy and x-Ray Diffraction to 12.5 A. J. Cell Biol. 1989, 109, 755–774. [Google Scholar] [CrossRef]
  53. Brejc, K.; van Dijk, W.J.; Klaassen, R.V.; Schuurmans, M.; van Der Oost, J.; Smit, A.B.; Sixma, T.K. Crystal Structure of an ACh-Binding Protein Reveals the Ligand-Binding Domain of Nicotinic Receptors. Nature 2001, 411, 269–276. [Google Scholar] [CrossRef] [PubMed]
  54. Unwin, N. Refined Structure of the Nicotinic Acetylcholine Receptor at 4A Resolution. J. Mol. Biol. 2005, 346, 967–989. [Google Scholar] [CrossRef] [PubMed]
  55. Hilf, R.J.C.; Dutzler, R. X-Ray Structure of a Prokaryotic Pentameric Ligand-Gated Ion Channel. Nature 2008, 452, 375–379. [Google Scholar] [CrossRef]
  56. Bocquet, N.; Nury, H.; Baaden, M.; Le Poupon, C.; Changeux, J.-P.; Delarue, M.; Corringer, P.-J. X-Ray Structure of a Pentameric Ligand-Gated Ion Channel in an Apparently Open Conformation. Nature 2009, 457, 111–114. [Google Scholar] [CrossRef]
  57. Morales-Perez, C.L.; Noviello, C.M.; Hibbs, R.E. X-Ray Structure of the Human A4β2 Nicotinic Receptor. Nature 2016, 538, 411–415. [Google Scholar] [CrossRef]
  58. 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]
  59. Huang, S.; Li, S.-X.; Bren, N.; Cheng, K.; Gomoto, R.; Chen, L.; Sine, S.M. Complex between α-Bungarotoxin and an A7 Nicotinic Receptor Ligand-Binding Domain Chimaera. Biochem. J. 2013, 454, 303–310. [Google Scholar] [CrossRef]
  60. Dellisanti, C.D.; Yao, Y.; Stroud, J.C.; Wang, Z.-Z.; Chen, L. Crystal Structure of the Extracellular Domain of NAChR Alpha1 Bound to Alpha-Bungarotoxin at 1.94 A Resolution. Nat. Neurosci. 2007, 10, 953–962. [Google Scholar] [CrossRef]
  61. Zouridakis, M.; Giastas, P.; Zarkadas, E.; Chroni-Tzartou, D.; Bregestovski, P.; Tzartos, S.J. Crystal Structures of Free and Antagonist-Bound States of Human A9 Nicotinic Receptor Extracellular Domain. Nat. Struct. Mol. Biol. 2014, 21, 976–980. [Google Scholar] [CrossRef]
  62. 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.e5. [Google Scholar] [CrossRef]
  63. Noviello, C.M.; Gharpure, A.; Mukhtasimova, N.; Cabuco, R.; Baxter, L.; Borek, D.; Sine, S.M.; Hibbs, R.E. Structure and Gating Mechanism of the A7 Nicotinic Acetylcholine Receptor. Cell 2021, 184, 2121–2134.e13. [Google Scholar] [CrossRef] [PubMed]
  64. Nys, M.; Zarkadas, E.; Brams, M.; Mehregan, A.; Kambara, K.; Kool, J.; Casewell, N.R.; Bertrand, D.; Baenziger, J.E.; Nury, H.; et al. The Molecular Mechanism of Snake Short-Chain α-Neurotoxin Binding to Muscle-Type Nicotinic Acetylcholine Receptors. Nat. Commun. 2022, 13, 4543. [Google Scholar] [CrossRef]
  65. Tsetlin, V.I.; Karlsson, E.; Arseniev, A.S.; Utkin, Y.N.; Surin, A.M.; Pashkov, V.S.; Pluzhnikov, K.A.; Ivanov, V.T.; Bystrov, V.F.; Ovchinnikov, Y.A. EPR and Fluorescence Study of Interaction of Naja naja oxiana Neurotoxin II and Its Derivatives with Acetylcholine Receptor Protein from Torpedo Marmorata. FEBS Lett. 1979, 106, 47–52. [Google Scholar] [CrossRef]
  66. Kreienkamp, H.J.; Utkin, Y.N.; Weise, C.; Machold, J.; Tsetlin, V.I.; Hucho, F. Investigation of Ligand-Binding Sites of the Acetylcholine Receptor Using Photoactivatable Derivatives of Neurotoxin II from Naja naja oxiana. Biochemistry 1992, 31, 8239–8244. [Google Scholar] [CrossRef] [PubMed]
  67. Machold, J.; Utkin, Y.; Kirsch, D.; Kaufmann, R.; Tsetlin, V.; Hucho, F. Photolabeling Reveals the Proximity of the Alpha-Neurotoxin Binding Site to the M2 Helix of the Ion Channel in the Nicotinic Acetylcholine Receptor. Proc. Natl. Acad. Sci. USA 1995, 92, 7282–7286. [Google Scholar] [CrossRef]
  68. Utkin, Y.N.; Krivoshein, A.V.; Davydov, V.L.; Kasheverov, I.E.; Franke, P.; Maslennikov, I.V.; Arseniev, A.S.; Hucho, F.; Tsetlin, V.I. Labeling of Torpedo Californica Nicotinic Acetylcholine Receptor Subunits by Cobratoxin Derivatives with Photoactivatable Groups of Different Chemical Nature at Lys23. Eur. J. Biochem. 1998, 253, 229–235. [Google Scholar] [CrossRef] [PubMed]
  69. 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]
  70. Kasaragod, V.B.; Mortensen, M.; Hardwick, S.W.; Wahid, A.A.; Dorovykh, V.; Chirgadze, D.Y.; Smart, T.G.; Miller, P.S. Mechanisms of Inhibition and Activation of Extrasynaptic Aβ GABAA Receptors. Nature 2022, 602, 529–533. [Google Scholar] [CrossRef]
  71. 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. [Google Scholar] [CrossRef] [PubMed]
  72. Utkin, Y.N.; Kuch, U.; Kasheverov, I.E.; Lebedev, D.S.; Cederlund, E.; Molles, B.E.; Polyak, I.; 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]
  73. Kasheverov, I.E.; Kuzmenkov, A.I.; Kudryavtsev, D.S.; Chudetskiy, I.S.; Shelukhina, I.V.; Barykin, E.P.; Ivanov, I.A.; Siniavin, A.E.; Ziganshin, R.H.; Baranov, M.S.; et al. Snake Toxins Labeled by Green Fluorescent Protein or Its Synthetic Chromophore Are New Probes for Nicotinic Acetylcholine Receptors. Front. Mol. Biosci. 2021, 8, 753283. [Google Scholar] [CrossRef]
  74. Dement’eva, D.V.; Utkin, I.N.; Arsen’ev, A.S. [Secondary structure and conformational heterogeneity of Naja naja oxiana cytotoxin II]. Bioorg. Khim. 1996, 22, 339–352. [Google Scholar] [PubMed]
  75. Dementieva, D.V.; Bocharov, E.V.; Arseniev, A.S. Two Forms of Cytotoxin II (Cardiotoxin) from Naja naja oxiana in Aqueous Solution: Spatial Structures with Tightly Bound Water Molecules. Eur. J. Biochem. 1999, 263, 152–162. [Google Scholar] [CrossRef]
  76. Dubovskii, P.V.; Lesovoy, D.M.; Dubinnyi, M.A.; Utkin, Y.N.; Arseniev, A.S. Interaction of the P-Type Cardiotoxin with Phospholipid Membranes. Eur. J. Biochem. 2003, 270, 2038–2046. [Google Scholar] [CrossRef] [PubMed]
  77. Dubovskii, P.V.; Dubova, K.M.; Bourenkov, G.; Starkov, V.G.; Konshina, A.G.; Efremov, R.G.; Utkin, Y.N.; Samygina, V.R. Variability in the Spatial Structure of the Central Loop in Cobra Cytotoxins Revealed by X-Ray Analysis and Molecular Modeling. Toxins 2022, 14, 149. [Google Scholar] [CrossRef]
  78. Dubovskii, P.V.; Ignatova, A.A.; Alekseeva, A.S.; Starkov, V.G.; Boldyrev, I.A.; Feofanov, A.V.; Utkin, Y.N. Membrane-Disrupting Activity of Cobra Cytotoxins Is Determined by Configuration of the N-Terminal Loop. Toxins 2022, 15, 6. [Google Scholar] [CrossRef]
  79. Feofanov, A.V.; Sharonov, G.V.; Dubinnyi, M.A.; Astapova, M.V.; Kudelina, I.A.; Dubovskii, P.V.; Rodionov, D.I.; Utkin, Y.N.; Arseniev, A.S. Comparative Study of Structure and Activity of Cytotoxins from Venom of the Cobras Naja oxiana, Naja kaouthia, and Naja haje. Biochemistry 2004, 69, 1148–1157. [Google Scholar] [CrossRef]
  80. Kuleshina, O.N.; Kruykova, E.V.; Cheremnykh, E.G.; Kozlov, L.V.; Andreeva, T.V.; Starkov, V.G.; Osipov, A.V.; Ziganshin, R.H.; Tsetlin, V.I.; Utkin, Y.N. Screening Snake Venoms for Toxicity to Tetrahymena Pyriformis Revealed Anti-Protozoan Activity of Cobra Cytotoxins. Toxins 2020, 12, 325. [Google Scholar] [CrossRef] [PubMed]
  81. Averin, A.S.; Nenov, M.N.; Starkov, V.G.; Tsetlin, V.I.; Utkin, Y.N. Effects of Cardiotoxins from Naja oxiana Cobra Venom on Rat Heart Muscle and Aorta: A Comparative Study of Toxin-Induced Contraction Mechanisms. Toxins 2022, 14, 88. [Google Scholar] [CrossRef]
  82. Averin, A.S.; Goltyaev, M.V.; Andreeva, T.V.; Starkov, V.G.; Tsetlin, V.I.; Utkin, Y.N. S- and P-Type Cobra Venom Cardiotoxins Differ in Their Action on Isolated Rat Heart. J. Venom. Anim. Toxins Incl. Trop. Dis. 2022, 28, e20210110. [Google Scholar] [CrossRef] [PubMed]
  83. Averin, A.S.; Berezhnov, A.V.; Pimenov, O.Y.; Galimova, M.H.; Starkov, V.G.; Tsetlin, V.I.; Utkin, Y.N. Effects of Cobra Cardiotoxins on Intracellular Calcium and the Contracture of Rat Cardiomyocytes Depend on Their Structural Types. Int. J. Mol. Sci. 2023, 24, 9259. [Google Scholar] [CrossRef]
  84. Utkin, Y.N.; Kukhtina, V.V.; Kryukova, E.V.; Chiodini, F.; Bertrand, D.; Methfessel, C.; Tsetlin, V.I. “Weak Toxin” from Naja Kaouthia Is a Nontoxic Antagonist of Alpha 7 and Muscle-Type Nicotinic Acetylcholine Receptors. J. Biol. Chem. 2001, 276, 15810–15815. [Google Scholar] [CrossRef] [PubMed]
  85. Mordvintsev, D.Y.; Polyak, Y.L.; Rodionov, D.I.; Jakubik, J.; Dolezal, V.; Karlsson, E.; Tsetlin, V.I.; Utkin, Y.N. Weak Toxin WTX from Naja Kaouthia Cobra Venom Interacts with Both Nicotinic and Muscarinic Acetylcholine Receptors. FEBS J. 2009, 276, 5065–5075. [Google Scholar] [CrossRef]
  86. Ogay, A.Y.; Rzhevsky, D.I.; Murashev, A.N.; Tsetlin, V.I.; Utkin, Y.N. Weak Neurotoxin from Naja Kaouthia Cobra Venom Affects Haemodynamic Regulation by Acting on Acetylcholine Receptors. Toxicon 2005, 45, 93–99. [Google Scholar] [CrossRef] [PubMed]
  87. Miwa, J.M.; Anderson, K.R.; Hoffman, K.M. Lynx Prototoxins: Roles of Endogenous Mammalian Neurotoxin-Like Proteins in Modulating Nicotinic Acetylcholine Receptor Function to Influence Complex Biological Processes. Front. Pharmacol. 2019, 10, 343. [Google Scholar] [CrossRef]
  88. Tsetlin, V.I. Three-Finger Snake Neurotoxins and Ly6 Proteins Targeting Nicotinic Acetylcholine Receptors: Pharmacological Tools and Endogenous Modulators. Trends Pharmacol. Sci. 2015, 36, 109–123. [Google Scholar] [CrossRef]
  89. Lyukmanova, E.N.; Shenkarev, Z.O.; Shulepko, M.A.; Mineev, K.S.; D’Hoedt, D.; Kasheverov, I.E.; Filkin, S.Y.; Krivolapova, A.P.; Janickova, H.; Dolezal, V.; et al. NMR Structure and Action on Nicotinic Acetylcholine Receptors of Water-Soluble Domain of Human LYNX1. J. Biol. Chem. 2011, 286, 10618–10627. [Google Scholar] [CrossRef]
  90. Shulepko, M.A.; Lyukmanova, E.N.; Paramonov, A.S.; Lobas, A.A.; Shenkarev, Z.O.; Kasheverov, I.E.; Dolgikh, D.A.; Tsetlin, V.I.; Arseniev, A.S.; Kirpichnikov, M.P. Human Neuromodulator SLURP-1: Bacterial Expression, Binding to Muscle-Type Nicotinic Acetylcholine Receptor, Secondary Structure, and Conformational Heterogeneity in Solution. Biochemistry 2013, 78, 204–211. [Google Scholar] [CrossRef]
  91. Lyukmanova, E.N.; Shulepko, M.A.; Kudryavtsev, D.; Bychkov, M.L.; Kulbatskii, D.S.; Kasheverov, I.E.; Astapova, M.V.; Feofanov, A.V.; Thomsen, M.S.; Mikkelsen, J.D.; et al. Human Secreted Ly-6/UPAR Related Protein-1 (SLURP-1) Is a Selective Allosteric Antagonist of A7 Nicotinic Acetylcholine Receptor. PLoS ONE 2016, 11, e0149733. [Google Scholar] [CrossRef]
  92. Lyukmanova, E.N.; Shulepko, M.A.; Buldakova, S.L.; Kasheverov, I.E.; Shenkarev, Z.O.; Reshetnikov, R.V.; Filkin, S.Y.; Kudryavtsev, D.S.; Ojomoko, L.O.; Kryukova, E.V.; et al. Water-Soluble LYNX1 Residues Important for Interaction with Muscle-Type and/or Neuronal Nicotinic Receptors. J. Biol. Chem. 2013, 288, 15888–15899. [Google Scholar] [CrossRef] [PubMed]
  93. Apsalon, U.R.; Shamborant, O.G.; Miroshnikov, A.I. Isolation and Some Properties of Phospholipase A2 from Venom of Middle-Asian Cobra Naja-Naja-Oxiana. Bioorg. Khim. 1977, 3, 1553–1559. [Google Scholar]
  94. Ovchinnikov, Y.A.; Miroshnikov, A.I.; Nazimov, I.V.; Apsalon, U.R.; Soldatova, L.N. Complete Amino-Acid Sequence of Phospholipase-A2 (Isoenzyme E3) from the Venom of Middle Asian Cobra Naja-Naja-Oxiana. Bioorg. Khim. 1979, 5, 805–813. [Google Scholar]
  95. Osipov, A.V.; Filkin, S.Y.; Makarova, Y.V.; Tsetlin, V.I.; Utkin, Y.N. A New Type of Thrombin Inhibitor, Noncytotoxic Phospholipase A2, from the Naja Haje Cobra Venom. Toxicon 2010, 55, 186–194. [Google Scholar] [CrossRef] [PubMed]
  96. Vulfius, C.A.; Kasheverov, I.E.; Starkov, V.G.; Osipov, A.V.; Andreeva, T.V.; Filkin, S.Y.; Gorbacheva, E.V.; Astashev, M.E.; Tsetlin, V.I.; Utkin, Y.N. Inhibition of Nicotinic Acetylcholine Receptors, a Novel Facet in the Pleiotropic Activities of Snake Venom Phospholipases A2. PLoS ONE 2014, 9, e115428. [Google Scholar] [CrossRef]
  97. Vulfius, C.A.; Kasheverov, I.E.; Kryukova, E.V.; Spirova, E.N.; Shelukhina, I.V.; Starkov, V.G.; Andreeva, T.V.; Faure, G.; Zouridakis, M.; Tsetlin, V.I.; et al. Pancreatic and Snake Venom Presynaptically Active Phospholipases A2 Inhibit Nicotinic Acetylcholine Receptors. PLoS ONE 2017, 12, e0186206. [Google Scholar] [CrossRef]
  98. Tran, T.V.; Siniavin, A.E.; Hoang, A.N.; Le, M.T.T.; Pham, C.D.; Phung, T.V.; Nguyen, K.C.; Ziganshin, R.H.; Tsetlin, V.I.; Weng, C.-F.; et al. Phospholipase A2 from Krait Bungarus Fasciatus Venom Induces Human Cancer Cell Death in Vitro. PeerJ 2019, 7, e8055. [Google Scholar] [CrossRef]
  99. Siniavin, A.E.; Streltsova, M.A.; Nikiforova, M.A.; Kudryavtsev, D.S.; Grinkina, S.D.; Gushchin, V.A.; Mozhaeva, V.A.; Starkov, V.G.; Osipov, A.V.; Lummis, S.C.R.; et al. Snake Venom Phospholipase A2s Exhibit Strong Virucidal Activity against SARS-CoV-2 and Inhibit the Viral Spike Glycoprotein Interaction with ACE2. Cell. Mol. Life Sci. 2021, 78, 7777–7794. [Google Scholar] [CrossRef]
  100. Siniavin, A.; Grinkina, S.; Osipov, A.; Starkov, V.; Tsetlin, V.; Utkin, Y. Anti-HIV Activity of Snake Venom Phospholipase A2s: Updates for New Enzymes and Different Virus Strains. Int. J. Mol. Sci. 2022, 23, 1610. [Google Scholar] [CrossRef]
  101. Shelukhina, I.V.; Zhmak, M.N.; Lobanov, A.V.; Ivanov, I.A.; Garifulina, A.I.; Kravchenko, I.N.; Rasskazova, E.A.; Salmova, M.A.; Tukhovskaya, E.A.; Rykov, V.A.; et al. Azemiopsin, a Selective Peptide Antagonist of Muscle Nicotinic Acetylcholine Receptor: Preclinical Evaluation as a Local Muscle Relaxant. Toxins 2018, 10, 34. [Google Scholar] [CrossRef]
  102. Kozlov, S.A.; Vassilevski, A.A.; Feofanov, A.V.; Surovoy, A.Y.; Karpunin, D.V.; Grishin, E.V. Latarcins, Antimicrobial and Cytolytic Peptides from the Venom of the Spider Lachesana Tarabaevi (Zodariidae) That Exemplify Biomolecular Diversity. J. Biol. Chem. 2006, 281, 20983–20992. [Google Scholar] [CrossRef] [PubMed]
  103. Dubovskii, P.V.; Vassilevski, A.A.; Kozlov, S.A.; Feofanov, A.V.; Grishin, E.V.; Efremov, R.G. Latarcins: Versatile Spider Venom Peptides. Cell. Mol. Life Sci. 2015, 72, 4501–4522. [Google Scholar] [CrossRef]
  104. Gray, W.R.; Luque, A.; Olivera, B.M.; Barrett, J.; Cruz, L.J. Peptide Toxins from Conus Geographus Venom. J. Biol. Chem. 1981, 256, 4734–4740. [Google Scholar] [CrossRef] [PubMed]
  105. McManus, O.B.; Musick, J.R.; Gonzalez, C. Peptides Isolated from the Venom of Conus Geographus Block Neuromuscular Transmission. Neurosci. Lett. 1981, 25, 57–62. [Google Scholar] [CrossRef]
  106. Lebbe, E.K.M.; Peigneur, S.; Wijesekara, I.; Tytgat, J. Conotoxins Targeting Nicotinic Acetylcholine Receptors: An Overview. Mar. Drugs 2014, 12, 2970–3004. [Google Scholar] [CrossRef]
  107. Abraham, N.; Lewis, R.J. Neuronal Nicotinic Acetylcholine Receptor Modulators from Cone Snails. Mar. Drugs 2018, 16, 208. [Google Scholar] [CrossRef] [PubMed]
  108. Kasheverov, I.; Kudryavtsev, D.; Shelukhina, I.; Nikolaev, G.; Utkin, Y.; Tsetlin, V. Marine Origin Ligands of Nicotinic Receptors: Low Molecular Compounds, Peptides and Proteins for Fundamental Research and Practical Applications. Biomolecules 2022, 12, 189. [Google Scholar] [CrossRef]
  109. Maslennikov, I.V.; Sobol, A.G.; Gladky, K.V.; Lugovskoy, A.A.; Ostrovsky, A.G.; Tsetlin, V.I.; Ivanov, V.T.; Arseniev, A.S. Two Distinct Structures of Alpha-Conotoxin GI in Aqueous Solution. Eur. J. Biochem. 1998, 254, 238–247. [Google Scholar] [CrossRef]
  110. Utkin, Y.N.; Zhmak, M.N.; Methfessel, C.; Tsetlin, V.I. Aromatic Substitutions in Alpha-Conotoxin ImI. Synthesis of Iodinated Photoactivatable Derivative. Toxicon 1999, 37, 1683–1695. [Google Scholar] [CrossRef]
  111. Kasheverov, I.; Zhmak, M.; Chivilyov, E.; Saez-Brionez, P.; Utkin, Y.; Hucho, F.; Tsetlin, V. Benzophenone-Type Photoactivatable Derivatives of Alpha-Neurotoxins and Alpha-Conotoxins in Studies on Torpedo Nicotinic Acetylcholine Receptor. J. Recept. Signal Transduct. Res. 1999, 19, 559–571. [Google Scholar] [CrossRef] [PubMed]
  112. Kasheverov, I.; Rozhkova, A.; Zhmak, M.; Utkin, Y.; Ivanov, V.; Tsetlin, V.I. Photoactivatable Alpha-Conotoxins Reveal Contacts with All Subunits as Well as Antagonist-Induced Rearrangements in the Torpedo Californica Acetylcholine Receptor. Eur. J. Biochem. 2001, 268, 3664–3673. [Google Scholar] [CrossRef]
  113. Kasheverov, I.E.; Chiara, D.C.; Zhmak, M.N.; Maslennikov, I.V.; Pashkov, V.S.; Arseniev, A.S.; Utkin, Y.N.; Cohen, J.B.; Tsetlin, V.I. Alpha-Conotoxin GI Benzoylphenylalanine Derivatives. (1)H-NMR Structures and Photoaffinity Labeling of the Torpedo Californica Nicotinic Acetylcholine Receptor. FEBS J. 2006, 273, 1373–1388. [Google Scholar] [CrossRef] [PubMed]
  114. Kasheverov, I.E.; Zhmak, M.N.; Vulfius, C.A.; Gorbacheva, E.V.; Mordvintsev, D.Y.; Utkin, Y.N.; van Elk, R.; Smit, A.B.; Tsetlin, V.I. Alpha-Conotoxin Analogs with Additional Positive Charge Show Increased Selectivity towards Torpedo Californica and Some Neuronal Subtypes of Nicotinic Acetylcholine Receptors. FEBS J. 2006, 273, 4470–4481. [Google Scholar] [CrossRef]
  115. Kasheverov, I.E.; Zhmak, M.N.; Khruschov, A.Y.; Tsetlin, V.I. Design of New α-Conotoxins: From Computer Modeling to Synthesis of Potent Cholinergic Compounds. Mar. Drugs 2011, 9, 1698–1714. [Google Scholar] [CrossRef]
  116. Kasheverov, I.E.; Chugunov, A.O.; Kudryavtsev, D.S.; Ivanov, I.A.; Zhmak, M.N.; Shelukhina, I.V.; Spirova, E.N.; Tabakmakher, V.M.; Zelepuga, E.A.; Efremov, R.G.; et al. High-Affinity α-Conotoxin PnIA Analogs Designed on the Basis of the Protein Surface Topography Method. Sci. Rep. 2016, 6, 36848. [Google Scholar] [CrossRef]
  117. Vulfius, C.A.; Tumina, O.B.; Kasheverov, I.E.; Utkin, Y.N.; Tsetlin, V.I. Diversity of Nicotinic Receptors Mediating Cl- Current in Lymnaea Neurons Distinguished with Specific Agonists and Antagonist. Neurosci. Lett. 2005, 373, 232–236. [Google Scholar] [CrossRef] [PubMed]
  118. Pivovarov, A.S.; Palikhova, T.A.; Nikolaev, G.M.; Velikanov, A.N.; Vasilieva, N.A.; Kasheverov, I.E.; Utkin, Y.N.; Tsetlin, V.I. Atypical Acetylcholine Receptors on the Neurons of the Turkish Snail. Dokl. Biochem. Biophys. 2020, 491, 81–84. [Google Scholar] [CrossRef]
  119. Koval, L.; Lykhmus, O.; Zhmak, M.; Khruschov, A.; Tsetlin, V.; Magrini, E.; Viola, A.; Chernyavsky, A.; Qian, J.; Grando, S.; et al. Differential Involvement of A4β2, A7 and A9α10 Nicotinic Acetylcholine Receptors in B Lymphocyte Activation in Vitro. Int. J. Biochem. Cell Biol. 2011, 43, 516–524. [Google Scholar] [CrossRef]
  120. Safronova, V.G.; Vulfius, C.A.; Shelukhina, I.V.; Mal’tseva, V.N.; Berezhnov, A.V.; Fedotova, E.I.; Miftahova, R.G.; Kryukova, E.V.; Grinevich, A.A.; Tsetlin, V.I. Nicotinic Receptor Involvement in Regulation of Functions of Mouse Neutrophils from Inflammatory Site. Immunobiology 2016, 221, 761–772. [Google Scholar] [CrossRef]
  121. Safronova, V.G.; Vulfius, C.A.; Astashev, M.E.; Tikhonova, I.V.; Serov, D.A.; Jirova, E.A.; Pershina, E.V.; Senko, D.A.; Zhmak, M.N.; Kasheverov, I.E.; et al. A9α10 Nicotinic Acetylcholine Receptors Regulate Murine Bone Marrow Granulocyte Functions. Immunobiology 2021, 226, 152047. [Google Scholar] [CrossRef]
  122. Terpinskaya, T.I.; Osipov, A.V.; Balashevich, T.V.; Yanchanka, T.L.; Tamashionik, E.A.; Tsetlin, V.I.; Utkin, Y.N. Blockers of Nicotinic Acetylcholine Receptors Delay Tumor Growth and Increase Antitumor Activity of Mouse Splenocytes. Dokl. Biochem. Biophys. 2020, 491, 89–92. [Google Scholar] [CrossRef] [PubMed]
  123. Terpinskaya, T.I.; Osipov, A.V.; Kuznetsova, T.E.; Ryzhkovskaya, E.L.; Ulaschik, V.S.; Ivanov, I.A.; Tsetlin, V.I.; Utkin, Y.N. α-Conotoxins Revealed Different Roles of Nicotinic Cholinergic Receptor Subtypes in Oncogenesis of Ehrlich Tumor and in the Associated Inflammation. Dokl. Biochem. Biophys. 2015, 463, 216–219. [Google Scholar] [CrossRef] [PubMed]
  124. Osipov, A.V.; Terpinskaya, T.I.; Yanchanka, T.; Balashevich, T.; Zhmak, M.N.; Tsetlin, V.I.; Utkin, Y.N. α-Conotoxins Enhance Both the In Vivo Suppression of Ehrlich Carcinoma Growth and In Vitro Reduction in Cell Viability Elicited by Cyclooxygenase and Lipoxygenase Inhibitors. Mar. Drugs 2020, 18, 193. [Google Scholar] [CrossRef]
  125. Terpinskaya, T.I.; Osipov, A.V.; Kryukova, E.V.; Kudryavtsev, D.S.; Kopylova, N.V.; Yanchanka, T.L.; Palukoshka, A.F.; Gondarenko, E.A.; Zhmak, M.N.; Tsetlin, V.I.; et al. α-Conotoxins and α-Cobratoxin Promote, While Lipoxygenase and Cyclooxygenase Inhibitors Suppress the Proliferation of Glioma C6 Cells. Mar. Drugs 2021, 19, 118. [Google Scholar] [CrossRef]
  126. 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 Alpha-Conotoxin PnIA Variant. Nat. Struct. Mol. Biol. 2005, 12, 582–588. [Google Scholar] [CrossRef] [PubMed]
  127. Ulens, C.; Hogg, R.C.; Celie, P.H.; Bertrand, D.; Tsetlin, V.; Smit, A.B.; Sixma, T.K. Structural Determinants of Selective Alpha-Conotoxin Binding to a Nicotinic Acetylcholine Receptor Homolog AChBP. Proc. Natl. Acad. Sci. USA 2006, 103, 3615–3620. [Google Scholar] [CrossRef]
  128. Lin, B.; Xu, M.; Zhu, X.; Wu, Y.; Liu, X.; Zhangsun, D.; Hu, Y.; Xiang, S.-H.; Kasheverov, I.E.; Tsetlin, V.I.; et al. From Crystal Structure of α-Conotoxin GIC in Complex with Ac-AChBP to Molecular Determinants of Its High Selectivity for A3β2 NAChR. Sci. Rep. 2016, 6, 22349. [Google Scholar] [CrossRef] [PubMed]
  129. Yu, J.; Zhu, X.; Zhang, L.; Kudryavtsev, D.; Kasheverov, I.; Lei, Y.; Zhangsun, D.; Tsetlin, V.; Luo, S. Species Specificity of Rat and Human A7 Nicotinic Acetylcholine Receptors towards Different Classes of Peptide and Protein Antagonists. Neuropharmacology 2018, 139, 226–237. [Google Scholar] [CrossRef]
  130. 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 A3β2 Nicotinic Acetylcholine Receptors Is Mediated by Β2 Functionally Important Residues. J. Med. Chem. 2020, 63, 13656–13668. [Google Scholar] [CrossRef]
  131. Zouridakis, M.; Papakyriakou, A.; Ivanov, I.A.; Kasheverov, I.E.; Tsetlin, V.; Tzartos, S.; Giastas, P. Crystal Structure of the Monomeric Extracellular Domain of A9 Nicotinic Receptor Subunit in Complex With α-Conotoxin RgIA: Molecular Dynamics Insights Into RgIA Binding to A9α10 Nicotinic Receptors. Front. Pharmacol. 2019, 10, 474. [Google Scholar] [CrossRef]
  132. Romero, H.K.; Christensen, S.B.; Di Cesare Mannelli, L.; Gajewiak, J.; Ramachandra, R.; Elmslie, K.S.; Vetter, D.E.; Ghelardini, C.; Iadonato, S.P.; Mercado, J.L.; et al. Inhibition of A9α10 Nicotinic Acetylcholine Receptors Prevents Chemotherapy-Induced Neuropathic Pain. Proc. Natl. Acad. Sci. USA 2017, 114, E1825–E1832. [Google Scholar] [CrossRef] [PubMed]
  133. Zheng, N.; Christensen, S.B.; Dowell, C.; Purushottam, L.; Skalicky, J.J.; McIntosh, J.M.; Chou, D.H. Discovery of Methylene Thioacetal-Incorporated α-RgIA Analogues as Potent and Stable Antagonists of the Human A9α10 Nicotinic Acetylcholine Receptor for the Treatment of Neuropathic Pain. J. Med. Chem. 2021, 64, 9513–9524. [Google Scholar] [CrossRef] [PubMed]
  134. 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] [PubMed]
  135. Dyachenko, I.A.; Palikova, Y.A.; Palikov, V.A.; Korolkova, Y.V.; Kazakov, V.A.; Egorova, N.S.; Garifulina, A.I.; Utkin, Y.N.; Tsetlin, V.I.; Kryukova, E.V. α-Conotoxin RgIA and Oligoarginine R8 in the Mice Model Alleviate Long-Term Oxaliplatin Induced Neuropathy. Biochimie 2022, 194, 127–136. [Google Scholar] [CrossRef]
  136. Gladkikh, I.N.; Klimovich, A.A.; Kalina, R.S.; Kozhevnikova, Y.V.; Khasanov, T.A.; Osmakov, D.I.; Koshelev, S.G.; Monastyrnaya, M.M.; Andreev, Y.A.; Leychenko, E.V.; et al. Anxiolytic, Analgesic and Anti-Inflammatory Effects of Peptides Hmg 1b-2 and Hmg 1b-4 from the Sea Anemone Heteractis Magnifica. Toxins 2023, 15, 341. [Google Scholar] [CrossRef]
  137. Kalina, R.; Gladkikh, I.; Dmitrenok, P.; Chernikov, O.; Koshelev, S.; Kvetkina, A.; Kozlov, S.; Kozlovskaya, E.; Monastyrnaya, M. New APETx-like Peptides from Sea Anemone Heteractis Crispa Modulate ASIC1a Channels. Peptides 2018, 104, 41–49. [Google Scholar] [CrossRef]
  138. Kalina, R.S.; Koshelev, S.G.; Zelepuga, E.A.; Kim, N.Y.; Kozlov, S.A.; Kozlovskaya, E.P.; Monastyrnaya, M.M.; Gladkikh, I.N. APETx-Like Peptides from the Sea Anemone Heteractis Crispa, Diverse in Their Effect on ASIC1a and ASIC3 Ion Channels. Toxins 2020, 12, 266. [Google Scholar] [CrossRef]
  139. Kalina, R.S.; Kasheverov, I.E.; Koshelev, S.G.; Sintsova, O.V.; Peigneur, S.; Pinheiro-Junior, E.L.; Popov, R.S.; Chausova, V.E.; Monastyrnaya, M.M.; Dmitrenok, P.S.; et al. Nicotinic Acetylcholine Receptors Are Novel Targets of APETx-like Toxins from the Sea Anemone Heteractis Magnifica. Toxins 2022, 14, 697. [Google Scholar] [CrossRef]
  140. Andreev, Y.A.; Kozlov, S.A.; Koshelev, S.G.; Ivanova, E.A.; Monastyrnaya, M.M.; Kozlovskaya, E.P.; Grishin, E.V. Analgesic Compound from Sea Anemone Heteractis Crispa Is the First Polypeptide Inhibitor of Vanilloid Receptor 1 (TRPV1). J. Biol. Chem. 2008, 283, 23914–23921. [Google Scholar] [CrossRef]
  141. Kozlov, S.A.; Andreev, I.A.; Murashev, A.N.; Skobtsov, D.I.; D’iachenko, I.A.; Grishin, E.V. [New polypeptide components from the Heteractis crispa sea anemone with analgesic activity]. Bioorg. Khim. 2009, 35, 789–798. [Google Scholar] [CrossRef]
  142. Logashina, Y.A.; Mosharova, I.V.; Korolkova, Y.V.; Shelukhina, I.V.; Dyachenko, I.A.; Palikov, V.A.; Palikova, Y.A.; Murashev, A.N.; Kozlov, S.A.; Stensvåg, K.; et al. Peptide from Sea Anemone Metridium Senile Affects Transient Receptor Potential Ankyrin-Repeat 1 (TRPA1) Function and Produces Analgesic Effect. J. Biol. Chem. 2017, 292, 2992–3004. [Google Scholar] [CrossRef]
  143. Logashina, Y.A.; Solstad, R.G.; Mineev, K.S.; Korolkova, Y.V.; Mosharova, I.V.; Dyachenko, I.A.; Palikov, V.A.; Palikova, Y.A.; Murashev, A.N.; Arseniev, A.S.; et al. New Disulfide-Stabilized Fold Provides Sea Anemone Peptide to Exhibit Both Antimicrobial and TRPA1 Potentiating Properties. Toxins 2017, 9, 154. [Google Scholar] [CrossRef] [PubMed]
  144. Osmakov, D.I.; Kozlov, S.A.; Andreev, Y.A.; Koshelev, S.G.; Sanamyan, N.P.; Sanamyan, K.E.; Dyachenko, I.A.; Bondarenko, D.A.; Murashev, A.N.; Mineev, K.S.; et al. Sea Anemone Peptide with Uncommon β-Hairpin Structure Inhibits Acid-Sensing Ion Channel 3 (ASIC3) and Reveals Analgesic Activity. J. Biol. Chem. 2013, 288, 23116–23127. [Google Scholar] [CrossRef] [PubMed]
  145. Andreev, Y.A.; Osmakov, D.I.; Koshelev, S.G.; Maleeva, E.E.; Logashina, Y.A.; Palikov, V.A.; Palikova, Y.A.; Dyachenko, I.A.; Kozlov, S.A. Analgesic Activity of Acid-Sensing Ion Channel 3 (ASIC3) Inhibitors: Sea Anemones Peptides Ugr9-1 and APETx2 versus Low Molecular Weight Compounds. Mar. Drugs 2018, 16, 500. [Google Scholar] [CrossRef]
  146. Sintsova, O.; Gladkikh, I.; Klimovich, A.; Palikova, Y.; Palikov, V.; Styshova, O.; Monastyrnaya, M.; Dyachenko, I.; Kozlov, S.; Leychenko, E. TRPV1 Blocker HCRG21 Suppresses TNF-α Production and Prevents the Development of Edema and Hypersensitivity in Carrageenan-Induced Acute Local Inflammation. Biomedicines 2021, 9, 716. [Google Scholar] [CrossRef]
  147. Antonov, S.M.; Grishin, E.V.; Magazanik, L.G.; Shupliakov, O.V.; Vesselkin, N.P.; Volkova, T.M. Argiopin Blocks the Glutamate Responses and Sensorimotor Transmission in Motoneurones of Isolated Frog Spinal Cord. Neurosci. Lett. 1987, 83, 179–184. [Google Scholar] [CrossRef]
  148. Kiskin, N.I.; Kryshtal, O.A.; Tsyndrenko, A.; Volkova, T.M.; Grishin, E.V. [Argiopine, Argiopinines and Pseudoargiopinines--Blockers of the Glutamate Receptors in Hippocampal Neurons]. Neirofiziologiia Neurophysiol. 1989, 21, 748–756. [Google Scholar] [CrossRef]
  149. Twomey, E.C.; Yelshanskaya, M.V.; Vassilevski, A.A.; Sobolevsky, A.I. Mechanisms of Channel Block in Calcium-Permeable AMPA Receptors. Neuron 2018, 99, 956–968.e4. [Google Scholar] [CrossRef] [PubMed]
  150. Grishin, E.V.; Savchenko, G.A.; Vassilevski, A.A.; Korolkova, Y.V.; Boychuk, Y.A.; Viatchenko-Karpinski, V.Y.; Nadezhdin, K.D.; Arseniev, A.S.; Pluzhnikov, K.A.; Kulyk, V.B.; et al. Novel Peptide from Spider Venom Inhibits P2X3 Receptors and Inflammatory Pain. Ann. Neurol. 2010, 67, 680–683. [Google Scholar] [CrossRef]
  151. Männikkö, R.; Shenkarev, Z.O.; Thor, M.G.; Berkut, A.A.; Myshkin, M.Y.; Paramonov, A.S.; Kulbatskii, D.S.; Kuzmin, D.A.; Sampedro Castañeda, M.; King, L.; et al. Spider Toxin Inhibits Gating Pore Currents Underlying Periodic Paralysis. Proc. Natl. Acad. Sci. USA 2018, 115, 4495–4500. [Google Scholar] [CrossRef]
  152. Dudina, E.E.; Korolkova, Y.V.; Bocharova, N.E.; Koshelev, S.G.; Egorov, T.A.; Huys, I.; Tytgat, J.; Grishin, E.V. OsK2, a New Selective Inhibitor of Kv1.2 Potassium Channels Purified from the Venom of the Scorpion Orthochirus Scrobiculosus. Biochem. Biophys. Res. Commun. 2001, 286, 841–847. [Google Scholar] [CrossRef]
  153. Kuzmenkov, A.I.; Nekrasova, O.V.; Peigneur, S.; Tabakmakher, V.M.; Gigolaev, A.M.; Fradkov, A.F.; Kudryashova, K.S.; Chugunov, A.O.; Efremov, R.G.; Tytgat, J.; et al. KV1.2 Channel-Specific Blocker from Mesobuthus Eupeus Scorpion Venom: Structural Basis of Selectivity. Neuropharmacology 2018, 143, 228–238. [Google Scholar] [CrossRef]
  154. Kuzmenkov, A.I.; Nekrasova, O.V.; Kudryashova, K.S.; Peigneur, S.; Tytgat, J.; Stepanov, A.V.; Kirpichnikov, M.P.; Grishin, E.V.; Feofanov, A.V.; Vassilevski, A.A. Fluorescent Protein-Scorpion Toxin Chimera Is a Convenient Molecular Tool for Studies of Potassium Channels. Sci. Rep. 2016, 6, 33314. [Google Scholar] [CrossRef]
  155. Billen, B.; Vassilevski, A.; Nikolsky, A.; Debaveye, S.; Tytgat, J.; Grishin, E. Unique Bell-Shaped Voltage-Dependent Modulation of Na+ Channel Gating by Novel Insect-Selective Toxins from the Spider Agelena Orientalis. J. Biol. Chem. 2010, 285, 18545–18554. [Google Scholar] [CrossRef] [PubMed]
  156. Mikov, A.N.; Fedorova, I.M.; Potapieva, N.N.; Maleeva, E.E.; Andreev, Y.A.; Zaitsev, A.V.; Kim, K.K.; Bocharov, E.V.; Bozin, T.N.; Altukhov, D.A.; et al. ω-Tbo-IT1-New Inhibitor of Insect Calcium Channels Isolated from Spider Venom. Sci. Rep. 2015, 5, 17232. [Google Scholar] [CrossRef]
  157. Dubovskii, P.V.; Lesovoy, D.M.; Dubinnyi, M.A.; Konshina, A.G.; Utkin, Y.N.; Efremov, R.G.; Arseniev, A.S. Interaction of Three-Finger Toxins with Phospholipid Membranes: Comparison of S- and P-Type Cytotoxins. Biochem. J. 2005, 387, 807–815. [Google Scholar] [CrossRef] [PubMed]
  158. Filippov, A.K.; Kozlov, S.A.; Pluzhnikov, K.A.; Grishin, E.V.; Brown, D.A. M-Type K+ Current Inhibition by a Toxin Fron the Scorpion Buthus Eupeus. FEBS Lett. 1996, 384, 277–280. [Google Scholar] [CrossRef] [PubMed]
  159. Grishin, E.V.; Volkova, T.M.; Arseniev, A.S. Isolation and Structure Analysis of Components from Venom of the Spider Argiope Lobata. Toxicon 1989, 27, 541–549. [Google Scholar] [CrossRef]
  160. Dubovskii, P.V.; Volynsky, P.E.; Polyansky, A.A.; Karpunin, D.V.; Chupin, V.V.; Efremov, R.G.; Arseniev, A.S. Three-Dimensional Structure/Hydrophobicity of Latarcins Specifies Their Mode of Membrane Activity. Biochemistry 2008, 47, 3525–3533. [Google Scholar] [CrossRef]
  161. Dubovskii, P.V.; Volynsky, P.E.; Polyansky, A.A.; Chupin, V.V.; Efremov, R.G.; Arseniev, A.S. Spatial Structure and Activity Mechanism of a Novel Spider Antimicrobial Peptide. Biochemistry 2006, 45, 10759–10767. [Google Scholar] [CrossRef] [PubMed]
  162. Kudryashova, K.S.; Nekrasova, O.V.; Kuzmenkov, A.I.; Vassilevski, A.A.; Ignatova, A.A.; Korolkova, Y.V.; Grishin, E.V.; Kirpichnikov, M.P.; Feofanov, A.V. Fluorescent System Based on Bacterial Expression of Hybrid KcsA Channels Designed for Kv1.3 Ligand Screening and Study. Anal. Bioanal. Chem. 2013, 405, 2379–2389. [Google Scholar] [CrossRef] [PubMed]
  163. Kryukova, E.V.; Ivanov, I.A.; Lebedev, D.S.; Spirova, E.N.; Egorova, N.S.; Zouridakis, M.; Kasheverov, I.E.; Tzartos, S.J.; Tsetlin, V.I. Orthosteric and/or Allosteric Binding of α-Conotoxins to Nicotinic Acetylcholine Receptors and Their Models. Mar. Drugs 2018, 16, 460. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Selected amino acid sequences of animal toxins discovered at IBCh.
Figure 1. Selected amino acid sequences of animal toxins discovered at IBCh.
Ijms 24 13884 g001
Figure 2. NMR structures of selected representatives of animal toxins obtained in different years at IBCh. (A) Short-type neurotoxin NTII from cobra N. oxiana (PDB ID: 2MJ4); (B) [P33A]-mutant of non-conventional toxin WTX from cobra N. kaouthia (2MJ0); (C) dimeric form of long-type α-cobratoxin from cobra N. kaouthia (4AEA); (D) OSK1 from scorpion O. scrobiculosus (1SCO); (E) Ms11a-3 from sea anemone M. senile (6XYI); (F) purotoxin-2 from spider A. marikovskyi (2MZF); (G) apamin from bee A. mellifera (7OXF); (H) α-conotoxin ImI from C. imperialis (1IMI). Disulfides in all toxins are shown in red. Roman numerals indicate loop numbers in TFTs: I—N-terminal loop, II—central loop, III—C-terminal loop.
Figure 2. NMR structures of selected representatives of animal toxins obtained in different years at IBCh. (A) Short-type neurotoxin NTII from cobra N. oxiana (PDB ID: 2MJ4); (B) [P33A]-mutant of non-conventional toxin WTX from cobra N. kaouthia (2MJ0); (C) dimeric form of long-type α-cobratoxin from cobra N. kaouthia (4AEA); (D) OSK1 from scorpion O. scrobiculosus (1SCO); (E) Ms11a-3 from sea anemone M. senile (6XYI); (F) purotoxin-2 from spider A. marikovskyi (2MZF); (G) apamin from bee A. mellifera (7OXF); (H) α-conotoxin ImI from C. imperialis (1IMI). Disulfides in all toxins are shown in red. Roman numerals indicate loop numbers in TFTs: I—N-terminal loop, II—central loop, III—C-terminal loop.
Ijms 24 13884 g002
Figure 3. Three-dimensional structures of toxin–receptor complexes in the determination of which scientists from IBCh took part. (A,B)—side and top views, respectively, of X-ray structure of Aplysia californica AChBP in complex with α-conotoxin PnIA[L10,K14] (PDB ID: 2BR8); (C,D)—side and top views, respectively, of X-ray structure of extracellular domain of human α9 nAChR in complex with α-conotoxin RgIA (6HY7); (E)—side view of cryo-electron structure of calcium-permeable constructed AMPA receptor (GluA2Q-STZ) in complex with argiopin, also known as argiotoxin-636 (AgTx-636) (6O9G). All toxins are shown in red. Identical five AChBP protomers as well as four AMPA receptor subunits are colored in different variants of green for clarity.
Figure 3. Three-dimensional structures of toxin–receptor complexes in the determination of which scientists from IBCh took part. (A,B)—side and top views, respectively, of X-ray structure of Aplysia californica AChBP in complex with α-conotoxin PnIA[L10,K14] (PDB ID: 2BR8); (C,D)—side and top views, respectively, of X-ray structure of extracellular domain of human α9 nAChR in complex with α-conotoxin RgIA (6HY7); (E)—side view of cryo-electron structure of calcium-permeable constructed AMPA receptor (GluA2Q-STZ) in complex with argiopin, also known as argiotoxin-636 (AgTx-636) (6O9G). All toxins are shown in red. Identical five AChBP protomers as well as four AMPA receptor subunits are colored in different variants of green for clarity.
Ijms 24 13884 g003
Table 1. Some representatives of animal toxins discovered and investigated at IBCh.
Table 1. Some representatives of animal toxins discovered and investigated at IBCh.
Toxin(s)SpeciesStructure TypeStructural and Functional Characterization:
Sequence/Spatial Structure/Functional Characterization
UniProt CodeYear of DiscoveryRef.
Snake venom-derived toxins
Neurotoxin IINaja oxianaTFT: Short-type α-neurotoxin+/+/+3S11_NAJOX1973[3,31,34,36,69]
Neurotoxin IN. oxianaTFT: Long-type α-neurotoxin+/−/+3L21_NAJOX1974[4,69,71]
CytotoxinsN. oxiana, N. kaouthiaTFT+/+/+3SA1(2)_NAJOX, 3SA3(7A,8)_NAJKA1974–2022[5,6,9,38,74,75,76,77,78,79,80,81,82,83,157]
Muscarinic toxin-like protein 1N. kaouthia, N. melanoleucaTFT+/−/+3SUC1_NAJKA, 3SUC1_NAJME2000–2021[8,71]
WTXN. kaouthiaNon-conventional TFT+/+/+3NO2_NAJKA2001[7,35,37,69,84,85,86]
Homodimer of α-cobratoxinN. kaouthiaTFT: Dimer of long-type α-neurotoxin+/+/+3L21_NAJKA2008[38,39]
Heterodimeric neurotoxic phospholipase A2 Vipera nikolskiiSecreted PLA2 of group II+/−/+PA2B1(2)_VIPBN
PA2H_VIPBN
2008[10]
Phospholipases A2V. ursinii renardi, Bungarus fasciatus, Bitis arietans,
N. oxiana, N. haje
Secreted PLA2 of group I and II+/−/+PA2A2(A3,B)_VIPRE PA2B_BITAR PA2B6_BUNFA PA2TI_NAJHH1977–2011[16,93,94]
AzemiopsinAzemiops feaeLinear peptide+/−/+AON_AZEFE2012[11,101]
αδ-Bungarotoxin (αδ-BgTx-1)Bungarus candidusTFT: Long-type α-neurotoxin+/−/+3L21_BUNCA2019[72]
Bradykinin-
potentiating peptides
A. feaeLinear peptide+/−/+No code2020[12]
Tx-NM2, Tx-NM3-1 N. melanoleucaTFT: Long-type α-neurotoxin+/−/+ 3L23(24)_NAJME 2021[71]
Scorpion venom-derived toxins
Neurotoxin BeKm-1Mesobuthus eupeusα-helix and a triple-stranded antiparallel β-sheet+/+/+KGX21_MESEU1996[45,158]
Spider venom-derived toxins
β/δ-Agatoxin-5Agelena orientalisICK+/−/+T5G1A_AGEOR2010[155]
Argiopin, argiopinins, and pseudoargiopininsArgiope lobataAcylpolyamines+/+/+CID: 122294
189479
189486 *
1989[159]
Latarcin-1Lachesana tarabaeviLinear peptide+/+/+LAT1_LACTA2006[102,160]
Latarcin-2aL. tarabaeviLinear peptide+/+/+LAT2A_LACTA2006[102,161]
Purotoxin-1 (PT1)Alopecosa marikovskyiICK+/+/+TXPR1_ALOMR2010[150]
α-Latrototoxin-Lt1aLatrodectus mactans tredecimguttatusMultidomain organization+/−/+LATA_LATTR1990[27]
α-Latroinsectotoxin-
Lt1a
L. mactans tredecimguttatusMultidomain organization+/−/+LITA_LATTR1993[28]
Sea anemone venom-derived toxins
Analgesic polypeptide HC1 (τPI-SHTX-Hcr2b) Heteractis crispaKunitz-type+/−/+VKT2B_HETCR2008[140]
π-AnmTX Ugr 9a-1Urticina grebelnyiβ-hairpin structure+/+/+TX9A_URTGR2013[144]
τ-AnmTX Ms 9a-1Metridium senileβ-hairpin structure+/−/+TX91O_METSE2017[142]
Ms11a-1/4M. senileICK+/+/+No code2023[21]
* These are references to the structures in PubChem. The toxins are of non-peptide nature.
Table 2. Animal toxins and their derivatives developed as tools in IBCh.
Table 2. Animal toxins and their derivatives developed as tools in IBCh.
Native Toxin
(Species)
ModificationsTarget(s)/ActivityTasksRef.
α-Bungarotoxin
(Bungarus multicinctus)
Radiolabeled and fluorescent derivativesαβδγ/ε-, α7-, α9(α10) nAChRs;
AChBPs;
GABA-A
Detection of respective targets in different preparations, cells, tissues
Radioligand in radioligand assay
[7,11,35,38,39,69,71,73]
Other long-type α-neurotoxins Radiolabeled, photoactivatable, and fluorescent derivatives, analogsαβδγ/ε-, α7-, α9(α10) nAChRs;
AChBPs;
GABA-A
Detection of respective targets in different preparations, cells, tissues
Radioligand in radioligand assay
Structure–function characterization
Mapping of ligand-binding sites of T. californica nAChR
[4,34,68,69,71,73]
κ-Bungarotoxin (Bungarus multicinctus) α7-, α3β2 nAChRsDetection of respective targets in different preparations[38]
Short-type α-neurotoxins Radiolabeled and photoactivatable derivatives, analogsαβδγ/ε nAChRsDetection of respective targets in different preparations
Determination of the spatial structure
Structure–function characterization
Mapping of ligand-binding sites of T. californica nAChR
[3,34,65,66,67,69]
Phospholipases A2 Cytotoxicity, antiviral activity, thrombin inhibitor, αβδγ/ε-, α7-nAChRsDetection of respective targets in different preparations, cells, tissues
Revealing the different biological activity
[95,96,97,98,99,100]
CytotoxinsFluorescent derivativesCytotoxicityDetermination of the spatial structure
Determination of the cytotoxicity mechanism
[138]
Azemiopsin
(Azemiops feae)
Ala analogs, fluorescent derivativeαβδγ/ε nAChRsBinding and functional characterization
Structure–function characterization
Preclinical trials as myorelaxant
[11,101]
Agitoxin-2 and OSK-1Toxin fused with fluorescent proteinsVoltage-gated potassium channels (KV)Ligand screening in the spheroplast binding assay[154]
Agitoxin-2Fluorescently labeled derivativesSpheroplasts with the embedded KcsA-Kv1.3 hybrid proteinScreening of prospective compounds recognized by Kv1.3[162]
Toxin PT1Pure pharmacological substanceP2X3 Preclinical trials as analgesic-
APHC3Pure pharmacological substanceTRPV1 Preclinical trials as analgesic-
Different α-conotoxins
(Conidae)
Different analogs, radiolabeled, photoactivatable, and fluorescent derivativesDifferent nAChR subtypes;
AChBPs;
GABA-A
Determination of the spatial structure
Detection of respective targets in different preparations, cells, tissues
Radioligand in radioligand assay
Design of new analogs
Mapping of ligand-binding sites of T. californica nAChR
X-ray studies of AChPB complexes
In vivo studies on neuropathy models
[40,69,72,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,135]
αO-Conotoxin GXIVA
(Conus geographus)
Isomers, radiolabeled derivativeα9(α10) nAChR;
AChBPs
Binding and functional characterization[163]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tsetlin, V.; Shelukhina, I.; Kozlov, S.; Kasheverov, I. Fifty Years of Animal Toxin Research at the Shemyakin–Ovchinnikov Institute of Bioorganic Chemistry RAS. Int. J. Mol. Sci. 2023, 24, 13884. https://doi.org/10.3390/ijms241813884

AMA Style

Tsetlin V, Shelukhina I, Kozlov S, Kasheverov I. Fifty Years of Animal Toxin Research at the Shemyakin–Ovchinnikov Institute of Bioorganic Chemistry RAS. International Journal of Molecular Sciences. 2023; 24(18):13884. https://doi.org/10.3390/ijms241813884

Chicago/Turabian Style

Tsetlin, Victor, Irina Shelukhina, Sergey Kozlov, and Igor Kasheverov. 2023. "Fifty Years of Animal Toxin Research at the Shemyakin–Ovchinnikov Institute of Bioorganic Chemistry RAS" International Journal of Molecular Sciences 24, no. 18: 13884. https://doi.org/10.3390/ijms241813884

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