Next Article in Journal
Computational Modelling Suggests Bacteriostatic Saline Does Not Reverse Botulinum Toxin-Induced Brow Ptosis
Previous Article in Journal
Compromised Regeneration, Damage to Blood Vessels and the Endomysium Underpin Permanent Muscle Damage Induced by Puff Adder (Bitis arietans) Venom
Previous Article in Special Issue
The Therapeutic Potential of Antimicrobial Peptides Isolated from the Skin Secretions of Anurans of the Genus Boana in the Face of the Global Antimicrobial Resistance Crisis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Properties and Pharmacology of Scorpion Toxins and Their Biotechnological Potential in Agriculture and Medicine

by
Cháriston André Dal Belo
1,2,*,
Stephen Hyslop
2 and
Célia Regina Carlini
3,4,*
1
Departamento Multidisciplinar, Escola Paulista de Política, Economia e Negócios (EPPEN), Universidade Federal de São Paulo (UNIFESP), Rua General Newton Estilac Leal, 932, Quitaúna, Osasco 06190-170, SP, Brazil
2
Departamento de Farmacologia, Faculdade de Ciências Médicas, Universidade Estadual de Campinas (UNICAMP), Rua Vital Brazil, 80, Cidade Universitária Zeferino Vaz, Campinas 13083-888, SP, Brazil
3
Programa de Pós-Graduação em Biociências, Universidade Federal de Ciências da Saúde de Porto Alegre (UFCSPA), Rua Sarmento Leite 245, Porto Alegre 90050-170, RS, Brazil
4
Centro de Biotecnologia e Departamento de Bioquímica, Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre 91501-970, RS, Brazil
*
Authors to whom correspondence should be addressed.
Toxins 2025, 17(10), 497; https://doi.org/10.3390/toxins17100497
Submission received: 13 July 2025 / Revised: 23 September 2025 / Accepted: 26 September 2025 / Published: 7 October 2025

Abstract

Scorpion venoms contain a wide range of toxins that interact with a variety of target molecules (ion channels, receptors and enzymes) associated with synaptic transmission, action potential propagation, cardiac function, hemostasis and other physiological systems. Scorpion toxins are also active towards bacteria, viruses, fungi and parasites. Such interactions make scorpion toxins useful lead molecules for developing compounds with biotechnological and therapeutic applications, and as tools for cell biology. In addition, scorpion toxins act as insectotoxins, with promising applications as insecticides. This review describes the range of scorpion toxins and discusses their usefulness for the development of insecticides and therapeutic drugs.
Key Contribution: This article provides a general overview of the variety of scorpion toxins and their potential biotechnological and therapeutic applications.

1. Introduction

Scorpions (phylum Arthropoda, class Arachnida, order Scorpiones) are a highly diverse assemblage of approximately 2900 species classified into 24 families [1]. Scorpions have a worldwide distribution, but are absent from many temperate countries, certain Pacific islands and Antarctica [2]. The greatest diversity of scorpions occurs in tropical and subtropical regions and coincides with the distribution of most cases of scorpionism worldwide [3,4]. This extensive species diversity and wide geographic distribution attest to the biological adaptability and ecological success of scorpions [5]. Figure 1 shows examples of medically important scorpion species from around the world.

2. Scorpion Venoms: An Overview of Composition and Diversity

Scorpion venoms are a complex mixture of metal ions, amino acids and amines (e.g., serotonin), and low molecular mass peptides, neurotoxins, and enzymes such as hyaluronidase, metalloproteinases, and phospholipases A2 (PLA2) [6,7,8,9,10] that serve to capture and subdue prey and to provide defense against predation.
Numerous transcriptomic and proteomic studies over the last 20 years [11,12,13,14,15,16,17] have revealed the complexity of scorpion venoms, particularly with regard to their peptide content [10,18,19,20]. These peptides have been extensively studied because of their structural diversity, their wide-range of pharmacological effects, and their potential to serve as pharmacological tools, such as ion channel [21,22,23,24] and G-protein-coupled receptor [25,26] modulators, as well as inhibitors of enzymatic activity, including angiotensin-converting enzyme (ACE) [27,28,29,30,31,32,33,34] and matrix metalloproteinases [35]. Scorpion peptides have also been investigated as lead compounds for the development of novel therapeutic agents [10,36,37,38,39,40] to treat conditions such as bacterial [41,42,43,44,45] and viral [46,47] infections, cancer [47,48,49,50,51,52], pain [53,54], inflammation [39,55], diabetes [56] and hypertension [25,57,58]. In addition to peptides, scorpion venom enzymes, such as PLA2 [59,60] and hyaluronidases [61] have been studied for potential therapeutic applications.

3. Evolving Paradigms in Scorpion Toxin Classification

3.1. General Classification of Toxin Families

Historically, scorpion toxins have been classified based on their interactions with ion channels, particularly voltage-gated sodium (Nav), potassium (Kv), calcium (Cav), and chloride (ClC) channels, as a reflection of the central roles of these channels in neurophysiology and venom pathology, and also based on their primary amino acid sequences [7,19,20,21,22,23,24]. While this general approach is still applied today, the advent of high-throughput sequencing, structural biology techniques, and, more recently, artificial intelligence [62] has greatly expanded our understanding of toxin diversity such that scorpion toxins can now be classified based not only on their primary sequences, but also on their three-dimensional folding patterns and the conservation of key functional residues [54]. For instance, while many scorpion toxins have an α-helix/β-sheet fold configuration, variations in their disulfide bridge patterns and loop regions dictate their stability and their precise binding sites and pharmacological profiles [19]. This allows for the identification of new subfamilies within previously established groups and reveals a finer level of functional specialization, even among toxins targeting the same ion channel. This approach is particularly valuable for uncovering toxins with novel mechanisms of action or unusual target specificities. The primary general classification of scorpion peptide toxins includes (1) Disulfide-bridged peptides (DBP) and (2) Non-disulfide-bridged peptides (NDBP), although these venoms also contain peptides with other important biological functions [10]. Figure 2 provides a general classification of the principal components of scorpion venoms that includes proteins, peptides, and low molecular mass compounds.

3.2. General Organization of Scorpion Peptide Genes

Scorpion venom diversity is very high (estimated to be ~300,000 components across all species), with peptides being the principal components involved [63]. However, compared to the extensive biochemical and pharmacological characterization of numerous purified toxins, much less is known of scorpion toxins at the genomic level and of how variations in gene expression can influence venom composition and activity. While transcriptomic and proteomic studies have provided important insights into the general composition of scorpion venoms [11,12,13,14,15,16,17], detailed analyses of scorpion toxin gene organization and expression, and the possible types and mechanisms of post-translational modifications, have also been helpful in explaining the high diversity of toxins in these venoms at the genic level and the factors that influence this.
Scorpion peptide genes have a basic minimal organization consisting of coding regions for a signal peptide and the mature peptide of interest, often with the presence of introns in the signal peptide and mature peptide coding regions [64,65,66,67,68,69]. In some cases, a pro-peptide may also be coded [66,68,69], and such peptides, once removed, may persist in the venom, although it is unclear whether they exert any biological function [11,70]. Peptide toxins are generally coded for by their own specific genes rather than resulting from the post-translational processing of a large precursor protein to release mature peptides [64,65,66,68]. Specific events that can affect toxin diversity, in terms of sequence and biological activities, include single point mutations, deletions, insertions, alternative- and trans-splicing events, gene duplications, and post-translational modifications [17,67,71,72,73]. These mechanisms can lead to variations in venom peptide content and molecular size that can be detected by chromatographic analyses (especially RP-HPLC), mass spectrometry, immunoassays and other biochemical approaches, as well as biological assays.
Despite considerable advances in our knowledge of the structure and function of a wide variety of scorpion toxins in recent decades, for a much larger number of these toxins their phylogenetic and structural relationships with specific peptide or protein families, the occurrence of non-toxic homologs, their biological activities towards prey species, and their precise contribution to envenomation remain unclear. For the main toxin families discussed here, further details on their gene organization, protein sequences and structures, and biological activities are available through the UniProt Protein Family (PF) or Prosite (PS) codes indicated alongside the corresponding toxin groups, and from pertinent publications [12,64,65,66,68,69,74,75,76,77]. All peptide and protein sizes (indicated either as the number of amino acids or the molecular mass) mentioned in the main text and tables refer to the final size of the mature molecules, and not to precursor forms.

3.3. Scorpion Disulfide-Bridged Peptides (DBPs)

3.3.1. Cystine-Stabilized α/β (CSα/β) Scaffold

This is the most prevalent and functionally diverse structural category of scorpion peptides and involves an α-helix connected to a β-sheet (often triple-stranded), stabilized by multiple disulfide bonds [78]. Most classic neurotoxins that target ion channels fall into this group. These molecules contain two completely conserved disulfide bonds at the Ci–Cj and Ci + 4–Cj + 2 positions, although some of them also exhibit an extra link connecting the two ends of the peptide chain [79]. These conserved disulfide bridges are vital for the structural integrity and biological activity of these peptides. By locking the structure into a precise and rigid three-dimensional conformation, these covalent bonds enable the peptide to effectively bind to its specific ion channel target and produce its characteristic pharmacological effects [80]. All scorpion peptides containing CS α/β motifs act in a similar way, namely, by interacting with ion channels to block or modulate normal channel function [36,81]. Members of this superfamily can be subdivided into long or short scorpion toxins, depending on their respective structures.
Long-Chain Scorpion Toxins (Protein Family PF14866)
These peptides are responsible for the severe neurotoxic effects observed in scorpion envenomation, primarily by disrupting the function of voltage-gated ion channels, particularly Nav channels [36]. Their potency and specificity have made these toxins invaluable tools in neuropharmacology and drug discovery [82]. The typical gene organization for scorpion Na+ channel toxins (NaTx) includes two exons (exons 1 and 2) separated by an intron [64,65,66]. Exon 1 contains the 5′ non-coding region (untranscribed region, UTR) and the N-terminal region of the signal peptide, whereas exon 2 consists of the C-terminal of the signal peptide, the region coding the mature peptide, and the 3′ non-coding region. The overall premRNA size is 650–950 nucleotides with an intron of 300–620 nucleotides. Processing of this premRNA results in mature mRNA of 300–350 nucleotides that is translated to produce peptides of 82–96 amino acids from which the signal peptide (18–21 amino acids) is subsequently removed [64]. Further processing results in mature toxins of 60–76 amino acid residues in length [20,21,22,23,36,83,84]. The structural integrity and biological activity of NaTx are critically dependent on a conserved three-dimensional fold stabilized by three or, more commonly, four disulfide bridges [59]. The precise arrangement of these disulfide bonds and the configuration of specific loops within this fold dictate their specificity for different ion channel subtypes and their distinct mechanisms of action. Based on their electrophysiological effects on Nav, long-chain scorpion neurotoxins are broadly divided into two main functional classes, namely, α- and β-toxins. While both of these classes ultimately lead to neuronal hyperexcitability, they achieve this by binding to distinct receptor sites on the Nav channel.
Alpha-toxins, often referred to as classic α-toxins, bind to receptor site 3 on the extracellular surface of Nav channels, specifically interacting with elements in domains I and IV. Their primary effect is to inhibit or significantly slow the fast inactivation of these channels, thereby prolonging the open state of the channels, resulting in a persistent inward sodium current [85]. Physiologically, this action results in extended action potential duration and repetitive firing of neurons that causes hyperexcitability, muscle spasms, and paralysis [54]. AaH II from Androctonus australis Hector [86], BmK1 from Buthus martensi Karsch [87], CvIV4 from Centruroides vittatus [88], Lqh αIT from Leiurus quinquestriatus hebraeus [89] and Lqq III from Leiurus quinquestriatus quinquestriatus [90] are examples of scorpion α-toxins.
Beta-toxins bind to receptor site 4 of the Nav channel, located within domain II. The primary effect of these toxins is to shift the voltage-dependence of activation to more negative (hyperpolarizing) potentials [91]. This means the Nav channels become easier to open, with opening being possible even at resting membrane potentials, when the channels would normally be closed. This leads to spontaneous and repetitive action potentials and an increased excitability of nerve and muscle cells. Some β-toxins can also reduce the peak sodium current. Ts1 [92], Ts2 [93] and TsTX-I [94] from Tityus serrulatus, Cn2 [95] and Cn12 [96] from Centruroides noxius, BmK IT2 [97] from Buthus martensii Karsch, and Lqh-dprIT(3) [98] from Leiurus quinquestriatus hebraeus are representatives of scorpion β-toxins. However, not all NaTxs fit neatly into this two-category system, e.g., AaH IT4, from Androctonus australis hector, that exerts both α and β-NaTx effects [79].
Short-Chain Scorpion Toxins (Protein Family PF00451)
Short-chain scorpion toxins, also known as K+ channel toxins (KTx), represent a diverse group of scorpion venom peptides, generally 30–40 amino acids long and typically containing 3–4 disulfide bonds [20,21,36,99]. The overall gene organization for KTx is similar to that for NaTx in that it consists of a 5′-UTR, a signal peptide coding region, a mature peptide coding region and a 3′-UTR [12,64,65,68]. Introns are frequently observed in the C-terminal portion of the signal peptide coding region and can vary in size from relatively small, e.g., 78–94 bp for BmKTX, BmTX1, and BmTX2 [75] and BmP01, BmP03 and BmP05 [74], to quite large (500 bp) for BmKαTx11 and BmKαTx15 [100]. However, there is considerable variation in the location of introns among the different types of KTx. For example, the genes for α-KTx, κ-KTx, and γ-KTx, such as BmKTX (from B. martensii), BmKK7 (B. martensii), and HeTx203 (Heterometrus spinifer), respectively, all have a single intron that interrupts the signal peptide coding region [68]. For β-KTx, the signal peptide coding region may contain two introns, e.g., TtrKIK (Tityus trivittatus), or none; in the latter case, the intron (887 bp) is located in the region coding for the mature peptide, e.g., BmTXKβ2 (B. martensii), while for δ-KTx (toxins with a Kunitz-type fold) that also block KTx, such as LmKTT-1a (Lychas mucronatus) and BmKTT-2 (B. martensii), there are two introns in the mature peptide coding region and none in the signal peptide coding region [66,68]. In all cases, removal of the signal peptide (~20–28 amino acids) results in mature peptides ranging in size from 23 to 75 amino acids, depending on the subclass of KTx involved (Table 1) [12,68,74,75,76,100]. Some of these peptides may subsequently undergo post-translational modifications at the N-terminal, e.g., Gln modified to pyro-Glu in BmTX1 and BmTX2, and C-terminal amidation following removal of the terminal Gly in BmKTX [75].
KTxs exert a wide range of pharmacological activities through their ability to selectively block different K+ channel subtypes [101], including voltage-gated K+ channels (Kv) [102] and Ca2+-activated K+ channels (KCa) [103]. With the exception of the peptide MeuTXKβ1 from Mesobuthus eupeus venom [104], no other scorpion toxins capable of modulating two-pore-domain potassium (K2P) channels have been identified [105]. Several families and subfamilies of KTxs are recognized, based on their amino acid sequences, structure, and specific ion channel targets [23], with the main subfamilies being α, β, γ, κ, δ, λ, and ε-KTx [63,101,106,107,108].
Alpha-KTx, which generally contain 23–42 amino acid residues and three or four disulfide bridges, are potent blockers of Kv channels, with high selectivity for the subtypes Kv1.1, Kv1.2 and Kv1.3 [102,109]. Beta-KTx consist of approximately 45–75 amino acid residues and two disulfide bridges [106,110] and also target various Kv channels, including Kv1.3, Kv4.2, Kv4.3, and Kv1.7 [81]. However, studying the modulation of Kv by these toxins is challenging because of their cytolytic effect [111]. Gamma-KTxs, with 35–45 amino acid residues and 3–4 disulfide bridges [106], occur in the genera Centruroides, Mesobuthus, and Buthus, and mainly block human Ether-à-go-go-Related Gene (hERG) channels [112,113,114]. Toxins with cysteine-stabilized helix-loop-helix folding (CSα/α) are classified as part of the κ-KTx family (25–30 residues and two disulfide bridges) [115,116]. Delta-KTx, also known as Kunitz toxins, have about 60–70 residues and 3–4 disulfide bridges with a double-stranded antiparallel β-sheet flanked by an α-helix in both the C-terminal and N-terminal segments [117]. These toxins modulate both protease and K+ channel activities [117,118,119]. Finally, the λ-KTx (35–40 residues, three disulfide bridges) and ε-KTx (29 residues, four disulfide bridges) families are characterized by a shared inhibitory cystine knot (ICK) motif [107]. The ε-KTx subfamily includes Ts11 and Ts12 isolated from Tityus serrulatus venom [107]. Although Ts11 shows less than 50% identity with KTxs from other subfamilies, it shares the ICK motif found in λ-KTxs, but is distinguished from λ-KTxs by possessing four disulfide bonds, whereas λ-KTxs have only three [23]. Table 1 summarizes the general properties of short-chain (KTx) toxins. Overall, the diversity of K+ channel subtypes and the specificity of the scorpion toxins acting on them indicates that these molecules have considerable therapeutic potential [21].
Table 1. Comparison of the main characteristics of the subfamilies of scorpion short-chain potassium channel-blocking toxins (KTxs).
Table 1. Comparison of the main characteristics of the subfamilies of scorpion short-chain potassium channel-blocking toxins (KTxs).
KTx SubfamilyLength (Amino Acids)Disulfide BridgesKey Structural FeaturesPrimary Target/FunctionNoteworthy CharacteristicsReferences
General Short KTxs23–643–4-K+ channel blockade-[101]
α-KTx23–423 or 4CS α/β motif (α-helix + β-sheet)K+ channel blockadeLargest subgroup of short scorpion toxins.[63,120]
β-KTx50–75-CS α/β motifK+ channel blockadeComprises longer chain peptides within the short toxin family.[63,110]
γ-KTx--CS α/β motifHuman Ether-à-go-go-Related Gene (hERG) channelsFound in the genera Centruroides, Mesobuthus, and Buthus.[112,121]
κ-KTx-2Two parallel short α-helices connected by a β-turnK+ channel blockadeInteraction with K+ channels similar to α-KTx, despite structural difference.[115]
δ-KTx59–703Kunitz-type structural fold (double-stranded antiparallel β-sheet flanked by α-helix)K+ channel blockade; Serine protease inhibitionExerts antiprotease and K+ channel-blocking properties.[68]
λ-KTx29–493Inhibitor cystine knot (ICK) motif (triple-stranded antiparallel β-sheet)-Related to calcins[122,123]
ε-KTx-4ICK motif (unique pattern)-Only two members (Ts11 and Ts12 from T. serrulatus venom); Ts11 shows <50% identity with other KTxs.[23,107]

3.3.2. Calcium Channel-Modulating Peptides (Calcins) (Protein Family PF08099)

Calcium channel-modulating peptides or calcins are relatively short peptides, typically containing 33–35 amino acids [124], that primarily affect ryanodine receptors (RyRs), intracellular calcium channels found in the endoplasmic/sarcoplasmic reticulum membrane [19,23]. Thus, whereas many scorpion toxins target Kv or KCa channels in the plasma membrane, calcins specifically target intracellular RyRs to modulate calcium release from internal stores rather than directly affecting the membrane potential or action potential firing as do KTx [23]. Well-known examples of calcins include imperacalcin (imperatoxin A) from Pandinus imperator (the Emperor scorpion) [125], maurocalcin from Scorpio maurus palmatus [126], hemicalcin from Hemiscorpius lepturus [127], hadrucalcin from Hadrurus gertschi [128,129], opicalcin from Opistophthalmus carinatus (African yellow leg scorpion) [124], urocalcin from Urodacus yaschenkoi [109,130] and vejocalcin from Vaejovis mexicanus [131].
Few calcins have been studied at the genic level. Zhijian et al. [132] reported that the gene for BmCa1, a ryanodine Ca2+ channel toxin, consists of three exons separated by two introns, with one of the introns (72 bp) located at the end of the signal peptide coding region and the other much larger intron (1076 bp) located within the mature peptide coding region; both introns begin with GT and end with AC, an arrangement typical of other scorpion introns. Removal of the 27-amino acid signal peptide from the precursor protein yields the mature peptide with 37 amino acids and three disulfide bonds. This gene organization is similar to that of opicalcins (opicalcine 1 and 2), from Opistophthalmus carinatus [133]. In this case, the gene consists of three exons each coding for a specific region (signal peptide, propeptide, and mature peptide) and two intros (487 and 544 bp, located at the N-terminal portion of the propeptide and mature peptide regions, respectively). A cDNA for opicalcin coded for a 22 amino acid signal peptide, an 11 amino acid propeptide rich in negatively charged amino acids, and a 33 amino acid mature peptide. The gene organization described above is shared by a variety of calcins and other scorpion toxins with an ICK motif (see below) [122]. In contrast to this gene organization (and to that for NaTx, KTx and ClTx-like peptides), there are no introns in the genomic DNA sequence of the depressant insect toxin BmK AS (and its homolog BmK AS-1), which acts on skeletal muscle ryanodine receptors [134]. In this case, the region coding for the 19 amino acid signal peptide sequence is followed immediately by that coding for the 66 amino acid mature peptide.
Structurally, calcins are characterized by an ICK motif, also known as a cystine stabilized α-helix-loop-loop-motif, that involves three disulfide bonds (S-S linkages, typically CysI-CysIV, CysII-CysV, and CysIII-CysVI) to form the ICK. The ICK confers a highly stable three-dimensional conformation that is resistant to proteolytic degradation [107] and is fundamental for calcin stability and biological activity [123,135].
Ryanodine receptors are ligand-activated channels involved in the rapid intracellular release of Ca2+ from endoplasmic reticulum and sarcoplasmic reticulum [124], with this release being fundamental for excitation-contraction (EC) coupling in muscle cells and excitation-secretion coupling in neurons [136,137]. Calcins bind to an extraluminal (cytoplasmic) domain of the RyR [138] to allosterically induce conformational changes that facilitate or prolong the open state of the channel [139]. The high affinity and selectivity of calcines for specific RyR isoforms, e.g., RyR1, are key aspects of their pharmacological profile [138].

3.3.3. Chloride Channel Toxins (Protein Subfamily PS51200)

Cell membrane permeability to chloride (Cl) ions is mediated by a range of membrane proteins that include voltage-gated chloride channels (ClC-1, ClC-2, ClKa and ClKb), chloride/proton (2Cl/H+)-exchangers (ClC-3 to ClC-7), chloride channels activated by an elevation in intracellular Ca2+ (CaCC or ClCa), volume-regulated chloride channels, ligand (γ-aminobutyric acid—GABA, glycine)-gated chloride channels, and the cystic fibrosis transmembrane conductance regulator (CFTR) that is modulated via phosphorylation by protein kinase A (cAMP-dependent protein kinase or PKA) and ATP hydrolysis [140,141,142] (Figure 3 [143,144,145,146,147]). Whereas most of these proteins function as channels located in the plasma membrane, the 2Cl/H+-exchangers are antiporters (importing 2 Cl for every H+ exported) located primarily intracellularly, in endosomes and lysosomes [142]. Chloride channels conduct chloride ions (Cl) and other anions (Br, I, HCO3, NO3, SCN and small organic acids) across the cell membranes of tissues such as cardiac and skeletal muscle, neurons, and intestinal, pulmonary and renal epithelia, and are involved in activities such as membrane repolarization (hyperpolarization) and stabilization, intracellular pH and volume regulation, and the regulation of electrolyte balance and fluid transport [140]. Mutations in these different channels can result in a variety of diseases, including congenital myotonias, cystic fibrosis (involving CTRF), deafness, epilepsy, osteoporosis, renal dysfunction (formation of kidney stones, salt-wasting and reduction in the volume of extracellular fluid), and other conditions [141,142,148].
Transcriptomic, proteomic and genomic analyses of scorpion venoms have indicated that toxins active on Cl channels are much less abundant than those acting on K+ and Na+ channels, generally accounting for <5% of venom peptides [17,149,150,151,152]. Structurally, Cl channel toxins are short-chain toxins (mature peptides have 35–37 amino acids) containing eight cysteine residues arranged in the consensus sequence …CXXC…CXXCC…C…CXC… that form four disulfide bonds [153], but few of these toxins, and from only two venoms (L. q. hebraeus and L. q. quinquestriatus), have been characterized electrophysiologically and pharmacologically in any detail (Table 2 [143,144,145,146,147,154,155,156,157,158]).
The few reports that have examined the gene structure of ClTx-like toxins indicate that the overall organization is similar to that for other scorpion peptide toxins, i.e., coding regions for the signal and mature peptides, with an intron (88–93 bp) located in the region coding for 24–25 amino acid signal peptides [65,77,159,160,161]. Thus, for example, the gene (364 bp) for Bm-12, a 35 amino acid ClTx-like peptide from B. martensii venom (identical to BmKCT cloned by Zeng et al. [160] and subsequently renamed as BmKClTx3 [77]), contains a 93 bp intron in the 24 amino acid signal peptide coding region; this intron size is similar to that of other short-chain toxins (80–100 bp), but smaller than for long-chain peptides such mammalian α-neurotoxin, neurotoxins that affect mammal and insects, and insect excitatory neurotoxins (introns > 400 bp) [159]. Mature ClTx-like peptides have no terminal glycine, thus precluding post-translational amidation [66,161].
Although several short-chain toxins, later found to be structurally related to ClTx, were identified in the 1970s and 1980s [153], DeBin and Strichartz [143] were the first to pharmacologically demonstrate the ability of a scorpion venom (L. quinquestriatus) to block slow conductance ClC of embryonic rat brain growth cones and rat colonic epithelial cells, with a stoichiometry of one toxin molecule per channel; the molecular mass of the toxin was estimated to be ~5 kDa. The toxin responsible for this blockade (chlorotoxin, ClTx) was subsequently purified and sequenced (molecular mass: 4.07 kDa, with eight cysteines that formed four disulfide bonds) and shown to account for ~4.3% of the venom protein content. ClTx selectively blocked open ClC in a voltage-dependent manner and produced progressive, reversible paralysis in crayfish (crustaceans) and cockroaches [146]. Nuclear magnetic resonance (NMR) experiments later defined the α-helical and β-sheet structure and disulfide bonds of ClTx [162]. Subsequent studies indicated that L. quinquestriatus venom contained a toxin(s) capable of blocking ClC-2 (but not CIC-0 or CIC-1) channels [145] and CFTR [155,156]. Since the initial description of ClTx, various peptides structurally related to ClTx have been described from other scorpion species [153], including AaCt (Androctonus australis) [163], BmKCT (Buthus martensii Karsch) [160], Bs-Tx7 (Buthus sindicus) [164], BtITx3 [165] and ButaIT [166] (both from Buthus tamulus), GaTX1 [157] and GaTX2 [154] (both from L. q. hebraeus), Lqh2-2, Lqh7-1 and Lqh8-6 (L. q. hebraeus) [158], meuCl14, meuCl15 and meuCl16 (Mesobuthus eupeus) [167], OdClTx1 (Ondobuthus doriae) [168], and Ce-1 (Compsobuthus egyptiensis) [169]. However, the pharmacology of most of these peptides remains unknown or poorly characterized and the extent to which they interact with Cl channels remains to be determined. ClTx itself is not a general Cl channel blocker as native toxin applied extracellularly does not block volume-regulated, Ca2+-regulated, CFTR (cAMP-regulated) Cl channels, and glioma-specific Cl channels [147,155]. GaTX1 selectively blocks CFTR [145,157] and Lqh7-1 blocks Ca2+-activated Cl channels in vascular myocytes [158]. An important limitation in the electrophysiological study of toxins such as ClTx and GaTX1 is that they are apparently active only when applied to the intracellular (cytosolic) surface of their target channels such that their true extracellular sites of action on ion channels remain to be determined [144,147,157].
Table 2. Scorpion venoms and toxins active on chloride (Cl) channels.
Table 2. Scorpion venoms and toxins active on chloride (Cl) channels.
SpeciesToxin NameMolecular Mass (kDa)Target(s)Biological ActivityReferences
Leiurus quinquestriatus quintestriatusVenom
(peptide(s) not identified)
-Slow conductance Cl channels (ClC)Venom caused voltage-dependent reversible blockade of small conductance Cl channels from embryonic rat brain growth cones (200 μg/mL) and anion channels from rat colonic epithelial cells (enterocytes) (100 and 400 μg/mL) reconstituted in artificial membranes (stoichiometry of one toxin molecule per channel). Channels blocked in open state. Large conductance Cl channels of blood cells (predominantly platelets and white blood cells) were unaffected by venom (200 μg/mL). Specific Cl channel type not identified in these studies.[143,144]
Chlorotoxin (ClTx)~5 (4.07)Slow conductance Cl channels (ClC)Purified toxin (ClTx) reproduced the venom activity initially reported by DeBin and Strichartz [143] and blocked CIC with a KD of ~600 nM. ClTx caused progressive, reversible paralysis (recovery depended on venom dose) in crayfish and cockroaches, and accounted for ~4.3% of the venom protein content. Blockade seen only when ClTx was applied to the cytoplasmic surface. A later study reported that ClTx applied extracellularly did not block volume-regulated, CFTR (cAMP-regulated) and glioma-specific Cl channels [147]. [144,147]
Ca2+-regulated Cl
channels
Ca2+-regulated Cl channels in astrocytes were potently blocked by ClTx [146], whereas similar channels in a human colon carcinoma (T84) cell line were not blocked by ClTx [147].[146,147]
Leiurus quinquestriatus hebraeusVenom (peptide(s) not identified)----ClC-2 Cl channelsA peptide-enriched fraction of venom (obtained by filtering through 10 kDa-cutoff filters) caused concentration-dependent, progressive, reversible blockade of rabbit CIC-2 (but not Torpedo CIC-0 or human CIC-1) channels when applied extracellularly. Inhibition was unaffected by boiling the venom fraction, but was prevented by incubation with trypsin, indicating involvement of a heat-stable peptide. Blockade produced by modulating the channel gating mechanism (slower activation, but unaltered deactivation). Purified ClTx from L. q. quinquestriatus was inactive on CIC-2 channels.[145]
GaTX2
(Gating modifier of anion channels 2)
3.2ClC-2 Cl channelsA continuation of the investigation reported in [145] identified a toxin with three disulfide bonds and a sequence unrelated to ClTx but identical to previously identified leiuropeptide II from this same venom. Structurally related to toxins that inhibit K+ channels (apamin-sensitive K+ channels, Ca2+-activated K+ channels and Kv1.2 channels). Causes voltage-dependent blockade of ClC-2 channels with an apparent KD of ~20 pM. Slows channel activation but does not block open channels. No blockade of CIC-1, CIC-3 and CIC-4 channels or transporters, nor of CFTR (when applied extracellularly or intracellularly). No effect on GABAC receptors when applied to the extracellular surface, or when applied to the cytosolic side of endogenous Ca2+-activated chloride Cl channels. No effect on Kv1.2 channels.[154]
Venom
(peptide(s) not identified)
----Cystic fibrosis transmembrane conductance regulator (CFTR)Initial experiments showed that venom reversibly inhibited the CFTR channel in a voltage-dependent manner via a pore-block mechanism. Rapid, all-or-none blockade involving high affinity interaction with the nucleotide binding site of the channel in an interburst closed state, with a reduction in channel burst duration and open probability. No effect on ATP-dependent macroscopic opening rate. Only active when applied to the cytoplasmic side of phosphorylated channels. Activity was unaffected by boiling but was abolished by incubating venom with trypsin, suggesting peptide involvement. No effect on Xenopus oocyte Ca2+-activated Cl channels (ClCa) when added to the extracellular or cytosolic side. Purfied ClTx from L. q. quinquestriatus was inactive on CFTR. [155,156]
Georgia anion toxin 1 (GaTX1)3.67CFTRPotent, state-dependent (closed channel), reversible blockade of CFTR from the cytosolic side, with KD = 41.5 nM and IC50 = 48 nM. Reduced the open probability and increased the closed time of the channel. Blockade was reduced by high [ATP]. Possibly acts as a non-competitive blocker of CFTR. Greater than 94% identity with cDNA-derived sequences of ClTx-a, -b and -c and >62% sequence identity with various other ClTx-like peptides. No effect when applied to the extracellular surface of CIC-1 and CIC-2 channels or the chloride/H+ exchanger (antiporter) CIC-3, or the cytoplasmic surface of CIC-2 channels. GaTX1 was therefore different from the venom peptide that inhibits CIC-2 channels (see above). At concentrations that affected CFTR, GaTX1 had no effect on GABAC receptors when applied to the extracellular surface, or when applied to the cytosolic side of endogenous Ca2+-activated chloride Cl channels. Also did not affect the ABC transporters MRP1, MRP2, and MRP3.[157]
Lqh7-13.65Ca2+-regulated Cl channelsOf three ClTx-related peptides identified (Lqh2-2, Lqh7-1 and Lqh8-6), Lqh7-1 blocked Ca2+-regulated Cl channels in rat portal vein myocytes (IC50 63 ± 13 nM). Synthetic Lqh-1 caused similar blockade (IC50 49 ± 5 nM). Lqh2-2 caused only 50% blockade at 1 μM, whereas Lqh8-6 was inactive. Lqh7-1 had no effect on L-type and T-type voltage-dependent Ca2+ channels, on intracellular Ca2+ release via ryanodine-sensitive channels, or on Ca2+-activated and voltage-activated K+ currents.[158]
Only toxins for which electrophysiological studies on Cl channels have been reported are listed here. As indicated in the main text, various ClTx-related peptides have been identified in scorpion venoms, but very few of these have been rigorously examined with regard to their ion channel pharmacology. ABC—ATP-binding cassette, MRP—multidrug resistance-associated protein.
The biological functions of Cl channel toxins present in scorpion venoms remain largely unknown, mainly because few studies have examined the action of these peptides on the natural prey of most scorpion species. A few studies have examined the effects of these toxins on invertebrates. ClTx is paralytic in crayfish (Procambarus clarkii) and cockroaches (Periplaneta americana) [144], and lethal to aphids (Acyrthosiphon pisum) [153], while structurally related peptides (I1, I3, I4, and I5) from M. eupeus cause paralysis in Nauphoeta cinerea cockroaches [153] and peptides BTChl2 [144] and BtITx3 [165] from B. tamulus are lethal to tobacco bollworms (Heliothis virescens) and cotton bollworms (Helicoverpa armigera), respectively. Peptide ButaIT, also from B. tamulus, causes selective progressive irreversible flaccid paralysis of H. virescens but is not toxic to blowfly larvae or mice [166], while Ce-1 from Compsobuthus egyptiensis is lethal to crickets (Acheta domesticus) [169]. Although Cl channel modulation is assumed to be involved in the physiological effects of these toxins in invertebrates, especially insects, the precise molecular mechanisms involved remain unknown.
ClTx remains the best-characterized Cl channel toxin from scorpion venoms [170] and, over the years, has been proposed as a potentially useful molecule for treating certain tumors [171,172], particularly brain gliomas [173,174,175,176]. More recently, conjugates containing ClTx have been developed for imaging, diagnosing and treating glioblastoma brain tumors [51,177,178,179]. Interactions of ClTx with ClC-3 channel proteins and a variety of non-channel proteins, such as annexin 2, matrix metalloproteinase-2 (MMP-2), neuropilin 1 and others [164,180] may be important in the antitumor activity of ClTx and related peptides, possibly mediated via the inhibition of MMP-2 activity [153,181]. Indeed, such interactions, rather than Cl channel blockade, may be the most relevant physiological actions of ClTx [153,170].

3.4. Non-Disulfide-Bridged Peptides (NDBPs)

In addition to the abundant disulfide-bridged peptides described above, scorpion venoms also contain peptides (13–56 amino acids) lacking these bonds (non-disulfide-bridged peptides, NDBPs) [36,62,182,183,184,185]. The structural diversity of NDBPs has made their classification difficult, with the simplest classification being based on their linear amino acid sequence, i.e., short (<20 amino acids), medium (20–35 amino acids) and large (>35 acids) peptides [184], or a variation in this [45], with most NDBPs being short-chain peptides. Secondary structure analyses of these peptides indicate that, with few exceptions, e.g., Peptide T from Tityus serrulatus [27] and peptide K12 from Buthus occitanus [28] (both of which are bradykinin-potentiating peptides), NDBPs exist in a random coil conformation in aqueous solution, but form cationic, amphipathic α-helical structures under appropriate conditions, e.g., in dodecylphosphocholine (DPC) micelles [62,183]. Based on this core structure, NDBPs have been classified into three major groups: (1) Peptides with a completely α-helical structure, (2) Peptides with a central α-helical domain and random coils in the N and C terminal regions, and (3) Peptides with two α-helical regions separated by a central random coil region [156]. More comprehensive classifications include a combination of peptide length, sequence similarity and physiological functions [182,183].
The structural diversity and flexibility of NDBPs allows them to interact with a range of molecular targets and helps explain the diversity of biological activities among these peptides (Figure 2), with many of them having more than one activity [36,182,183,184]. However, few specific targets have been identified so far for NDBPs, with many of the biological actions reported for these peptides being rather general or non-specific. The finding that many of the biological activities of NDBPs, e.g., antibacterial, antiviral, antiparasitic and anticancer activities, involve interaction with negatively charged biological membranes probably reflects the cationic nature of most NDBPs (with charges ranging from +1 to +7, and isoelectric points > 9.5) [36,44,45,183].

3.5. Non-Channel-Modulating Peptides

In addition to the complex diversity of peptides that modulate ion channels, scorpion venoms also contain peptides with other biological activities not directly related to ion channel function. These include bradykinin potentiating peptides (BPPs) that inhibit ACE activity [27,28,29,30,31,32,33], peptides that potentiate the hypotensive activity of bradykinin by stimulating nitric oxide release rather than by ACE inhibition [25,57,58], peptides that modulate G-protein-coupled receptors [25,26], angiogenesis and inflammation [186], intracellular signaling pathways [187], and inhibit metalloproteinases [35], and a wide variety of peptides with antibacterial, antiviral, antiparasitic and other activities. This diversity of peptides and their range of activities can be increased by recombination events at the level of transcription [29,30,188] and by post-translational modifications [18,31,72,73]. As a result, a given peptide may exert a range of biological effects [29], further amplifying the molecular targets in vivo.

3.6. Enzymes

Most scorpion toxins studied to date have been peptides, but scorpion venoms also contain enzymes, particularly hyaluronidases, phospholipase A2 (PLA2), and proteases, particularly metalloproteases [189,190,191,192,193], although not all scorpion venoms contain these enzymes [190,194]. Hyaluronidases (spreading factor) are widely distributed among scorpion venoms [61,195,196,197,198,199] and facilitate the dispersion of toxins from the sting site through their ability to damage the surrounding extracellular matrix and connective tissue [200,201,202,203,204]. Hyaluronidase activity can be inhibited by antivenoms [195] and flavonoids [201]. These enzymes have various applications, such as in enhancing drug dispersion and in cancer therapy [203], and scorpion hyaluronidases could have similar uses, although this has not been extensively studied.
Venom PLA2, particularly those of snake venoms, are well-known for their ability to cause tissue necrosis, hemorrhage, hemolysis, coagulation disturbances, inflammation and pain [205,206,207]. These enzymes have been identified and characterized from several scorpion venoms [59,60]. In contrast to snake venom PLA2 that belong to groups I (Elapidae and Hydrophidae) and II (Viperidae) in the classification of PLA2, scorpion PLA2 belong to group III that includes PLA2 from bee, bumblebee, jellyfish and Mexican beaded lizard (Heloderma horridum horridum) venoms. Structurally, scorpion venom PLA2 have a conserved calcium-binding loop and active site when compared to snake venom PLA2, although the calcium-binding loop is closer to the N-terminal than in group I and II PLA2. Scorpion PLA2 differ from groups I and II and certain group III PLA2 in that they are heterodimeric enzymes with a long enzymatic chain containing the calcium-binding loop and active site that is linked by a disulfide bridge to a short chain. This dimeric structure is produced by the release of a pentapeptide from the proenzyme during maturation [59]. Although scorpion venom PLA2 exert a variety of biological activities in experimental studies in vitro and in vivo, including anti-angiogenic, anti-tumoral, anticoagulant, hemolytic, and inflammatory actions, and possibly cardiotoxicity and neurotoxicity [59,60], their precise role in scorpion envenoming remains to be established. Specifically, scorpion PLA2 may contribute to venom-induced inflammation and pain, as demonstrated for snake venom PLA2 [208,209,210,211]. As such, these enzymes could be therapeutically interesting targets for inhibition by small molecules such as varespladib [212].
Proteases (metalloproteases and serine proteases) have been identified in a variety of scorpion venoms [114,189,191,192,198,213,214,215,216,217,218,219,220,221]. Most metalloproteases have been identified based solely on transcriptomic and proteomic analyses, and very few have been purified and characterized biochemically and evaluated for their biological activities, e.g., antarease [222,223]. Metalloproteases may be involved in the maturation of toxin precursors through post-translational processing [224] and in the metabolism of biologically active peptides [225]. These enzymes may also contribute to venom diffusion from the sting site through their ability to degrade extracellular matrix proteins, in addition to exerting other effects such as inhibiting platelet aggregation, altering cytokine release, activating the complement cascade [226,227,228], as well as mediating inflammation and pain [229,230]. The extent to which metalloproteases contribute to and modulate the effects of scorpion venoms has not been extensively investigated. In this regard, the use of selective low-molecular mass metalloprotease inhibitors and drug repurposing could be therapeutically relevant in attenuating the actions of scorpion metalloproteinases in the local (edema, inflammation and pain) and systemic (coagulation disturbances, hemodynamic alterations and pulmonary edema) effects of envenoming, in a manner analogous to the situation with snake venoms [231,232,233,234,235,236,237,238].

4. Insecticidal Activity of Scorpion Venoms and Toxins

4.1. Insecticidal Potential

Globally, insect pests reduce agricultural yields by 10–16% before harvest and consume a similar proportion following harvest [239], with the largest food-producing countries (China and the United States) experiencing the highest losses from invasive insects [240]. Overall, agricultural losses to invasive insects cost at least US$70 billion per year globally, with associated health costs exceeding US$6.9 billion per year [241,242].
The use of natural and synthetic chemical insecticides has revolutionized the management of insects that affect human health, agriculture and natural resources. However, the continued and often indiscriminate use of these chemicals, besides triggering the development of resistance in different populations of target and non-target insect species [243], also leads to significant sublethal effects on these organisms. Such effects can impact their physiology and behavior, even at doses lower than those typically considered lethal [244,245,246]. The ongoing challenge of insect resistance to insecticides is further complicated by the discovery of non-traditional resistance mechanisms and highlights the need for advanced monitoring and management strategies [247,248,249].
Scorpion venoms have gained increasing attention for their potent insecticidal properties. The insecticidal activity of scorpion venoms primarily stems from their neurotoxic components that exhibit marked specificity towards insect neuronal channels, particularly voltage-gated sodium, potassium, and calcium channels. By disrupting the normal ion flow, these toxins induce paralysis and ultimately lead to insect death. For example, specific scorpion toxins can bind to and modify the gating kinetics of insect sodium channels, causing persistent activation and neuronal hyperexcitability [250]. In addition to neurotoxicity, some scorpion venom components are cytotoxic through their ability to disrupt the cell membrane, leading to cell lysis and death [251]. These toxins may also interfere with other cellular processes, such as mitochondrial function and signal transduction pathways [252].
Insects are a primary food source for most scorpions and scorpion venoms contain insectotoxins with varying degrees of specificity towards different insect orders [253,254]. Examples of these insectotoxins include AaIT (Androctonus australis) [254], BjIT 1 and BjIT 2 (Buthotus judaicus) [255,256], and SmIT 1 and SmIT 2 (Scorpio maurus palmatus) [257]. Some toxins show broad-spectrum activity and affect a wide range of insects, while others display narrow selectivity towards specific pests. This specificity is crucial for developing targeted pest control strategies that minimize harm to non-target organisms. The potential applications of scorpion toxins in agriculture and public health are significant. For example, they could be used to control agricultural pests that damage crops, thereby reducing the need for synthetic pesticides [258]. Additionally, they could be used to combat disease-carrying insects, such as mosquitoes and flies, thereby contributing to public health initiatives [259]. Scorpion toxins have also been shown to be active against economically important pests such as lepidopteran larvae and dipteran vectors [64]. However, further research is needed to fully understand the mechanisms of action involved, identify specific toxins with high insecticidal activity, and develop safe and effective delivery systems. Furthermore, careful consideration of potential environmental impacts and the development of sustainable production methods are essential for achieving the full potential of scorpion toxins in pest control. Scorpion insectotoxins also have promising applications as insecticides. For example, the venom of the scorpion Bothriurus bonariensis (C.L. Koch, 1842) from southern Brazil (Figure 4A) exerts insecticidal activity in the lobster cockroach Nauphoeta cinerea [253]. In these cockroaches, the venom causes neuromuscular paralysis (Figure 4B,C) and has a depressant activity upon spontaneous neural compound action potentials (SNCAP) (Figure 4D). These pharmacological activities suggest the presence of insectotoxins capable of interacting with voltage-gated sodium channels [253], although the composition of this venom remains to be determined.
Scorpion toxins provide a rich diversity of lead molecules for developing highly specific bioinsecticides, particularly those targeting insect ion channels [9,260]. Sodium channel toxins (NaTx), the major components of scorpion venoms, specifically target Nav channels and cause insect paralysis or death. NaTx are long-chain scorpion toxins composed of 60 to 76 amino acid residues, most of which contain four disulfide bridges, although some contain three disulfide bridges [260]. Depending on the binding site and whether the toxin affects the opening or closing dynamics of Nav channels, NaTx can be further classified into two categories: α-NaTx and β-NaTx. α-NaTx, predominantly found in New World scorpions, inhibit or delay of Nav channel inactivation by binding to receptor site 3 (located in domain IV of the channel), whereas β-NaTx, primarily identified in Old World scorpions, modify the activation dynamics of these channels by binding to receptor site 4 (located in domain II of the channel) (reviewed in [9,99,261]. Based on their preference for mammalian or insect pharmacological targets, both α-NaTx and β-NaTx can be subdivided into three subgroups: classic α-NaTx and β-NaTx, insect α-toxins and β-toxins, and α-like toxins and β-like toxins [63,262,263].
Classic α-NaTx are highly toxic to mammals, whereas α-NaTx insectotoxins are specifically active against insect Nav channels. A third group, the α-NaTx-like toxins, affects both mammalian and insect Nav channels [63,99]. Although classic NaTx and insectotoxins differ in their primary structures, their three-dimensional folds are conserved. The selectivity of scorpion NaTx is likely determined by minor variations at the binding site in Nav channels [9].
Scorpion β-NaTx exhibit diverse actions on mammalian and insect Nav and are classified into four classes: (1) toxins with selective activity against mammalian Nav channels, (2) selective, excitatory toxins targeting insect Nav channels, (3) selective, depressant toxins targeting insect Nav channels, and (4) toxins affecting both mammalian and insect Nav channels [63]. Insect-selective excitatory β-NaTx are distinguished by a C-terminal α-helix and the conserved motif KKxGxxxDxxGKxxECx(4,9)YCxxxCTKVxYAxxGYCCxxxCYCxGLxDDKx(9)KxxCD (where ‘x’ represents any amino acid and the numbers in parentheses indicate the number of residues in that region) [9,63]. These toxins induce sustained muscle contraction. Conversely, insect-selective depressant β-NaTx, characterized by the motif DGY[IP][KR]x(2)[DNS]GC[KR]x[ADS]Cx(2,3)Nx(2,3)Cx(3)Cx(3)G[AG]x[FY]GYCW[AGT]WGLACWC[EQ][GN]LP[ADE] (where bracketed amino acids represent possible residues at that position) [63], lead to flaccid paralysis. Notably, certain insect-selective depressant β-NaTx, despite their high affinity for insect Nav channels, are also active on rat Nav channels [9].
AaIT, a 70-residue peptide with four disulfide linkages, was the first insect toxin identified in the venom of the scorpion Androctonus australis. Its excitatory nature and preference for insect targets are well documented [264,265,266]. Investigations into the cytotoxicity of this toxin using insect and human cell lines (Sf9 and MCF-7, respectively) highlighted a striking selectivity. The toxin was highly potent against insect cells (50% cytotoxicity at 0.13 μM), but showed no significant toxicity to human cells at 1.3 μM, suggesting potential applications where insect-specific toxicity is desired [267].
Valdez-Velázquez et al. [268] reported that chromatographic (RP-HPLC) fraction VII of Centruroides tecomanus venom was lethal to crickets and mice. Two components within this fraction, with molecular weights of 7013 Da and 7538 Da, were specifically toxic to crickets, but not to mice. More recently, two insect-specific toxins, Ct-IT1 and Ct-IT2, were characterized from venom of C. tecomanus and shown to be very active against crickets, even at low doses [269]. In contrast, they were non-toxic to mammals since high doses do not cause adverse effects in mice [258]. Ct-IT1, in particular, with an LD50 of 3.81 μg/100 mg, caused potent immediate paralysis in 80% of crickets injected with one LD50 and paralysis within 5 min in the remaining 20%; a lower dose of 0.8 μg/100 mg caused immediate paralysis in 50% and paralysis in the remaining 50% within 1 h. Structural analyses using homology modeling attributed this high toxicity to a “trapping apparatus” on the toxin’s surface that is thought to be crucial for its insecticidal activity, in a manner analogous to that described for LqhIT2, an insect-selective toxin from L. quinquestriatus hebraeus [270]. Other scorpion insectotoxins, including AaH IT5, BtITx3, and AaIT, have also been studied for their potential application in managing agricultural pests such as Bemisia tabaci (silverleaf whitefly) and Helicoverpa armigera (cotton bollworm) [165], Heliothis virescens (tobacco budworm) [271], Spodoptera littoralis [272] and Nilaparvata lugens (brown planthopper) [273].
Several insectotoxins have been identified in venoms of the scorpion genus Tityus that occurs in Central and South America. Ts1 (6879.4 Da) is a major component of Tityus serrulatus venom and contains the conserved motif GCK[FLV]xC[FV][IP][NR][NP][AES][EGS]x[CGN] characteristic of broad-spectrum β-NaTx, with residues Lys12, Trp39, and Trp54 playing a crucial role in the interaction of Ts1 with rat and cockroach synaptosomes [92]. C-terminal amidation is also essential for the toxin’s biological activity [274]. Ts1 exhibits a dual action on Na+ channels and affects both vertebrate and invertebrate systems. At a concentration of 100 nM, Ts1 functions as a typical β-NaTx by shifting the activation voltage of mammalian Nav1.2, Nav1.3, Nav1.4, and Nav1.6 channels to more negative potentials. Ts1 also reduces the peak Na+ current through Nav1.5 channels without affecting the channel’s activation or inactivation rates [275]. In Drosophila melanogaster, Ts1 irreversibly modifies the DmNav1 channel, resulting in an enhanced and sustained Na+ current, along with a similar hyperpolarizing shift in voltage dependence [275]. This toxin displays potent insecticidal activity against fly larvae and binds to synaptic components in house flies and cockroaches. Given its properties, Ts1 may represent an intermediate between traditional β-NaTx and insect-specific toxins [92].
Tityus serrulatus venom contains other toxins with insecticidal properties [114]. Notably, Ts5, a neurotoxin exhibiting similarity to α-NaTx, counteracts the inactivation of DmNav1, a Na+ channel found in D. melanogaster. The inhibitory effect of 1 μM Ts5 on DmNav1 inactivation was more pronounced than its effect on other Nav channels tested [120]. This venom also contains antarease, a metalloprotease. Recombinant antarease, expressed in Escherichia coli, induced paralysis at the neuromuscular junction of D. melanogaster larvae, most likely through a mechanism involving its metalloprotease activity and the cleavage of external substrates on the presynaptic membrane [223].
To1 (formerly Tc49b), a 7400 Da toxin isolated from Tityus obscurus venom, is a non-selective β-NaTx that shares 62.3% similarity in its amino acid sequence with Ts1 [276]. This toxin potently affects the Na+ channels of the German cockroach (Blattella germanica), specifically BgNav1. At concentrations as low as 70 nM, To1 shifts the BgNav1 activation potential towards more negative values, thereby enhancing the probability of channel opening [276].
Tityus bahiensis venom also contains insecticidal components [277]. Tb2-II has high toxicity in insects, with an LD50 of 40 ng per house fly, but is also toxic to mammals. Conversely, TbIT-I, while maintaining potent insecticidal activity (LD50 of 80 ng per house fly), has little effect on vertebrate tissues. The presence of arginine at position 10 in TbIT-I may contribute to its reduced mammalian toxicity, in contrast to the negative residues found at this position in α and β toxins. Both TbIT-I and Tb2-II are toxic to crickets and cockroaches, and share similarities with other non-selective Tityus β-NaTx toxins.
The venoms of non-Brazilian Tityus spp. also contain insectotoxins. The Panamanian Tityus spp., T. asthenes and T. championi, contain venom fractions capable of paralyzing and killing crickets (Gryllus bimaculatus) [278,279]. Two peptides, Tma2 and Tma3, isolated from the Colombian scorpion Tityus macrochirus, were shown to be lethal to Acheta domesticus crickets at a dose of 300 ng per cricket, while exhibiting no activity against human Nav1.1 to Nav1.7 channels; these peptides are structurally related to insect-specific Na+ channel toxins from T. obscurus [280]. Furthermore, bactridin-1, a peptide from T. discrepans, killed cockroaches (Periplaneta americana) within 48 h of injection, without harming mice [281]. In addition to these toxins, analyses of venoms from other Tityus spp. and other scorpion genera have identified numerous peptides with potential insecticidal properties, including in T. costatus [282], T. cisandinus [121], Chactas reticulatus (Chactidae), Opisthacanthus elatus (Hormuridae), Centruroides edwardsii (Buthidae) and T. asthenes (Buthidae) from Colombia [283] and Tityus pachyurus and T. obscurus from Colombia and the Brazilian Amazon, respectively [284]. Figure 5 summarizes the major effects of scorpion toxins in insects.

4.2. Practical Considerations Related to Scorpion Toxin-Based Insecticides

The potential of scorpion toxins to serve as lead molecules for novel insecticides raises important issues that need to be addressed during the development of an effective molecule, including matters related to selective toxicity, routes of toxin administration or delivery, product stability, and economic considerations such as large-scale production and costs, as discussed elsewhere for scorpion and spider insecticidal toxins [9,285,286,287,288].
Ideally, the toxin of interest should be highly insect-specific, or engineered to be insect-specific, with no adverse effect on vertebrates. As such, the insect-specific excitatory and depressant β-NaTx are particularly promising candidates as lead molecules [9,286], particularly given the greater structural similarity of voltage-gated Na+ channels (VGSC) among insects (≥87%) than between insects and humans (50–60%) [286]. However, even for insect-specific toxins, there is a need to tailor the specificity towards the target species so as not to adversely affect non-target beneficial species. Such targeted specificity needs to be addressed in order to overcome a major limitation of current chemical insecticides and pesticides, i.e., the indiscriminate deleterious effect they exert on non-target insects. In this regard, understanding the diversity of structural variations in voltage-gated Na+ channels (VGSC) among insects in general [286,289], and within target species, e.g., mosquitos (Aedes albopictus) [290], and how these can be tailored to increase potency is a fundamental step towards designing species-specific molecules. This is particularly important given the central role of VGSC structural diversity in the appearance of resistance to certain currently used pesticides, e.g., pyrethroids, in insects such as house flies (Musca domestica) [291], mosquitos (Aedes spp. [292,293]; Anopheles spp. [294]; Culex spp. [295]), buffalo lice (Haematopinus tuberculatus) [296], and fruit flies (Drosophila spp.) [297].
Adequate delivery of the molecule to the target species is fundamental for successful insecticidal activity, with the main possible routes of delivery being direct application through spraying, exposure to transgenic entomopathogens, and oral ingestion. Application by spraying is unlikely to be particularly efficient because of the difficulty that most peptide-based molecules have in penetrating the insect cuticle. This could possibly be overcome by engineering molecules with suitable cuticular permeability [298], or by coupling the toxin to nanoparticles to facilitate penetration [299,300,301]. A further difficulty with application by spraying is that it will not overcome a major limitation of the widespread spraying methods currently used in agricultural practice, namely, adverse effects on non-target species. Another issue related to application by spraying relates to the stability of the molecules to environmental conditions (exposure to heat, humidity and sunlight) that may have variable effects on molecular structure and biological activity. Studies with the insecticidal spider-venom peptides ω-Hv1a, ω/κ-Hv1a, Ta1a and Dc1a exposed to artificial sunlight for up to seven days revealed varying degrees of peptide degradation among the toxins, with 33–84% degradation after three days of continuous exposure; the main modifications involved oxidations, deamidations, and cysteine alterations [302]. Similar studies have yet to be performed for scorpion insecticidal toxins.
An alternative approach for toxin delivery to insects involves the use of infected entomopathogens, such as baculoviruses, fungi, e.g., Metarhizium and Beauveria spp., and the bacterium Bacillus thuringiensis, engineered to encode transgenes of the desired toxin that will later be expressed in the host insect [9,286]. Examples of this include the Drosophila X virus-like particle used to deliver the toxin AaIT (Androctonus australis Hector) to Drosophila suzukii [303], a fungus used to deliver this same toxin to malarial mosquitos [304], and toxin LqqIT1 expressed in the entomopathogenic fungus Beauveria bassiana (Balasmo) [305]. The presence of the scorpion toxins generally increases the lethality of these entomopathogens.
Since various scorpion toxins show oral toxicity in insects [9,306], oral ingestion can be a suitable route for toxin administration, although higher doses are required compared to direct injection into the body cavity. In the field, oral administration may be achieved by spraying plants with a toxin solution that would then be ingested by insects feeding on the plant parts or, alternatively, by creating transgenic plants to express the toxin of interest that would then be ingested by insects during feeding [9,286,287]. Scorpion toxin-containing transgenic plants have been produced for cotton (Gossypium hirsutum) [307], rapeseed (Brassica napus) [308], rice (Oryza sativa) [273] and tobacco (Nicotiana tabacum) [273,309] and show greater resistance to attack by pest insects. The extent of toxin absorption from the insect gut can vary considerably, but can be enhanced by fusing the toxin with a carrier protein such as Galanthus nivalis agglutinin (GNA) while retaining oral toxicity [9,286,287]. This approach has been applied to the insect toxins ButaIT (Mesobuthus tamulus) that is active against coleopteran, dipteran and lepidopteran pests [310,311], OdTx12, a β-excitatory toxin from Odontobothus doriae [312], and BjaIT (Buthotus judaicus) that is active against silkworms [313].
Other practical considerations for scorpion toxin-based pesticides include the need for large-scale industrial production of the toxin using expression systems such as Pichia pastoris, as described for the toxin LqhIT2 [314], determination of the ideal product formulation (as a liquid, or as a dry powder to be dissolved in appropriate solution immediately prior to use), packaging and distribution networks, product stability in the field (as commented above in relation to environmental conditions), the environmental safety of the product and how this will affect its mode of use, and the overall final cost of these insecticidal molecules to the user [286,287]. Clearly, while there is considerable potential for the use of scorpion toxins to develop novel insecticides, there are numerous aspects and stages that need to be addressed before the final product can be commercialized.

5. Therapeutic Applications of Scorpion Venoms and Toxins

Historically, scorpions and their venoms have been used in traditional medicine to address neurological disorders, rheumatism, and erectile dysfunction, with records dating back to the medieval era [315]. Biotechnological progress in recent decades has enabled the cloning, expression and synthesis of scorpion venom toxins for therapeutic applications. The therapeutic actions of these toxins often involve mechanisms distinct from conventional pharmaceuticals. Figure 6 summarizes the wide range of potential therapeutic applications of scorpion toxins.

5.1. Antibacterial Activity

Scorpion venoms from various genera, including Androctonus, Buthus, Leiurus, Parabuthus, Scorpio, and Tityus, exert antibacterial effects against Gram-positive and Gram-negative bacteria from a wide range of genera, including Acinetobacter, Bacillus, Enterobacter, Enterococcus, Escherichia, Klebsiella, Micrococcus, Pseudomonas, Salmonella, Staphylococcus, and Streptococcus (Table 3). This action is mediated primarily by DBP and NDBP that interact with cell membranes, with cationic peptides showing greater activity against Gram-negative bacteria, and hydrophobic peptides being more effective against Gram-positive bacteria [36,42,44,45,182,184].
The antibacterial mechanism of some amphipathic peptides involves insertion into bacterial lipid membranes, leading to electrostatic interactions and subsequent pore formation via concentration-dependent oligomerization, ultimately disrupting membrane integrity [328]. These peptides can also interfere with protein or DNA synthesis [318]. Notably, scorpions, like other arthropods, possess defensins that exert potent antibacterial activity through membrane permeabilization, particularly against Gram-positive bacteria. Scorpion defensins, found in hemolymph and venom [329], are characterized by an antiparallel β-sheet linked to an amphipathic α-helix and an extended N-terminal fragment via three disulfide bridges. They differ from other scorpion toxins in size, sequence, and biological activity [329,330].

5.2. Antiviral Activity

Antiviral activity has been described for several scorpion venoms, including scorpion species from the Middle East and North Africa (MENA) (Table 4 [331,332,333,334,335]). This antiviral activity may be mediated by cationic peptides and involves mechanisms such as the disruption of phospholipid membranes (specifically for enveloped viruses), a decrease in viral DNA and protein synthesis, modulation of cellular signaling pathways to inhibit viral replication, and interference with endosomal acidification, thereby preventing viral genome release [45,47]. The range of mechanisms involved in antiviral activity has led to the investigation of scorpion venoms as a source of compounds active against a broad spectrum of viral families [46,336].

5.3. Antifungal Activity

Fungal infections pose a significant challenge to global public health, and the limited availability of antifungal drugs underscores the importance of exploring novel natural sources [337]. Scorpion venoms provide such a source and have been shown to exert antifungal activity [45,338,339]. Scorpion venom peptides can impede fungal growth and induce cell lysis by interacting with the cell’s outer membrane [42], while others can target proteins in the fungal nuclear envelope, leading to the production of reactive oxygen species, ion efflux, and ATP depletion, ultimately resulting in apoptosis. Additional mechanisms include the disruption of membrane surface tension, the creation of pores to release intracellular ions, and interference with mitochondrial regulators [42]. Despite the potential for identifying novel compounds, research on the antifungal effects of scorpion venoms remains limited. Table 5 summarizes some studies that have investigated the antifungal activity of certain scorpion venoms. These compounds hold promise for pharmaceutical applications, either as direct natural drugs or as templates for the development of new antifungal agents.

5.4. Antiparasitic Activity

Given the challenges posed by parasitic outbreaks, especially in developing nations, where drug resistance and adverse effects are significant concerns, scorpion venoms offer a source of lead compounds for developing potentially novel antifungal agents. However, compared to the antibacterial and antiviral activities of scorpion venoms, considerably less is known of their antiparasitic activities. Most investigations have focused on the activities against species of Echinococcus, Plasmodium, and Toxoplasma, and have examined the ability to interfere with parasite growth and replication as a means of controlling their dissemination [341]. Table 6 summarizes selected examples of scorpion venoms that have shown antiparasitic activity.

5.5. Autoimmune Diseases

Scorpion venoms exert their toxic effects primarily by altering the function of excitable cells. This disruption of cellular communication involves the combined action of K+ channel inhibitors and Na+ channel activators, leading to cellular depolarization. However, non-excitable cells also possess ion channels that are similar or identical to those of excitable cells. Consequently, scorpion toxins can affect non-excitable cells if the targeted channels play a crucial role in their function. A notable example is T lymphocytes, the white blood cells responsible for cellular immune responses in humans [346]. The activation of cytotoxic T cells relies on a continuous influx of Ca2+ ions that necessitates a counterbalancing efflux of K+ ions to maintain the membrane potential. Different T cell subsets achieve this balance by increasing the expression of either voltage-gated or Ca2+-activated K+ channels (Kv1.3 and KCa3.1, respectively) [347]. Of particular interest are chronically activated effector memory T cells (TEM) that contribute to the tissue damage observed in various autoimmune diseases, including multiple sclerosis, rheumatoid arthritis, and type-I diabetes mellitus [348]. TEM cells overexpress Kv1.3 channels that are susceptible to blockade by specific blockers [349], whereas other T cell subsets are unaffected by these blockers as their membrane potential is primarily regulated by KCa3.1 channels. In some cases, KCa3.1 channels can compensate for dysfunctional or absent Kv1.3 channels, although both types of channels are often required for complete normal cell functioning [347]. Given that several scorpion CSα/β toxins are high-affinity blockers of these specific K+ channels, these toxins can be promising candidates for developing drugs to treat autoimmune disorders.

5.6. Antidiabetic Activity

Several scorpion venoms, particularly from Middle Eastern and North African species, have been studied for their antidiabetic activity, but the number of such studies is small (Table 7). The findings so far suggest that only a few venom components exert any significant antidiabetic activity. The toxins implicated in these effects represent a promising source for the development of novel therapeutic agents, but further studies are necessary to fully characterize the venom components involved.

5.7. Anticancer Activity

Scorpion venoms are recognized as potential sources for anticancer drug development [353,354,355,356]. Scorpion venom peptides exhibit diverse anticancer mechanisms, including ion channel modulation, induction of apoptosis, and membrane disruption (Table 8 [357,358,359,360,361,362,363,364,365,366,367,368,369,370,371,372,373,374,375,376,377,378,379,380,381,382,383,384]). Peptides that target ion channels and signaling pathways, such as chlorotoxin and cationic antimicrobial peptides [382], are of particular interest because of their selective toxicity and ability to disrupt cancer cell homeostasis. For example, meuCl14 (from the venom of Mesobuthus eupeus) [167] and Smp43 (from the venom of Scorpio maurus palmatus) [383] efficiently inhibit tumor proliferation and migration by targeting tumor-associated chloride channels and inducing mitochondrial dysfunction. Additionally, phospholipase A2, such as leptulipin, can influence apoptotic pathways and cell cycle arrest [368]. These findings indicate that scorpion venom peptides and proteins provide promising avenues for innovative cancer therapies.

5.8. Analgesic Activity

Pain is a frequent finding in clinical envenoming by scorpions [385,386,387,388], with Nav channel activation by venom peptides being the major mechanism involved [20,52,53,389,390], although K+ channel inhibition [53,389,390], TRPV1 channel activation [390,391,392], and the release of pro-inflammatory mediators, including cytokines (e.g., IFN-γ, IL-1β, TNF-α) [392,393] also contribute to this phenomenon.
Paradoxically, scorpion venoms also contain peptides with analgesic activity, although fewer such molecules have been identified and characterized pharmacologically compared to pain-inducing toxins [19,37,47,84,390,394,395,396]. The vast majority of these peptides act by blocking Na+ channels, although the blockade of selected Ca2+ channels, e.g., N-type Ca2+ currents and T-type Cav3.2 and Cav3.3 [397], activation of K+ channels, e.g., Kv1.1 and Kv1.3 channels by hetlaxin from Heterometrus laoticus [398], indirect [399] and direct (by interaction with opioid receptors) [400,401] activation of the endogenous opioid system, the inhibition of inflammatory pathways and alterations in neurotransmitter release may also be involved [20,69,394,395]. Many of the analgesic peptides identified so far (>20) are from the venom of Buthus (Mesobuthus) mertensii Karsch, a species widely used in Chinese traditional medicine [20,37,47,84,390,394,395,396,402,403], although molecules from other genera (Androctonus, Hemiscorpius, Leiurus, Tityus) have also been characterized. Table 9 [404,405,406,407,408,409,410,411,412,413,414,415,416,417,418,419] summarizes the characteristics of several of these analgesic peptides. It should be noted that the blockade of Na+ channels by a given toxin may involve several channel subtypes, e.g., Nav1.3, Nav1.4, Nav1.5, Nav1.7 and Nav1.8 for toxins such as BmK AGAP, BmK AS and BmK M9, or may be selective for a specific subtype, e.g., Nav1.7 for toxins DKK-SP2, BmKBTx and BmNaL-3SS2, and Nav1.8 for Syb-prII (all toxins from B. mertensii venom).
Analgesic peptides are of particular interest for developing new compounds for the clinical treatment of pain [20,47,390,395,396,402], and the fact that various of these peptides are insectotoxins devoid of toxic effects in vertebrates (mammals) makes them particularly amenable for use as lead molecules for therapeutically useful analgesics [390]. In addition to the classic approach of identifying analgesic molecules and then proceeding with structure-activity analyses to design novel therapeutic compounds, other approaches that have been used include site-directed mutagenesis to identify important amino acids and ‘core’ or minimal structures necessary for analgesic activity, as done for Bmk AGAP [420], the use of computational methods and molecular biology to produce chimeric analgesic molecules from non-toxic and toxic peptides, e.g., a combination of the non-toxic β-neurotoxin CeIIB from Centruroides elegans and the toxic β-neurotoxin CssII from Centruroides suffusus suffuses [421], and the use of machine learning (artificial intelligence) to identify potential analgesic molecules starting with non-toxic peptides such as defensin 4 from B. martensii Karsch [401].

5.9. Therapeutic Potential of Scorpion Toxins in Neurological and Neurodegenerative Diseases

The high specificity and stability of scorpion toxins make them promising lead compounds for the development of therapies for a range of central nervous system (CNS) disorders. For example, peptides targeting voltage-gated Ca2+, K+ and Na+ channels are of interest for treating epilepsy, with engineered scorpion toxins acting as selective channel blockers or modulators to dampen hyperexcitability [422]. Similarly, the scorpion venom peptide chlorotoxin (ClTx) selectively binds glioma and glial cells (see Section 3.3.3 above), and modified analogs are being explored for imaging and therapeutic purposes in neuroinflammatory diseases like multiple sclerosis [423].
Neurodegenerative diseases such as Parkinson’s disease and Alzheimer’s disease involve complex mechanisms, including neuroinflammation, oxidative stress, mitochondrial dysfunction, and protein aggregation. The venoms of some scorpion species, e.g., A. australis and B. martensii, contain peptides that have neuroprotective, anti-inflammatory, and antioxidant properties, and offer promise as novel molecules for disease-modifying therapies. Figure 7 illustrates structural aspects of three B. martensii Karsch toxins. Table 10 summarizes the therapeutic potential of selected scorpion toxins for neurodegenerative diseases.

6. Challenges and Future Perspectives

As discussed above, scorpion venoms and their toxins have been investigated with regard to their potential biotechnological use as pesticides in agriculture and as therapeutic agents for treating a variety of medical conditions.
The insecticidal activities of scorpion venom peptides hold considerable promise for developing novel pesticides. However, the application of these toxins faces several challenges that need to be addressed to ensure their safe and sustainable use. A significant concern regarding scorpion toxin-based insecticides is their potential toxicity to non-target organisms. While some toxins exhibit high specificity towards target insects, others may pose risks to beneficial insects, such as pollinators and predators. Pollinators, like bees, are crucial for agricultural productivity and ecosystem health. The indiscriminate use of scorpion toxins could negatively impact these organisms, leading to ecological imbalances. Studies are needed to thoroughly assess the toxicity of specific scorpion toxins to a wide range of non-target organisms and to develop strategies for minimizing off-target effects [429].
The environmental impact of scorpion toxin-based pest control strategies is another crucial consideration. The production and application of these toxins must be sustainable to avoid adverse effects on ecosystems. This includes minimizing the use of solvents and other chemicals during toxin extraction and purification, as well as developing environmentally friendly delivery systems. Furthermore, the long-term effects of scorpion toxins on soil microorganisms and other environmental components need to be investigated. The use of genetically modified microorganisms for toxin production or the synthesis of toxin mimics might represent a more sustainable alternative for obtaining pure toxin compared to conventional toxin purification procedures [62].
To overcome these challenges and maximize the potential of scorpion toxins, future research should focus on several key areas:
  • Mechanistic studies: A deeper understanding of the molecular mechanisms of scorpion toxin action in insects is essential for optimizing their insecticidal activity. This includes identifying specific target sites within insect neuronal channels and elucidating the interactions between toxins and these targets. Techniques such as electrophysiology, molecular modeling, and proteomics can be useful in reaching this goal. This should lead to the development of more specific toxins.
  • Structure-activity relationship studies: Research on the structure-activity relationships of scorpion toxins is crucial for designing more potent and selective insecticides. By identifying the structural features responsible for insecticidal activity, it should be possible to develop modified toxins with enhanced efficacy and reduced off-target effects. This can be achieved through peptide synthesis and combinatorial chemistry.
  • Development of novel delivery systems: Efficient and targeted delivery of scorpion toxin-based insecticides is essential for maximizing their efficacy and minimizing environmental impact. Novel delivery systems, such as nanoencapsulation, liposomes, and viral vectors, can be explored to achieve this goal. These systems can enhance toxin stability, improve target specificity, and reduce the amount of toxin required for pest control [264].

7. Conclusions

Scorpion venoms contain a complex mixture of peptides, enzymes, and other compounds that interact with various biological targets, including ion channels and receptors. These interactions underlie the potent insecticidal and therapeutic effects observed with scorpion venom components. Specifically, insectotoxins, primarily NaTxs and KTxs, have shown promise as novel bioinsecticides because of their high specificity and potency against insect pests. In addition, calcins and other peptides have potentially interesting therapeutic applications, including as carrier molecules for drug delivery. The potential of scorpion venoms and toxins as a source for developing novel insecticides and other biotechnological applications is substantial. The high specificity and potency of insectotoxins make them attractive candidates for targeted pest control strategies, offering an environmentally friendly alternative to synthetic pesticides. Furthermore, the diverse pharmacological properties of scorpion venom components suggest their potential for developing novel therapeutics for various diseases. Advances in biotechnology, including peptide synthesis and recombinant DNA technology, have facilitated the synthesis of scorpion toxins and derivatives, as well as the development of venom-derived therapeutics. Further research is warranted to explore the mechanisms of action of scorpion toxins and to optimize their applications in agriculture and medicine. This includes investigating the structure–activity relationships of scorpion toxins, developing novel delivery systems for sustainable application, and conducting clinical trials to assess the safety, potential risks, and efficacy of venom-derived therapeutics.

Author Contributions

Conceptualization, C.A.D.B., S.H. and C.R.C.; methodology, C.A.D.B., S.H. and C.R.C.; software, C.A.D.B.; validation, S.H. and C.R.C.; formal analysis, C.A.D.B., S.H. and C.R.C. investigation, C.A.D.B., S.H. and C.R.C.; resources, C.A.D.B., S.H. and C.R.C. data curation, C.A.D.B.; writing—original draft preparation, C.A.D.B.; writing—review and editing, S.H. and C.R.C.; visualization, C.A.D.B.; supervision, S.H. and C.R.C.; project administration, S.H. and C.R.C.; funding acquisition, C.A.D.B., S.H. and C.R.C. All authors have read and agreed to the published version of the manuscript.

Funding

CADB was supported by CAPES PDPG Consolidação 3,4 (2023), CRC was supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq; Universal grant no. 405810/2023-7), Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul (FAPERGS; Pesquisador Gaúcho grant no. 21/2551-0001946-8) and Instituto Nacional de Ciência e Tecnologia para Doenças Cerebrais, Excitotoxicidade e Neuroproteção (INCT-EN; CNPq grant no. 465671/2014-4; FAPERGS 17/2551-0000516-3), and SH was supported by INCT INOVATox (CNPq grant no. 406816/2022-0). CADB, CRC and SH are supported by research fellowships from CNPq (grant nos. 305933/2024-8, 305081/2019-5 and 312901/2022-4, respectively).

Institutional Review Board Statement

Not applicable as no animals or human subjects were used in this investigation.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors thank Mia Schezaro-Ramos for help in preparing the figures for this article.

Conflicts of Interest

The authors have no conflicts of interest with this work. None of the funding agencies cited above had any role in the design of the study, in the collection, analyses or interpretation of the data, in the writing of the manuscript, or in the decision to publish this work.

References

  1. Rein, J.O. The Scorpion Files (Version Apr 2025). In Catalogue of Life (Version 20/08/2025); Bánki, O., Roskov, Y., Döring, M., Ower, G., Hernández Robles, D.R., Plata Corredor, C.A., Stjernegaard Jeppesen, T., Örn, A., Pape, T., et al., Eds.; Catalogue of Life Foundation: Amsterdam, The Netherlands, 2025; Available online: https://www.catalogueoflife.org/data/dataset/1164 (accessed on 12 September 2025). [CrossRef]
  2. Lourenço, W.R. The evolution and distribution of noxious species of scorpions (Arachnida: Scorpiones). J. Venom. Anim. Toxins Incl. Trop. Dis. 2018, 24, 1. [Google Scholar] [CrossRef]
  3. Chippaux, J.P.; Goyffon, M. Epidemiology of scorpionism: A global appraisal. Acta Trop. 2008, 10, 71–79. [Google Scholar] [CrossRef]
  4. Santos, M.S.; Silva, C.G.; Neto, B.S.; Grangeiro Júnior, C.R.; Lopes, V.H.; Teixeira Júnior, A.G.; Bezerra, D.A.; Luna, J.V.; Cordeiro, J.B.; Júnior, J.G.; et al. Clinical and epidemiological aspects of scorpionism in the world: A systematic review. Wilderness Environ. Med. 2016, 27, 504–518. [Google Scholar] [CrossRef]
  5. Baradaran, M.; Salabi, F.; Mahdavinia, M.; Mohammadi, E.; Vazirianzadeh, B.; Avella, I.; Kazemi, S.M.; Lüddecke, T. ScorpDb: A novel open-access database for integrative scorpion toxinology. Toxins 2024, 16, 497. [Google Scholar] [CrossRef] [PubMed]
  6. Zlotkin, E.; Miranda, F.; Rochat, H. Chemistry and pharmacology of Buthinae scorpion venoms. In Arthropod Venoms. Handbook of Experimental Physiology; Bettini, S., Ed.; Springer: Berlin/Heidelberg, Germany, 1978; Volume 48, pp. 317–369. [Google Scholar]
  7. Possani, L.D. Structure of scorpion toxins. In Handbook of Natural Toxins: Insect Poisons, Allergens, and Other Invertebrate Venoms; Tu, A.T., Ed.; Marcel Dekker Inc.: New York, NY, USA, 1984; Volume 2, pp. 513–550. [Google Scholar]
  8. Gwee, M.C.E.; Nirthanan, S.; Khoo, H.E.; Gopalakrishnakone, P.; Kini, R.M.; Cheah, L.S. Autonomic effects of some scorpion venoms and toxins. Clin. Exp. Pharmacol. Physiol. 2002, 29, 795–801. [Google Scholar] [CrossRef] [PubMed]
  9. Ortiz, E.; Possani, L.D. The unfulfilled promises of scorpion insectotoxins. J. Venom. Anim. Toxins Incl. Trop. Dis. 2015, 21, 16. [Google Scholar] [CrossRef] [PubMed]
  10. Pashmforoosh, N.; Baradaran, M. Peptides with diverse functions from scorpion venom: A great opportunity for the treatment of a wide variety of diseases. Iran. Biomed. J. 2023, 27, 84–99. [Google Scholar] [CrossRef]
  11. Rodríguez de la Vega, R.C.; Schwartz, E.F.; Possani, L.D. Mining on scorpion venom biodiversity. Toxicon 2010, 56, 1155–1161. [Google Scholar] [CrossRef]
  12. Quintero-Hernández, V.; Ortiz, E.; Rendón-Anaya, M.; Schwartz, E.F.; Becerril, B.; Corzo, G.; Possani, L.D. Scorpion and spider venom peptides: Gene cloning and peptide expression. Toxicon 2011, 58, 644–663. [Google Scholar] [CrossRef]
  13. Cid-Uribe, J.I.; Veytia-Bucheli, J.I.; Romero-Gutierrez, T.; Ortiz, E.; Possani, L.D. Scorpion venomics: A 2019 overview. Expert Rev. Proteom. 2020, 17, 67–83. [Google Scholar] [CrossRef]
  14. Mouchbahani-Constance, S.; Sharif-Naeini, R. Proteomic and transcriptomic techniques to decipher the molecular evolution of venoms. Toxins 2021, 13, 154. [Google Scholar] [CrossRef] [PubMed]
  15. Vonk, F.J.; Bittenbinder, M.A.; Kerkkamp, H.M.I.; Grashof, D.G.B.; Archer, J.P.; Afonso, S.; Richardson, M.K.; Kool, J.; van der Meijden, A. A non-lethal method for studying scorpion venom gland transcriptomes, with a review of potentially suitable taxa to which it can be applied. PLoS ONE 2021, 16, e0258712. [Google Scholar] [CrossRef] [PubMed]
  16. Marchi, F.C.; Mendes-Silva, E.; Rodrigues-Ribeiro, L.; Bolais-Ramos, L.G.; Verano-Braga, T. Toxinology in the proteomics era: A review on arachnid venom proteomics. J. Venom. Anim. Toxins Incl. Trop. Dis. 2022, 28, 20210034. [Google Scholar] [CrossRef]
  17. Solano-Godoy, J.A.; Betancourt-Osorio, M.; Orjuela-Rodriguez, M.; Guerrero-Vargas, J.A.; Sepulveda-Arias, J.C. Scorpion venom gland transcriptomics: A systematic review. Toxicon 2025, 267, 108563. [Google Scholar] [CrossRef]
  18. Verano-Braga, T.; Dutra, A.A.; León, I.R.; Melo-Braga, M.N.; Roepstorff, P.; Pimenta, A.M.; Kjeldsen, F. Moving pieces in a venomic puzzle: Unveiling post-translationally modified toxins from Tityus serrulatus. J. Proteome Res. 2013, 12, 3460–3470. [Google Scholar] [CrossRef]
  19. Ahmadi, S.; Knerr, J.M.; Argemi, L.; Bordon, K.C.F.; Pucca, M.B.; Cerni, F.A.; Arantes, E.C.; Çalışkan, F.; Laustsen, A.H. Scorpion venom: Detriments and benefits. Biomedicines 2020, 8, 118. [Google Scholar] [CrossRef]
  20. Xin, K.; Sun, R.; Xiao, W.; Lu, W.; Sun, C.; Lou, J.; Xu, Y.; Chen, T.; Wu, D.; Gao, Y. Short peptides from Asian scorpions: Bioactive molecules with promising therapeutic potential. Toxins 2025, 17, 114. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Chen, Z.; Cao, Z.; Li, W.; Wu, Y. Diverse structural features of potassium channels characterized by scorpion toxins as molecular probes. Molecules 2019, 24, 2045. [Google Scholar] [CrossRef] [PubMed]
  22. Tajti, G.; Wai, D.C.C.; Panyi, G.; Norton, R.S. The voltage-gated potassium channel Kv1.3 as a therapeutic target for venom-derived peptides. Biochem. Pharmacol. 2020, 181, 114146. [Google Scholar] [CrossRef]
  23. Mendes, L.C.; Viana, G.M.M.; Nencioni, A.L.A.; Pimenta, D.C.; Beraldo-Neto, E. Scorpion peptides and ion channels: An insightful review of mechanisms and drug development. Toxins 2023, 15, 238. [Google Scholar] [CrossRef]
  24. Rojas-Palomino, J.; Gómez-Restrepo, A.; Salinas-Restrepo, C.; Segura, C.; Giraldo, M.A.; Calderón, J.C. Electrophysiological evaluation of the effect of peptide toxins on voltage-gated ion channels: A scoping review on theoretical and methodological aspects with focus on the Central and South American experience. J. Venom. Anim. Toxins Incl. Trop. Dis. 2024, 30, e20230048. [Google Scholar] [CrossRef]
  25. Verano-Braga, T.; Figueiredo-Rezende, F.; Melo, M.N.; Lautner, R.Q.; Gomes, E.R.; Mata-Machado, L.T.; Murari, A.; Rocha-Resende, C.; Lima, M.E.; Guatimosim, S.; et al. Structure-function studies of Tityus serrulatus hypotensin-I (TsHpt-I): A new agonist of B2 kinin receptor. Toxicon 2010, 56, 1162–1171. [Google Scholar] [CrossRef]
  26. Rocha-Resende, C.; Leão, N.M.; de Lima, M.E.; Santos, R.A.; Pimenta, A.M.C.; Verano-Braga, T. Moving pieces in a cryptomic puzzle: Cryptide from Tityus serrulatus Ts3 Nav toxin as potential agonist of muscarinic receptors. Peptides 2017, 98, 70–77. [Google Scholar] [CrossRef]
  27. Ferreira, L.A.; Alves, E.W.; Henriques, O.B. Peptide T, a novel bradykinin potentiator isolated from Tityus serrulatus scorpion venom. Toxicon 1993, 31, 941–947. [Google Scholar] [CrossRef] [PubMed]
  28. Meki, A.R.; Nassar, A.Y.; Rochat, H. A bradykinin-potentiating peptide (peptide K12) isolated from the venom of Egyptian scorpion Buthus occitanus. Peptides 1995, 16, 1359–1365. [Google Scholar] [CrossRef] [PubMed]
  29. Zeng, X.C.; Wang, S.; Nie, Y.; Zhang, L.; Luo, X. Characterization of BmKbpp, a multifunctional peptide from the Chinese scorpion Mesobuthus martensii Karsch: Gaining insight into a new mechanism for the functional diversification of scorpion venom peptides. Peptides 2012, 33, 44–51. [Google Scholar] [CrossRef] [PubMed]
  30. Sunagar, K.; Undheim, E.A.; Chan, A.H.; Koludarov, I.; Muñoz-Gómez, S.A.; Antunes, A.; Fry, B.G. Evolution stings: The origin and diversification of scorpion toxin peptide scaffolds. Toxins 2013, 5, 2456–2487. [Google Scholar] [CrossRef]
  31. Pucca, M.B.; Cerni, F.A.; Pinheiro-Junior, E.L.; Zoccal, K.F.; Bordon, K.C.F.; Amorim, F.G.; Peigneur, S.; Vriens, K.; Thevissen, K.; Cammue, B.P.A.; et al. Non-disulfide-bridged peptides from Tityus serrulatus venom: Evidence for proline-free ACE-inhibitors. Peptides 2016, 82, 44–51. [Google Scholar] [CrossRef]
  32. Setayesh-Mehr, Z.; Asoodeh, A. The inhibitory activity of HL-7 and HL-10 peptide from scorpion venom (Hemiscorpius lepturus) on angiotensin converting enzyme: Kinetic and docking study. Bioorg. Chem. 2017, 75, 30–37. [Google Scholar] [CrossRef]
  33. Goudarzi, H.R.; Salehi Najafabadi, Z.; Movahedi, A.; Noofeli, M. Bradykinin-potentiating factors of venom from Iranian medically important scorpions. Arch. Razi Inst. 2019, 74, 385–394. [Google Scholar] [CrossRef]
  34. Hasan, H.F.; Mohmed, H.K.; Galal, S.M. Scorpion bradykinin potentiating factor mitigates lung damage induced by γ-irradiation in rats: Insights on AngII/ACE/Ang(1-7) axis. Toxicon 2021, 203, 58–65. [Google Scholar] [CrossRef]
  35. Sharma, A.; Balde, A.; Nazeer, R.A. A review on animal venom-based matrix metalloproteinase modulators and their therapeutic implications. Int. Immunopharmacol. 2025, 157, 114703. [Google Scholar] [CrossRef]
  36. Ortiz, E.; Gurrola, G.B.; Schwartz, E.F.; Possani, L.D. Scorpion venom components as potential candidates for drug development. Toxicon 2015, 93, 125–135. [Google Scholar] [CrossRef] [PubMed]
  37. Kamau, P.M.; Zhong, J.; Yao, B.; Lai, R.; Luo, L. Bioactive peptides from scorpion venoms: Therapeutic scaffolds and pharmacological tools. Chin. J. Nat. Med. 2023, 21, 19–35. [Google Scholar] [CrossRef]
  38. Fong-Coronado, P.A.; Ramirez, V.; Quintero-Hernández, V.; Balleza, D. A critical review of short antimicrobial peptides from scorpion venoms, their physicochemical attributes, and potential for the development of new drugs. J. Membr. Biol. 2024, 257, 165–205. [Google Scholar] [CrossRef] [PubMed]
  39. Hemajha, L.; Singh, S.; Biji, C.A.; Balde, A.; Benjakul, S.; Nazeer, R.A. A review on inflammation modulating venom proteins/peptide therapeutics and their delivery strategies: A review. Int. Immunopharmacol. 2024, 142, 113130. [Google Scholar] [CrossRef] [PubMed]
  40. Lafnoune, A.; Darkaoui, B.; Chbel, A.; Nait Irahal, I. Emerging therapeutic applications of scorpion venom peptides in the Middle East and North Africa: A comprehensive review. Toxicon 2025, 256, 108270. [Google Scholar] [CrossRef]
  41. Amorim-Carmo, B.; Parente, A.M.S.; Souza, E.S.; Silva-Junior, A.A.; Araújo, R.M.; Fernandes-Pedrosa, M.F. Antimicrobial peptide analogs from scorpions: Modifications and structure-activity. Front. Mol. Biosci. 2022, 9, 887763. [Google Scholar] [CrossRef]
  42. Rincón-Cortés, C.A.; Bayona-Rojas, M.A.; Reyes-Montaño, E.A.; Vega-Castro, N.A. Antimicrobial activity developed by scorpion venoms and its peptide component. Toxins 2022, 14, 740. [Google Scholar] [CrossRef]
  43. Nasr, S.; Borges, A.; Sahyoun, C.; Nasr, R.; Roufayel, R.; Legros, C.; Sabatier, J.M.; Fajloun, Z. Scorpion venom as a source of antimicrobial peptides: Overview of biomolecule separation, analysis and characterization methods. Antibiotics 2023, 12, 1380. [Google Scholar] [CrossRef]
  44. Panayi, T.; Diavoli, S.; Nicolaidou, V.; Papaneophytou, C.; Petrou, C.; Sarigiannis, Y. Short-chained linear scorpion peptides: A pool for novel antimicrobials. Antibiotics 2024, 13, 422. [Google Scholar] [CrossRef] [PubMed]
  45. Xia, Z.; Xie, L.; Li, B.; Lv, X.; Zhang, H.; Cao, Z. Antimicrobial potential of scorpion-venom-derived peptides. Molecules 2024, 29, 5080. [Google Scholar] [CrossRef] [PubMed]
  46. El Hidan, M.A.; Laaradia, M.A.; El Hiba, O.; Draoui, A.; Aimrane, A.; Kahime, K. Scorpion-derived antiviral peptides with a special focus on medically important viruses: An update. Biomed Res. Int. 2021, 2021, 9998420. [Google Scholar] [CrossRef]
  47. Xia, Z.; He, D.; Wu, Y.; Kwok, H.F.; Cao, Z. Scorpion venom peptides: Molecular diversity, structural characteristics, and therapeutic use from channelopathies to viral infections and cancers. Pharmacol. Res. 2023, 197, 106978. [Google Scholar] [CrossRef] [PubMed]
  48. Srairi-Abid, N.; Othman, H.; Aissaoui, D.; BenAissa, R. Anti-tumoral effect of scorpion peptides: Emerging new cellular targets and signaling pathways. Cell Calcium 2019, 80, 160–174. [Google Scholar] [CrossRef]
  49. Díaz-García, A.; Varela, D. Voltage-gated K+/Na+ channels and scorpion venom toxins in cancer. Front. Pharmacol. 2020, 11, 913. [Google Scholar] [CrossRef]
  50. Mikaelian, A.G.; Traboulay, E.; Zhang, X.M.; Yeritsyan, E.; Pedersen, P.L.; Ko, Y.H.; Matalka, K.Z. Pleiotropic anticancer properties of scorpion venom peptides: Rhopalurus princeps venom as an anticancer agent. Drug Des. Devel Ther. 2020, 14, 881–893. [Google Scholar] [CrossRef]
  51. Boltman, T.; Meyer, M.; Ekpo, O. Diagnostic and therapeutic approaches for glioblastoma and neuroblastoma cancers using chlorotoxin nanoparticles. Cancers 2023, 15, 3388. [Google Scholar] [CrossRef]
  52. El-Qassas, J.; Abd El-Atti, M.; El-Badri, N. Harnessing the potency of scorpion venom-derived proteins: Applications in cancer therapy. Bioresour. Bioprocess. 2024, 11, 93. [Google Scholar] [CrossRef]
  53. Wang, X.; Luo, H.; Peng, X.; Chen, J. Spider and scorpion knottins targeting voltage-gated sodium ion channels in pain signaling. Biochem. Pharmacol. 2024, 227, 116465. [Google Scholar] [CrossRef]
  54. He, D.; Lei, Y.; Qin, H.; Cao, Z.; Kwok, H.F. Deciphering scorpion toxin-induced pain: Molecular mechanisms and ion channel dynamics. Int. J. Biol. Sci. 2025, 21, 2921–2934. [Google Scholar] [CrossRef] [PubMed]
  55. Balde, A.; Benjakul, S.; Nazeer, R.A. A review on NLRP3 inflammasome modulation by animal venom proteins/peptides: Mechanisms and therapeutic insights. Inflammopharmacology 2025, 33, 1013–1031. [Google Scholar] [CrossRef]
  56. Coulter-Parkhill, A.; McClean, S.; Gault, V.A.; Irwin, N. Therapeutic potential of peptides derived from animal venoms: Current views and emerging drugs for diabetes. Clin. Med. Insights Endocrinol. Diabetes 2021, 14, 11795514211006071. [Google Scholar] [CrossRef]
  57. Verano-Braga, T.; Rocha-Resende, C.; Silva, D.M.; Ianzer, D.; Martin-Eauclaire, M.F.; Bougis, P.E.; de Lima, M.E.; Santos, R.A.; Pimenta, A.M. Tityus serrulatus hypotensins: A new family of peptides from scorpion venom. Biochem. Biophys. Res. Commun. 2008, 371, 515–520. [Google Scholar] [CrossRef]
  58. Machado, R.J.; Junior, L.G.; Monteiro, N.K.; Silva-Júnior, A.A.; Portaro, F.C.; Barbosa, E.G.; Braga, V.A.; Fernandes-Pedrosa, M.F. Homology modeling, vasorelaxant and bradykinin-potentiating activities of a novel hypotensin found in the scorpion venom from Tityus stigmurus. Toxicon 2015, 101, 11–18. [Google Scholar] [CrossRef]
  59. Krayem, N.; Gargouri, Y. Scorpion venom phospholipases A2: A minireview. Toxicon 2020, 184, 48–54. [Google Scholar] [CrossRef]
  60. Soltan-Alinejad, P.; Alipour, H.; Meharabani, D.; Azizi, K. Therapeutic potential of bee and scorpion venom phospholipase A2 (PLA2): A narrative review. Iran. J. Med. Sci. 2022, 47, 300–313. [Google Scholar]
  61. Bordon, K.C.F.; Wiezel, G.A.; Amorim, F.G.; Arantes, E.C. Arthropod venom hyaluronidases: Biochemical properties and potential applications in medicine and biotechnology. J. Venom. Anim. Toxins Incl. Trop. Dis. 2015, 21, 43. [Google Scholar] [CrossRef]
  62. Psenicnik, A.; Ojanguren-Affilastro, A.A.; Graham, M.R.; Hassan, M.K.; Abdel-Rahman, M.A.; Sharma, P.P.; Santibáñez-López, C.E. Optimizing scorpion toxin processing through artificial intelligence. Toxins 2024, 16, 437. [Google Scholar] [CrossRef] [PubMed]
  63. Quintero-Hernández, V.; Jiménez-Vargas, J.M.; Gurrola, G.B.; Valdivia, H.H.; Possani, L.D. Scorpion venom components that affect ion-channels function. Toxicon 2013, 76, 328–342. [Google Scholar] [CrossRef] [PubMed]
  64. Possani, L.D.; Becerril, B.; Delepierre, M.; Tytgat, J. Scorpion toxins specific for Na+-channels. Eur. J. Biochem. 1999, 264, 287–300. [Google Scholar] [CrossRef]
  65. Froy, O.; Sagiv, T.; Poreh, M.; Urbach, D.; Zilberberg, N.; Gurevitz, M. Dynamic diversification from a putative common ancestor of scorpion toxins affecting sodium, potassium, and chloride channels. J. Mol. Evol. 1999, 48, 187–196. [Google Scholar] [CrossRef] [PubMed]
  66. Goudet, C.; Chi, C.-W.; Tytgat, J. An overview of toxins and genes from the venom of the Asian scorpion Buthus martensi Karsch. Toxicon 2002, 40, 1239–1258. [Google Scholar] [CrossRef]
  67. Zhijian, C.; Feng, L.; Yingliang, W.; Xin, M.; Wenxin, L. Genetic mechanisms of scorpion venom peptide diversification. Toxicon 2006, 47, 348–355. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, Z.; Luo, F.; Feng, J.; Yang, W.; Zeng, D.; Zhao, R.; Cao, Z.; Liu, M.; Li, W.; Jiang, L.; et al. Genomic and structural characterization of Kunitz-type peptide LmKTT-1a highlights diversity and evolution of scorpion potassium channel toxins. PLoS ONE 2013, 8, e60201. [Google Scholar] [CrossRef]
  69. Wang, X.; Gao, B.; Zhu, S. Exon shuffling and origin of scorpion venom biodiversity. Toxins 2017, 9, 10. [Google Scholar] [CrossRef]
  70. Rates, B.; Ferraz, K.K.; Borges, M.H.; Richardson, M.; De Lima, M.E.; Pimenta, A.M. Tityus serrulatus venom peptidomics: Assessing venom peptide diversity. Toxicon 2008, 52, 611–618. [Google Scholar] [CrossRef]
  71. Wang, X.; Gao, B.; Zhu, S. A single-point mutation enhances dual functionality of a scorpion toxin. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2016, 179, 72–78. [Google Scholar] [CrossRef]
  72. Delgado-Prudencio, G.; Possani, L.D.; Becerril, B.; Ortiz, E. The dual α-amidation system in scorpion venom glands. Toxins 2019, 11, 425. [Google Scholar] [CrossRef]
  73. Zeng, L.; Zhang, C.; Yang, M.; Sun, J.; Lu, J.; Zhang, H.; Qin, J.; Zhang, W.; Jiang, Z. Unveiling the diversity and modifications of short peptides in Buthus martensii scorpion venom through liquid chromatography-high resolution mass spectrometry. Toxins 2024, 16, 155. [Google Scholar] [CrossRef] [PubMed]
  74. Wu, J.J.; Dai, L.; Lan, Z.D.; Chi, C.W. Genomic organization of three neurotoxins active on small conductance Ca2+-activated potassium channels from the scorpion Buthus martensi Karsch. FEBS Lett. 1999, 452, 360–364. [Google Scholar] [CrossRef]
  75. Dai, L.; Wu, J.J.; Gu, Y.H.; Lan, Z.D.; Ling, M.H.; Chi, C.W. Genomic organization of three novel toxins from the scorpion Buthus martensi Karsch that are active on potassium channels. Biochem. J. 2000, 346, 805–809. [Google Scholar] [CrossRef]
  76. Chen, Z.-Y.; Zeng, D.-Y.; Hu, Y.-T.; Ya-Wen He, Y.-W.; Pan, N.; Ding, J.-P.; Cao, Z.-J.; Liu, M.-L.; Wen-Xin Li, W.-X.; Yi, H.; et al. Structural and functional diversity of acidic scorpion potassium channel toxins. PLoS ONE 2012, 7, e35154. [Google Scholar] [CrossRef]
  77. Cao, Z.; Yu, Y.; Wu, Y.; Hao, P.; Di, Z.; He, Y.; Chen, Z.; Yang, W.; Shen, Z.; He, X.; et al. The genome of Mesobuthus martensii reveals a unique adaptation model of arthropods. Nat. Commun. 2013, 4, 2602. [Google Scholar] [CrossRef]
  78. Saucedo, A.L.; Flores-Solis, D.; Rodríguez de la Vega, R.C.; Ramírez-Cordero, B.; Hernández-López, R.; Cano-Sánchez, P.; Noriega Navarro, R.; García-Valdés, J.; Coronas-Valderrama, F.; de Roodt, A.; et al. New tricks of an old pattern: Structural versatility of scorpion toxins with common cysteine spacing. J. Biol. Chem. 2012, 287, 12321–12330. [Google Scholar] [CrossRef] [PubMed]
  79. Rodríguez de la Vega, R.C.; Vidal, N.; Possani, L.D. Scorpion peptides. In Handbook of Biologically Active Peptides, 2nd ed.; Kastin, A.J., Ed.; Academic Press: Boston, MA, USA, 2013; pp. 423–429. [Google Scholar]
  80. Gregory, A.J.; Voit-Ostricki, L.; Lovas, S.; Watts, C.R. Effects of selective substitution of cysteine residues on the conformational properties of chlorotoxin explored by molecular dynamics simulations. Int. J. Mol. Sci. 2019, 20, 1261. [Google Scholar] [CrossRef]
  81. Titaux-Delgado, G.; Lopez-Giraldo, A.E.; Carrillo, E.; Cofas-Vargas, L.F.; Carranza, L.E.; López-Vera, E.; García-Hernández, E.; Del Rio-Portilla, F. β-KTx14.3, a scorpion toxin, blocks the human potassium channel KCNQ1. Biochim. Biophys. Acta Proteins Proteom. 2023, 1871, 140906. [Google Scholar] [CrossRef]
  82. Ghosh, A.; Roy, R.; Nandi, M.; Mukhopadhyay, A. Scorpion venom-toxins that aid in drug development: A review. Int. J. Pept. Res. Ther. 2019, 25, 27–37. [Google Scholar] [CrossRef] [PubMed]
  83. Valdivia, H.H.; Martin, B.M.; Ramírez, A.N.; Fletcher, P.L.; Possani, L.D. Isolation and pharmacological characterization of four novel Na+ channel-blocking toxins from the scorpion Centruroides noxius Homann. J. Biochem. 1994, 116, 1383–1391. [Google Scholar] [CrossRef]
  84. Wang, X.; Zhang, S.; Zhu, Y.; Zhang, Z.; Sun, M.; Cheng, J.; Xiao, Q.; Li, G.; Tao, J. Scorpion toxins from Buthus martensii Karsch (BmK) as potential therapeutic agents for neurological disorders: State of the art and beyond. In Medical Toxicology; Erkekoglu, P., Ogawa, T., Eds.; IntechOpen: London, UK, 2020; pp. 1–20. [Google Scholar] [CrossRef]
  85. Wisedchaisri, G.; Gamal El-Din, T.M. Druggability of voltage-gated sodium channels—Exploring old and new drug receptor sites. Front. Pharmacol. 2022, 13, 858348. [Google Scholar] [CrossRef] [PubMed]
  86. Abbas, F.; Blömer, L.A.; Millet, H.; Montnach, J.; De Waard, M.; Canepari, M. Analysis of the effect of the scorpion toxin AaH-II on action potential generation in the axon initial segment. Sci. Rep. 2024, 14, 4967. [Google Scholar] [CrossRef]
  87. Sun, H.Y.; Zhu, H.F.; Ji, Y.H. BmK I, an α-like scorpion neurotoxin, specifically modulates isolated rat cardiac mechanical and electrical activity. Sheng Li Xue Bao. 2003, 55, 530–534. [Google Scholar] [PubMed]
  88. Rowe, A.H.; Xiao, Y.; Scales, J.; Linse, K.D.; Rowe, M.P.; Cummins, T.R.; Zakon, H.H. Isolation and characterization of CvIV4: A pain inducing α-scorpion toxin. PLoS ONE 2011, 6, e23520. [Google Scholar] [CrossRef] [PubMed]
  89. Chen, H.; Gordon, D.; Heinemann, S.H. Modulation of cloned skeletal muscle sodium channels by the scorpion toxins Lqh II, Lqh III, and Lqh αIT. Pflug. Arch. 2000, 439, 423–432. [Google Scholar] [CrossRef]
  90. Cestèle, S.; Gordon, D.; Kopeyan, C.; Rochat, H. Toxin III from Leiurus quinquestriatus quinquestriatus: A specific probe for receptor site 3 on insect sodium channels. Insect Biochem. Mol. Biol. 1997, 27, 523–528. [Google Scholar] [CrossRef]
  91. Goyffon, M.; Tournier, J.-N. Scorpions: A presentation. Toxins 2014, 6, 2137–2148. [Google Scholar] [CrossRef]
  92. Martin-Eauclaire, M.F.; Bougis, P.E.; de Lima, M.E. Ts1 from the Brazilian scorpion Tityus serrulatus: A half-century of studies on a multifunctional β-like toxin. Toxicon 2018, 152, 106–120. [Google Scholar] [CrossRef] [PubMed]
  93. Cologna, C.T.; Peigneur, S.; Rustiguel, J.K.; Nonato, M.C.; Tytgat, J.; Arantes, E.C. Investigation of the relationship between the structure and function of Ts2, a neurotoxin from Tityus serrulatus venom. FEBS J. 2012, 279, 1495–1504. [Google Scholar] [CrossRef]
  94. Sampaio, S.V.; Arantes, E.C.; Prado, W.A.; Riccioppo Neto, F.; Giglio, J.R. Further characterization of toxins T1IV (TsTX-III) and T2IV from Tityus serrulatus scorpion venom. Toxicon 1991, 29, 663–672. [Google Scholar] [CrossRef]
  95. Israel, M.R.; Tanaka, B.S.; Castro, J.; Thongyoo, P.; Robinson, S.D.; Zhao, P.; Deuis, J.R.; Craik, D.J.; Durek, T.; Brierley, S.M.; et al. Nav 1.6 regulates excitability of mechanosensitive sensory neurons. J. Physiol. 2019, 597, 3751–3768. [Google Scholar] [CrossRef]
  96. del Río-Portilla, F.; Hernández-Marín, E.; Pimienta, G.; Coronas, F.V.; Zamudio, F.Z.; Rodríguez de la Vega, R.C.; Wanke, E.; Possani, L.D. NMR solution structure of Cn12, a novel peptide from the Mexican scorpion Centruroides noxius with a typical β-toxin sequence but with α-like physiological activity. Eur. J. Biochem. 2004, 271, 2504–2516. [Google Scholar] [CrossRef]
  97. He, H.; Liu, Z.; Dong, B.; Zhang, J.; Shu, X.; Zhou, J.; Ji, Y. Localization of receptor site on insect sodium channel for depressant β-toxin BmK IT2. PLoS ONE 2011, 6, e14510. [Google Scholar] [CrossRef]
  98. Song, W.; Du, Y.; Liu, Z.; Luo, N.; Turkov, M.; Gordon, D.; Gurevitz, M.; Goldin, A.L.; Dong, K. Substitutions in the domain III voltage-sensing module enhance the sensitivity of an insect sodium channel to a scorpion β-toxin. J. Biol. Chem. 2011, 286, 15781–15788. [Google Scholar] [CrossRef]
  99. Housley, D.M.; Housley, G.D.; Liddell, M.J.; Jennings, E.A. Scorpion toxin peptide action at the ion channel subunit level. Neuropharmacology 2017, 127, 46–78. [Google Scholar] [CrossRef] [PubMed]
  100. Xu, X.; Cao, Z.; Sheng, J.; Wu, W.; Luo, F.; Sha, Y.; Mao, X.; Liu, H.; Jiang, D.; Li, W. Genomic sequence analysis and organization of BmKαTx11 and BmKαTx15 from Buthus martensii Karsch: Molecular evolution of α-toxin genes. J. Biochem. Mol. Biol. 2005, 38, 386–390. [Google Scholar] [CrossRef] [PubMed]
  101. Rodríguez de la Vega, R.C.; Possani, L.D. Current views on scorpion toxins specific for K+-channels. Toxicon 2004, 43, 865–875. [Google Scholar] [CrossRef]
  102. Naseem, M.U.; Gurrola-Briones, G.; Romero-Imbachi, M.R.; Borrego, J.; Carcamo-Noriega, E.; Beltrán-Vidal, J.; Zamudio, F.Z.; Shakeel, K.; Possani, L.D.; Panyi, G. Characterization and chemical synthesis of Cm39 (α-KTx 4.8): A scorpion toxin that inhibits voltage-gated K+ channel KV1.2 and small- and intermediate-conductance Ca2+-activated K+ channels KCa2.2 and KCa3.1. Toxins 2023, 15, 41. [Google Scholar] [CrossRef] [PubMed]
  103. Xu, C.Q.; He, L.L.; Brône, B.; Martin-Eauclaire, M.F.; Van Kerkhove, E.; Zhou, Z.; Chi, C.W. A novel scorpion toxin blocking small conductance Ca2+ activated K+ channel. Toxicon 2004, 43, 961–971. [Google Scholar] [CrossRef] [PubMed]
  104. Zhu, S.; Gao, B.; Aumelas, A.; Rodríguez, M.C.; Lanz-Mendoza, H.; Peigneur, S.; Diego-Garcia, E.; Martin-Eauclaire, M.F.; Tytgat, J.; Possani, L.D. MeuTXKβ1, a scorpion venom-derived two-domain potassium channel toxin-like peptide with cytolytic activity. Biochim. Biophys. Acta 2010, 1804, 872–883. [Google Scholar] [CrossRef]
  105. Bajaj, S.; Han, J. Venom-derived peptide modulators of cation-selective channels: Friend, foe or frenemy. Front. Pharmacol. 2019, 10, 58. [Google Scholar] [CrossRef]
  106. Tytgat, J.; Chandy, K.G.; Garcia, M.L.; Gutman, G.A.; Martin-Eauclaire, M.F.; van der Walt, J.J.; Possani, L.D. A unified nomenclature for short-chain peptides isolated from scorpion venoms: α-KTx molecular subfamilies. Trends Pharmacol. Sci. 1999, 20, 444–447. [Google Scholar] [CrossRef]
  107. Cremonez, C.M.; Maiti, M.; Peigneur, S.; Cassoli, J.S.; Dutra, A.A.A.; Waelkens, E.; Lescrinier, E.; Herdewijn, P.; de Lima, M.E.; Pimenta, A.M.C.; et al. Structural and functional elucidation of peptide Ts11 shows evidence of a novel subfamily of scorpion venom toxins. Toxins 2016, 8, 288. [Google Scholar] [CrossRef]
  108. Bergeron, Z.L.; Bingham, J.-P. Scorpion toxins specific for potassium (K+) channels: A historical overview of peptide bioengineering. Toxins 2012, 4, 1082–1119. [Google Scholar] [CrossRef]
  109. Luna-Ramirez, K.; Csoti, A.; McArthur, J.R.; Chin, Y.K.Y.; Anangi, R.; Najera, R.D.C.; Possani, L.D.; King, G.F.; Panyi, G.; Yu, H.; et al. Structural basis of the potency and selectivity of urotoxin, a potent Kv1 blocker from scorpion venom. Biochem. Pharmacol. 2020, 174, 113782. [Google Scholar] [CrossRef]
  110. Cerni, F.A.; Pucca, M.B.; Amorim, F.G.; Bordon, K.C.F.; Echterbille, J.; Quinton, L.; De Pauw, E.; Peigneur, S.; Tytgat, J.; Arantes, E.C. Isolation and characterization of Ts19 Fragment II, a new long-chain potassium channel toxin from Tityus serrulatus venom. Peptides 2016, 80, 9–17. [Google Scholar] [CrossRef] [PubMed]
  111. Zhu, S.; Li, W.; Zeng, X.; Jiang, D.; Mao, X.; Liu, H. Molecular cloning and sequencing of two ‘short chain’ and two ‘long chain’ K+ channel-blocking peptides from the Chinese scorpion Buthus martensii Karsch. FEBS Lett. 1999, 457, 509–514. [Google Scholar] [CrossRef] [PubMed]
  112. Jiménez-Vargas, J.M.; Restano-Cassulini, R.; Possani, L.D. Toxin modulators and blockers of hERG K+ channels. Toxicon 2012, 60, 492–501. [Google Scholar] [CrossRef] [PubMed]
  113. Olamendi-Portugal, T.; Bartok, A.; Zamudio-Zuñiga, F.; Balajthy, A.; Becerril, B.; Panyi, G.; Possani, L.D. Isolation, chemical and functional characterization of several new K+-channel blocking peptides from the venom of the scorpion Centruroides tecomanus. Toxicon 2016, 115, 1–12. [Google Scholar] [CrossRef]
  114. Kalapothakis, Y.; Miranda, K.; Pereira, A.H.; Witt, A.S.A.; Marani, C.; Martins, A.P.; Leal, H.G.; Campos-Júnior, E.; Pimenta, A.M.C.; Borges, A.; et al. Novel components of Tityus serrulatus venom: A transcriptomic approach. Toxicon 2021, 189, 91–104. [Google Scholar] [CrossRef]
  115. Srinivasan, K.N.; Sivaraja, V.; Huys, I.; Sasaki, T.; Cheng, B.; Kumar, T.K.; Sato, K.; Tytgat, J.; Yu, C.; San, B.C.; et al. κ-Hefutoxin1, a novel toxin from the scorpion Heterometrus fulvipes with unique structure and function. Importance of the functional diad in potassium channel selectivity. J. Biol. Chem. 2002, 277, 30040–30047. [Google Scholar] [CrossRef]
  116. Chagot, B.; Pimentel, C.; Dai, L.; Pil, J.; Tytgat, J.; Nakajima, T.; Corzo, G.; Darbon, H.; Ferrat, G. An unusual fold for potassium channel blockers: NMR structure of three toxins from the scorpion Opisthacanthus madagascariensis. Biochem. J. 2005, 388, 263–271. [Google Scholar] [CrossRef]
  117. Chen, Z.Y.; Hu, Y.T.; Yang, W.S.; He, Y.W.; Feng, J.; Wang, B.; Zhao, R.M.; Ding, J.P.; Cao, Z.J.; Li, W.X.; et al. Hg1, novel peptide inhibitor specific for Kv1.3 channels from first scorpion Kunitz-type potassium channel toxin family. J. Biol. Chem. 2012, 287, 13813–13821. [Google Scholar] [CrossRef] [PubMed]
  118. Schwartz, E.F.; Diego-Garcia, E.; Rodríguez de la Veja, R.C.; Possani, L.D. Transcriptome analysis of the venom gland of the Mexican scorpion Hadrurus gertschi (Arachnida: Scorpiones). BMC Genom. 2007, 8, 119. [Google Scholar] [CrossRef]
  119. Zhao, R.; Dai, H.; Qiu, S.; Li, T.; He, Y.; Ma, Y.; Chen, Z.; Wu, Y.; Li, W.; Cao, Z. SdPI, the first functionally characterized Kunitz-type trypsin inhibitor from scorpion venom. PLoS ONE 2011, 6, e27548. [Google Scholar] [CrossRef] [PubMed]
  120. Pucca, M.B.; Peigneur, S.; Cologna, C.T.; Cerni, F.A.; Zoccal, K.F.; Bordon, K.C.; Faccioli, L.H.; Tytgat, J.; Arantes, E.C. Electrophysiological characterization of the first Tityus serrulatus α-like toxin, Ts5: Evidence of a pro-inflammatory toxin on macrophages. Biochimie 2015, 115, 8–16. [Google Scholar] [CrossRef]
  121. Kalapothakis, Y.; Miranda, K.; Molina, D.A.M.; Conceição, I.M.C.A.; Larangote, D.; Op den Camp, H.J.M.; Kalapothakis, E.; Chávez-Olórtegui, C.; Borges, A. An overview of Tityus cisandinus scorpion venom: Transcriptome and mass fingerprinting reveal conserved toxin homologs across the Amazon region and novel lipolytic components. Int. J. Biol. Macromol. 2023, 225, 1246–1266. [Google Scholar] [CrossRef] [PubMed]
  122. Gao, B.; Harvey, P.J.; Craik, D.J.; Ronjat, M.; De Waard, M.; Zhu, S. Functional evolution of scorpion venom peptides with an inhibitor cystine knot fold. Biosci. Rep. 2013, 33, e00047. [Google Scholar] [CrossRef]
  123. Smith, J.J.; Hill, J.M.; Little, M.J.; Nicholson, G.M.; King, G.F.; Alewood, P.F. Unique scorpion toxin with a putative ancestral fold provides insight into evolution of the inhibitor cystine knot motif. Proc. Natl. Acad. Sci. USA 2011, 108, 10478–10483. [Google Scholar] [CrossRef]
  124. Xiao, L.; Gurrola, G.B.; Zhang, J.; Valdivia, C.R.; SanMartin, M.; Zamudio, F.Z.; Zhang, L.; Possani, L.D.; Valdivia, H.H. Structure-function relationships of peptides forming the calcin family of ryanodine receptor ligands. J. Gen. Physiol. 2016, 147, 375–394. [Google Scholar] [CrossRef]
  125. Simeoni, I.; Rossi, D.; Zhu, X.; García, J.; Valdivia, H.H.; Sorrentino, V. Imperatoxin A (IpTxa) from Pandinus imperator stimulates [3H]ryanodine binding to RyR3 channels. FEBS Lett. 2001, 508, 5–10. [Google Scholar] [CrossRef]
  126. Fajloun, Z.; Kharrat, R.; Chen, L.; Lecomte, C.; Di Luccio, E.; Bichet, D.; El Ayeb, M.; Rochat, H.; Allen, P.D.; Pessah, I.N.; et al. Chemical synthesis and characterization of maurocalcine, a scorpion toxin that activates Ca2+ release channel/ryanodine receptors. FEBS Lett. 2000, 469, 179–185. [Google Scholar] [CrossRef] [PubMed]
  127. Shahbazzadeh, D.; Srairi-Abid, N.; Feng, W.; Ram, N.; Borchani, L.; Ronjat, M.; Akbari, A.; Pessah, I.N.; De Waard, M.; El Ayeb, M. Hemicalcin, a new toxin from the Iranian scorpion Hemiscorpius lepturus which is active on ryanodine-sensitive Ca2+ channels. Biochem. J. 2007, 404, 89–96. [Google Scholar] [CrossRef] [PubMed]
  128. Schwartz, E.F.; Capes, E.M.; Diego-García, E.; Zamudio, F.Z.; Fuentes, O.; Possani, L.D.; Valdivia, H.H. Characterization of hadrucalcin, a peptide from Hadrurus gertschi scorpion venom with pharmacological activity on ryanodine receptors. Br. J. Pharmacol. 2009, 157, 392–403. [Google Scholar] [CrossRef]
  129. Zamaleeva, A.I.; Collot, M.; Bahembera, E.; Tisseyre, C.; Rostaing, P.; Yakovlev, A.V.; Oheim, M.; de Waard, M.; Mallet, J.M.; Feltz, A. Cell-penetrating nanobiosensors for pointillistic intracellular Ca2+-transient detection. Nano Lett. 2014, 14, 2994–3001. [Google Scholar] [CrossRef]
  130. Luna-Ramírez, K.; Bartok, A.; Restano-Cassulini, R.; Quintero-Hernández, V.; Coronas, F.I.; Christensen, J.; Wright, C.E.; Panyi, G.; Possani, L.D. Structure, molecular modeling, and function of the novel potassium channel blocker urotoxin isolated from the venom of the Australian scorpion Urodacus yaschenkoi. Mol. Pharmacol. 2014, 86, 28–41. [Google Scholar] [CrossRef]
  131. Vargas-Jaimes, L.; Xiao, L.; Zhang, J.; Possani, L.D.; Valdivia, H.H.; Quintero-Hernández, V. Recombinant expression of intrepicalcin from the scorpion Vaejovis intrepidus and its effect on skeletal ryanodine receptors. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 936–946. [Google Scholar] [CrossRef]
  132. Zhijian, C.; Yun, X.; Chao, D.; Shunyi, Z.; Shijin, Y.; Yingliang, W.; Wenxin, L. Cloning and characterization of a novel calcium channel toxin-like gene BmCa1 from Chinese scorpion Mesobuthus martensii Karsch. Peptides 2006, 27, 1235–1240. [Google Scholar] [CrossRef]
  133. Zhu, S.; Darbon, H.; Dyason, K.; Verdonck, F.; Tytgat, J. Evolutionary origin of inhibitor cystine knot peptides. FASEB J. 2003, 17, 1765–1767. [Google Scholar] [CrossRef]
  134. Lan, Z.D.; Dai, L.; Zhou, X.L.; Feng, J.C.; Xu, K.; Chi, C.W. Gene cloning and sequencing of BmK AS and BmK AS-1, two novel neurotoxins from the scorpion Buthus martensi Karsch. Toxicon 1999, 37, 815–823. [Google Scholar] [CrossRef]
  135. McCarthy, S.; Robinson, J.; Thalassinos, K.; Tabor, A.B. A chemical biology approach to probing the folding pathways of the inhibitory cystine knot (ICK) peptide ProTx-II. Front. Chem. 2020, 8, 228. [Google Scholar] [CrossRef] [PubMed]
  136. Lanner, J.T.; Georgiou, D.K.; Joshi, A.D.; Hamilton, S.L. Ryanodine receptors: Structure, expression, molecular details, and function in calcium release. Cold Spring Harb. Perspect. Biol. 2010, 2, a003996. [Google Scholar] [CrossRef]
  137. Rufenach, B.; Van Petegem, F. Structure and function of STAC proteins: Calcium channel modulators and critical components of muscle excitation-contraction coupling. J. Biol. Chem. 2021, 297, 100874. [Google Scholar] [CrossRef]
  138. Haji-Ghassemi, O.; Chen, Y.S.; Woll, K.; Gurrola, G.B.; Valdivia, C.R.; Cai, W.; Li, S.; Valdivia, H.H.; Van Petegem, F. Cryo-EM analysis of scorpion toxin binding to ryanodine receptors reveals subconductance that is abolished by PKA phosphorylation. Sci Adv. 2023, 9, eadf4936. [Google Scholar] [CrossRef]
  139. Gurrola, G.B.; Arévalo, C.; Sreekumar, R.; Lokuta, A.J.; Walker, J.W.; Valdivia, H.H. Activation of ryanodine receptors by imperatoxin A and a peptide segment of the II-III loop of the dihydropyridine receptor. J. Biol. Chem. 1999, 274, 7879–7886. [Google Scholar] [CrossRef]
  140. Hille, B. Ion Channels of Excitable Membranes, 3rd ed.; Sinauer Associates: Sunderland, MA, USA, 2001; p. 814. [Google Scholar]
  141. Verkman, A.; Galietta, L. Chloride channels as drug targets. Nat. Rev. Drug. Discov. 2009, 8, 153–171. [Google Scholar] [CrossRef] [PubMed]
  142. Jentsch, T.J.; Pusch, M. CLC chloride channels and transporters: Structure, function, physiology, and disease. Physiol. Rev. 2018, 98, 1493–1590. [Google Scholar] [CrossRef] [PubMed]
  143. DeBin, J.A.; Strichartz, G.R. Chloride channel inhibition by the venom of the scorpion Leiurus quinquestriatus. Toxicon 1991, 29, 1403–1408. [Google Scholar] [CrossRef]
  144. DeBin, J.A.; Maggio, J.E.; Strichartz, G.R. Purification and characterization of chlorotoxin, a chloride channel ligand from the venom of the scorpion. Am. J. Physiol. 1993, 264, C361–C369. [Google Scholar] [CrossRef] [PubMed]
  145. Thompson, C.H.; Fields, D.M.; Olivetti, P.R.; Fuller, M.D.; Zhang, Z.R.; Kubanek, J.; McCarty, N.A. Inhibition of ClC-2 chloride channels by a peptide component or components of scorpion venom. J. Membr. Biol. 2005, 208, 65–76. [Google Scholar] [CrossRef]
  146. Dalton, S.; Gerzanich, V.; Chen, M.; Dong, Y.; Shuba, Y.; Simardi, J.M. Chlorotoxin-sensitive Ca2+-activated Cl- channel in type R2 reactive astrocytes from adult rat brain. Glia 2003, 42, 325–339. [Google Scholar] [CrossRef]
  147. Maertens, C.; Wei, L.; Tytgat, J.; Droogmans, G.; Nilius, B. Chlorotoxin does not inhibit volume-regulated, calcium-activated and cyclic AMP-activated chloride channels. Br. J. Pharmacol. 2000, 129, 791–801. [Google Scholar] [CrossRef]
  148. Ni, M.-M.; Jie-Yu Sun, J.-Y.; Li, Z.Q.; Qiu, J.C.; Wu, C.-F. Role of voltage-gated chloride channels in epilepsy: Current insights and future directions. Front. Pharmacol. 2025, 16, 1560392. [Google Scholar] [CrossRef]
  149. Xu, X.; Duan, Z.; Di, Z.; He, Y.; Li, J.; Li, Z.; Xie, C.; Zeng, X.; Cao, Z.; Wu, Y.; et al. Proteomic analysis of the venom from the scorpion Mesobuthus martensii. J. Proteom. 2014, 106, 162–180. [Google Scholar] [CrossRef] [PubMed]
  150. Mille, B.G.; Peigneur, S.; Predel, R.; Tytgat, J. Trancriptomic approach reveals the molecular diversity of Hottentotta conspersus (Buthidae) venom. Toxicon 2015, 99, 73–79. [Google Scholar] [CrossRef]
  151. Daoudi, K.; Malosse, C.; Lafnoune, A.; Darkaoui, B.; Chakir, S.; Sabatier, J.M.; Chamot-Rooke, J.; Cadi, R.; Oukkache, N. Mass spectrometry-based top-down and bottom-up approaches for proteomic analysis of the Moroccan Buthus occitanus scorpion venom. FEBS Open Biol. 2021, 11, 1867–1892. [Google Scholar] [CrossRef] [PubMed]
  152. Borges, A.; Lomonte, B. Proteomic analysis and lethality of the venom of Aegaeobuthus nigrocinctus, a scorpion of medical significance in the Middle East. Acta Trop. 2024, 255, 107230. [Google Scholar] [CrossRef]
  153. Arzamasov, A.A.; Vassilevski, A.A.; Grishin, E.V. Chlorotoxin and related peptides: Short insect toxins from scorpion venom. Russ. J. Bioorg. Chem. 2014, 40, 359–369. [Google Scholar] [CrossRef] [PubMed]
  154. Thompson, C.H.; Olivetti, P.R.; Fuller, M.D.; Freeman, C.S.; McMaster, D.; French, R.J.; Pohl, J.; Kubanek, J.; McCarty, N.A. Isolation and characterization of a high affinity peptide inhibitor of ClC-2 chloride channels. J. Biol. Chem. 2009, 284, 26051–26062. [Google Scholar] [CrossRef]
  155. Fuller, M.D.; Zhang, Z.R.; Cui, G.; Kubanek, J.; McCarty, N.A. Inhibition of CFTR channels by a peptide toxin of scorpion venom. Am. J. Physiol. Cell Physiol. 2004, 287, C1328–C1341. [Google Scholar] [CrossRef]
  156. Fuller, M.D.; Zhang, Z.R.; Cui, G.; McCarty, N.A. The block of CFTR by scorpion venom is state-dependent. Biophys. J. 2005, 89, 3960–3975. [Google Scholar] [CrossRef]
  157. Fuller, M.D.; Thompson, C.H.; Zhang, Z.R.; Freeman, C.S.; Schay, E.; Szakács, G.; Bakos, E.; Sarkadi, B.; McMaster, D.; French, R.J.; et al. State-dependent inhibition of cystic fibrosis transmembrane conductance regulator chloride channels by a novel peptide toxin. J. Biol. Chem. 2007, 282, 37545–37555. [Google Scholar] [CrossRef] [PubMed]
  158. Morel, J.-L.; Mokrzycki, N.; Lippens, G.; Drobecq, H.; Sautière, P.; Hugues, M. Characterization of a family of scorpion toxins modulating Ca2+-activated Cl current in vascular myocytes. Toxins 2022, 14, 780. [Google Scholar] [CrossRef]
  159. Wu, J.J.; Dai, L.; Lan, Z.D.; Chi, C.W. The gene cloning and sequencing of Bm-12, a chlorotoxin-like peptide from the scorpion Buthus martensi Karsch. Toxicon 2000, 38, 661–668. [Google Scholar] [CrossRef] [PubMed]
  160. Zeng, X.C.; Li, W.X.; Zhu, S.Y.; Peng, F.; Zhu, Z.H.; Wu, K.L.; Yiang, F.H. Cloning and characterization of a cDNA sequence encoding the precursor of a chlorotoxin-like peptide from the Chinese scorpion Buthus martensii Karsch. Toxicon 2000, 38, 1009–1014. [Google Scholar] [CrossRef]
  161. Diego-García, E.; Caliskan, F.; Tytgat, J. The Mediterranean scorpion Mesobuthus gibbosus (Scorpiones, Buthidae): Transcriptome analysis and organization of the genome encoding chlorotoxin-like peptides. BMC Genom. 2014, 15, 295. [Google Scholar] [CrossRef]
  162. Lippens, G.; Najib, J.; Wodak, S.J.; Tartar, A. NMR sequential assignments and solution structure of chlorotoxin, a small scorpion toxin that blocks chloride channels. Biochemistry 1995, 34, 13–21. [Google Scholar] [CrossRef]
  163. Rjeibi, I.; Mabrouk, K.; Mosrati, H.; Berenguer, C.; Mejdoub, H.; Villard, C.; Laffitte, D.; Bertin, D.; Ouafik, L.; Luis, J.; et al. Purification, synthesis and characterization of AaCtx, the first chlorotoxin-like peptide from Androctonus australis scorpion venom. Peptides 2011, 32, 656–663. [Google Scholar] [CrossRef]
  164. Ali, S.A.; Alam, M.; Abbasi, A.; Undheim, E.A.; Fry, B.G.; Kalbacher, H.; Voelter, W. Structure-activity relationship of chlorotoxin-like peptides. Toxins 2016, 8, 36. [Google Scholar] [CrossRef]
  165. Dhawan, R.D.; Joseph, S.; Sethi, A.; Lala, A.K. Purification and characterization of a short insect toxin from the venom of the scorpion Buthus tamulus. FEBS Lett. 2002, 528, 261–266. [Google Scholar] [CrossRef]
  166. Wudayagiri, R.; Inceoglu, B.; Herrmann, R.; Derbel, M.; Choudary, P.V.; Hammock, B.D. Isolation and characterization of a novel lepidopteran-selective toxin from the venom of South Indian red scorpion, Mesobuthus tamulus. BMC Biochem. 2001, 2, 16. [Google Scholar] [CrossRef] [PubMed]
  167. Baradaran, M.; Jalali, A.; Soorki, M.N.; Jokar, M.; Galehdari, H. Three new scorpion chloride channel toxins as potential anti-cancer drugs: Computational prediction of the interactions with Hmmp-2 by docking and steered molecular dynamics simulations. Iran. J. Pharm. Res. 2019, 18, 720–734. [Google Scholar] [CrossRef] [PubMed]
  168. Soorki, M.N.; Jalali, A.; Galehdari, H. Molecular characterization and biodiversity of a putative chlorotoxin from the Iranian yellow scorpion Odontobuthus doriae. Iran. Biomed. J. 2017, 21, 342–346. [Google Scholar] [CrossRef]
  169. Megaly, A.M.A.; Nakamichi, R.; Wakayu, M.; Nakagawa, Y.; Abdel-Wahab, M.; Miyashita, M. Identification of a novel insecticidal chlorotoxin-like peptide from the venom of the Compsobuthus egyptiensis scorpion. Toxicon 2025, 267, 108556. [Google Scholar] [CrossRef]
  170. Ojeda, P.G.; Wang, C.K.; Craik, D.J. Chlorotoxin: Structure, activity, and potential uses in cancer therapy. Pept. Sci. 2016, 106, 25–36. [Google Scholar] [CrossRef]
  171. Soroceanu, L.; Gillespie, Y.; Khazaeli, M.B.; Sontheimer, H. Use of chlorotoxin for targeting of primary brain tumors. Cancer Res. 1998, 58, 4871–4879. [Google Scholar]
  172. Soroceanu, L.; Manning, T.J., Jr.; Sontheimer, H. Modulation of glioma cell migration and invasion using Cl- and K+ ion channel blockers. J. Neurosci. 1999, 19, 5942–5954. [Google Scholar] [CrossRef]
  173. Fu, Y.J.; Yin, L.T.; Liang, A.H.; Zhang, C.F.; Wang, W.; Chai, B.F.; Yang, J.Y.; Fan, X.J. Therapeutic potential of chlorotoxin-like neurotoxin from the Chinese scorpion for human gliomas. Neurosci. Lett. 2007, 412, 62–67. [Google Scholar] [CrossRef]
  174. Fan, S.; Sun, Z.; Jiang, D.; Dai, C.; Ma, Y.; Zhao, Z.; Liu, H.; Wu, Y.; Cao, Z.; Li, W. BmKCT toxin inhibits glioma proliferation and tumor metastasis. Cancer Lett. 2010, 291, 158–166. [Google Scholar] [CrossRef]
  175. Xiang, Y.; Liang, L.; Wang, X.; Wang, J.; Zhang, X.; Zhang, Q. Chloride channel-mediated brain glioma targeting of chlorotoxin-modified doxorubicine-loaded liposomes. J. Control. Release 2011, 152, 402–410. [Google Scholar] [CrossRef] [PubMed]
  176. Fu, Y.; An, N.; Li, K.; Zheng, Y.; Liang, A. Chlorotoxin-conjugated nanoparticles as potential glioma-targeted drugs. J. Neurooncol. 2012, 107, 457–462. [Google Scholar] [CrossRef] [PubMed]
  177. Cohen, G.; Burks, S.R.; Frank, J.A. Chlorotoxin—A multimodal imaging platform for targeting glioma tumors. Toxins 2018, 10, 496. [Google Scholar] [CrossRef]
  178. Khanyile, S.; Masamba, P.; Oyinloye, B.E.; Mbatha, L.S.; Kappo, A.P. Current biochemical applications and future prospects of chlorotoxin in cancer diagnostics and therapeutics. Adv. Pharm. Bull. 2019, 9, 510–520. [Google Scholar] [CrossRef]
  179. de Paula, G.A.; de Paula, M.C.; Dutra, J.A.P.; Carvalho, S.G.; Di Filippo, L.D.; Villanova, J.C.O.; Chorilli, M. Targeted polymeric nanoparticles as a strategy for the treatment of glioblastoma: A review. Curr. Drug Deliv. 2025, 22, 413–430. [Google Scholar] [CrossRef] [PubMed]
  180. Farkas, S.; Cioca, D.; Murányi, J.; Hornyák, P.; Brunyánszki, A.; Szekér, P.; Boros, E.; Horváth, P.; Hujber, Z.; Rácz, G.Z.; et al. Chlorotoxin binds to both matrix metalloproteinase 2 and neuropilin 1. J. Biol. Chem. 2023, 299, 104998. [Google Scholar] [CrossRef] [PubMed]
  181. Othman, H.; Wieninger, S.A.; ElAyeb, M.; Nilges, M.; Srairi-Abid, N. In silico prediction of the molecular basis of ClTx and AaCTx interaction with matrix metalloproteinase-2 (MMP-2) to inhibit glioma cell invasion. J. Biomol. Struct. Dyn. 2017, 35, 2815–2829. [Google Scholar] [CrossRef]
  182. Zeng, X.-C.; Corzo, G.; Hahin, R. Scorpion venom peptides without disulfide bridges. IUBMB Life 2005, 57, 13–21. [Google Scholar] [CrossRef]
  183. Almaaytah, A.; Albalas, Q. Scorpion venom peptides with no disulfide bridges: A review. Peptides 2014, 51, 35–45. [Google Scholar] [CrossRef]
  184. Harrison, P.L.; Abdel-Rahman, M.A.; Miller, K.; Strong, P.N. Antimicrobial peptides from scorpion venoms. Toxicon 2014, 88, 115–137. [Google Scholar] [CrossRef] [PubMed]
  185. Dias, N.B.; de Souza, B.M.; Cocchi, F.K.; Chalkidis, H.M.; Dorce, V.A.C.; Palma, M.S. Profiling the short, linear, non-disulfide bond-containing peptidome from the venom of the scorpion Tityus obscurus. J. Proteom. 2018, 170, 70–79. [Google Scholar] [CrossRef]
  186. Cassini-Vieira, P.; Felipetto, M.; Prado, L.B.; Verano-Braga, T.; Andrade, S.P.; Santos, R.A.S.; Teixeira, M.M.; de Lima, M.E.; Pimenta, A.M.C.; Barcelos, L.S. Ts14 from Tityus serrulatus boosts angiogenesis and attenuates inflammation and collagen deposition in sponge-induced granulation tissue in mice. Peptides 2017, 98, 63–69. [Google Scholar] [CrossRef]
  187. Goméz-Mendoza, D.P.; Lemos, R.P.; Jesus, I.C.G.; Gorshkov, V.; McKinnie, S.M.K.; Vederas, J.C.; Kjeldsen, F.; Guatimosim, S.; Santos, R.A.; Pimenta, A.M.C.; et al. Moving pieces in a cellular puzzle: A cryptic peptide from the scorpion toxin Ts14 activates AKT and ERK signaling and decreases cardiac myocyte contractility via dephosphorylation of phospholamban. J. Proteome Res. 2020, 19, 3467–3477. [Google Scholar] [CrossRef] [PubMed]
  188. Zeng, X.C.; Zhang, L.; Nie, Y.; Luo, X. Identification and molecular characterization of three new K+-channel specific toxins from the Chinese scorpion Mesobuthus martensii Karsch revealing intronic number polymorphism and alternative splicing in duplicated genes. Peptides 2012, 34, 311–323. [Google Scholar] [CrossRef] [PubMed]
  189. Venancio, E.J.; Portaro, F.C.; Kuniyoshi, A.K.; Carvalho, D.C.; Pidde-Queiroz, G.; Tambourgi, D.V. Enzymatic properties of venoms from Brazilian scorpions of Tityus genus and the neutralisation potential of therapeutical antivenoms. Toxicon 2013, 69, 180–190. [Google Scholar] [CrossRef]
  190. Mendoza-Tobar, L.L.; Clement, H.; Arenas, I.; Sepulveda-Arias, J.C.; Vargas, J.A.G.; Corzo, G. An overview of some enzymes from buthid scorpion venoms from Colombia: Centruroides margaritatus, Tityus pachyurus, and Tityus n. sp. aff. metuendus. J. Venom. Anim. Toxins Incl. Trop. Dis. 2024, 30, e20230063. [Google Scholar] [CrossRef]
  191. de Oliveira, I.S.; Alano-da-Silva, N.M.; Ferreira, I.G.; Cerni, F.A.; Sachett, J.A.G.; Monteiro, W.M.; Pucca, M.B.; Arantes, E.C. Understanding the complexity of Tityus serrulatus venom: A focus on high molecular weight components. J. Venom. Anim. Toxins Incl. Trop. Dis. 2024, 30, e20230046. [Google Scholar] [CrossRef]
  192. Díaz, C.; Lomonte, B.; Chang-Castillo, A.; Bonilla, F.; Alfaro-Chinchilla, A.; Triana, F.; Angulo, D.; Fernández, J.; Sasa, M. Venomics of scorpion Ananteris platnicki (Lourenco, 1993), a New World buthid that inhabits Costa Rica and Panama. Toxins 2024, 16, 327. [Google Scholar] [CrossRef]
  193. Wiezel, G.A.; Oliveira, I.S.; Reis, M.B.; Ferreira, I.G.; Cordeiro, K.R.; Bordon, K.C.F.; Arantes, E.C. The complex repertoire of Tityus spp. venoms: Advances on their composition and pharmacological potential of their toxins. Biochimie 2024, 220, 144–166. [Google Scholar] [CrossRef]
  194. Das, B.; Patra, A.; Mukherjee, A.K. Correlation of venom toxinome composition of Indian red scorpion (Mesobuthus tamulus) with clinical manifestations of scorpion stings: Failure of commercial antivenom to immune-recognize the abundance of low molecular mass toxins of this venom. J. Proteome Res. 2020, 19, 1847–1856. [Google Scholar] [CrossRef]
  195. Guerra-Duarte, C.; Horta, C.C.R.; Oliveira-Mendes, B.B.R.; Magalhães, B.F.; Costal-Oliveira, F.; Stransky, S.; de Freitas, C.F.; Campolina, D.; Pardal, P.P.O.; Lira-da-Silva, R.; et al. Determination of hyaluronidase activity in Tityus spp. scorpion venoms and its inhibition by Brazilian antivenoms. Toxicon 2019, 167, 134–143. [Google Scholar] [CrossRef]
  196. Abreu, C.B.; Bordon, K.C.F.; Cerni, F.A.; Oliveira, I.S.; Balenzuela, C.; Alexandre-Silva, G.M.; Zoccal, K.F.; Reis, M.B.; Wiezel, G.A.; Peigneur, S.; et al. Pioneering study on Rhopalurus crassicauda scorpion venom: Isolation and characterization of the major toxin and hyaluronidase. Front. Immunol. 2020, 11, 2011. [Google Scholar] [CrossRef] [PubMed]
  197. Valdez-Velázquez, L.L.; Cid-Uribe, J.; Romero-Gutierrez, M.T.; Olamendi-Portugal, T.; Jimenez-Vargas, J.M.; Possani, L.D. Transcriptomic and proteomic analyses of the venom and venom glands of Centruroides hirsutipalpus, a dangerous scorpion from Mexico. Toxicon 2020, 179, 21–32. [Google Scholar] [CrossRef]
  198. Magalhães, A.C.M.; de Santana, C.J.C.; Melani, R.D.; Domont, G.B.; Castro, M.S.; Fontes, W.; Roepstorff, P.; Júnior, O.R.P. Exploring the biological activities and proteome of Brazilian scorpion Rhopalurus agamemnon venom. J. Proteom. 2021, 237, 104119. [Google Scholar] [CrossRef] [PubMed]
  199. Ghezellou, P.; Jakob, K.; Atashi, J.; Ghassempour, A.; Spengler, B. Mass-spectrometry-based lipidome and proteome profiling of Hottentotta saulcyi (Scorpiones: Buthidae) venom. Toxins 2022, 14, 370. [Google Scholar] [CrossRef]
  200. Silva de França, F.; Tambourgi, D.V. Hyaluronan breakdown by snake venom hyaluronidases: From toxins delivery to immunopathology. Front. Immunol. 2023, 14, 1125899. [Google Scholar] [CrossRef] [PubMed]
  201. Pessini, A.C.; Takao, T.T.; Cavalheiro, E.C.; Vichnewski, W.; Sampaio, S.V.; Giglio, J.R.; Arantes, E.C. A hyaluronidase from Tityus serrulatus scorpion venom: Isolation, characterization and inhibition by flavonoids. Toxicon 2001, 39, 1495–1504. [Google Scholar] [CrossRef] [PubMed]
  202. Girish, K.S.; Kemparaju, K. The magic glue hyaluronan and its eraser hyaluronidase: A biological overview. Life Sci. 2007, 80, 1921–1943. [Google Scholar] [CrossRef]
  203. Khan, N.; Niazi, Z.R.; Rehman, F.U.; Akhtar, A.; Khan, M.M.; Khan, S.; Baloch, N.; Khan, S. Hyaluronidases: A therapeutic enzyme. Protein Pept. Lett. 2018, 25, 663–676. [Google Scholar] [CrossRef]
  204. De Oliveira-Mendes, B.B.R.; Miranda, S.E.M.; Sales-Medina, D.F.; de Magalhães, B.F.; Kalapothakis, Y.; deSouza, R.P.; Cardoso, V.N.; de Barros, A.L.B.; Guerra-Duarte, C.; Kalapothakis, E.; et al. Inhibition of Tityus serrulatus venom hyaluronidase affects venom biodistribution. PLoS Negl. Trop. Dis. 2019, 13, e0007048. [Google Scholar] [CrossRef]
  205. Gutiérrez, J.M.; Lomonte, B. Phospholipases A2: Unveiling the secrets of a functionally versatile group of snake venom toxins. Toxicon 2013, 62, 27–39. [Google Scholar] [CrossRef]
  206. Lomonte, B. Lys49 myotoxins, secreted phospholipase A2-like proteins of viperid venoms: A comprehensive review. Toxicon 2023, 224, 107024. [Google Scholar] [CrossRef]
  207. Sampat, G.H.; Hiremath, K.; Dodakallanavar, J.; Patil, V.S.; Harish, D.R.; Biradar, P.; Mahadevamurthy, R.K.; Barvaliya, M.; Roy, S. Unraveling snake venom phospholipase A2: An overview of its structure, pharmacology, and inhibitors. Pharmacol. Rep. 2023, 75, 1454–1473. [Google Scholar] [CrossRef]
  208. Zambelli, V.O.; Picolo, G.; Fernandes, C.A.H.; Fontes, M.R.M.; Cury, Y. Secreted phospholipases A2 from animal venoms in pain and analgesia. Toxins 2017, 9, 406. [Google Scholar] [CrossRef]
  209. Mamede, C.C.N.; Simamoto, B.B.S.; Pereira, D.F.C.; Costa, J.O.; Ribeiro, M.S.M.; de Oliveira, F. Edema, hyperalgesia and myonecrosis induced by Brazilian bothropic venoms: Overview of the last decade. Toxicon 2020, 187, 10–18. [Google Scholar] [CrossRef]
  210. Moreira, V.; Leiguez, E.; Janovits, P.M.; Maia-Marques, R.; Fernandes, C.M.; Teixeira, C. Inflammatory effects of Bothrops phospholipases A2: Mechanisms involved in biosynthesis of lipid mediators and lipid accumulation. Toxins 2021, 13, 868. [Google Scholar] [CrossRef]
  211. Tonello, F. Secretory phospholipases A2, from snakebite envenoming to a myriad of inflammation associated human diseases—What is the secret of their activity? Int. J. Mol. Sci. 2023, 24, 1579. [Google Scholar] [CrossRef]
  212. Salvador, G.H.M.; Cardoso, F.F.; Lomonte, B.; Fontes, M.R.M. Inhibitors and activators for myotoxic phospholipase A2-like toxins from snake venoms—A structural overview. Biochimie 2024, 227, 231–247. [Google Scholar] [CrossRef]
  213. Almeida, F.M.; Pimenta, A.M.C.; De Figueiredo, S.G.; Santoro, M.M.; Martin-Eauclaire, M.F.; Diniz, C.R.; De Lima, M.E. Enzymes with gelatinolytic activity can be found in Tityus bahiensis and Tityus serrulatus venoms. Toxicon 2002, 40, 1041–1045. [Google Scholar] [CrossRef]
  214. Brazón, J.; Guerrero, B.; D’Suze, G.; Sevcik, C.; Arocha-Piñango, C.L. Fibrin(ogen)olytic enzymes in scorpion (Tityus discrepans) venom. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2014, 168, 62–69. [Google Scholar] [CrossRef] [PubMed]
  215. Estrada-Gómez, S.; Gomez-Rave, L.; Vargas-Muñoz, L.J.; van der Meijden, A. Characterizing the biological and biochemical profile of six different scorpion venoms from the Buthidae and Scorpionidae family. Toxicon 2017, 130, 104–115. [Google Scholar] [CrossRef] [PubMed]
  216. Romero-Gutiérrez, M.T.; Santibáñez-López, C.E.; Jiménez-Vargas, J.M.; Batista, C.V.F.; Ortiz, E.; Possani, L.D. Transcriptomic and proteomic analyses reveal the diversity of venom components from the vaejovid scorpion Serradigitus gertschi. Toxins 2018, 10, 359. [Google Scholar] [CrossRef] [PubMed]
  217. Kazemi, S.M.; Sabatier, J.-M. Venoms of Iranian scorpions (Arachnida, Scorpiones) and their potential for drug discovery. Molecules 2019, 24, 2670. [Google Scholar] [CrossRef]
  218. Rojas-Azofeifa, D.; Sasa, M.; Lomonte, B.; Diego-García, E.; Ortiz, N.; Bonilla, F.; Murillo, R.; Tytgat, J.; Díaz, C. Biochemical characterization of the venom of Central American scorpion Didymocentrus krausi Francke, 1978 (Diplocentridae) and its toxic effects in vivo and in vitro. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2019, 217, 54–67. [Google Scholar] [CrossRef]
  219. Rami, A.; Damizadeh, B.; Behdani, M.; Kazemi-Lomedasht, F. Insights into the evolutionary dynamics: Characterization of disintegrin and metalloproteinase proteins in the venom gland transcriptome of the Hemiscorpius lepturus scorpion. Protein Pept. Lett. 2024, 31, 639–656. [Google Scholar] [CrossRef]
  220. Salabi, F.; Jafari, H.; Mahdavinia, M.; Azadnasab, R.; Shariati, S.; Baghal, M.L.; Tebianian, M.; Baradaran, M. First transcriptome analysis of the venom glands of the scorpion Hottentotta zagrosensis (Scorpions: Buthidae) with focus on venom lipolysis activating peptides. Front. Pharmacol. 2024, 15, 1464648. [Google Scholar] [CrossRef] [PubMed]
  221. Alano-da-Silva, N.M.; Oliveira, I.S.; Cardoso, I.A.; Bordon, K.C.F.; Arantes, E.C. Exploring high molecular weight components in Tityus serrulatus venom. Toxicon 2025, 255, 108240. [Google Scholar] [CrossRef]
  222. Fletcher, P.L., Jr.; Fletcher, M.D.; Weninger, K.; Anderson, T.E.; Martin, B.M. Vesicle associated membrane protein (VAMP) cleavage by a new metalloprotease from the Brazilian scorpion Tityus serrulatus. J. Biol. Chem. 2010, 285, 7405–7416. [Google Scholar] [CrossRef] [PubMed]
  223. Zornetta, I.; Scorzeto, M.; Mendes dos Reis, P.V.; de Lima, M.E.; Montecucco, C.; Megighian, A.; Rossetto, O. Electrophysiological characterization of the antarease metalloprotease from Tityus serrulatus venom. Toxins 2017, 9, 81. [Google Scholar] [CrossRef]
  224. Carmo, A.O.; Oliveira-Mendes, B.B.R.; Horta, C.C.R.; Magalhães, B.F.; Dantas, A.E.; Chaves, L.M.; Chávez-Olórtegui, C.; Kalapothakis, E. Molecular and functional characterization of metalloserrulases, new metalloproteases from the Tityus serrulatus venom gland. Toxicon 2014, 90, 45–55. [Google Scholar] [CrossRef] [PubMed]
  225. Cajado-Carvalho, D.; Kuniyoshi, A.K.; Duzzi, B.; Iwai, L.K.; Oliveira, Ú.C.; Junqueira-de-Azevedo, I.L.; Kodama, R.T.; Portaro, F.V. Insights into the hypertensive effects of Tityus serrulatus scorpion venom: Purification of an angiotensin-converting enzyme-like peptidase. Toxins 2016, 8, 348. [Google Scholar] [CrossRef]
  226. Gutiérrez, J.M.; Rucavado, A. Snake venom metalloproteinases: Their role in the pathogenesis of local tissue damage. Biochimie 2000, 82, 841–850. [Google Scholar] [CrossRef]
  227. Gutiérrez, J.M.; Rucavado, A.; Escalante, T.; Díaz, C. Hemorrhage induced by snake venom metalloproteinases: Biochemical and biophysical mechanisms involved in microvessel damage. Toxicon 2005, 45, 997–1011. [Google Scholar] [CrossRef]
  228. Gutiérrez, J.M.; Escalante, T.; Rucavado, A.; Herrera, C.; Fox, J.W. A comprehensive view of the structural and functional alterations of extracellular matrix by snake venom metalloproteinases (SVMPs): Novel perspectives on the pathophysiology of envenoming. Toxins 2016, 8, 304. [Google Scholar] [CrossRef]
  229. Luo, P.; Ji, Y.; Liu, X.; Zhang, W.; Cheng, R.; Zhang, S.; Qian, X.; Huang, C. Affected inflammation-related signaling pathways in snake envenomation: A recent insight. Toxicon 2023, 234, 107288. [Google Scholar] [CrossRef]
  230. Rao, S.; Reghu, N.; Nair, B.G.; Vanuopadath, M. The role of snake venom proteins in inducing inflammation post-evenomation: An overview on mechanistic insights and treatment strategies. Toxins 2024, 16, 519. [Google Scholar] [CrossRef]
  231. Arias, A.S.; Rucavado, A.; Gutiérrez, J.M. Peptidomimetic hydroxamate metalloproteinase inhibitors abrogate local and systemic toxicity induced by Echis ocellatus (saw-scaled) snake venom. Toxicon 2017, 132, 40–49. [Google Scholar] [CrossRef] [PubMed]
  232. Soeiro, P.A.; Romanelli, M.A.; Cesar, M.O.; Nogueira-Souza, P.D.; Monteiro-Machado, M.; Oliveira, S.S.C.; Santos, A.L.S.; Melo, P.A.; Lara, L.S. Doxycycline treatment reestablishes renal function of Wistar rats in experimental envenomation with Bothrops jararacussu venom. Toxicon 2021, 99, 20–30. [Google Scholar] [CrossRef] [PubMed]
  233. Clare, R.H.; Dawson, C.A.; Westhorpe, A.; Albulescu, L.O.; Woodley, C.M.; Mosallam, N.; Chong, D.J.W.; Kool, J.; Berry, N.G.; O’Neill, P.M.; et al. Snakebite drug discovery: High-throughput screening to identify novel snake venom metalloproteinase toxin inhibitors. Front. Pharmacol. 2024, 14, 1328950. [Google Scholar] [CrossRef]
  234. Smith, C.F.; Modahl, C.M.; Galindo, D.C.; Larson, K.Y.; Maroney, S.P.; Bahrabadi, L.; Brandehoff, N.P.; Perry, B.W.; McCabe, M.C.; Petras, D.; et al. Assessing target specificity of the small molecule inhibitor MARIMASTAT to snake venom toxins: A novel application of thermal proteome profiling. Mol. Cell. Proteom. 2024, 23, 100779. [Google Scholar] [CrossRef] [PubMed]
  235. Arens, D.K.; Rose, M.A.; Salazar, E.M.; Harvey, M.A.; Huh, E.Y.; Ford, A.A.; Thompson, D.W.; Sanchez, E.E.; Hwang, Y.Y. Doxycycline-mediated inhibition of snake venom phospholipase and metalloproteinase. Mil. Med. 2024, 189, e2430–e2438. [Google Scholar] [CrossRef]
  236. Kadler, R.; Morrison, B.; Yanagihara, A.A. Assessing the utility of broad-acting inhibitors as therapeutics in diverse venoms. Toxins 2025, 17, 188. [Google Scholar] [CrossRef]
  237. Rudresha, G.V.; Khochare, S.; Casewell, N.R.; Sunagar, K. Preclinical evaluation of small molecule inhibitors as early intervention therapeutics against Russell’s viper envenoming in India. Commun. Med. 2025, 5, 226. [Google Scholar] [CrossRef] [PubMed]
  238. Gutiérrez, J.M.; Casewell, N.R.; Laustsen, A.H. Progress and challenges in the field of snakebite envenoming therapeutics. Annu. Rev. Pharmacol. Toxicol. 2025, 6, 465–485. [Google Scholar] [CrossRef]
  239. Bebber, D.P.; Ramotowski, M.A.T.; Gurr, S.J. Crop pests and pathogens move polewards in a warming world. Nat. Clim. Change 2013, 3, 985–988. [Google Scholar] [CrossRef]
  240. Paini, D.R.; Sheppard, A.W.; Cook, D.C.; Thomas, M.B. Global threat to agriculture from invasive species. Proc. Natl. Acad. Sci. USA 2016, 113, 7575–7579. [Google Scholar] [CrossRef]
  241. Bradshaw, C.J.; Leroy, B.; Bellard, C.; Roiz, D.; Albert, C.; Fournier, A.; Barbet-Massin, M.; Salles, J.M.; Simard, F.; Courchamp, F. Massive yet grossly underestimated global costs of invasive insects. Nat. Commun. 2016, 7, 12986. [Google Scholar] [CrossRef]
  242. Gula, L.T. Researchers Helping Protect Crops From Pests; The National Institute of Food and Agriculture. U.S. Department of Agriculture: Washington, DC, USA, 2023; pp. 1–5. Available online: https://www.nifa.usda.gov/about-nifa/blogs/researchers-helping-protect-crops-pests (accessed on 1 July 2005).
  243. Pu, J.; Chung, H. New and emerging mechanisms of insecticide resistance. Curr. Opin. Insect Sci. 2024, 63, 101184. [Google Scholar] [CrossRef] [PubMed]
  244. Rosa, M.E.; Campos, L.; Borges, B.T.; Santos, S.; Barreto, Y.C.; de Assis, D.R.; Hyslop, S.; de Souza, V.Q.; Vinadé, L.; Dal Belo, C.A. Fipronil affects cockroach behavior and olfactory memory. J. Exp. Biol. 2023, 226, jeb245239. [Google Scholar] [CrossRef] [PubMed]
  245. Rosa, M.E.; Oliveira, R.S.; Barbosa, R.F.; Hyslop, S.; Dal Belo, C.A. Recent advances on the influence of fipronil on insect behavior. Curr. Opin. Insect Sci. 2024, 65, 101251. [Google Scholar] [CrossRef]
  246. Ferreira, L.C.; Rosa, M.E.; Rodrigues, L.G.S.; Dias, D.R.C.; Guimarães, M.P.; Valsecchi, C.; Queiroz de Souza, V.; Barbosa, R.F.; Vinadé, L.H.; Hyslop, S.; et al. Disruption of exploratory behavior and olfactory memory in cockroaches exposed to sublethal doses of the neonicotinoid thiamethoxam. Pestic. Biochem. Physiol. 2024, 205, 106167. [Google Scholar] [CrossRef]
  247. Whalon, M.; Mota-Sanchez, D.; Hollingworth, R. Analysis of global pesticide resistance in arthropods. In Global Pesticide Resistance in Arthropods; Whalon, M.E., Ed.; CABI: Wallingford, UK, 2008; pp. 5–31. [Google Scholar]
  248. Hafeez, M.; Li, X.; Zhang, Z.; Huang, J.; Wang, L.; Zhang, J.; Shah, S.; Khan, M.M.; Xu, F.; Fernández-Grandon, G.M.; et al. De novo transcriptomic analyses revealed some detoxification genes and related pathways responsive to Noposion Yihaogong® 5% EC (Lambda-Cyhalothrin 5%) exposure in Spodoptera frugiperda third-instar larvae. Insects 2021, 12, 132. [Google Scholar] [CrossRef]
  249. Kariyanna, B.; Sowjanya, M. Unravelling the use of artificial intelligence in management of insect pests. Smart Agric. Technol. 2024, 8, 100517. [Google Scholar] [CrossRef]
  250. Zlotkin, E. The insect voltage-gated sodium channel as target of insecticides. Annu. Rev. Entomol. 1999, 44, 429–455. [Google Scholar] [CrossRef]
  251. Abdel-Rahman, M.A.; Omran, M.A.A.; Abdel-Nabi, I.M.; Nassier, O.A.; Schemerhorn, B.J. Neurotoxic and cytotoxic effects of venom from different populations of the Egyptian Scorpio maurus palmatus. Toxicon 2010, 55, 298–306. [Google Scholar] [CrossRef]
  252. Huang, L.; Xu, J.; Duan, K.; Bao, T.; Cheng, Y.; Zhang, H.; Zhang, Y.; Lin, Y.; Li, F. Scorpion venom heat-resistant peptide alleviates mitochondrial dynamics imbalance induced by PM2.5 exposure by downregulating the PGC-1α/SIRT3 signaling pathway. Toxicol. Res. 2023, 12, 756–764. [Google Scholar] [CrossRef] [PubMed]
  253. Dos Santos, D.S.; Carvalho, E.L.; de Lima, J.C.; Breda, R.V.; Oliveira, R.S.; de Freitas, T.C.; Salamoni, S.D.; Domingues, M.F.; Piovesan, A.R.; Boldo, J.T.; et al. Bothriurus bonariensis scorpion venom activates voltage-dependent sodium channels in insect and mammalian nervous systems. Chem. Biol. Interact. 2016, 258, 1–9. [Google Scholar] [CrossRef]
  254. Deng, S.Q.; Chen, J.T.; Li, W.W.; Chen, M.; Peng, H.J. Application of the scorpion neurotoxin AaIT against insect pests. Int. J. Mol. Sci. 2019, 20, 3467. [Google Scholar] [CrossRef]
  255. Zilberberg, N.; Zlotkin, E.; Gurevitz, M. The cDNA sequence of a depressant insect selective neurotoxin from the scorpion Buthotus judaicus. Toxicon 1991, 29, 1155–1158. [Google Scholar] [CrossRef]
  256. Oren, D.A.; Froy, O.; Amit, E.; Kleinberger-Doron, N.; Gurevitz, M.; Shaanan, B. An excitatory scorpion toxin with a distinctive feature: An additional a helix at the C terminus and its implications for interaction with insect sodium channels. Structure 1998, 6, 1095–1103. [Google Scholar] [CrossRef]
  257. Zlotkin, E. Insect selective toxins derived from scorpion venoms: An approach to insect neuropharmacology. Insect Biochem. 1983, 13, 219–236. [Google Scholar] [CrossRef]
  258. Bermúdez-Guzmán, M.J.; Buenrostro-Nava, M.T.; Valdez-Velázquez, L.; Lino-López, G.J.; García-Villalvazo, P.E.; Orozco-Santos, M.; Michel-López, C.Y. Scorpion venom insectotoxins: A sustainable alternative for pest control in agriculture. Phytoparasitica 2024, 52, 83. [Google Scholar] [CrossRef]
  259. Riaz, N.; Zubair, F.; Amjad, N.; Ashraf, S.; Asghar, S.; Awan, M.Z.; Javaid, S. Acetylcholinesterase inhibitory potential of scorpion venom in Aedes aegypti (Diptera: Culicidae). Braz. J. Biol. 2022, 84, e259506. [Google Scholar] [CrossRef]
  260. Bosmans, F.; Tytgat, J. Voltage-gated sodium channel modulation by scorpion α-toxins. Toxicon 2007, 49, 142–158. [Google Scholar] [CrossRef] [PubMed]
  261. Cestèle, S.; Catterall, W.A. Molecular mechanisms of neurotoxin action on voltage-gated sodium channels. Biochimie 2000, 82, 883–892. [Google Scholar] [CrossRef]
  262. Pedraza Escalona, M.; Possani, L.D. Scorpion β-toxins and voltage-gated sodium channels: Interactions and effects. Front. Biosci. 2013, 18, 572–587. [Google Scholar] [CrossRef] [PubMed]
  263. Zhorov, B.S.; Du, Y.; Song, W.; Luo, N.; Gordon, D.; Gurevitz, M.; Dong, K. Mapping the interaction surface of scorpion β-toxins with an insect sodium channel. Biochem. J. 2021, 478, 2843–2869. [Google Scholar] [CrossRef]
  264. Zlotkin, E.; Rochat, H.; Kopeyan, C.; Miranda, F.; Lissitzky, S. Purification and properties of the insect toxin from the venom of the scorpion Androctonus australis Hector. Biochimie 1971, 53, 1073–1078. [Google Scholar] [CrossRef]
  265. Zlotkin, E.; Fishman, Y.; Elazar, M. AaIT: From neurotoxin to insecticide. Biochimie 2000, 82, 869–881. [Google Scholar] [CrossRef]
  266. Darbon, H.; Weber, C.; Brawn, W. Two dimensional 1H nuclear magnetic resonance study of AaH IT, an anti-insect toxin from the scorpion Androctonus australis Hector. Sequential resonance assignments and folding of the polypeptide chain. Biochemistry 1991, 30, 1836–1844. [Google Scholar] [CrossRef]
  267. Ji, S.J.; Liu, F.; Li, E.Q.; Zhu, Y.X. Recombinant scorpion insectotoxin AaIT kills specifically insect cells but not human cells. Cell Res. 2002, 12, 143–150. [Google Scholar] [CrossRef]
  268. Valdez-Velázquez, L.L.; Romero-Gutierrez, M.T.; Delgado-Enciso, I.; Dobrovinskaya, O.; Melnikov, V.; Quintero-Hernández, V.; Ceballos-Magaña, S.G.; Gaitan-Hinojosa, M.A.; Coronas, F.I.; Puebla-Perez, A.M.; et al. Comprehensive analysis of venom from the scorpion Centruroides tecomanus reveals compounds with antimicrobial, cytotoxic, and insecticidal activities. Toxicon 2016, 118, 95–103. [Google Scholar] [CrossRef] [PubMed]
  269. Bermúdez-Guzmán, M.J.; Jiménez-Vargas, J.M.; Possani, L.D.; Zamudio, F.; Orozco-Gutiérrez, G.; Oceguera Contreras, E.; Enríquez-Vara, J.N.; Vazquez-Vuelvas, O.F.; García-Villalvazo, P.E.; Valdez-Velázquez, L.L. Biochemical characterization and insecticidal activity of isolated peptides from the venom of the scorpion Centruroides tecomanus. Toxicon 2022, 206, 90–102. [Google Scholar] [CrossRef]
  270. Zhu, S.; Gao, B.; Peigneur, S.; Tytgat, J. How a scorpion toxin selectively captures a prey sodium channel: The molecular and evolutionary basis uncovered. Mol. Biol. Evol. 2020, 37, 3149–3164. [Google Scholar] [CrossRef] [PubMed]
  271. Nakagawa, Y.; Lee, Y.M.; Lehmberg, E.; Herrmann, R.; Herrmann, R.; Moskowitz, H.; Jones, A.D.; Hammock, B.D. Anti-insect toxin 5 (Aalts) from Androctonus australis. Eur. J. Biochem. 1997, 246, 496–501. [Google Scholar] [CrossRef] [PubMed]
  272. Bel Haj Rhouma, R.; Cérutti-Duonor, M.; Benkhadir, K.; Goudey-Perrière, F.; El Ayeb, M.; Lopez-Ferber, M.; Karoui, H. Insecticidal effects of Buthus occitanus tunetanus BotIT6 toxin expressed in Escherichia coli and baculovirus/insect cells. J. Insect Physiol. 2005, 51, 1376–1383. [Google Scholar] [CrossRef]
  273. Liu, S.M.; Li, J.; Zhu, J.Q.; Wang, X.W.; Wang, C.S.; Liu, S.S.; Chen, X.X.; Li, S. Transgenic plants expressing the AaIT/GNA fusion protein show increased resistance and toxicity to both chewing and sucking pests. Insect Sci. 2016, 23, 265–276. [Google Scholar] [CrossRef]
  274. Coelho, V.A.; Cremonez, C.M.; Anjolette, F.A.; Aguiar, J.F.; Varanda, W.A.; Arantes, E.C. Functional and structural study comparing the C-terminal amidated β-neurotoxin Ts1 with its isoform Ts1-G isolated from Tityus serrulatus venom. Toxicon 2014, 83, 15–21. [Google Scholar] [CrossRef]
  275. Peigneur, S.; Cologna, C.T.; Cremonez, C.M.; Mille, B.G.; Pucca, M.B.; Cuypers, E.; Arantes, E.C.; Tytgat, J. A gamut of undiscovered electrophysiological effects produced by Tityus serrulatus toxin 1 on NaV-type isoforms. Neuropharmacology 2015, 95, 269–277. [Google Scholar] [CrossRef]
  276. Tibery, D.V.; Campos, L.A.; Mourão, C.B.F.; Peigneur, S.; Cruz e Carvalho, A.; Tytgat, J.; Schwartz, E.F. Electrophysiological characterization of Tityus obscurus β toxin 1 (To1) on Na+-channel isoforms. Biochim. Biophys. Acta Biomembr. 2019, 1861, 142–150. [Google Scholar] [CrossRef]
  277. Pimenta, A.M.; Martin-Eauclaire, M.; Rochat, H.; Figueiredo, S.G.; Kalapothakis, E.; Afonso, L.C.; De Lima, M.E. Purification, amino-acid sequence and partial characterization of two toxins with anti-insect activity from the venom of the South American scorpion Tityus bahiensis (Buthidae). Toxicon 2001, 39, 1009–1019. [Google Scholar] [CrossRef]
  278. Salazar, M.H.; Arenas, I.; Corrales-García, L.L.; Miranda, R.; Vélez, S.; Sánchez, J.; Mendoza, K.; Cleghorn, J.; Zamudio, F.Z.; Castillo, A.; et al. Venoms of Centruroides and Tityus species from Panama and their main toxic fractions. Toxicon 2018, 141, 79–87. [Google Scholar] [CrossRef] [PubMed]
  279. Salazar, M.H.; Ortíz, M.H.; Encarnación, S.; Zamudio, F.; Possani, L.D.; Cleghorn, J.; Morán, M.; Acosta, H.; Corzo, G. A proteomic overview of the major venom components from Tityus championi from Panama. Toxicon 2023, 227, 107082. [Google Scholar] [CrossRef]
  280. Rincón-Cortés, C.A.; Olamendi-Portugal, T.; Carcamo-Noriega, E.N.; Santillán, E.G.; Zuñiga, F.Z.; Reyes-Montaño, E.A.; Castro, N.A.V.; Possani, L.D. Structural and functional characterization of toxic peptides purified from the venom of the Colombian scorpion Tityus macrochirus. Toxicon 2019, 169, 5–11. [Google Scholar] [CrossRef]
  281. Díaz, P.; D’Suze, G.; Salazar, V.; Sevcik, C.; Shannon, J.D.; Sherman, N.E.; Fox, J.W. Antibacterial activity of six novel peptides from Tityus discrepans scorpion venom. A fluorescent probe study of microbial membrane Na+ permeability changes. Toxicon 2009, 54, 802–817. [Google Scholar] [CrossRef]
  282. Diego-García, E.; Batista, C.V.; García-Gómez, B.I.; Lucas, S.; Candido, D.M.; Gómez-Lagunas, F.; Possani, L.D. The Brazilian scorpion Tityus costatus Karsch: Genes, peptides and function. Toxicon 2005, 45, 273–283. [Google Scholar] [CrossRef]
  283. Estrada-Gómez, S.; Vargas-Muñoz, L.J.; Saldarriaga-Córdoba, M.M.; van der Meijden, A. MS/MS analysis of four scorpion venoms from Colombia: A descriptive approach. J. Venom. Anim. Toxins Incl. Trop. Dis. 2021, 27, e20200173. [Google Scholar] [CrossRef] [PubMed]
  284. Guerrero-Vargas, J.A.; Mourão, C.B.F.; Quintero-Hernández, V.; Possani, L.D.; Schwartz, E.F. Identification and phylogenetic analysis of Tityus pachyurus and Tityus obscurus novel putative Na+-channel scorpion toxins. PLoS ONE 2012, 7, e30478. [Google Scholar] [CrossRef] [PubMed]
  285. Windley, M.J.; Herzig, V.; Dziemborowicz, S.A.; Hardy, M.C.; King, G.F.; Nicholson, G.M. Spider-venom peptides as bioinsecticides. Toxins 2012, 4, 191–227. [Google Scholar] [CrossRef] [PubMed]
  286. Smith, J.J.; Herzig, V.; King, G.F.; Alewood, P.F. The insecticidal potential of venom peptides. Cell. Mol. Life. Sci. 2013, 70, 3665–3693. [Google Scholar] [CrossRef]
  287. Saez, N.J.; Herzig, V. Versatile spider venom peptides and their medical and agricultural applications. Toxicon 2019, 158, 109–126. [Google Scholar] [CrossRef]
  288. Fassolo, E.M.; Guo, S.; Wang, Y.; Rosa, S.; Herzig, V. Genetically encoded libraries and spider venoms as emerging sources for crop protective peptides. J. Pept. Sci. 2024, 30, e3600. [Google Scholar] [CrossRef]
  289. Chen, L.; Wang, Y.; Zhang, K.; Wu, S. Functional diversity of sodium channel variants in common eastern bumblebee, Bombus impatiens. Arch. Insect Biochem. Physiol. 2023, 114, e22052. [Google Scholar] [CrossRef]
  290. Endersby-Harshman, N.M.; Schmidt, T.L.; Hoffmann, A.A. Diversity and distribution of sodium channel mutations in Aedes albopictus (Diptera: Culicidae). J. Med. Entomol. 2024, 61, 630–643. [Google Scholar] [CrossRef] [PubMed]
  291. Alzabib, A.A.; Al-Sarar, A.S.; Abobakr, Y.; Saleh, A.A. Investigating the molecular mechanisms of deltamethrin resistance in Musca domestica populations from Saudi Arabia. Parasites Vectors 2025, 18, 55. [Google Scholar] [CrossRef]
  292. Mashlawi, A.M.; Al-Nazawi, A.M.; Noureldin, E.M.; Alqahtani, H.; Mahyoub, J.A.; Saingamsook, J.; Debboun, M.; Kaddumukasa, M.; Al-Mekhlafi, H.M.; Walton, C. Molecular analysis of knockdown resistance (kdr) mutations in the voltage-gated sodium channel gene of Aedes aegypti populations from Saudi Arabia. Parasites Vectors 2022, 15, 375. [Google Scholar] [CrossRef]
  293. Nascimento, G.J.; Cosme, L.V.; Torres, A.L.Q.; Lima, J.B.P.; Martins, A.J. High voltage-gated sodium channel gene diversity in Aedes albopictus across Brazil. Sci. Rep. 2025, 15, 20832. [Google Scholar] [CrossRef]
  294. Zhang, Y.; Zhang, C.; Wu, L.; Luo, C.; Guo, X.; Yang, R.; Zhang, Y. Population genetic structure and evolutionary genetics of Anopheles sinensis based on knockdown resistance (kdr) mutations and mtDNA-COII gene in China-Laos, Thailand-Laos, and Cambodia-Laos borders. Parasites Vectors 2022, 15, 229. [Google Scholar] [CrossRef]
  295. Zang, C.; Wang, X.; Cheng, P.; Liu, L.; Guo, X.; Wang, H.; Lou, Z.; Lei, J.; Wang, W.; Wang, Y.; et al. Evaluation of the evolutionary genetics and population structure of Culex pipiens pallens in Shandong province, China based on knockdown resistance (kdr) mutations and the mtDNA-COI gene. BMC Genom. 2023, 24, 145. [Google Scholar] [CrossRef]
  296. Seydi-Gazafi, K.; Seidi, S.; Ahmadabad, A.E.; Belkahia, H.; Tavassoli, M.; Said, M.B. Phylogenetic analysis and pyrethroid resistance mutation profiling of Haematopinus tuberculatus in buffaloes from Naqadeh, Iran. Parasitol. Int. 2025, 109, 103101. [Google Scholar] [CrossRef] [PubMed]
  297. Yuan, L.; Zhang, K.; Wang, Z.; Xian, L.; Liu, K.; Wu, S. Functional diversity of voltage-gated sodium channel in Drosophila suzukii (Matsumura). Pest Manag. Sci. 2024, 80, 592–601. [Google Scholar] [CrossRef] [PubMed]
  298. Kerkis, I.; Prieto da Silva, A.R.B.; Pompeia, C.; Tytgat, J.; de Sá Junior, P.L. Toxin bioportides: Exploring toxin biological activity and multifunctionality. Cell. Mol. Life Sci. 2017, 74, 647–661. [Google Scholar] [CrossRef]
  299. Torres-Rêgo, M.; Gláucia-Silva, F.; Rocha Soares, K.S.; de Souza, L.B.F.C.; Damasceno, I.Z.; Santos-Silva, E.D.; Lacerda, A.F.; Chaves, G.M.; Silva-Júnior, A.A.D.; Fernandes-Pedrosa, M.F. Biodegradable cross-linked chitosan nanoparticles improve anti-Candida and anti-biofilm activity of TistH, a peptide identified in the venom gland of the Tityus stigmurus scorpion. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 103, 109830. [Google Scholar] [CrossRef]
  300. Rebbouh, F.; Martin-Eauclaire, M.F.; Laraba-Djebari, F. Chitosan nanoparticles as a delivery platform for neurotoxin II from Androctonus australis hector scorpion venom: Assessment of toxicity and immunogenicity. Acta Trop. 2020, 205, 105353. [Google Scholar] [CrossRef]
  301. Gláucia-Silva, F.; Torres, J.V.P.; Torres-Rêgo, M.; Daniele-Silva, A.; Furtado, A.A.; Ferreira, S.S.; Chaves, G.M.; Xavier-Júnior, F.H.; Rocha Soares, K.S.; Silva-Júnior, A.A.D.; et al. . Tityus stigmurus-venom-loaded cross-linked chitosan nanoparticles improve antimicrobial activity. Int. J. Mol. Sci. 2024, 25, 9893. [Google Scholar] [CrossRef]
  302. Herzig, V.; Ahabh, A.; Jones, A.; King, G.F. Shining a light on the photochemical stability of peptidic bioinsecticides. Toxicon 2025, 262, 108381. [Google Scholar] [CrossRef]
  303. Xue, Q.; Swevers, L.; Taning, C.N.T. Drosophila X virus-like particles as delivery carriers for improved oral insecticidal efficacy of scorpion Androctonus australis peptide against the invasive fruit fly, Drosophila suzukii. Insect Sci. 2024, 31, 847–858. [Google Scholar] [CrossRef] [PubMed]
  304. Bilgo, E.; Lovett, B.; Fang, W.; Bende, N.; King, G.F.; Diabate, A.; St Leger, R.J. Improved efficacy of an arthropod toxin expressing fungus against insecticide-resistant malaria-vector mosquitoes. Sci. Rep. 2017, 7, 3433. [Google Scholar] [CrossRef] [PubMed]
  305. Rajput, S.; Suroshe, S.S.; Yadav, P.R.; Kumar, A.; Saini, G.K. Bioefficacy of engineered Beauveria bassiana with scorpion neurotoxin, LqqIT1 against cotton mealybug, Phenacoccus solenopsis and cowpea aphid, Aphis craccivora. Peer J. 2023, 11, e16030. [Google Scholar] [CrossRef] [PubMed]
  306. Zlotkin, E.; Fishman, L.; Shapiro, J.P. Oral toxicity to flesh flies of a neurotoxic polypeptide. Arch. Insect Biochem. Physiol. 1992, 21, 41–52. [Google Scholar] [CrossRef]
  307. Wu, J.; Luo, X.; Wang, Z.; Tian, Y.; Liang, A.; Sun, Y. Transgenic cotton expressing synthesized scorpion insect toxin AaHIT gene confers enhanced resistance to cotton bollworm (Heliothis armigera) larvae. Biotechnol. Lett. 2008, 30, 547–554. [Google Scholar] [CrossRef]
  308. Wang, J.; Chen, Z.; Du, J.; Sun, Y.; Liang, A. Novel insect resistance in Brassica napus developed by transformation of chitinase and scorpion toxin genes. Plant Cell Rep. 2005, 24, 549–555. [Google Scholar] [CrossRef]
  309. Yao, B.; Fan, Y.; Zeng, Q.; Zhao, R. Insect-resistant tobacco plants expressing insect-specific neurotoxin AaIT. Chin. J. Biotechnol. 1996, 12, 67–72. [Google Scholar]
  310. Trung, N.P.; Fitches, E.; Gatehouse, J.A. A fusion protein containing a lepidopteran-specific toxin from the South Indian red scorpion (Mesobuthus tamulus) and snowdrop lectin shows oral toxicity to target insects. BMC Biotechnol. 2006, 6, 18. [Google Scholar] [CrossRef]
  311. Fitches, E.C.; Bell, H.A.; Powell, M.E.; Back, E.; Sargiotti, C.; Weaver, R.J.; Gatehouse, J.A. Insecticidal activity of scorpion toxin (ButaIT) and snowdrop lectin (GNA) containing fusion proteins towards pest species of different orders. Pest Manag. Sci. 2010, 6, 74–83. [Google Scholar] [CrossRef]
  312. Nezamiha, F.K.; Imani, S.; Mianroodi, R.A.; Tirgari, S.; Shahbazzadeh, D. OdTx12/GNA, a chimeric variant of a β excitatory toxin from Odontobuthus doriae, reveals oral toxicity towards Locusta migratoria and Tenebrio molitor. Toxicon 2021, 202, 13–19. [Google Scholar] [CrossRef]
  313. Li, H.; Tian, C.; Chen, J.; Xia, Y. The fusion protein of scorpion neurotoxin BjαIT and Galanthus nivalis agglutinin (GNA) enhanced the injection insecticidal activity against silkworms, but only has lethal activity against newly hatched larva when administered orally. World J. Microbiol. Biotechnol. 2024, 40, 326. [Google Scholar] [CrossRef]
  314. Li, H.; Tian, C.; Chen, J.; Xia, Y. Scorpion insect neurotoxin LqhIT2 is a promising oral biopesticide: High-level preparation in Pichia pastoris and bioactivity assays. Pest Manag. Sci. 2025, 81, 2266–2276. [Google Scholar] [CrossRef] [PubMed]
  315. Tobassum, S.; Tahir, H.M.; Arshad, M.; Zahid, M.T.; Ali, S.; Ahsan, M.M. Nature and applications of scorpion venom: An overview. Toxin Rev. 2020, 39, 214–225. [Google Scholar] [CrossRef]
  316. Koç, H.; Avşar, C.; Bayrakci, Y. Antimicrobial activity of hemolymph and venom obtained from some scorpion species. Hacet. J. Biol. Chem. 2020, 48, 349–354. [Google Scholar] [CrossRef]
  317. Du, Q.; Hou, X.; Wang, L.; Zhang, Y.; Xi, X.; Wang, H.; Zhou, M.; Duan, J.; Wei, M.; Chen, T.; et al. AaeAP1 and AaeAP2: Novel antimicrobial peptides from the venom of the scorpion, Androctonus aeneas: Structural characterisation, molecular cloning of biosynthetic precursor-encoding cDNAs and engineering of analogues with enhanced antimicrobial and anticancer activities. Toxins 2015, 7, 219–237. [Google Scholar] [CrossRef] [PubMed]
  318. Almaaytah, A.; Zhou, M.; Wang, L.; Chen, T.; Walker, B.; Shaw, C. Antimicrobial/cytolytic peptides from the venom of the North African scorpion, Androctonus amoreuxi: Biochemical and functional characterization of natural peptides and a single site-substituted analog. Peptides 2012, 35, 291–299. [Google Scholar] [CrossRef]
  319. Song, C.; Wen, R.; Zhou, J.; Zeng, X.; Kou, Z.; Zhang, J.; Wang, T.; Chang, P.; Lv, Y.; Wu, R. Antibacterial and antifungal properties of a novel antimicrobial Peptide GK-19 and its application in skin and soft tissue infections induced by MRSA or Candida albicans. Pharmaceutics 2022, 14, 1937. [Google Scholar] [CrossRef]
  320. Zerouti, K.; Khemili, D.; Laraba-Djebari, F.; Hammoudi-Triki, D. Nontoxic fraction of scorpion venom reduces bacterial growth and inflammatory response in a mouse model of infection. Toxin Rev. 2021, 40, 310–324. [Google Scholar] [CrossRef]
  321. Alajmi, R.; Al-ghamdi, S.; Barakat, I.; Mahmoud, A.; Abdon, N.; Al-Ahidib, M.; AbdelGaber, R. Antimicrobial activity of two novel venoms from Saudi Arabian scorpions (Leiurus quinquestriatus and Androctonus crassicauda). Int. J. Pept. Res. Therapeut. 2020, 26, 67–74. [Google Scholar] [CrossRef]
  322. Salama, W.; Geasa, N. Investigation of the antimicrobial and hemolytic activity of venom of some Egyptian scorpion. J. Microbiol. Antimicrob. 2014, 6, 21–28. [Google Scholar] [CrossRef]
  323. Al-Asmari, A.K.; Alamri, M.A.; Almasoudi, A.S.; Abbasmanthiri, R.; Mahfoud, M. Evaluation of the in vitro antimicrobial activity of selected Saudi scorpion venoms tested against multidrug-resistant micro-organisms. J. Glob. Antimicrob. Resist. 2017, 10, 14–18. [Google Scholar] [CrossRef] [PubMed]
  324. Jahangirian, E.; Zargan, J. Investigating the antibacterial effects of 3 novel peptides isolated from the venom of Iranian Odontobuthus doriae and Buthotus saulcyi scorpions on Escherichia coli (UTI89) and Enterococcus faecalis causing urinary tract infection. Int. J. Pept. Res. Ther. 2023, 29, 55. [Google Scholar] [CrossRef]
  325. Conde, R.; Zamudio, F.Z.; Rodríguez, M.H.; Possani, L.D. Scorpine, an anti-malaria and anti-bacterial agent purified from scorpion venom. FEBS Lett. 2000, 471, 165–168. [Google Scholar] [CrossRef] [PubMed]
  326. Tawfik, M.M.; Bertelsen, M.; Abdel-Rahman, M.A.; Strong, P.N.; Miller, K. Scorpion venom antimicrobial peptides induce siderophore biosynthesis and oxidative stress responses in Escherichia coli. mSphere 2021, 6, e00267-21. [Google Scholar] [CrossRef]
  327. Harrison, P.L.; Abdel-Rahman, M.A.; Strong, P.N.; Tawfik, M.M.; Miller, K. Characterisation of three α-helical antimicrobial peptides from the venom of Scorpio maurus palmatus. Toxicon 2016, 117, 30–36. [Google Scholar] [CrossRef]
  328. Gagandeep, K.R.; Narasingappa, R.B.; Vyas, G.V. Unveiling mechanisms of antimicrobial peptide: Actions beyond the membranes disruption. Heliyon 2024, 10, e38079. [Google Scholar] [CrossRef]
  329. Meng, L.; Zhao, Y.; Qu, D.; Xie, Z.; Guo, X.; Zhu, Z.; Chen, Z.; Zhang, L.; Li, W.; Cao, Z.; et al. Ion channel modulation by scorpion hemolymph and its defensin ingredients highlights origin of neurotoxins in telson formed in Paleozoic scorpions. Int. J. Biol. Macromol. 2020, 148, 351–363. [Google Scholar] [CrossRef]
  330. Di, Z.; Qiao, S.; Liu, X.; Xiao, S.; Lei, C.; Li, Y.; Li, S.; Zhang, F. Transcriptome sequencing and comparison of venom glands revealed intraspecific differentiation and expression characteristics of toxin and defensin genes in Mesobuthus martensii populations. Toxins 2022, 14, 630. [Google Scholar] [CrossRef] [PubMed]
  331. Jlassi, A.; Mekni-Toujani, M.; Ferchichi, A.; Gharsallah, C.; Malosse, C.; Chamot-Rooke, J.; ElAyeb, M.; Ghram, A.; Srairi-Abid, N.; Daoud, S. BotCl, the first chlorotoxin-like peptide inhibiting Newcastle disease virus: The emergence of a new scorpion venom AMPs family. Molecules 2023, 28, 4355. [Google Scholar] [CrossRef]
  332. Mahnam, K.; Lotfi, M.; Shapoorabadi, F.A. Examining the interactions scorpion venom peptides (HP1090, Meucin-13, and Meucin-18) with the receptor binding domain of the coronavirus spike protein to design a mutated therapeutic peptide. J. Mol. Graph. Model. 2021, 107, 107952. [Google Scholar] [CrossRef]
  333. Soorki, M.N. In silico antiviral effect assessment of some venom gland peptides from Odontobuthus doriae scorpion against SARS-CoV-2. Toxicon 2025, 255, 108229. [Google Scholar] [CrossRef]
  334. El-Bitar, A.M.H.; Sarhan, M.M.H.; Aoki, C.; Takahara, Y.; Komoto, M.; Deng, L.; Moustafa, M.A.; Hotta, H. Virocidal activity of Egyptian scorpion venoms against hepatitis C virus. Virol. J. 2015, 12, 47. [Google Scholar] [CrossRef]
  335. El-Bitar, A.M.H.; Sarhan, M.; Abdel-Rahman, M.A.; Quintero-Hernandez, V.; AokiUtsubo, C.; Moustafa, M.A.; Possani, L.D.; Hotta, H. Smp76, a scorpine-like peptide isolated from the venom of the scorpion Scorpio maurus palmatus, with a potent antiviral activity against hepatitis C virus and dengue virus. Int. J. Pept. Res. Therapeut. 2020, 26, 811–821. [Google Scholar] [CrossRef]
  336. Keikha, M.; Kamali, H.; Ghazvini, K.; Karbalaei, M. Antimicrobial peptides: Natural or synthetic defense peptides against HBV and HCV infections. Virusdisease 2022, 33, 445–455. [Google Scholar] [CrossRef] [PubMed]
  337. Agbadamashi, D.J.; Price, C.L. Novel strategies for preventing fungal infections—Outline. Pathogens 2025, 14, 126. [Google Scholar] [CrossRef] [PubMed]
  338. Guilhelmelli, F.; Vilela, N.; Smidt, K.S.; de Oliveira, M.A.; Álvares, A.C.M.; Rigonatto, M.C.; Costa, P.H.S.; Tavares, A.H.; de Freitas, S.M.; Nicola, A.M.; et al. Activity of scorpion venom-derived antifungal peptides against planktonic cells of Candida spp. and Cryptococcus neoformans and Candida albicans biofilms. Front. Microbiol. 2016, 7, 1844. [Google Scholar] [CrossRef]
  339. Park, J.; Kim, H.; Kang, D.D.; Park, Y. Exploring the therapeutic potential of scorpion-derived Css54 peptide against Candida albicans. J. Microbiol. 2024, 62, 101–112. [Google Scholar] [CrossRef] [PubMed]
  340. Ehret-Sabatier, L.; Loew, D.; Goyffon, M.; Fehlbaum, P.; Hoffmann, J.A.; van Dorsselaer, A.; Bulet, P. Characterization of novel cysteine-rich antimicrobial peptides from scorpion blood. J. Biol. Chem. 1996, 271, 29537–29544. [Google Scholar] [CrossRef]
  341. Rostamkolaie, L.K.; Hamidinejat, H.; Jalali, M.H.R.; Varzi, H.N.; Abadshapouri, M.R.S.; Jafari, H. Inhibitory effect of Hemiscorpius lepturus scorpion venom fractions on tachyzoites of Toxoplasma gondii. Iran. J. Parasitol. 2022, 17, 79–89. [Google Scholar] [CrossRef] [PubMed]
  342. Alshammari, A.; Anwar, F.A.S.; Mohamed, S.A.-A.; Abdelsater, N. Antihelimentic effect of Androctonus crassicauda scorpion venom against Trichuris arvicolae isolated from Psammomys obesus in Egypt. Saudi J. Biol. Sci. 2023, 30, 103713. [Google Scholar] [CrossRef]
  343. Al-Malki, E.S.; Abdelsater, N. In vitro scolicidal effects of Androctonus crassicauda (Olivier, 1807) venom against the protoscolices of Echinococcus granulosus. Saudi J. Biol. Sci. 2020, 27, 1760–1765. [Google Scholar] [CrossRef] [PubMed]
  344. Gao, B.; Xu, J.; Rodriguez, M.C.; Lanz-Mendoza, H.; Hernandez-Rivas, R.; Du, W.; Zhu, S. Characterization of two linear cationic antimalarial peptides in the scorpion Mesobuthus eupeus. Biochimie 2010, 92, 350–359. [Google Scholar] [CrossRef]
  345. Jafari, H.; Nemati, M.; Molayan, P.H.; Rostamkolaie, L.K.; Nejat, H. Scolicidal activity of Mesobuthus eupeus venom against the protoscolices of Echinococcus granulosus. Arch. Razi Inst. 2019, 74, 183–189. [Google Scholar] [CrossRef]
  346. Casella-Martins, A.; Ayres, L.R.; Burin, S.M.; Morais, F.R.; Pereira, J.C.; Faccioli, L.H.; Sampaio, S.V.; Arantes, E.C.; Castro, F.A.; Pereira-Crott, L.S. Immunomodulatory activity of Tityus serrulatus scorpion venom on human T lymphocytes. J. Venom. Anim. Toxins Incl. Trop. Dis. 2015, 21, 46. [Google Scholar] [CrossRef]
  347. Chiang, E.Y.; Li, T.; Jeet, S.; Peng, I.; Zhang, J.; Lee, W.P.; DeVoss, J.; Caplazi, P.; Chen, J.; Warming, S.; et al. Potassium channels Kv1.3 and KCa3.1 cooperatively and compensatorily regulate antigen-specific memory T cell functions. Nat. Commun. 2017, 8, 14644. [Google Scholar] [CrossRef]
  348. Varga, Z.; Gurrola-Briones, G.; Papp, F.; Rodríguez de la Veja, R.C.; Pedraza-Alva, G.; Tajhya, R.B.; Gaspar, R.; Cardenas, L.; Rosenstein, Y.; Beeton, C.; et al. Vm24, a natural immunosuppressant peptide potently and selectively blocks Kv1.3 potassium channels of human T cells. Mol. Pharmacol. 2012, 82, 372–382. [Google Scholar] [CrossRef]
  349. Balozet, L. Scorpionism in the Old World. In Venomous Animals and Their Venoms; Bücherl, W., Buckley, E., Eds.; Academic Press: New York, NY, USA, 1971; Volume 3, pp. 349–371. [Google Scholar]
  350. Bouafir, Y.; Ait-Lounis, A.; Laraba-Djebari, F. Improvement of function and survival of pancreatic beta-cells in streptozotocin-induced diabetic model by the scorpion venom fraction F1. Toxin Rev. 2017, 36, 101–108. [Google Scholar] [CrossRef]
  351. Salama, W.M.; El-Naggar, S.A.; Tabl, G.A.E.; El Shefiey, L.M.M.; El-Desouki, N.I. Treatment with Leiurus quinquestratus scorpion venom ameliorates the histopathological changes of type-2 diabetic rats’ splenic tissues. J. Biosci. Appl. Res. 2023, 9, 356–365. [Google Scholar] [CrossRef]
  352. Salama, W.M.; El-Naggar, S.A.; Tabl, G.A.; El-Desouki, N.I.; El Shefiey, L.M. Leiurus quinquestratus venom promotes beta islets regeneration and restores glucose level in streptozotocin induced type 2 diabetes mellitus in rats. Sci. Rep. 2025, 15, 11841. [Google Scholar] [CrossRef] [PubMed]
  353. Roy, A.; Bharadvaja, N. Venom-derived bioactive compounds as potential anticancer agents: A review. Int. J. Pept. Res. Ther. 2021, 27, 129–147. [Google Scholar] [CrossRef]
  354. Naeem, A.; Hu, P.; Yang, M.; Zhang, J.; Liu, Y.; Zhu, W.; Zheng, Q. Natural products as anticancer agents: Current status and future perspectives. Molecules 2022, 27, 8367. [Google Scholar] [CrossRef]
  355. Lafnoune, A.; Chakir, S.; Darkaoui, B.; Cadi, R.; Oukkache, N. Moroccan Naja haje venom and its peptides: In vivo toxicity and in vitro antiproliferative effect on hepatocellular carcinoma HepG2 cells. Int. J. Pept. Res. Ther. 2024, 30, 62. [Google Scholar] [CrossRef]
  356. Lafnoune, A.; Chbel, A.; Darkaoui, B.; Irahal, I.N.; Oukkache, N. Cobra venom: From envenomation syndromes to therapeutic innovations. Int. J. Pept. Res. Ther. 2024, 30, 71. [Google Scholar] [CrossRef]
  357. Mlayah-Bellalouna, S.; Aissaoui-Zid, D.; Chantome, A.; Jebali, J.; Souid, S.; Ayedi, E.; Mejdoub, H.; Belghazi, M.; Marrakchi, N.; Essafi-Benkhadir, K.; et al. Insights into the mechanisms governing P01 scorpion toxin effect against U87 glioblastoma cells oncogenesis. Front. Pharmacol. 2023, 14, 1203247. [Google Scholar] [CrossRef]
  358. Aissaoui-Zid, D.; Saada, M.C.; Moslah, W.; Potier-Cartereau, M.; Lemettre, A.; Othman, H.; Gaysinski, M.; Abdelkafi-Koubaa, Z.; Souid, S.; Marrakchi, N.; et al. AaTs-1: A tetrapeptide from Androctonus australis scorpion venom, inhibiting U87 glioblastoma cells proliferation by p53 and FPRL-1 up-regulations. Molecules 2021, 26, 7610. [Google Scholar] [CrossRef]
  359. Moslah, W.; Aissaoui-Zid, D.; Aboudou, S.; Abdelkafi-Koubaa, Z.; Potier-Cartereau, M.; Lemettre, A.; ELBini-Dhouib, I.; Marrakchi, N.; Gigmes, D.; Vandier, C.; et al. Strengthening anti-glioblastoma effect by multi-branched dendrimers design of a scorpion venom tetrapeptide. Molecules 2022, 27, 806. [Google Scholar] [CrossRef]
  360. Béchohra, L.; Laraba-Djebari, F.; Hammoudi-Triki, D. Cytotoxic activity of Androctonus australis hector venom and its toxic fractions on human lung cancer cell line. J. Venom. Anim. Toxins Incl. Trop. Dis. 2016, 22, 29. [Google Scholar] [CrossRef]
  361. Issaad, N.; Ait-Lounis, A.; Laraba-Djebari, F. Cytotoxicity and actin cytoskeleton damage induced in human alveolar epithelial cells by Androctonus australis hector venom. Toxin Rev. 2018, 37, 67–74. [Google Scholar] [CrossRef]
  362. Caliskan, F.; Ergene, E.; Sogut, I.; Hatipoglu, I.; Basalp, A.; Sivas, H.; Kanbak, G. Biological assays on the effects of Acra3 peptide from Turkish scorpion Androctonus crassicauda venom on a mouse brain tumor cell line (BC3H1) and production of specific monoclonal antibodies. Toxicon 2013, 76, 350–361. [Google Scholar] [CrossRef]
  363. Al-Asmari, A.K.; Riyasdeen, A.; Islam, M. Scorpion venom causes upregulation of p53 and downregulation of Bcl-xL and BID protein expression by modulating signaling proteins Erk1/2 and STAT3, and DNA damage in breast and colorectal cancer cell lines. Integr. Cancer Ther. 2018, 17, 271–281. [Google Scholar] [CrossRef]
  364. Al-Asmari, A.K.; Riyasdeen, A.; Islam, M. Scorpion venom causes apoptosis by increasing reactive oxygen species and cell cycle arrest in MDA-MB-231 and HCT-8 cancer cell lines. J. Evid. Based Integr. Med. 2018, 23, 2156587217751796. [Google Scholar] [CrossRef]
  365. Lafnoune, A.; Lee, S.-Y.; Heo, J.-Y.; Daoudi, K.; Darkaoui, B.; Chakir, S.; Cadi, R.; Mounaji, K.; Shum, D.; Seo, H.-R.; et al. Anti-cancer activity of Buthus occitanus venom on hepatocellular carcinoma in 3D cell culture. Molecules 2022, 27, 2219. [Google Scholar] [CrossRef]
  366. Khamessi, O.; Ben Mabrouk, H.; ElFessi-Magouri, R.; Kharrat, R. RK1, the first very short peptide from Buthus occitanus tunetanus inhibits tumor cell migration, proliferation and angiogenesis. Biochem. Biophys. Res. Commun. 2018, 499, 1–7. [Google Scholar] [CrossRef] [PubMed]
  367. Kayhan, N.Y.; Ocal, I.Ç.; Akdeniz, M. Investigation the cytotoxic and antiproliferative effects of crude venom of Euscorpius mingrelicus (Scorpiones: Euscorpiidae) scorpion. Eur. J. Biol. 2022, 81, 117–124. [Google Scholar] [CrossRef]
  368. Rezaei, A.; Asgari, S.; Komijani, S.; Sadat, S.N.; Sabatier, J.M.; Nasrabadi, D.; Pooshang Bagheri, K.; Shahbazzadeh, D.; Akbari Eidgahi, M.R.; De Waard, M.; et al. Discovery of leptulipin, a new anticancer protein from the Iranian scorpion, Hemiscorpius lepturus. Molecules 2022, 27, 2056. [Google Scholar] [CrossRef] [PubMed]
  369. Nosouhian, M.; Rastegari, A.A.; Shahanipour, K.; Ahadi, A.M.; Sajjadieh, M.S. Anticancer potentiality of Hottentotta saulcyi scorpion curd venom against breast cancer: An in vitro and in vivo study. Sci. Rep. 2024, 14, 24607. [Google Scholar] [CrossRef] [PubMed]
  370. Dezianian, S.; Zargan, J.; Goudarzi, H.R.; Noormohamadi, A.H.; Mousavi, M.; Alikhani, H.K.; Johari, B. In vitro study of Hottentotta schach crude venom anticancer effects on MCF-7 and Vero cell lines. Iran. J. Pharm. Res. 2020, 19, 192–202. [Google Scholar] [CrossRef]
  371. Zidan, S.A.; Ghany, S.A.; Abdel-Hakeem, M.A. Leiurus quinquestriatus venom gold nanoparticle conjugates: A novel approach for modulating ion channel gene expression in liver cancer. Egypt. J. Basic Appl. Sci. 2025, 12, 1–17. [Google Scholar] [CrossRef]
  372. Cohen-Inbar, O.; Zaaroor, M. Glioblastoma multiforme targeted therapy: The chlorotoxin story. J. Clin. Neurosci. 2016, 33, 52–58. [Google Scholar] [CrossRef]
  373. Ahmed, O.A.A.; Badr-Eldin, S.M.; Caruso, G.; Fahmy, U.A.; Alharbi, W.S.; Almehmady, A.M.; Alghamdi, S.A.; Alhakamy, N.A.; Mohamed, A.I.; Aldawsari, H.M.; et al. Colon targeted Eudragit coated beads loaded with optimized fluvastatin-scorpion venom conjugate as a potential approach for colon cancer therapy: In vitro anticancer activity and in vivo colon imaging. J. Pharmaceut. Sci. 2022, 111, 3304–3317. [Google Scholar] [CrossRef]
  374. Asfour, H.Z.; Fahmy, U.A.; Alharbi, W.S.; Almehmady, A.M.; Alamoudi, A.J.; Tima, S.; Mansouri, R.A.; Omar, U.M.; Ahmed, O.A.A.; Zakai, S.A.; et al. Phyto-phospholipid conjugated scorpion venom nanovesicles as promising carrier that improves efficacy of thymoquinone against adenocarcinoma human alveolar basal epithelial cells. Pharmaceutics 2021, 13, 2144. [Google Scholar] [CrossRef]
  375. Gandomkari, M.S.; Ayat, H.; Ahadi, A.M. Recombinantly expressed MeICT, a new toxin from Mesobuthus eupeus scorpion, inhibits glioma cell proliferation and downregulates Annexin A2 and FOXM1 genes. Biotechnol. Lett. 2022, 44, 703–712. [Google Scholar] [CrossRef]
  376. Alikhani, H.K.; Bidmeshkipour, A.; Zargan, J. Cytotoxic and apoptotic induction effects of the venom of Iranian scorpion (Odontobuthus bidentatus) in the hepatocellular carcinoma cell line (HepG2). Int. J. Pept. Res. Therapeut. 2020, 26, 2475–2484. [Google Scholar] [CrossRef]
  377. Alikhani, H.K.; Zargan, J.; Bidmeshkipour, A.; Zamani, E.; Mosavi, M.; Heidari, A.; Hajinoormohammadi, A. Iranian scorpion (Odontobuthus bidentatus) crude venom change the redox potential of MCF-7 breast cancer cell line and induce apoptosis. Mod. Med. Lab. J. 2021, 4, 28–35. [Google Scholar] [CrossRef]
  378. Zargan, J.; Sajad, M.; Umar, S.; Naime, M.; Ali, S.; Khan, H.A. Scorpion (Odontobuthus doriae) venom induces apoptosis and inhibits DNA synthesis in human neuroblastoma cells. Mol. Cell. Biochem. 2011, 348, 173–181. [Google Scholar] [CrossRef] [PubMed]
  379. Zargan, J.; Umar, S.; Sajad, M.; Naime, M.; Ali, S.; Khan, H.A. Scorpion venom (Odontobuthus doriae) induces apoptosis by depolarization of mitochondria and reduces S-phase population in human breast cancer cells (MCF-7). Toxicol. In Vitro 2011, 25, 1748–1756. [Google Scholar] [CrossRef]
  380. Dogan, T.; Igci, N.; Biber, A.; Gerekci, S.; Husnugil, H.; Izbirak, A.; Ozen, C. Peptidomic characterization and bioactivity of Protoiurus kraepelini (Scorpiones: Iuridae) venom. Turk. J. Biol. 2018, 42, 490–497. [Google Scholar] [CrossRef]
  381. Guo, R.; Liu, J.; Chai, J.; Gao, Y.; Abdel-Rahman, M.A.; Xu, X. Scorpion peptide Smp24 exhibits a potent antitumor effect on human lung cancer cells by damaging the membrane and cytoskeleton in vivo and in vitro. Toxins 2022, 14, 438. [Google Scholar] [CrossRef]
  382. Nguyen, T.; Guo, R.; Chai, J.; Wu, J.; Liu, J.; Chen, X.; Abdel-Rahman, M.A.; Xia, H.; Xu, X. Smp24, a scorpion-venom peptide, exhibits potent antitumor effects against hepatoma HepG2 cells via multi-mechanisms in vivo and in vitro. Toxins 2022, 14, 717. [Google Scholar] [CrossRef]
  383. Deng, Z.; Gao, Y.; Nguyen, T.; Chai, J.; Wu, J.; Li, J.; Abdel-Rahman, M.A.; Xu, X.; Chen, X. The potent antitumor activity of Smp43 against non-small-cell lung cancer A549 cells via inducing membranolysis and mitochondrial dysfunction. Toxins 2023, 15, 347. [Google Scholar] [CrossRef]
  384. Chai, J.; Yang, W.; Gao, Y.; Guo, R.; Peng, Q.; Abdel-Rahman, M.A.; Xu, X. Antitumor effects of scorpion peptide Smp43 through mitochondrial dysfunction and membrane disruption on hepatocellular carcinoma. J. Nat. Prod. 2021, 84, 3147–3160. [Google Scholar] [CrossRef]
  385. Bucaretchi, F.; Fernandes, L.C.; Fernandes, C.B.; Branco, M.M.; Prado, C.C.; Vieira, R.J.; De Capitani, E.M.; Hyslop, S. Clinical consequences of Tityus bahiensis and Tityus serrulatus scorpion stings in the region of Campinas, southeastern Brazil. Toxicon 2014, 89, 17–25. [Google Scholar] [CrossRef] [PubMed]
  386. Rodrigo, C.; Gnanathasan, A. Management of scorpion envenoming: A systematic review and meta-analysis of controlled clinical trials. Syst. Rev. 2017, 6, 74. [Google Scholar] [CrossRef]
  387. Amr, Z.S.; Abu Baker, M.A.; Al-Saraireh, M.; Warrell, D.A. Scorpions and scorpion sting envenoming (scorpionism) in the Arab countries of the Middle East. Toxicon 2021, 191, 83–103. [Google Scholar] [CrossRef] [PubMed]
  388. Vaucel, J.A.; Larréché, S.; Paradis, C.; Courtois, A.; Pujo, J.M.; Elenga, N.; Résière, D.; Caré, W.; de Haro, L.; Gallart, J.C.; et al. French scorpionism (mainland and oversea territories): Narrative review of scorpion species, scorpion venom, and envenoming management. Toxins 2022, 14, 719. [Google Scholar] [CrossRef] [PubMed]
  389. Jami, S.; Erickson, A.; Brierley, S.M.; Vetter, I. Pain-causing venom peptides: Insights into sensory neuron pharmacology. Toxins 2017, 10, 15. [Google Scholar] [CrossRef]
  390. Diochot, S. Pain-related toxins in scorpion and spider venoms: A face to face with ion channels. J. Venom. Anim. Toxins Incl. Trop. Dis. 2021, 27, e20210026. [Google Scholar] [CrossRef]
  391. Geron, M.; Hazan, A.; Priel, A. Animal toxins providing insights into TRPV1 activation mechanism. Toxins 2017, 9, 326. [Google Scholar] [CrossRef]
  392. Ferraz, C.R.; Manchope, M.F.; Bertozzi, M.M.; Saraiva-Santos, T.; Andrade, K.C.; Franciosi, A.; Zaninelli, T.H.; Bagatim-Souza, J.; Borghi, S.M.; Cândido, D.M.; et al. Tityus serrulatus scorpion venom-induced nociceptive responses depend on TRPV1, immune cells, and pro-inflammatory cytokines. Toxins 2025, 17, 332. [Google Scholar] [CrossRef]
  393. Haddad, L.; Chender, A.; Roufayel, R.; Accary, C.; Borges, A.; Sabatier, J.M.; Fajloun, Z.; Karam, M. Effect of Hottentotta judaicus scorpion venom on nociceptive response and inflammatory cytokines in mice using experimental hyperalgesia. Molecules 2025, 30, 2750. [Google Scholar] [CrossRef] [PubMed]
  394. Xu, Y.; Sun, J.; Liu, H.; Sun, J.; Yu, Y.; Su, Y.; Cui, Y.; Zhao, M.; Zhang, J. Scorpion toxins targeting voltage-gated sodium channels associated with pain. Curr. Pharm. Biotechnol. 2018, 19, 848–855. [Google Scholar] [CrossRef] [PubMed]
  395. Pereira, A.F.M.; Cavalcante, J.S.; Angstmam, D.G.; Almeida, C.; Soares, G.S.; Pucca, M.B.; Ferreira Junior, R.S. Unveiling the pain relief potential: Harnessing analgesic peptides from animal venoms. Pharmaceutics 2023, 15, 2766. [Google Scholar] [CrossRef]
  396. Bagheri-Ziari, S.; Bagheri, K.P. From discovery to the future medical applications of venom-derived analgesic peptides for the treatment of peripheral pains. Curr. Pharm. Des. 2025; in press. [Google Scholar] [CrossRef]
  397. Wang, D.; Herzig, V.; Dekan, Z.; Rosengren, K.J.; Payne, C.D.; Hasan, M.M.; Zhuang, J.; Bourinet, E.; Ragnarsson, L.; Alewood, P.F.; et al. Novel scorpion toxin ω-buthitoxin-Hf1a selectively inhibits calcium influx via CaV3.3 and CaV3.2 and alleviates allodynia in a mouse model of acute postsurgical pain. Int. J. Mol. Sci. 2024, 25, 4745. [Google Scholar] [CrossRef] [PubMed]
  398. Hoang, A.N.; Vo, H.D.; Vo, N.P.; Kudryashova, K.S.; Nekrasova, O.V.; Feofanov, A.V.; Kirpichnikov, M.P.; Andreeva, T.V.; Serebryakova, M.V.; Tsetlin, V.I.; et al. Vietnamese Heterometrus laoticus scorpion venom: Evidence for analgesic and anti-inflammatory activity and isolation of new polypeptide toxin acting on Kv1.3 potassium channel. Toxicon 2014, 77, 40–48. [Google Scholar] [CrossRef]
  399. Martin-Eauclaire, M.-F.; Abbas, H.; Sauze, N.; Mercier, L.; Berge-Lefranc, J.-L.; Condo, J.; Bougis, P.E.; Guieu, R. Involvement of endogenous opioid system in scorpion toxin-induced antinociception in mice. Neurosci Lett. 2010, 482, 45–50. [Google Scholar] [CrossRef]
  400. Zhang, Y.; Xu, J.; Wang, Z.; Zhang, X.; Liang, X.; Civelli, O. BmK-YA, an enkephalin-like peptide in scorpion venom. PLoS ONE 2012, 7, e40417. [Google Scholar] [CrossRef]
  401. Aliakbari, F.; Rahmati, S.; Ghanbari, A.; Madanchi, H.; Rashidy-Pour, A. Identification and designing an analgesic opioid cyclic peptide from Defensin 4 of Mesobuthus martensii Karsch scorpion venom with more effectiveness than morphine. Biomed. Pharmacother. 2025, 188, 118139. [Google Scholar] [CrossRef]
  402. Li, Z.; Hu, P.; Wu, W.; Wang, Y. Peptides with therapeutic potential in the venom of the scorpion Buthus martensii Karsch. Peptides 2019, 115, 43–50. [Google Scholar] [CrossRef]
  403. Zheng, Y.; Wen, Q.; Huang, Y.; Guo, D. The significant therapeutic effects of Chinese scorpion: Modern scientific exploration of ion channels. Pharmaceuticals 2024, 17, 1735. [Google Scholar] [CrossRef]
  404. Shoukry, N.M.; Salem, M.L.; Teleb, W.K.; Abdel Daim, M.M.; Abdel-Rahman, M.A. Antinociceptive, antiinflammatory, and antipyretic effects induced by the venom of Egyptian scorpion Androctonus amoreuxi. J. Basic Appl. Zool. 2020, 81, 56. [Google Scholar] [CrossRef]
  405. Abbas, N.; Gaudioso-Tyzra, C.; Bonnet, C.; Gabriac, M.; Amsalem, M.; Lonigro, A.; Padilla, F.; Crest, M.; Martin-Eauclaire, M.F.; Delmas, P. The scorpion toxin Amm VIII induces pain hypersensitivity through gain-of-function of TTX-sensitive Na+ channels. Pain 2013, 154, 1204–1215. [Google Scholar] [CrossRef]
  406. Maatoug, R.; Jebali, J.; Guieu, R.; De Waard, M.; Kharrat, R. BotAF, a new Buthus occitanus tunetanus scorpion toxin, produces potent analgesia in rodents. Toxicon 2018, 149, 72–85. [Google Scholar] [CrossRef] [PubMed]
  407. Bagheri-Ziari, S.; Shahbazzadeh, D.; Sardari, S.; Sabatier, J.-M.; Pooshang Bagheri, K. Discovery of a new analgesic peptide, leptucin, from the Iranian scorpion, Hemiscorpius lepturus. Molecules 2021, 26, 2580. [Google Scholar] [CrossRef] [PubMed]
  408. Han, T.; Ming, H.; Deng, L.; Zhu, H.; Liu, Z.; Zhang, J.; Song, Y. A novel expression vector for the improved solubility of recombinant scorpion venom in Escherichia coli. Biochem. Biophys. Res. Commun. 2017, 482, 120–125. [Google Scholar] [CrossRef]
  409. Song, Y.; Liu, Z.; Zhang, Q.; Li, C.; Jin, W.; Liu, L.; Zhang, J.; Zhang, J. Investigation of binding modes and functional surface of scorpion toxins ANEP to sodium channels 1.7. Toxins 2017, 9, 387. [Google Scholar] [CrossRef]
  410. Mao, Q.; Ruan, J.; Cai, X.; Lu, W.; Ye, J.; Yang, J.; Yang, Y.; Sun, X.; Cao, J.; Cao, P. Antinociceptive effects of analgesic-antitumor peptide (AGAP), a neurotoxin from the scorpion Buthus martensii Karsch, on formalin-induced inflammatory pain through a mitogen-activated protein kinases-dependent mechanism in mice. PLoS ONE 2013, 8, e78239. [Google Scholar] [CrossRef] [PubMed]
  411. Ruan, J.-P.; Mao, Q.-H.; Lu, W.-G.; Cai, X.-T.; Chen, J.; Li, Q.; Fu, Q.; Yan, H.-J.; Cao, J.L.; Cao, P. Inhibition of spinal MAPKs by scorpion venom peptide BmK AGAP produces a sensory-specific analgesic effect. Mol. Pain 2018, 14, 1–11. [Google Scholar] [CrossRef]
  412. Richard, S.A.; Kampo, S.; Sackey, M.; Hechavarria, M.E.; Buunaaim, A.D.B. The pivotal potentials of scorpion Buthus martensii Karsch-analgesic-antitumor peptide in pain management and cancer. Evid. Based Complement. Alternat. Med. 2020, 2020, 4234273. [Google Scholar] [CrossRef]
  413. Kampo, S.; Cui, Y.; Yu, J.; Anabah, T.W.; Falagán, A.A.; Bayor, M.T.; Wen, Q.P. Scorpion venom peptide, AGAP inhibits TRPV1 and potentiates the analgesic effect of lidocaine. Heliyon 2021, 7, e08560. [Google Scholar] [CrossRef] [PubMed]
  414. Chen, J.; Feng, X.H.; Shi, J.; Tan, Z.Y.; Bai, Z.T.; Liu, T.; Ji, Y.H. The anti-nociceptive effect of BmK AS, a scorpion active polypeptide, and the possible mechanism on specifically modulating voltage-gated Na+ currents in primary afferent neurons. Peptides 2006, 27, 2182–2192. [Google Scholar] [CrossRef]
  415. Cao, Z.Y.; Mi, Z.M.; Cheng, G.F.; Shen, W.Q.; Xiao, X.; Liu, X.M.; Liang, X.T.; Yu, D.Q. Purification and characterization of a new peptide with analgesic effect from the scorpion Buthus martensi Karch. J. Pept. Res. 2004, 64, 33–41. [Google Scholar] [CrossRef]
  416. Liu, Y.; Li, Y.; Zhu, Y.; Zhang, L.; Ji, J.; Gui, M.; Li, C.; Song, Y. Study of anti-inflammatory and analgesic activity of scorpion toxins DKK-SP1/2 from scorpion Buthus martensii Karsch (BmK). Toxins 2021, 13, 498–518. [Google Scholar] [CrossRef] [PubMed]
  417. Lin, S.; Wang, X.; Hu, X.; Zhao, Y.; Zhao, M.; Zhang, J.; Cui, Y. Recombinant expression, functional characterization of two scorpion venom toxins with three disulfide bridges from the Chinese scorpion Buthus martensii Karsch. Protein Pept. Lett. 2017, 24, 235–240. [Google Scholar] [CrossRef] [PubMed]
  418. Bai, F.; Song, Y.; Cao, Y.; Ban, M.; Zhang, Z.; Sun, Y.; Feng, Y.; Li, C. Scorpion neurotoxin Syb-prII-1 exerts analgesic effect through Nav1.8 channel and MAPKs pathway. Int. J. Mol. Sci. 2022, 23, 7065–7088. [Google Scholar] [CrossRef]
  419. Rigo, F.K.; Bochi, G.V.; Pereira, A.L.; Adamante, G.; Ferro, P.R.; Dal-Toé De Prá, S.; Milioli, A.M.; Damiani, A.P.; da Silveira Prestes, G.; Dalenogare, D.P.; et al. TsNTxP, a non-toxic protein from Tityus serrulatus scorpion venom, induces antinociceptive effects by suppressing glutamate release in mice. Eur. J. Pharm. 2019, 855, 65–74. [Google Scholar] [CrossRef]
  420. Ma, R.; Cui, Y.; Zhou, Y.; Bao, Y.M.; Yang, W.Y.; Liu, Y.F.; Wu, C.F.; Zhang, J.H. Location of the analgesic domain in scorpion toxin BmK AGAP by mutagenesis of disulfide bridges. Biochem. Biophys. Res. Commun. 2010, 394, 330–334. [Google Scholar] [CrossRef]
  421. Montero-Dominguez, P.A.; Restano-Cassulini, R.; Magaña-Ávila, L.C.; Almanza, A.; Mercado, F.; Corzo, G. Design of antinociceptive peptide by grafting domains between scorpion β-neurotoxins. Bioorg. Chem. 2025, 162, 108592. [Google Scholar] [CrossRef]
  422. Lima, L.S.; Loyola, V.; Bicca, J.V.M.L.; Faro, L.; Vale, C.L.C.; Denucci, B.L.; Mortari, M.R. Innovative treatments for epilepsy: Venom peptides, cannabinoids, and neurostimulation. J. Neurosci. Res. 2022, 100, 1969–1986. [Google Scholar] [CrossRef]
  423. Veiseh, M.; Gabikian, P.; Bahrami, S.-B.; Veiseh, O.; Zhang, M.; Hackman, R.C.; Ravanpay, A.C.; Stroud, M.R.; Kusuma, Y.; Hansen, S.J.; et al. Tumor paint: A chlorotoxin:Cy5.5 bioconjugate for intraoperative visualization of cancer foci. Cancer Res. 2007, 67, 6882–6888. [Google Scholar] [CrossRef]
  424. Beeton, C.; Wulff, H.; Standifer, N.E.; Azam, P.; Mullen, K.M.; Pennington, M.W.; Kolski-Andreaco, A.; Wei, E.; Grino, A.; Counts, D.R.; et al. Kv1.3 channels are a therapeutic target for T cell-mediated autoimmune diseases. Proc. Natl. Acad. Sci. USA 2005, 102, 9860–9865. [Google Scholar] [CrossRef]
  425. Guan, R.J.; Wang, M.; Wang, D.; Wang, D.C. A new insect neurotoxin AngP1 with analgesic effect from the scorpion Buthus martensii Karsch: Purification and characterization. J. Pept. Res. 2001, 58, 27–35. [Google Scholar] [CrossRef]
  426. Sui, A.R.; Piao, H.; Xiong, S.T.; Zhang, P.; Guo, S.Y.; Kong, Y.; Gao, C.Q.; Wang, Z.X.; Yang, J.; Ge, B.Y.; et al. Scorpion venom heat-resistant synthesized peptide ameliorates epileptic seizures and imparts neuroprotection in rats mediated by NMDA receptors. Eur. J. Pharmacol. 2024, 978, 176704. [Google Scholar] [CrossRef] [PubMed]
  427. Du, J.; Wang, R.; Yin, L.; Fu, Y.; Cai, Y.; Zhang, Z.; Liang, A. BmK CT enhances the sensitivity of temozolomide-induced apoptosis of malignant glioma U251 cells in vitro through blocking the AKT signaling pathway. Oncol. Lett. 2018, 15, 1537–1544. [Google Scholar] [CrossRef] [PubMed]
  428. Zhao, F.; Zou, X.; Li, S.; He, J.; Xi, C.; Tang, Q.; Wang, Y.; Cao, Z. BmK NSPK, a potent potassium channel inhibitor from scorpion Buthus martensii Karsch, promotes neurite outgrowth via NGF/TrkA signaling pathway. Toxins 2021, 13, 33. [Google Scholar] [CrossRef] [PubMed]
  429. Ismail, M. The scorpion envenoming syndrome. Toxicon 1995, 33, 825–858. [Google Scholar] [CrossRef]
Figure 1. Representative scorpion species from around the world. (A). Androctonus crassicauda (Arabian fat-tailed scorpion; North Africa and Middle East). (B). Centruroides sculpturatus (Arizona bark scorpion; North America). (C). Hottentotta tamulus (Indian red scorpion; Indian subcontinent). (D). Leiurus quinquestriatus (Deathstalker scorpion; North Africa, Middle East and western Asia). (E). Parabuthus transvaalicus (South African thick-tailed scorpion; Southeastern Africa). (F). Tityus serrulatus (Brazilian yellow scorpion; South America). Photographs from Wikimedia Commons.
Figure 1. Representative scorpion species from around the world. (A). Androctonus crassicauda (Arabian fat-tailed scorpion; North Africa and Middle East). (B). Centruroides sculpturatus (Arizona bark scorpion; North America). (C). Hottentotta tamulus (Indian red scorpion; Indian subcontinent). (D). Leiurus quinquestriatus (Deathstalker scorpion; North Africa, Middle East and western Asia). (E). Parabuthus transvaalicus (South African thick-tailed scorpion; Southeastern Africa). (F). Tityus serrulatus (Brazilian yellow scorpion; South America). Photographs from Wikimedia Commons.
Toxins 17 00497 g001
Figure 2. A general classification of scorpion venom components consisting of proteins (including several enzymes; green pathway), peptides with and without disulfide bonds (blue pathway), and low molecular mass compounds such as amines and amino acids (orange pathway). Of these components, DBP, consisting primarily of ion channel toxins, have been the most extensively studied, followed by NDBP and, in recent years, proteins (enzymes). The wide range of biological activities associated with NDBP reflects the marked structural diversity of these peptides. For specific details on the various groups of toxins, see the following Sections and corresponding tables (where applicable): Section 3.3 for DBP, Section 3.4 and Section 3.5 for NDBP, and Section 3.6 for proteins. For DBP subclasses, see Section 3.3.1 for CSα/β scaffold toxins (Na+ and K+ channel toxins are summarized in the Sections on Long-Chain Scorpion Toxins (Protein Family PF14866) and Short-Chain Scorpion Toxins (Protein Family PF00451), respectively), Section 3.3.2 for Ca2+ channel toxins and Section 3.3.3 for Cl channel toxins.
Figure 2. A general classification of scorpion venom components consisting of proteins (including several enzymes; green pathway), peptides with and without disulfide bonds (blue pathway), and low molecular mass compounds such as amines and amino acids (orange pathway). Of these components, DBP, consisting primarily of ion channel toxins, have been the most extensively studied, followed by NDBP and, in recent years, proteins (enzymes). The wide range of biological activities associated with NDBP reflects the marked structural diversity of these peptides. For specific details on the various groups of toxins, see the following Sections and corresponding tables (where applicable): Section 3.3 for DBP, Section 3.4 and Section 3.5 for NDBP, and Section 3.6 for proteins. For DBP subclasses, see Section 3.3.1 for CSα/β scaffold toxins (Na+ and K+ channel toxins are summarized in the Sections on Long-Chain Scorpion Toxins (Protein Family PF14866) and Short-Chain Scorpion Toxins (Protein Family PF00451), respectively), Section 3.3.2 for Ca2+ channel toxins and Section 3.3.3 for Cl channel toxins.
Toxins 17 00497 g002
Figure 3. Diversity of chloride (Cl) channels and the sites of action (extracellular, intracellular) of scorpion venoms and toxins that interact with these channels. See Table 2 for details of these interactions and toxin channel specificities. Lqh and Lqq venoms and GaTX1 have no effect on CFTR when applied to the extracellular surface. Note that the original studies by Debin and colleagues [143,144] only showed that ClTx blocked ClC from the cytoplasmic side, without identifying the specific channel type. Pure ClTx was subsequently shown not to block ClC-2 channels [145]. Note also that although ClTx has been shown to block Ca2+-regulated Cl channels [146], another study reported no such blockade [147]. CFTR—cystic fibrosis transmembrane conductance regulator, ClTx—chlorotoxin, GABA—γ-aminobutyric acid, Lqh—L. q. hebraeus, Lqq—L. q. quinquestriatus, P—phosphorylation.
Figure 3. Diversity of chloride (Cl) channels and the sites of action (extracellular, intracellular) of scorpion venoms and toxins that interact with these channels. See Table 2 for details of these interactions and toxin channel specificities. Lqh and Lqq venoms and GaTX1 have no effect on CFTR when applied to the extracellular surface. Note that the original studies by Debin and colleagues [143,144] only showed that ClTx blocked ClC from the cytoplasmic side, without identifying the specific channel type. Pure ClTx was subsequently shown not to block ClC-2 channels [145]. Note also that although ClTx has been shown to block Ca2+-regulated Cl channels [146], another study reported no such blockade [147]. CFTR—cystic fibrosis transmembrane conductance regulator, ClTx—chlorotoxin, GABA—γ-aminobutyric acid, Lqh—L. q. hebraeus, Lqq—L. q. quinquestriatus, P—phosphorylation.
Toxins 17 00497 g003
Figure 4. The neuromuscular and neuronal activity of Bothriurus bonariensis scorpion venom (BBV) in Nauphoeta cinerea cockroaches. (A) Bothriurus bonariensis. (B) Neuromuscular paralysis (seen as a decrease in muscle twitch-tension) caused by BBV in cockroach neuromuscular preparations. (C) Representative recording of cockroach neuromuscular twitches showing the blockade by BBV. (D) Extracellular recordings of spontaneous neural compound action potentials (SNCAP) from cockroach sensilla showing a decrease in the frequency of potentials after incubation with BBV (recording 2) compared with the saline control (recording 1). The points in (B) represent the mean ± SEM (n = 6). * p < 0.05 compared to saline controls. Modified from dos Santos et al. [253].
Figure 4. The neuromuscular and neuronal activity of Bothriurus bonariensis scorpion venom (BBV) in Nauphoeta cinerea cockroaches. (A) Bothriurus bonariensis. (B) Neuromuscular paralysis (seen as a decrease in muscle twitch-tension) caused by BBV in cockroach neuromuscular preparations. (C) Representative recording of cockroach neuromuscular twitches showing the blockade by BBV. (D) Extracellular recordings of spontaneous neural compound action potentials (SNCAP) from cockroach sensilla showing a decrease in the frequency of potentials after incubation with BBV (recording 2) compared with the saline control (recording 1). The points in (B) represent the mean ± SEM (n = 6). * p < 0.05 compared to saline controls. Modified from dos Santos et al. [253].
Toxins 17 00497 g004
Figure 5. Primary actions of scorpion venoms and toxins in insects. Venom or toxin inoculation leads to a variety of effects, as indicated on the right, including manifestations of neurotoxicity (indicated in red) and cytotoxicity (indicated in blue). Note that the paralysis resulting from neurotoxicity may be either spastic paralysis caused by highly insect-selective excitatory β-NaTx such as AaHIT (Androctonus australis hector), Bj-xtrIT (Hottenttota judaicus), Lqh-xtrIT (L. hebraeus) and LqqIT1 (L. quinquestriatus), or flaccid paralysis caused by insect-selective depressant β-NaTx such as BaIT2 (Buthacus arenicola), BjIT2 (H. judaicus), BotIT2 (B. tunetanus) and LqhIT2 (L. hebraeus) [9]. Additional sites of action and effects (indicated in black) have been documented to varying degrees in vertebrates but much less studied in insects.
Figure 5. Primary actions of scorpion venoms and toxins in insects. Venom or toxin inoculation leads to a variety of effects, as indicated on the right, including manifestations of neurotoxicity (indicated in red) and cytotoxicity (indicated in blue). Note that the paralysis resulting from neurotoxicity may be either spastic paralysis caused by highly insect-selective excitatory β-NaTx such as AaHIT (Androctonus australis hector), Bj-xtrIT (Hottenttota judaicus), Lqh-xtrIT (L. hebraeus) and LqqIT1 (L. quinquestriatus), or flaccid paralysis caused by insect-selective depressant β-NaTx such as BaIT2 (Buthacus arenicola), BjIT2 (H. judaicus), BotIT2 (B. tunetanus) and LqhIT2 (L. hebraeus) [9]. Additional sites of action and effects (indicated in black) have been documented to varying degrees in vertebrates but much less studied in insects.
Toxins 17 00497 g005
Figure 6. The therapeutic arsenal and potential of scorpion venom. This figure summarizes the multifaceted pharmacological potential of scorpion venoms, as discussed in the text.
Figure 6. The therapeutic arsenal and potential of scorpion venom. This figure summarizes the multifaceted pharmacological potential of scorpion venoms, as discussed in the text.
Toxins 17 00497 g006
Figure 7. Structures of three B. martensii Karsch toxins. (A) BmK M1 (also known as BmK 1), a long-chain cardiotoxic Na+ channel blocker and one of the most abundant and best characterized long-chain toxins of this venom, showing the disulfide bridges (indicated in yellow and by arrows) at Cys12–63, 16–36, 22–46 and 26–48. (B) BmK αTx11, an α-toxin (Na+ channel blocker) homolog, showing the disulfide bridges at Cys12–63, 16–36, 22–46 and 22–48. (C) BmBKTtx1, a short-chain Ca2+-activated K+ channel (BKCa) blocker, showing the disulfide bridges at Cys3–22, 8–27 and 12–29. The structural models shown here are based on experimentally determined sequences. References [66,84,402,403] provide detailed discussion of B. martensii Karsch toxins, including those indicated here.
Figure 7. Structures of three B. martensii Karsch toxins. (A) BmK M1 (also known as BmK 1), a long-chain cardiotoxic Na+ channel blocker and one of the most abundant and best characterized long-chain toxins of this venom, showing the disulfide bridges (indicated in yellow and by arrows) at Cys12–63, 16–36, 22–46 and 26–48. (B) BmK αTx11, an α-toxin (Na+ channel blocker) homolog, showing the disulfide bridges at Cys12–63, 16–36, 22–46 and 22–48. (C) BmBKTtx1, a short-chain Ca2+-activated K+ channel (BKCa) blocker, showing the disulfide bridges at Cys3–22, 8–27 and 12–29. The structural models shown here are based on experimentally determined sequences. References [66,84,402,403] provide detailed discussion of B. martensii Karsch toxins, including those indicated here.
Toxins 17 00497 g007
Table 3. Antibacterial activities of some scorpion venoms and toxins.
Table 3. Antibacterial activities of some scorpion venoms and toxins.
SpeciesOriginMolecule(s)NatureActivityReferences
Aegaeobuthus
gibbosus
TurkeyVenomVenomVenom was active against Bacillus sphaericus, Bacillus subtilis, Bacillus cereus and Staphylococcus aureus (all Gram-positive) and Escherichia coli (Gram-negative), but not against Enterococcus faecalis and Micrococcus luteus (both Gram-positive) [316]
Androctonus
aeneas
North AfricaAaeAP1,
AaeAP2
Antimicrobial peptidesBoth peptides inhibited S. aureus (MIC: 16 μg/mL), but were much less effective against E. coli (MIC: >512 μg/mL).[317]
Androctonus
amoreuxi
North AfricaAamAP1, AamAP2Antimicrobial peptidesAamP1 and AamAP2 inhibited S. aureus (MIC: 20 and 48 μM, respectively) more effectively than E. coli (MIC: 120 and 150 μM, respectively).[318]
GK-19 (AamAP1
derivative)
Antimicrobial peptideBroad-spectrum antibacterial activity against E. coli, E. faecalis, K. pneumoniae, P. aeruginosa and S. aureus (MIC: 3–10 μM). Disrupted bacterial membranes.[319]
Androctonus
australis
AlgeriaVenom and peptide G-TISodium channel inhibitorVenom was active against B. cereus, E. coli, Microcossus spp. and S. aureus (MIC: 75, 150, 125, and 125 μg/mL, respectively). Disrupted bacterial membranes, leading to cell death. The peptide G-TI was active against B. cereus. [320]
Androctonus
crassicauda
Saudi
Arabia
VenomVenomInhibited the growth of E. coli, Salmonella spp., S. aureus, and Paenibacillus larvae.[321]
TurkeyVenomVenomVenom was active against Bacillus cereus, Bacillus sphaericus, Bacillus subtilis, Escherichia coli, and Micrococcus luteus, but not against Enterococcus faecalis and Staphylococcus aureus.[316]
Hottentotta
saulcyi
TurkeyVenomVenomVenom was active against Bacillus cereus, Bacillus sphaericus, Bacillus subtilis, Enterococcus faecalis and Staphylococcus aureus, but not against Escherichia coli or Micrococcus luteus.[316]
Leiurus
abdullahbayrami
TurkeyVenomVenomVenom was active against Bacillus sphaericus, but not against Bacillus cereus, Bacillus subtilis, Enterococcus faecalis, Escherichia coli, Micrococcus luteus or Staphylococcus aureus.[316]
Leiurus
quinquestriatus
EgyptVenomVenomInhibited B. subtilis and C. freundii; no effect on B. cereus and K. pneumoniae. [322]
Saudi
Arabia
VenomVenomInhibited the growth of E. coli, Salmonella spp., S. aureus, and Paenibacillus larvae. More effective than A. crassicauda venom [321]
Saudi
Arabia
VenomVenomConcentration-dependent inhibition of A. baumannii, E. coli, E. faecalis, K. pneumoniae, P. aeruginosa, and S. aureus.[323]
Odontobuthus doriaeIranPeptide 3Venom fractionInhibited Enterococcus faecalis and Escherichia coli UT189 (IC50: 160 and 80 μg/mL, respectively).[324]
Pandinus
imperator
North AfricaScorpineAntimicrobial peptideInhibited B. subtilis and K. pneumoniae (MIC: 1 and 10 μM, respectively).[325]
Protoiurus
kraepelini
TurkeyVenomVenomHigh bactericidal activity, with inhibition zones of 9–20 mm against Bacillus sphaericus, Bacillus cereus, Bacillus subtilis, Enterococcus faecalis, Escherichia coli, Micrococcus luteus and Staphylococcus aureus.[316]
Scorpio
maurus
palmatus
EgyptSmp24, Smp43Antimicrobial peptidesDisrupted bacterial membranes, with activity against Bacillus subtilis, Escherichia coli, Klebsiella pneumoniae, Pseudomonas aeruginosa, Staphylococcus aureus and Staphylococcus epidermidis. Smp43 also interfered with B. subtilis DNA synthesis. Caused pore formation and induced oxidative stress in E. coli (MIC: 4–128 μg/mL). Greater efficacy against Gram-positive bacteria.[326,327]
MIC: Minimum inhibitory concentration.
Table 4. Antiviral activities of some scorpion venoms and toxins.
Table 4. Antiviral activities of some scorpion venoms and toxins.
SpeciesOriginMolecule(s)NatureActivityReferences
Buthus occitanus tunetanusTunisiaBotClChlorotoxin-like peptideConcentration-dependent inhibition of Newcastle disease virus in avian species, with an IC50 of 0.69 μM. Mechanism involves direct disruption of viral particle structure, preventing cellular entry and proliferation.[331]
Mesobuthus eupeusIranMucin 13, Mucin 18Antimicrobial peptidesPotential antiviral activity against SARS-CoV-2 by targeting the receptor-binding domain (RBD) of the spike protein. Mucin-18 shows stronger binding affinity, with the A9T mutation enhancing its effectiveness.[332]
Odontobuthus doriaeIranODAMP2, ODAMP5Antimicrobial peptidesComputational studies suggest high binding affinities to the SARS-CoV-2 spike protein’s receptor binding domain, with binding energies of −59.2 and −51.8 kcal/mol for ODAMP2 and ODAMP5, respectively, indicating potential efficacy.[333]
Scorpio maurus
palmatus
EgyptVenomVenomSignificant activity against hepatitis C virus (HCV), interfering with viral entry. IC50 of 6.3 μg/mL, selectivity index of 15.8. Activity is independent of enzymatic processes and specific to Flaviviridae viruses such as HCV and Dengue virus (DENV).[334]
EgyptSmp76Scorpine-like peptidePotent antiviral activity against HCV and DENV, inhibiting early infection stages, likely Via direct viral particle interaction. IC50 of 0.01 μg/mL for both viruses. No cytotoxic or hemolytic effects at concentrations exceeding 1000 times the antiviral dose.[335]
Table 5. Antifungal activities of some scorpion venoms and toxins.
Table 5. Antifungal activities of some scorpion venoms and toxins.
SpeciesOriginMolecule(s)NatureActivityReferences
Androctonus
aeneas
North
Africa
AaeAP1, AaeAP2Antimicrobial peptidesInhibits Candida albicans (MIC: 32 μg/mL).[317]
Androctonus
amoreuxi
North
Africa
GK-19Derived from AamAP1Shows strong antifungal activity against Candida albicans, Candida glabrata and Candida krusei (MIC: 5–10 µM). Disrupts fungal membranes, causing structural damage.[319]
Androctonus
australis
North
Africa
AamAP1, AamP2Antimicrobial peptidesInhibits Candida albicans (MIC: 64 µM).[42]
North
Africa
AndroctoninCysteine-rich antimicrobial peptideExhibits potent antifungal activity against various fungal species, particularly Verticillium torelis and Fusarium oxysporum (MIC: <4 µM). Completely inhibits Neurospora crassa spore growth at ≥12 µM, without regrowth.[340]
Leiurus
quinquestriatus
Saudi
Arabia
VenomVenomReduces growth and survival of Candida albicans (by 31.2%) and Candida glabrata (by 39.0%).[323]
MIC: Minimum inhibitory concentration.
Table 6. Antiparasitic activities of some scorpion venoms, venom fractions and toxins.
Table 6. Antiparasitic activities of some scorpion venoms, venom fractions and toxins.
SpeciesOriginMolecule(s)NatureActivityReferences
Androctonus
crassicauda
EgyptVenomVenomAntihelminthic activity against Trichuris arvicolae that involved significant ultrastructural changes. Potential use in the treatment of gastrointestinal nematodes resistant to conventional drugs.[342]
Saudi ArabiaVenomVenomComplete destruction of Echinococcus granulosus protoscolices after 4 h of exposure at 100 μg/mL. The damage involved apoptosis and structural alterations. Potential use in the non-surgical treatment for hydatidosis, a significant public health problem.[343]
Hemiscorpius
lepturus
IranFraction 5Venom peptide fraction (<10 kDa)Reduced the viability of Toxoplasma gondii tachyzoites at 100 μg after 2 h.[341]
Mesobuthus eupeusIranMucin 24, Mucin 25Antimicrobial peptidesInhibited Plasmodium falciparum without harming human cells (erythrocytes and GC-2 cells), and prevented P. berghei ookinete growth at 10–20 μM, possibly via membrane disruption.[344]
Mesobuthus eupeusIranFraction 8Venom peptide fraction (<10 kDa)Scolicidal activity against Echinococcus granulosus protoscolices within 30 min of exposure to venom fraction.[345]
Pandinus
imperator
North
Africa
ScorpineAntimicrobial peptideActivity against Plasmodium berghei, involving disruption of sexual stage development in mosquito midguts. Inhibition of fertilization and ookinete formation with ED50 of 10 μM and 0.7 μM, respectively, an effect attributed to membrane disruption.[325]
Table 7. Antidiabetic activities of some scorpion venoms and venom fractions.
Table 7. Antidiabetic activities of some scorpion venoms and venom fractions.
SpeciesOriginMolecule(s)ActivityReferences
Androctonus australis
hector
AlgeriaVenom
Fraction F1
Treatment with Fraction F1 (10 mg/kg, i.p.; daily for 11 days) restored the body weight gain, attenuated the diabetes-induced hyperglycemia and improved the glucose tolerance in mice with streptozotocin-induced diabetes. Fraction F1 improved β-cell function and survival and increased the mitotic activity of these cells. Characterization of Fraction F1 revealed the presence of hyaluronidase and peptides, potentially contributing to glucose homeostasis and β-cell survival.[350]
Leiurus
quinquestriatus
EgyptVenomVenom (daily injection of 1/10 of the lethal dose for eight weeks) normalized the blood glucose levels and body weight in rats with streptozotocin-induced diabetes. The venom prevented the histopathological and immunohistochemical changes caused by diabetes in splenic tissues.[351]
EgyptVenomVenom (daily injection of 1/40 of the sublethal dose for eight weeks) prevented body weight loss, normalized hematological parameters, blood cell counts and blood glucose levels, reduced C-peptide levels to slightly below normal, normalized indicators of hepatic function and indicators of oxidative stress, and promoted β islets regeneration in rats with streptozotocin-induced diabetes.[352]
Table 8. Anticancer activities of some scorpion venoms and toxins.
Table 8. Anticancer activities of some scorpion venoms and toxins.
SpeciesOriginMolecule(s)NatureActivityReferences
Androctonus australisTunisiaP01K+ channel toxinInhibits proliferation, adhesion, and migration of U87 glioblastoma cells.[357]
TunisiaAaTs-1TetrapeptideInhibits glioblastoma U87 cell proliferation, modulates kinase expression, and enhances p53 and FPRL-1 expression.[358]
TunisiaAaTs-1-4BDendrimer multi-branched moleculesInhibits U87 cell proliferation and migration, enhances ERK1/2 and AKT phosphorylation, and increases p53 expression.[359]
Androctonus australis hectorAlgeriaF3 fractionVenom fractionInduces apoptosis in lung cancer cells (NCI-H358) via the mitochondrial pathway.[360]
AlgeriaVenomVenomAlters alveolar epithelial cell (A549) integrity, disrupts cytoskeleton, and reduces cell migration.[361]
Androctonus
crassicauda
TurkeyAcra3Na+ channel toxinCytotoxic to BC3H1 cells and induces necrosis.[362]
Androctonus
crassicauda,
Androctonus bicolor,
Leiurus
quinquestriatus
Saudi ArabiaVenomVenomInhibits cell proliferation, motility, and colony formation in colorectal and breast cancer cells. [363,364]
Buthus occitanusMoroccoα-Insect toxin Lqq3,
α-Like toxin Bom4
Na+ channel toxinInhibits hepatocellular carcinoma cell proliferation (Huh 7.5 cells in 3D culture).[365]
Buthus occitanus tunetanusTunisiaRK1Short peptide
(14 amino acids)
Inhibits cell proliferation, migration, and angiogenesis in U87 and IGR39 cells.[366]
Euscorpius
mingrelicus
TurkeyVenomVenomCytotoxicity towards breast and lung cancer cells by inducing apoptosis and necrosis.[367]
Hemiscorpius lepturusIranLeptulipinPhospholipase A2Inhibits proliferation, alters morphology, induces DNA fragmentation, and cell cycle arrest in HT-29 and MDA-MB-231 cells.[368]
Hottentotta saulcyiIranVenomVenomInduces apoptosis in MCF-7 cells, reduces tumor density in vivo, and upregulates pro-apoptotic genes.[369]
Hottentotta SchachIranVenomVenomAntiproliferative activity in MCF-7 cells by inducing oxidative stress leading to apoptosis.[370]
Leiurus
quinquestriatus
EgyptGNPs-VVenom conjugated with gold nanoparticlesAnticancer activity against liver cancer cell lines by inhibiting migration, inducing cell cycle arrest, and promoting apoptosis.[371]
IsraelChlorotoxinCl channel toxinTargets glioblastoma multiform cells, inhibits angiogenesis, and binds to specific channels.[372]
Saudi ArabiaFLV-SVFluvastatin-scorpion venom peptide nano-conjugateAntiproliferative activity in colorectal adenocarcinoma cells.[373]
Saudi ArabiaTHQ–PL–SVThymoquinone-phospholipon-scorpion venom peptide nanovesiclesAntiproliferative activity in lung cancer cells through modulation of gene expression.[374]
Mesobuthus eupeusIranmeuCl14ChlorotoxinInhibits hMMP-2 and limits tumor progression.[167]
IranMeICTCl channel toxinInhibits glioma cell proliferation and migration; downregulates Annexin A2 and FOXM1.[375]
Odontobuthus
bidentatus
IranVenomVenomInduces apoptosis in HepG2 cells via increased nitric oxide levels. Antiproliferative activity in MCF-7 cells.[376,377]
Odontobuthus doriaeIranVenomVenomApoptotic and antiproliferative activity in human neuroblastoma (SH-SY5Y) and human breast cancer (MCF-7) cells[378,379]
Protoiurus kraepeliniTurkeyVenomVenomConcentration-dependent cytotoxicity in Jurkat cells.[380]
Scorpio maurus
palmatus
EgyptSmp24Cationic antimicrobial peptideSuppresses lung cancer cell growth, and induces apoptosis, cell cycle arrest, and autophagy in HepG2 cells.[381,382]
EgyptSmp43Cationic antimicrobial peptideInhibits lung cancer cell proliferation, causes membrane rupture and mitochondrial dysfunction, and alters apoptosis, cell cycle, and autophagy. Inhibits hepatocellular carcinoma cells.[383,384]
Table 9. Analgesic activities of some scorpion venoms and toxins.
Table 9. Analgesic activities of some scorpion venoms and toxins.
SpeciesOriginMolecule(s)Toxin Class ActivityReferences
Androctonus
amoreuxi
EgyptVenom- Dose-dependent reduction in acetic acid-induced writhing (peripheral pain) and increased latency in the tail flick test (central pain).[404]
Androctonus
mauretanicusmauretanicus
Moroccoanatoxin Amm VIIINaTxDose-dependent analgesia in hot plate and tail flick tests. Antinociception mediated by blockade of Nav1.2, activation of endogenous opioid system, and activation of mechanisms involving diffuse noxious inhibitory controls (DNIC).[399,405]
Buthus
occitanus
tunetanus
TunisiaBotAFNaTxNon-toxic β-toxin-like peptide with no effect on motor activity. Reversible low blockade (≤20%) of Na+ channels. Peripheral or spinal mechanisms involved in attenuating the pain associated with the acetic acid writhing, formalin, hot plate, and tail-flick assays. Analgesic activity independent of opioid system. More potent than β-endorphin or morphine in these tests. Effective when given i.p., but not when given i.v. or i.c.v. Low activity on TTX-S Na+ channels of DRG and does not bind to rat brain synaptosomes. Stimulates lumbar spinal cord c-fos/c-jun mRNA up regulation. [406]
Hemiscorpius
lepturus
IranLeptucin NDAnalgesic effect against acute thermal pain (hot plate and tail-flick tests) at doses of 0.32 and 0.64 mg/kg, i.p., with similar or greater activity than morphine. No cytotoxicity or hemolysis. No histopathological alterations in heart, kidney or liver. LD50 (mice) > 4 mg/kg.[407]
Heterometrus
laoticus
VietnamVenomKTxVenom showed analgesic activity (9.5 and 19 mg/kg) in the tail immersion (55 °C water) and tail flick assays that was less potent than morphine (5 mg/kg). A short K+ channel toxin (Hetlaxin) that blocks Kv1.1 and Kv1.3, with high affinity for the latter (Ki = 59 nM), was isolated from the venom, but its analgesic activity in the assays indicated above was not tested.[398]
Leiurus q.
quinquestriatus
SudanLqqIT2NaTx Dose-dependent analgesia in hot plate and tail flick tests. Antinociception mediated by blockade of Nav1.2, activation of endogenous opioid system, and activation of mechanisms involving diffuse noxious inhibitory controls (DNIC).[398]
Mesobuthus
martensii
ChinaANEP (Anti-neuroexcitation peptide)NaTxβ-Toxin that blocks Nav1.7. Analgesic activity in the mouse acetic acid writhing test and hot plate test. Recombinant peptide has same activity as native peptide. Several mutants showed greater activity than the recombinant peptide.[408,409]
ChinaBmK-YA 8Short-chain NDBPStructurally related to enkephalin. Activates μ, κ and δ opioid receptors, with selectivity for δ opioid receptors ~7- and 12-fold greater than for µ and κ receptors, respectively. Full agonist at δ opioid receptors, and partial agonist with lower efficacy and potency on μ and κ receptors. Activity at δ opioid receptors antagonized by naloxone.[400]
BmK AGAP (Analgesic-antitumor peptide)NaTxBlockade of Nav1.4, Nav1.5, Nav1.7, Nav1.8, TRPV1, and KCNQ2/3 currents. Analgesic activity in the writhing, hot-plate and formalin tests. Attenuation of pain in the formalin-induced spontaneous nociceptive behavior assay involved inhibition of the expression of peripheral and spinal mitogen-activated protein kinases (MAPK), including p-p38, p-ERK, p-JNK and spinal Fos. Intrathecal administration of AGAP inhibited and reversed pain from chronic constrictive injury (CCI) of the sciatic nerve, from thermal hyperalgesia, and mechanical allodynia. Potentiated the analgesic effect of lidocaine.[410,411,412,413]
ChinaBmK ASNaTxBlockade of TTX-R (Nav1.8, 1.9) and TTX-S (Nav1.3), with no effect on voltage-dependent IK and KCl or caffeine-induced Ca2+ influx in neurons. Decreased the number of action potentials in DRG neurons by ~50% at 0.5 μM. Analgesic activity in formalin and carrageenan nociceptive assays.[414]
ChinaBmK AngM1NaTxLittle or no toxicity at doses up to 50 mg/kg, i.v. Irreversible (by washing) blockade of Nav and delayed rectifier K+ (IK) currents, but no effect on transient K+ currents (IA) in rat hippocampal pyramidal neurons. Analgesic effect in acetic acid writhing assay in mice (63% inhibition at 0.8 mg/kg, compared to 83% with morphine 0.2 mg/kg, i.v.).[415]
ChinaDKK-SP2NaTxBlocked Nav1.7 channels expressed in Chinese hamster ovary cells and reduced the expression of this channel in trigeminal neurons. Reduced writhing in the acetic acid test in mice and pain associated with chronic constriction injury of the infraorbital nerve in rats was also attenuated. At the highest doses tested, the potency in both tests was similar to or better than morphine.[416]
ChinaBmKBTx,
BmNaL-3SS2
NaTxBlockade of Nav1.7. BmKBTx and BmNal-3SS2 reducing writhing in the acetic acid test in mice and show similar potency to morphine (reductions of 43% and 63% at 1 mg/kg i.p., respectively, compared to 49% with morphine 1.5 mg/kg i.p.).[417]
ChinaSyb-prIINaTxβ-Neurotoxin that blocks Nav1.8, but not Nav1.9. Attenuated pain associated with chronic constriction injury of the infraorbital nerve, probably by attenuating MAPK-activated pathways. Efficacy similar to morphine.[418]
Tityus serrulatusBrazilTsNTxPNaTxNon-toxic peptide structurally related to toxins TsVII (Ts1 or toxin-γ) and Ts3 (TsIV or tityustoxin). Non-toxic to mammals. Antinociceptive effect in tail-flick thermal test and intraplantar capsaisin injection. Attenuates neuropathic pain caused by constriction injury to sciatic nerve and paclitaxel administration. Caused reduced glutamate release from mouse spinal cord synaptosomes.[419]
DRG—dorsal root ganglion, i.c.v.—intracerebroventricular, i.p.—intraperitoneal, i.v.—intravenous, KTx—potassium channel toxin, NaTx—sodium channel toxin, ND—not determined, NDBP—non-disulfide bridged peptide, TTX-S—tetrodotoxin-sensitive channel.
Table 10. Characteristics of some scorpion toxins with therapeutic potential in neurodegenerative diseases.
Table 10. Characteristics of some scorpion toxins with therapeutic potential in neurodegenerative diseases.
SpeciesToxinToxin CharacteristicsTherapeutic Potential
Androctonus
australis
AaTxOne of a family of peptides that primarily targets Na+ channelsShows immunomodulatory activity, including specific targeting of Kv1.3 channels involved in T cell-mediated autoimmune responses [424]. As oxidative stress is a major contributor to dopaminergic neuron loss in Parkinson´s disease and amyloid pathology in Alzheimer´s disease, AaTx could represent a potentially useful lead compound for developing therapeutic drugs for these neurodegenerative diseases.
B. martensii Karsch BmK AngP1A 62-amino acid peptide with high sequence homology to neurotoxins that target ion channels, but it exerts non-toxic and neurotrophic effectsPromotes angiogenesis and neurogenesis in hippocampal neurons through modulation of pathways related to VEGF expression and PI3K/Akt signaling, both relevant to neuronal survival. In rodent models, this peptide reduces oxidative stress, neuroinflammation, and amyloid-β accumulation, suggesting a protective role against neurodegeneration [425,426].
B. martensii Karsch BmK CTA small peptide (~36–40 amino acids) structurally related to chlorotoxin (chlorotoxin-like peptide) and capable of crossing the Blood–Brain Barrier (BBB).Inhibits matrix metalloproteinase 2 (MMP-2), an enzyme overexpressed in neuroinflammation and tumors. Exhibits anti-inflammatory properties in models of neuroinflammation by suppressing IL-6 and TNF-α [427]. As chronic neuroinflammation is driven by cytokines and MMPs contributes to neurodegeneration, BmK CT could mitigate these pathways.
B. martensii KarschBmK NSPK
(Neuroprotective scorpion peptide Karsch
A low-molecular-weight peptide the structure of which has yet to be solved.Shows CNS bioactivity without the neurotoxicity of classic neurotoxins. As part of its protective mechanism of action, BmK NSPK improves mitochondrial membrane potential and ATP production in primary cortical neurons, and inhibits Aβ-induced neuronal apoptosis by regulating Bcl-2/Bax ratios and caspase-3 activity [428].
B. martensii KarschBmK AGAP
(Analgesic-antitumor peptide)
A 66-amino acid peptide originally isolated for its analgesic and anti-tumor effects.Suppresses microglial activation and reduces secretion of pro-inflammatory cytokines (IL-1β, TNF-α, IL-6). Inhibits p38 MAPK and ERK1/2 signaling pathways, key regulators of inflammatory and apoptotic signaling in activated microglia [413].
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

Dal Belo, C.A.; Hyslop, S.; Carlini, C.R. Properties and Pharmacology of Scorpion Toxins and Their Biotechnological Potential in Agriculture and Medicine. Toxins 2025, 17, 497. https://doi.org/10.3390/toxins17100497

AMA Style

Dal Belo CA, Hyslop S, Carlini CR. Properties and Pharmacology of Scorpion Toxins and Their Biotechnological Potential in Agriculture and Medicine. Toxins. 2025; 17(10):497. https://doi.org/10.3390/toxins17100497

Chicago/Turabian Style

Dal Belo, Cháriston André, Stephen Hyslop, and Célia Regina Carlini. 2025. "Properties and Pharmacology of Scorpion Toxins and Their Biotechnological Potential in Agriculture and Medicine" Toxins 17, no. 10: 497. https://doi.org/10.3390/toxins17100497

APA Style

Dal Belo, C. A., Hyslop, S., & Carlini, C. R. (2025). Properties and Pharmacology of Scorpion Toxins and Their Biotechnological Potential in Agriculture and Medicine. Toxins, 17(10), 497. https://doi.org/10.3390/toxins17100497

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