Next Article in Journal
Inhibition of CYP1A1 Alleviates Colchicine-Induced Hepatotoxicity
Next Article in Special Issue
Voltage-Gated Sodium Channel Inhibition by µ-Conotoxins
Previous Article in Journal
Gaston Ramon’s Big Four
Previous Article in Special Issue
Conotoxin Prediction: New Features to Increase Prediction Accuracy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diversity and Evolutionary Analysis of Venom Insulin Derived from Cone Snails

by
Qiqi Guo
1,
Meiling Huang
1,
Ming Li
1,
Jiao Chen
1,
Shuanghuai Cheng
1,
Linlin Ma
2,* and
Bingmiao Gao
1,*
1
Key Laboratory of Tropical Translational Medicine of Ministry of Education, Hainan Key Laboratory for Research and Development of Tropical Herbs, School of Pharmacy, Hainan Medical University, Haikou 571199, China
2
Griffith Institute for Drug Discovery (GRIDD), School of Environment and Science, Griffith University, Nathan, Brisbane, QLD 4111, Australia
*
Authors to whom correspondence should be addressed.
Submission received: 24 October 2023 / Revised: 4 January 2024 / Accepted: 5 January 2024 / Published: 9 January 2024
(This article belongs to the Special Issue Conotoxins: Evolution, Classifications and Targets)

Abstract

:
Cone snails possess a diverse array of novel peptide toxins, which selectively target ion channels and receptors in the nervous and cardiovascular systems. These numerous novel peptide toxins are a valuable resource for future marine drug development. In this review, we compared and analyzed the sequence diversity, three-dimensional structural variations, and evolutionary aspects of venom insulin derived from different cone snail species. The comparative analysis reveals that there are significant variations in the sequences and three-dimensional structures of venom insulins from cone snails with different feeding habits. Notably, the venom insulin of some piscivorous cone snails exhibits a greater similarity to humans and zebrafish insulins. It is important to emphasize that these venom insulins play a crucial role in the predatory strategies of these cone snails. Furthermore, a phylogenetic tree was constructed to trace the lineage of venom insulin sequences, shedding light on the evolutionary interconnections among cone snails with diverse diets.
Key Contribution: Conoinsulin serves not only as a natural, inherent predatory weapon but also as a valuable resource in marine medicine research. By exploring the homology and evolutionary lineage of diverse conoinsulin variants, along with examining the resemblance between conoinsulin and human insulin, we provide a theoretical framework to comprehend the diversity of venom insulin used in the predation strategies of cone snails with different feeding habits. This enhanced understanding paves the way for further scientific exploration in this domain.

1. Introduction

Cone snails thrive in various tropical and subtropical aquatic ecosystems, exhibiting a distribution that spans the coastal areas surrounding the South China Sea, Africa, the Red Sea, India, Ceylon, Japan, the Paracel Islands, and even extends as far north as the Croatian coast [1]. The number of identified species has increased significantly, from 500 species in 2004 to more than 900 species today, and research on Conus is still developing rapidly [2,3]. Cone snails have distinctive shell patterns and they can be classified. According to their dietary preferences by which they can be categorized into different groups, such as piscivorous (feeding on fish), vermivorous (feeding on worms), and molluscivorous (feeding on mollusks), thereby highlighting the various feeding habits within this snail species [2,4,5]. Figure 1 showcases several representative species of Conus with different feeding habits, including piscivorous cone snails such as Conus geographus, Conus tulipa, Conus kinoshitai, and Conus striatus, vermivorous cone snails such as Conus floridulus, Conus imperialis, Conus quercinus and Conus eburneus, and molluscivorous cone snails such as Conus textile, Conus marmoreus, Conus episcopatus and Conus victoriae.
Cone snails have evolved to produce intricate varieties of toxins, thus serving as a prominent focus of contemporary scientific research. They diversified across various species and eras, producing a toxic system that offers distinct advantages [6,7,8,9,10,11]. Conotoxins act on ion channels and receptors in the nervous systems, including sodium ion channels (Nav), potassium ion channels (Kv), and calcium ion channels (Cav). Moreover, they exhibit an affinity for hormone receptors, including oxytocin and antidiuretic hormone [12,13,14,15]. Each Conus species’ venom contains a distinct array of pharmacologically active peptides, predominantly specific to that particular species [3,16]. Cumulatively, the 900 species within the Conus genus produce a significant quantity of these peptides, potentially surpassing a remarkable total of 180,000 distinct variants [17,18,19,20,21,22]. Cone snails employ their venomous arsenal for prey capture and defense against competitors, exhibiting a wide range of hunting strategies that include predation upon worms, snails, and fishes [23,24]. Additionally, they harness their venoms as a means to deter potential predators [11].
Vertebrates such as snakes and invertebrates such as cone snails, spiders, and scorpions are all important sources of venoms [25,26,27,28]. Currently, pharmaceuticals derived from venoms exhibit a broad spectrum of applications encompassing analgesic, anti-inflammatory, and antihypertensive properties [29,30,31]. Clinically approved drugs derived from animal venom components include captopril, enalapril, ziconotide, eptifibatide, tirofiban, lepirudin, bivalirudin, batroxobin, apitox, cobratid, and exenatide [32,33]. Ziconotide, the groundbreaking pharmaceutical derived from conotoxins, has been utilized since 2004 to address chronic pain by selectively inhibiting voltage-gated calcium channels. It obtained FDA approval for the treatment of chronic severe pain in patients who display intolerance or inadequate response to systemic analgesics or intrathecal morphine [34,35].
Exenatide, derived from a peptide (exendin-4), found in the saliva of the Gila monster, is one of the most classical examples of venom toxin derivatives [32,36]. It effectively regulates glycemic levels by stimulating glucose-dependent insulin secretion [37], and has been developed for the treatment of diabetes [36,38]. In 2005, the FDA approved its utilization as an adjunct therapy, paired with metformin or sulfonylurea, in the management of type 2 diabetes mellitus. Apart from Heloderma’s exendin-4, a diverse array of bioactive substances derived from cone snails, sea anemones, scorpions, mosquitoes, bees, and spiders exhibit promising prospects in effectively regulating blood glucose levels [32,39,40]. Based on the findings of these investigations, Safavi-Hemami et al. [41] identified insulin-like peptides, namely Con-Ins G1 and Con-Ins G2, within the venom of C. geographus. Con-Ins-G1 was observed to possess predatory adaptations by effectively reducing blood glucose levels in prey organisms as a means of capturing and immobilizing them [41]. Therefore, the identification of conotoxin-like insulins (conoinsulins) sourced from Conus species holds paramount importance in advancing the field of marine drugs.
In this review, a comprehensive comparative analysis was conducted to investigate the sequence and structural characteristics of insulin derived from human, zebrafish, and cone snails. All available insulin sequences were retrieved from databases, followed by comparisons and subsequent analysis of their diversity. A rich array of conoinsulin sequences from Conus specimens exhibiting diverse feeding habits were collected, subsequently enabling the construction of a phylogenetic tree to unveil their intricate evolutionary relationships. The primary objective of this paper is to establish a solid foundation for the exploration of conoinsulin’s diversity and phylogenetic analysis, thus furnishing providing robust support for the advancement of novel marine drug development endeavors.

2. Conoinsulin as a Weapon for Predation

Cone snails are commonly found in coastal reefs, rocks, and sandy beaches [42]. The shells of the cone snails are visually striking, featuring a diverse range of colors and bearing a resemblance to their non-toxic Conus counterparts. Consequently, they are occasionally inadvertently picked up by humans, resulting in potentially life-threatening incidents. There have been several documented incidents of accidental Conus collections leading to human fatalities [43]. Such poisonings are attributed to various toxins contained in the cone snail. This particular occurrence of poisoning is predominantly ascribed to various toxins manufactured within the cone snails. C. geographus, a proficient piscivorous predator, further exemplifies one of the most dangerous species to humans. Its venom can cause approximately 50% of poisoning incidents, as it encompasses 50% of the overall number of fatalities in all poisoning incidents [43]. Furthermore, C. geographus tends to result in higher mortality rates in children compared to adults [43]. Regardless of the victim’s age, the poisoning incidents from larger snails are typically more prevalent than those from smaller snails. Additionally, humans have also suffered stings from other fish-feeding Conus species, leading to fatal injuries [43]. The venom produced by other Conus species that prey on gastropods also poses significant safety hazards to humans. However, most reported deaths have yet to be confirmed.
The unique predation methods of cone snails have attracted significant interest from many researchers [44,45]. Some cone snails employ a burying behavior, concealing their bodies within the sand while exposing their elongated proboscis. This adaptive strategy enables them to simultaneously acquire oxygen and observe the activities of neighboring organisms. When detecting the proximity of potential prey, the cone snail deploys its elongated siphon beak in the direction of the prey [23]. Subsequently, the cone snails employ a specialized radular tooth with a hollow, harpoon-like structure, to immobilize the fish. This tooth not only anchors firmly onto its prey, but also injects a potent venom [23,44,45], acting much like a flexible "hypodermic syringe" connected to a sac filled with toxins [46]. By means of muscular contraction, the cone snail expels its venom into the prey’s body within a fraction of a second [47]. The venom discharged by the Conus species possesses potent toxic properties, capable of inducing severe poisoning or even lethality in the affected organism.
Fish rely on their inherent biological nervous system to regulate and coordinate their physical movements. When penetrated by the cone snail’s harpoon-like structure, the fish experiences a transient period of uncontrolled locomotion, lasting for a fraction of a second, followed by an immediate and complete paralysis [46]. The mixture of conotoxins contained in the venom swiftly targets the chemical receptors and ion channels responsible for modulating the fish’s neural transmissions, causing the receptor to remain persistently open due to the influx of the toxin. The conotoxin, administered by the cone snail, induces muscular spasms in the fish. These toxins target the synaptic connections between the fish’s nerves and muscles, impeding the muscle tissue’s ability to receive command signals. Gradually, as the intensity of spasms diminishes, the fish typically succumbs to complete paralysis in the majority of observed instances [46,48]. Lastly, the cone snail retracts its proboscis and proceeds to maneuver the incapacitated prey towards its oral cavity, thereby concluding the entirety of the predation process (Figure 2A).
Furthermore, an alternative predatory tactic employed by certain Conus species, referred to as C. geographus, involves the conspicuous expansion of their oral aperture, colloquially referred to as their "bloody mouths." Once opened, these individuals commence the emission of venomous substances into the surrounding aqueous environment, effectively paralyzing an entire assemblage of fish. These immobilized fish are subsequently engulfed by the cone snail in a single ingestion motion, facilitating the swift consumption of multiple prey items (Figure 2B).
Within the venomous arsenal of C. geographus, a multitude of bioactive compounds are present, including insulin. This species uses specialized insulin in its venom to facilitate hunting [41]. Fast-acting neurotoxin and delivery systems are required for cone snails to use conoinsulin to capture prey [49]. After locking on to the prey, the insulin in the venom is rapidly propelled toward the bulbous base of the harpoon, where it is injected into the fish [50]. Relevant studies and data have shown a very fast delivery speed of cone neurotoxin [47]. Remarkably, C. geographus produces two distinct forms of insulin. One resembles molluscan insulin and is produced in the nerve ring and the esophagus, which play a role in the regulation of hemolymph glucose levels, memory and learning. Another type of insulin, similar to the prey’s insulin, known as "venom insulin", is exclusively expressed in proximity to the injection site within the venom duct, attaining a conspicuously high concentration, which is the main component of the venom used for predation [41,51].
The fish-hunting cone snail, C. geographus, was previously shown to use derived venom insulin Con-Ins G1 to capture prey [14]. In vivo, intraperitoneal injection of Con-Ins G1 into zebrafish lowers their blood glucose with the same potency as human insulin [41,52]. When applied to water, Con-Ins G1 reduces the overall locomotor activity of zebrafish larvae, demonstrating a significant decrease in the percentage of time spent swimming and movement frequency [41]. When tested on a mouse model of diabetes, Con-Ins G1 also lowered blood glucose with a 10-fold higher potency compared to human insulin [52]. Con-Ins G1 has been demonstrated to constitute a prominent constituent within the venom composition of this particular species. These insulin-like molecules contained therein exert a pronounced effect on the blood glucose levels of the prey organisms, resulting in a significant reduction that impairs their ability to evade capture or flee from predation [41,53]. The comprehensive literature review reveals that the crude venom of cone snails targets not only voltage-gated ion channels, such as Nav, Kv and Cav channels, but also acts on nAChR and insulin receptors (IRs) [Figure 2C]. Therefore, conoinsulins, like other conotoxins, exhibit robust biological activity and hold potential as drug candidates for the treatment of diabetes, infertility, and other diseases.

3. Diversity Analysis of Conoinsulin

Insulin-like peptides, identified in cone snail venom, exhibit species-specific variations regarding expression. Certain species exclusively express a single type of conoinsulin, whereas others possess multiple conoinsulin variants. The precursors of these conoinsulins typically feature a conserved N-terminal signal sequence in their amino acid sequences. Among these sequences, some contained a pro-peptide region, while others did not; however, all these sequences exhibit at least one amino acid difference in the mature region [54]. Cone snails deploy venoms containing conoinsulin, which acts within seconds to immobilize nearby fish, facilitating easier capture and consumption [41,55]. This rapid action of conoinsulin, in stark contrast to human insulin, has intrigued scientists. In the ConoServer and UniProt databases, 38 different insulin sequences and one insulin-like peptide have been reported in the venom of 18 types of Conus, and alignment analysis was performed using MEGA 7.0.14. Con-Ins G1, Con-Ins G1b, Con-Ins G1c, Con-Ins G2, Con-Ins G2b, Con-Ins G3, and Con-Ins G3b were all sourced from C. geographus; Con-Ins T1A, Con-Ins T1B, Con-Ins T2, Con-Ins T3, and Con-Ins T4 were all discovered in C. tulipa; Con-Ins K1 and Con-Ins K2 were derived from C. kinoshitai; Con-Ins Q1 and Con-Ins Q1b were both sourced from C. quercinus; Con-Ins F1, Con-Ins F2, Con-Ins F2b and Con-Ins F2c were derived from the species C. floridulus; Con-Ins Me1 was discovered in C. memiae; Con-Ins Im1 and Con-Ins Im2 were found in C. imperialis; Con-Ins Pa1, Con-Ins Ti1, Con-Ins Pu1, Con-Ins Vir and Con-Ins Bn1 were discovered in C. planorbis, C. tribblei, C. pulicarius, C. virgo and C. bandanus, respectively; Con-Ins Vr1 and Con-Ins Vr2 were found in C. varius; Con-Ins Ts1 and Con-Ins Ts2 were found in C. tessulatus; Con-Ins Eu1 and Con-Ins Eu2 were discovered in C. eburneus; Con-Ins Mr1 and Con-Ins Mr2 were discovered in C. marmoreus; Con-Ins Tx1 and Con-Ins Tx2 were discovered in C.textile and ILP was discovered in C. victoriae (Figure 3). The precursors of human insulin, zebrafish insulin, and conoinsulin are all composed of signal peptides and mature peptides [21,56]. The results of sequence comparison indicate that the signal peptide of insulin is highly conserved. Interestingly, we observed that unlike human insulin and zebrafish insulin, some conoinsulins possess a unique propeptide component, which might aid in protein folding during secretion, and prevent protein misfolding or aggregation [57,58].
Insulin, a natural hormone, has been proven to manifest an extensive assortment of aggregates with diverse structures and morphologies [51,59]. For instance, the mature region in insulin represents a functional domain encompassing both A, B and C chains, typically characterized by variability and interconnected through disulfide bonds, thereby facilitating the exertion of its pharmacological activity [56,60,61,62]. Like human insulin, the mature peptide of conoinsulin contains chains, albeit with notable differences. In insulin sequences, cysteine residues, which are largely conserved, predominantly cluster in the A and B chains. Typically, the A chain of insulin follows a CC-C-C cysteine pattern, while the B chain has a C-C pattern, leading to the formation of three disulfide bonds. In contrast, some conoinsulin A chains exhibit a C-CC-C-C pattern, while the B chain has a C-C-C pattern, resulting in four disulfide bonds. In addition, the human insulin B-chain contains a C-terminal segment, which plays an important role in the assembly of insulin dimers or hexamers [63,64,65,66]. Interestingly, some conoinsulins such as Con-Ins-G1, despite the noticeably short or partially absent C-terminus of the insulin B chain, still retain an affinity towards the human insulin receptor [52,67,68]. These findings imply that the C-terminal of the B-chain does not serve as a pivotal determinant influencing the binding affinity between insulin and the insulin receptor [67]. Moreover, the conoinsulin does not undergo dimerization or hexamerization, thereby allowing for rapid reaction kinetics surpassing those of current insulin medications. Cone snails produce conoinsulin not only for regulating blood sugar like most organisms, but also for predation purposes. This has led to the evolution of a wide variety of conoinsulins, akin to conotoxins, enhancing their efficacy on diverse prey. These attributes hold promising implications for the advancement of improved or novel therapeutic approaches for diabetes [52,68].
Figure 3. Alignment of insulin in humans, zebrafish, and different insulin sequences in the venom of various Conus. MEGA 7.0.14 software was used to create insulin sequence alignments and perform amino acid alignments on all peptide sequences, where MUSCLE algorithm was chosen to intelligently align amino acids [69,70]. Genedoc 2.7 software was used to export the sequence in FASTA format. B-chains and A-chains are enclosed within red and purple boxes, respectively. The conserved cysteine residues and small amino acids are highlighted in yellow. Aliphatic and aromatic amino acids are in red and blue fonts, respectively, on a grey background. Amphoteric and polar groups are in red and black fonts, respectively, on a green background. Negatively and positively charged amino acids are shown on a blue background in green and red. Proline and glycine amino acids are shown on a red background in blue and green. The hydrophobic amino acids are highlighted in black. ZF-Ins represents zebrafish insulin and ILP represents insulin-like peptide from C. victoriae.
Figure 3. Alignment of insulin in humans, zebrafish, and different insulin sequences in the venom of various Conus. MEGA 7.0.14 software was used to create insulin sequence alignments and perform amino acid alignments on all peptide sequences, where MUSCLE algorithm was chosen to intelligently align amino acids [69,70]. Genedoc 2.7 software was used to export the sequence in FASTA format. B-chains and A-chains are enclosed within red and purple boxes, respectively. The conserved cysteine residues and small amino acids are highlighted in yellow. Aliphatic and aromatic amino acids are in red and blue fonts, respectively, on a grey background. Amphoteric and polar groups are in red and black fonts, respectively, on a green background. Negatively and positively charged amino acids are shown on a blue background in green and red. Proline and glycine amino acids are shown on a red background in blue and green. The hydrophobic amino acids are highlighted in black. ZF-Ins represents zebrafish insulin and ILP represents insulin-like peptide from C. victoriae.
Toxins 16 00034 g003
The three-dimensional (3D) structure of Con-Ins G1 from C. geographus was solved using X-ray crystallography [71]. Con-Ins-G1 not only exhibits a high affinity for the human insulin receptor but also shares a striking similarity with the zebrafish insulin [41]. Biologically, Con-Ins G1 has the highest similarity to fish insulin [71], particularly in the A-chain, while the B-chain similarity is not as pronounced [41]. In addition, owing to its compact structure, Con-Ins G1 acts swiftly [72], which aligns well with the rapid predation strategy of cone snails. Using Con-Ins G1 (PDB 5JYQ) as a template, homologous modeling methods generated nine different conoinsulin variants corresponding to the dietary preferences of various cone snails, including G1b, G3, T1b, K1, F1, F2, Im1, Tx1, and Mr1. Meanwhile, 3D modeling of zebrafish insulin using human insulin (PDB 3I40) as a template revealed highly homologous structures (Figure 4) [63,73,74,75,76].
Through structural comparison, analysis reveals that all insulin molecules contain three α-helices, with two located in the A-chain and one in the B-chain. Additionally, they feature a hydrophobic core composed of non-polar residues, vital for their proper folding and structural integrity [77,78]. The conoinsulins from fish-hunting cone snails display a high degree of structural similarity with human insulin and zebrafish insulin. This resemblance likely correlates with their dietary characteristics, aiming to act on fish IRs to lower blood sugar levels for effective predation. However, conoinsulins from fish-hunting cone snails are significantly different from those of worm-hunting cone snails and mollusk-hunting cone snails, especially on the B-chain. The B-chain in the fish-hunting cone structure is shorter, which could be an adaptation for quicker prey capture.
Figure 4. Comparison of the 3D structures of human insulin, zebrafish insulin, and conoinsulins. The cartoon representations of models of insulin variants are depicted, with A-chains and B-chains of each insulin in cyan and orange, respectively. The structures of human insulin (PDB 3I40) and Con-Ins G1(PDB 5JYQ) were sourced from the PDB database. Additional structures were obtained from AlphaFold Protein Structure Database (https://alphafold.ebi.ac.uk/). Protein 3D structure is predicted using homology computational structure prediction modeling from amino acid sequence [79]. SWISS-MODEL, available through the Expasy web server or Deep View software (Swiss Pdb-Viewer), is a fully automated server for the homology modeling of protein structures. The homologous sequences with highest sequence identity were assigned as templates, and then cartoon mode was used to build the model.
Figure 4. Comparison of the 3D structures of human insulin, zebrafish insulin, and conoinsulins. The cartoon representations of models of insulin variants are depicted, with A-chains and B-chains of each insulin in cyan and orange, respectively. The structures of human insulin (PDB 3I40) and Con-Ins G1(PDB 5JYQ) were sourced from the PDB database. Additional structures were obtained from AlphaFold Protein Structure Database (https://alphafold.ebi.ac.uk/). Protein 3D structure is predicted using homology computational structure prediction modeling from amino acid sequence [79]. SWISS-MODEL, available through the Expasy web server or Deep View software (Swiss Pdb-Viewer), is a fully automated server for the homology modeling of protein structures. The homologous sequences with highest sequence identity were assigned as templates, and then cartoon mode was used to build the model.
Toxins 16 00034 g004

4. Evolutionary Relationship of Conoinsulin

To elucidate the evolutionary relationship of conoinsulins, we sourced conoinsulin sequences through a comprehensive search of the ConoServer and UniProt databases, followed by confirmation using the Blast algorithm. After removing duplicate sequences, a total of 38 different conoinsulin sequences and one insulin-like peptide were obtained, which were then analyzed using MEGA 7.0.14 software. Using the NJ method, we constructed a phylogenetic tree to facilitate the analysis of the evolutionary relationship among these conoinsulins (Figure 5). A diverse array of conoinsulins has been identified across Conus species: ranging from ancestral groups, such as worm-hunting Conus situated at the foundational levels of the food chain, to relatively advanced organisms, such as fish-hunting Conus.
Conoinsulin expression is ubiquitously observed in mollusk-hunting, worm-hunting, and fish-hunting Conus individuals. However, unlike mollusk-hunting Conus and worm-hunting Conus, only those fish-hunting Conus that use a net-hunting strategy express venom fish-like insulin [53]. Hence, among some fish-hunting Conus species, unique insulin variants serve as the principal components of their venomous arsenal, conferring a critical advantage in prey capture.
Con-Ins Ti1, found in the worm-hunter C. tribblei, represents the most primitive venom insulins of cone snails. Subsequently, the phylogenetic tree diverged to include venom insulins produced by cone snails which specialize in feeding on fish and those that prey on snails. Among these fish-hunting Conus, the earliest appearing insulins are Con-Ins G2 and Con-Ins G2b originating from C. geographus. The venom insulins produced by cone snails with a fish-based diet, particularly the Con-Ins G1 originating from C. geographus, have attracted significant attention in recent research. It is noteworthy that the abundance of venom insulins from fish-hunting cone snails surpasses others, constituting approximately half of the phylogenetic tree encompassing Conus venom insulin. Following this, venom insulins produced by cone snails specializing in worm consumption display a moderate presence, whereas the quantity of venom insulins from mollusk-hunting cone snails remains relatively minimal.
According to the literature research, it has been found that all mollusk-hunting cone snails, most but not all worm-hunters and only a small subset of fish-hunters express conoinsulins [53]. Differing prey capture strategies may explain this difference. Some fish-hunting cone snails appear to release conoinsulins into the water to make an entire school of small fish hypoglycemic, thereby enhancing the cone snail’s ability to engulf multiple fish. In contrast, some fish-hunting species with no conoinsulins mainly capture fish by producing complex conotoxins that by cause hyper-excitability of the nervous system and rapid onset of a tetanic paralysis. There may be no role for a conoinsulin in this prey-capture strategy. Additionally, several worm-hunting species exhibit low or no levels of conoinsulin expression, indicating that conoinsulin may no longer be important in these species [51].

5. Conclusions and Perspectives

This review employed a comprehensive retrieval of cone snail insulin sequences from databases to evaluate the homology of venom insulin and analyze their sequence diversity. Additionally, a detailed phylogenetic tree was constructed to elucidate the evolutionary relationships among insulin sequences derived from cone snails with diverse dietary preferences. The insights gained provide a fundamental framework for future research into the targets and hypoglycemic activities of conoinsulins.
Moreover, future advancements in sequencing technologies, particularly high-resolution tandem mass spectrometry, combined with venom gland transcriptome databases, hold great potential for enhancing our understanding. This integrated approach is expected to facilitate the discovery and identification of additional insulin sequences in cone snails. Such research will contribute significantly to our knowledge of the evolutionary patterns, diversity, and functional attributes of venom insulin in various subtypes of cone snails. Furthermore, such investigations open up promising avenues for the development and application of novel marine pharmaceuticals.

Author Contributions

These should be presented as follows: B.G. conceived and designed the project. Q.G. analyzed the data and wrote the paper. M.H. revised the manuscript and optimized the images. M.L. and S.C. revised the manuscript. All authors contributed to editorial changes in the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Hainan Provincial Key Point Research and Invention Program (no. ZDYF2022SHFZ309), Special scientific research project of the Hainan academician innovation platform (no. YSPTZX202132), Hainan Provincial Natural Science Foundation of China (no. 820RC636) and Hainan Medical University graduate innovation and entrepreneurship training program (no. HYYS2021A24).

Data Availability Statement

All insulin sequences are from UniProt and ConoServer databases (www.uniprot.org/; http://conoserver.org/).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gao, B.; Peng, C.; Yang, J.; Yi, Y.; Zhang, J.; Shi, Q. Cone Snails: A Big Store of Conotoxins for Novel Drug Discovery. Toxins 2017, 9, 397. [Google Scholar] [CrossRef] [PubMed]
  2. Himaya, S.W.; Jin, A.H.; Dutertre, S.; Giacomotto, J.; Mohialdeen, H.; Vetter, I.; Alewood, P.F.; Lewis, R.J. Comparative Venomics Reveals the Complex Prey Capture Strategy of the Piscivorous Cone Snail Conus catus. J. Proteome Res. 2015, 14, 4372–4381. [Google Scholar] [CrossRef] [PubMed]
  3. Abalde, S.; Dutertre, S.; Zardoya, R. A Combined Transcriptomics and Proteomics Approach Reveals the Differences in the Predatory and Defensive Venoms of the Molluscivorous Cone Snail Cylinder ammiralis (Caenogastropoda: Conidae). Toxins 2021, 13, 642. [Google Scholar] [CrossRef] [PubMed]
  4. Prashanth, J.R.; Dutertre, S.; Jin, A.H.; Lavergne, V.; Hamilton, B.; Cardoso, F.C.; Griffin, J.; Venter, D.J.; Alewood, P.F.; Lewis, R.J. The Role of Defensive Rcological Interactions in the Evolution of Conotoxins. Mol. Ecol. 2016, 25, 598–615. [Google Scholar] [CrossRef] [PubMed]
  5. Lewis, R.J.; Dutertre, S.; Vetter, I.; Christie, M.J. Conus Venom Peptide Pharmacology. Pharmacol. Rev. 2012, 64, 259–298. [Google Scholar] [CrossRef] [PubMed]
  6. Holford, M.; Daly, M.; King, G.F.; Norton, R.S. Venoms to the Rescue. Science 2018, 361, 842–844. [Google Scholar] [CrossRef] [PubMed]
  7. Modica, M.V.; Sunagar, K.; Holford, M.; Dutertre, S. Editorial: Diversity and Evolution of Animal Venoms: Neglected Targets, Ecological Interactions, Future Perspectives. Front. Ecol. Evol. 2020, 8, 65. [Google Scholar] [CrossRef]
  8. Pennington, M.W.; Czerwinski, A.; Norton, R.S. Peptide Therapeutics from Venom: Current Status and Potential. Bioorg. Med. Chem. 2018, 26, 2738–2758. [Google Scholar] [CrossRef]
  9. Utkin, Y.N. Animal Venom Studies: Current Benefits and Future Developments. World J. Biol. Chem. 2015, 6, 28–33. [Google Scholar] [CrossRef]
  10. Prator, C.A.; Murayama, K.M.; Schulz, J.R. Venom Variation During Prey Capture by the Cone Snail, Conus textile. PLoS ONE 2014, 9, e98991. [Google Scholar] [CrossRef]
  11. Dutertre, S.; Jin, A.H.; Vetter, I.; Hamilton, B.; Sunagar, K.; Lavergne, V.; Dutertre, V.; Fry, B.G.; Antunes, A.; Venter, D.J.; et al. Evolution of Separate Predation- and Defence-Evoked Venoms in Carnivorous Cone Snails. Nat. Commun. 2014, 5, 3521. [Google Scholar] [CrossRef] [PubMed]
  12. Lewis, R.J. Ion Channel Toxins and Therapeutics: From Cone Snail Venoms to Ciguatera. Ther. Drug Monit. 2000, 22, 61–64. [Google Scholar] [CrossRef] [PubMed]
  13. Favreau, P.; Gall, F.L.; Benoit, E.; Molgó, J. A Review on Conotoxins Targeting Ion Channels and Acetylcholine Receptors of the Vertebrate Neuromuscular Junction. Acta Physiol. Pharmacol. Et Ther. Latinoam. 1999, 49, 257–267. [Google Scholar]
  14. Robinson, S.D.; Li, Q.; Bandyopadhyay, P.K.; Gajewiak, J.; Yandell, M.; Papenfuss, A.T.; Purcell, A.W.; Norton, R.S.; Safavi-Hemami, H. Hormone-like Peptides in the Venoms of Marine Cone Snails. Gen. Comp. Endocrinol. 2017, 244, 11–18. [Google Scholar] [CrossRef]
  15. Norton, R.S.; Olivera, B.M. Conotoxins down under. Toxicon 2006, 48, 780–798. [Google Scholar] [CrossRef] [PubMed]
  16. Yin, X.; Guo, S.; Gao, J.; Luo, L.; Liao, X.; Li, M.; Su, H.; Huang, Z.; Xu, J.; Pei, J.; et al. Kinetic Analysis of Effects of Temperature and Time on the Regulation of Venom Expression in Bungarus multicinctus. Sci. Rep. 2020, 10, 14142. [Google Scholar] [CrossRef] [PubMed]
  17. Terlau, H.; Olivera, B.M. Conus Venoms: A Rich Source of Novel Ion Channel-Targeted Peptides. Physiol. Rev. 2004, 84, 41–68. [Google Scholar] [CrossRef]
  18. Olivera, B.M. Conus Peptides: Biodiversity-Based Discovery and Exogenomics. J. Biol. Chem. 2006, 281, 31173–31177. [Google Scholar] [CrossRef]
  19. Barghi, N.; Concepcion, G.P.; Olivera, B.M.; Lluisma, A.O. High Conopeptide Diversity in Conus tribblei Revealed Through Analysis of Venom Duct Transcriptome Using Two High-Throughput Sequencing Platforms. Mar. Biotechnol. 2015, 17, 81–98. [Google Scholar] [CrossRef]
  20. Lu, A.; Watkins, M.; Li, Q.; Robinson, S.D.; Concepcion, G.P.; Yandell, M.; Weng, Z.; Olivera, B.M.; Safavi-Hemami, H.; Fedosov, A.E. Transcriptomic Profiling Reveals Extraordinary Diversity of Venom Peptides in Unexplored Predatory Gastropods of the Genus Clavus. Genome Biol. Evol. 2020, 12, 684–700. [Google Scholar] [CrossRef]
  21. Robinson, S.D.; Norton, R.S. Conotoxin Gene Superfamilies. Mar. Drugs 2014, 12, 6058–6101. [Google Scholar] [CrossRef] [PubMed]
  22. Fedosov, A.; Zaharias, P.; Puillandre, N. A Phylogeny-Aware Approach Reveals Unexpected Venom Components in Divergent Lineages of Cone Snails. Proc. Biol. Sci. 2021, 288, 20211017. [Google Scholar] [CrossRef] [PubMed]
  23. Olivera, B.M.; Seger, J.; Horvath, M.P.; Fedosov, A.E. Prey-Capture Strategies of Fish-Hunting Cone Snails: Behavior, Neurobiology and Evolution. Brain Behav. Evol. 2015, 86, 58–74. [Google Scholar] [CrossRef] [PubMed]
  24. Phuong, M.A.; Mahardika, G.N.; Alfaro, M.E. Dietary Breadth is Positively Correlated with Venom Complexity in Cone Snails. BMC Genom. 2016, 17, 401. [Google Scholar] [CrossRef] [PubMed]
  25. Suryamohan, K.; Krishnankutty, S.P.; Guillory, J.; Jevit, M.; Schroder, M.S.; Wu, M.; Kuriakose, B.; Mathew, O.K.; Perumal, R.C.; Koludarov, I.; et al. The Indian Cobra Reference Genome and Transcriptome Enables Comprehensive Identification of Venom Toxins. Nat. Genet. 2020, 52, 106–117. [Google Scholar] [CrossRef] [PubMed]
  26. Pardos-Blas, J.R.; Irisarri, I.; Abalde, S.; Afonso, C.M.L.; Tenorio, M.J.; Zardoya, R. The Genome of the Venomous Snail Lautoconus ventricosus Sheds Light on the Origin of Conotoxin Diversity. Gigascience 2021, 10, giab037. [Google Scholar] [CrossRef]
  27. Sanggaard, K.W.; Bechsgaard, J.S.; Fang, X.; Duan, J.; Dyrlund, T.F.; Gupta, V.; Jiang, X.; Cheng, L.; Fan, D.; Feng, Y.; et al. Spider Genomes Provide Insight into Composition and Evolution of Venom and Silk. Nat. Commun. 2014, 5, 3765. [Google Scholar] [CrossRef]
  28. Murthy, K.R.; Hase, N.K. Scorpion Envenoming and the Role of Insulin. Toxicon 1994, 32, 1041–1044. [Google Scholar] [CrossRef]
  29. Herzig, V.; Cristofori-Armstrong, B.; Israel, M.R.; Nixon, S.A.; Vetter, I.; King, G.F. Animal Toxins-Nature’s Evolutionary-Refined Toolkit for Basic Research and Drug Discovery. Biochem. Pharmacol. 2020, 181, 114096. [Google Scholar] [CrossRef]
  30. Henney, N.C.; Glenn, F. King: Venoms to Drugs: Venom as a Source for the Development of Human Therapeutics. Chromatographia 2015, 78, 1507–1508. [Google Scholar] [CrossRef]
  31. Mohamed Abd El-Aziz, T.; Garcia Soares, A.; Stockand, J.D. Snake Venoms in Drug Discovery: Valuable Therapeutic Tools for Life Saving. Toxins 2019, 11, 564. [Google Scholar] [CrossRef] [PubMed]
  32. 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] [PubMed]
  33. Zambelli, V.O.; Pasqualoto, K.F.; Picolo, G.; Chudzinski-Tavassi, A.M.; Cury, Y. Harnessing the Knowledge of Animal Toxins to Generate Drugs. Pharmacol. Res. 2016, 112, 30–36. [Google Scholar] [CrossRef] [PubMed]
  34. Peigneur, S.; Tytgat, J. Toxins in Drug Discovery and Pharmacology. Toxins 2018, 10, 126. [Google Scholar] [CrossRef] [PubMed]
  35. Narain, S.; Al-Khoury, L.; Chang, E. Resolution of Chronic Migraine Headaches with Intrathecal Ziconotide: A Case Report. J. Pain. Res. 2015, 8, 603–606. [Google Scholar] [CrossRef] [PubMed]
  36. Iltz, J.L.; Baker, D.E.; Setter, S.M.; Keith Campbell, R. Exenatide: An Incretin Mimetic for the Treatment of Type 2 Diabetes Mellitus. Clin. Ther. 2006, 28, 652–665. [Google Scholar] [CrossRef] [PubMed]
  37. Moore, S.W.; Bhat, V.K.; Flatt, P.R.; Gault, V.A.; McClean, S. Isolation and Characterisation of Insulin-Releasing Compounds from Crotalus adamanteus, Crotalus vegrandis and Bitis nasicornis Venom. Toxicon 2015, 101, 48–54. [Google Scholar] [CrossRef]
  38. Yap, M.K.K.; Misuan, N. Exendin-4 from Heloderma suspectum Venom: From Discovery to its Latest Application as Type II Diabetes Combatant. Basic Clin. Pharmacol. Toxicol. 2019, 124, 513–527. [Google Scholar] [CrossRef]
  39. Robinson, S.D.; Safavi-Hemami, H. Venom Peptides as Pharmacological Tools and Therapeutics for Diabetes. Neuropharmacology 2017, 127, 79–86. [Google Scholar] [CrossRef]
  40. Mitchell, M.L.; Hossain, M.A.; Lin, F.; Pinheiro-Junior, E.L.; Peigneur, S.; Wai, D.C.C.; Delaine, C.; Blyth, A.J.; Forbes, B.E.; Tytgat, J.; et al. Identification, Synthesis, Conformation and Activity of an Insulin-like Peptide from a Sea Anemone. Biomolecules 2021, 11, 1785. [Google Scholar] [CrossRef]
  41. Safavi-Hemami, H.; Gajewiak, J.; Karanth, S.; Robinson, S.D.; Ueberheide, B.; Douglass, A.D.; Schlegel, A.; Imperial, J.S.; Watkins, M.; Bandyopadhyay, P.K.; et al. Specialized Insulin is Used for Chemical Warfare by Fish-Hunting Cone Snails. Proc. Natl. Acad. Sci. USA 2015, 112, 1743–1748. [Google Scholar] [CrossRef] [PubMed]
  42. Kohn, A. Maximal Species Richness in Conus: Diversity, Diet and Habitat on Reefs of Northeast Papua New Guinea. Coral Reefs 2001, 20, 25–38. [Google Scholar] [CrossRef]
  43. Kohn, A.J. Human Injuries and Fatalities Due to Venomous Marine Snails of the Family Conidae. Int. J. Clin. Pharmacol. Ther. 2016, 54, 524–538. [Google Scholar] [CrossRef] [PubMed]
  44. Terlau, H.; Shon, K.J.; Grilley, M.; Stocker, M.; Stuhmer, W.; Olivera, B.M. Strategy for Rapid Immobilization of Prey by a Fish-Hunting Marine Snail. Nature 1996, 381, 148–151. [Google Scholar] [CrossRef] [PubMed]
  45. Aman, J.W.; Imperial, J.S.; Ueberheide, B.; Zhang, M.M.; Aguilar, M.; Taylor, D.; Watkins, M.; Yoshikami, D.; Showers-Corneli, P.; Safavi-Hemami, H.; et al. Insights into the Origins of Fish hunting in Venomous Cone Snails from Studies of Conus tessulatus. Proc. Natl. Acad. Sci. USA 2015, 112, 5087–5092. [Google Scholar] [CrossRef] [PubMed]
  46. Salisbury, S.M.; Martin, G.G.; Kier, W.M.; Schulz, J.R. Venom Kinematics During Prey Capture in Conus: The Biomechanics of a Rapid Injection System. J. Exp. Biol. 2010, 213, 673–682. [Google Scholar] [CrossRef] [PubMed]
  47. Schulz, J.R.; Norton, A.G.; Gilly, W.F. The Projectile Tooth of a Fish-Hunting Cone Snail: Conus catus Injects Venom into Fish Prey Using a High-Speed Ballistic Mechanism. Biol. Bull. 2004, 207, 77–79. [Google Scholar] [CrossRef]
  48. Kelley, W.P.; Schulz, J.R.; Jakubowski, J.A.; Gilly, W.F.; Sweedler, J.V. Two Toxins from Conus striatus that Individually Induce Tetanic Paralysis. Biochemistry 2006, 45, 14212–14222. [Google Scholar] [CrossRef]
  49. Turesson, H.; Satta, A.; Domenici, P. Preparing for Escape: Anti-predator Posture and Fast-Start Performance in Gobies. J. Exp. Biol. 2009, 212, 2925–2933. [Google Scholar] [CrossRef]
  50. Schulz, J.R.; Jan, I.; Sangha, G.; Azizi, E. The High Speed Radular Prey Strike of a Fish-Hunting Cone Snail. Curr. Biol. 2019, 29, R788–R789. [Google Scholar] [CrossRef]
  51. Safavi-Hemami, H.; Lu, A.; Li, Q.; Fedosov, A.E.; Biggs, J.; Showers Corneli, P.; Seger, J.; Yandell, M.; Olivera, B.M. Venom Insulins of Cone Snails Diversify Rapidly and Track Prey Taxa. Mol. Biol. Evol. 2016, 33, 2924–2934. [Google Scholar] [CrossRef] [PubMed]
  52. Ahorukomeye, P.; Disotuar, M.M.; Gajewiak, J.; Karanth, S.; Watkins, M.; Robinson, S.D.; Florez Salcedo, P.; Smith, N.A.; Smith, B.J.; Schlegel, A.; et al. Fish-Hunting Cone Snail Venoms are a Rich Source of Minimized Ligands of the Vertebrate Insulin Receptor. Elife 2019, 8, e41574. [Google Scholar] [CrossRef] [PubMed]
  53. Robinson, S.D.; Safavi-Hemami, H. Insulin as a Weapon. Toxicon 2016, 123, 56–61. [Google Scholar] [CrossRef] [PubMed]
  54. Li, X.; Chen, W.; Zhangsun, D.; Luo, S. Diversity of Conopeptides and Their Precursor Genes of Conus Litteratus. Mar. Drugs 2020, 18, 464. [Google Scholar] [CrossRef] [PubMed]
  55. Kumar, P.S.; Kumar, D.S.; Umamaheswari, S. A Perspective on Toxicology of Conus Venom Peptides. Asian Pac. J. Trop. Med. 2015, 8, 337–351. [Google Scholar] [CrossRef] [PubMed]
  56. Puillandre, N.; Koua, D.; Favreau, P.; Olivera, B.M.; Stocklin, R. Molecular Phylogeny, Classification and Evolution of Conopeptides. J. Mol. Evol. 2012, 74, 297–309. [Google Scholar] [CrossRef] [PubMed]
  57. Wandersman, C. Secretion, Processing and Activation of Bacterial Extracellular Proteases. Mol. Microbiol. 2010, 3, 1825–1831. [Google Scholar] [CrossRef]
  58. González-Páez, G.E.; Roncase, E.J.; Wolan, D.W. X-ray Structure of an Inactive Zymogen Clostripain-like Protease from Parabacteroides distasonis. Acta Crystallogr. Sect. D 2019, 75, 325–332. [Google Scholar] [CrossRef]
  59. Shabanpoor, F.; Separovic, F.; Wade, J.D. The Human Insulin Superfamily of Polypeptide Hormones. Vitam. Horm. 2009, 80, 1–31. [Google Scholar] [CrossRef]
  60. Kaas, Q.; Westermann, J.C.; Craik, D.J. Conopeptide Characterization and Classifications: An Analysis Using ConoServer. Toxicon 2010, 55, 1491–1509. [Google Scholar] [CrossRef]
  61. Kordis, D.; Gubensek, F. Adaptive Evolution of Animal Toxin Multigene Families. Gene 2000, 261, 43–52. [Google Scholar] [CrossRef] [PubMed]
  62. Fassio, G.; Modica, M.V.; Mary, L.; Zaharias, P.; Fedosov, A.E.; Gorson, J.; Kantor, Y.I.; Holford, M.; Puillandre, N. Venom Diversity and Evolution in the Most Divergent Cone Snail Genus Profundiconus. Toxins 2019, 11, 623. [Google Scholar] [CrossRef] [PubMed]
  63. Harding, M.M.; Hodgkin, D.C.; Kennedy, A.F.; O’Conor, A.; Weitzmann, P.D. The Crystal Structure of Insulin. II. An Investigation of Rhombohedral Zinc Insulin Crystals and a Report of Other Crystalline Forms. J. Mol. Biol. 1966, 16, 212–226. [Google Scholar] [CrossRef] [PubMed]
  64. Dong, J.; Wan, Z.; Popov, M.; Carey, P.R.; Weiss, M.A. Insulin Assembly Damps Conformational Fluctuations: Raman Analysis of Amide I Linewidths in Native States and Fibrils. J. Mol. Biol. 2003, 330, 431–442. [Google Scholar] [CrossRef] [PubMed]
  65. Weiss, M.A.; Lawrence, M.C. A Thing of Beauty: Structure and Function of Insulin’s "Aromatic Triplet". Diabetes Obes. Metab. 2018, 20 (Suppl. S2), 51–63. [Google Scholar] [CrossRef] [PubMed]
  66. Derewenda, U.; Derewenda, Z.; Dodson, G.G.; Hubbard, R.E.; Korber, F. Molecular Structure of Insulin: The Insulin Monomer and Its Assembly. Br. Med. Bull. 1989, 45, 4–18. [Google Scholar] [CrossRef]
  67. Jiracek, J.; Zakova, L. From Venom Peptides to a Potential Diabetes Treatment. Elife 2019, 8, e44829. [Google Scholar] [CrossRef]
  68. Gorai, B.; Vashisth, H. Structures and Interactions of Insulin-like Peptides from Cone Snail Venom. Proteins 2022, 90, 680–690. [Google Scholar] [CrossRef]
  69. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef]
  70. Hall, B.G. Building Phylogenetic Srees from Molecular Data with MEGA. Mol. Biol. Evol. 2013, 30, 1229–1235. [Google Scholar] [CrossRef]
  71. Menting, J.G.; Gajewiak, J.; MacRaild, C.A.; Chou, D.H.; Disotuar, M.M.; Smith, N.A.; Miller, C.; Erchegyi, J.; Rivier, J.E.; Olivera, B.M.; et al. A Minimized Human Insulin-Receptor-Binding Motif Revealed in a Conus geographus Venom Insulin. Nat. Struct. Mol. Biol. 2016, 23, 916–920. [Google Scholar] [CrossRef] [PubMed]
  72. Xiong, X.; Menting, J.G.; Disotuar, M.M.; Smith, N.A.; Delaine, C.A.; Ghabash, G.; Agrawal, R.; Wang, X.; He, X.; Fisher, S.J.; et al. Author Correction: A Structurally Minimized yet Fully Active Insulin Based on Cone-Snail Venom Insulin Principles. Nat. Struct. Mol. Biol. 2020, 27, 683. [Google Scholar] [CrossRef] [PubMed]
  73. Sanger, F.; Thompson, E.O. The Amino-Acid Sequence in the Glycyl Chain of Insulin. I. The Identification of Lower Peptides from Partial Hydrolysates. Biochem. J. 1953, 53, 353–366. [Google Scholar] [CrossRef]
  74. Kahn, C.R.; Crettaz, M. Insulin Receptors and the Molecular Mechanism of Insulin Action. Diabetes Metab. Rev. 1985, 1, 5–32. [Google Scholar] [CrossRef] [PubMed]
  75. Nicol, D.S.; Smith, L.F. Amino-Acid Sequence of Human Insulin. Nature 1960, 187, 483–485. [Google Scholar] [CrossRef] [PubMed]
  76. Ostergaard, M.; Mishra, N.K.; Jensen, K.J. The ABC of Insulin: The Organic Chemistry of a Small Protein. Chemistry 2020, 26, 8341–8357. [Google Scholar] [CrossRef]
  77. Blundell, T.L.; Cutfield, J.F.; Cutfield, S.M.; Dodson, E.J.; Dodson, G.G.; Hodgkin, D.C.; Mercola, D.A. Three-Dimensional Atomic Structure of Insulin and Its Relationship to Activity. Diabetes 1972, 21, 492–505. [Google Scholar] [CrossRef]
  78. Baker, E.N.; Blundell, T.L.; Cutfield, J.F.; Cutfield, S.M.; Dodson, E.J.; Dodson, G.G.; Hodgkin, D.M.; Hubbard, R.E.; Isaacs, N.W.; Reynolds, C.D.; et al. The Structure of 2Zn Pig Insulin Crystals at 1.5 A Resolution. Philos. Trans. R Soc. Lond. B Biol. Sci. 1988, 319, 369–456. [Google Scholar] [CrossRef]
  79. Muhammed, M.T.; Aki-Yalcin, E. Homology Modeling in Drug Discovery: Overview, Current Applications, and Future Perspectives. Chem. Biol. Drug Des. 2019, 93, 12–20. [Google Scholar] [CrossRef]
Figure 1. Conus species exhibit variances in their predator-prey relationships. (A) Fish hunters. (B) Worm hunters. (C) Mollusk hunters.
Figure 1. Conus species exhibit variances in their predator-prey relationships. (A) Fish hunters. (B) Worm hunters. (C) Mollusk hunters.
Toxins 16 00034 g001
Figure 2. Roles as fishing weapons and medical value of conotoxins and conoinsulins. (A) Injection of the venom directly into the prey’s body to numb or poison it. (B) Release the venom into the surrounding water to numb the prey. (C) Diseases, receptors, and species related to different toxins and insulin. Representative cone snails are cited in the orange band. The light green band lists different targets for conotoxins or conoinsulins, including voltage-gated potassium ion channel (Kv), voltage-gated sodium channel (Nav), insulin receptor (IR), voltage-gated calcium channel (Cav), and nicotine acetylcholine receptor (nAChRs). The dark green band displays representative conotoxins or conoinsulins targeting different receptors. The blue band exemplifies typical diseases targeted by conotoxins or conoinsulins.
Figure 2. Roles as fishing weapons and medical value of conotoxins and conoinsulins. (A) Injection of the venom directly into the prey’s body to numb or poison it. (B) Release the venom into the surrounding water to numb the prey. (C) Diseases, receptors, and species related to different toxins and insulin. Representative cone snails are cited in the orange band. The light green band lists different targets for conotoxins or conoinsulins, including voltage-gated potassium ion channel (Kv), voltage-gated sodium channel (Nav), insulin receptor (IR), voltage-gated calcium channel (Cav), and nicotine acetylcholine receptor (nAChRs). The dark green band displays representative conotoxins or conoinsulins targeting different receptors. The blue band exemplifies typical diseases targeted by conotoxins or conoinsulins.
Toxins 16 00034 g002
Figure 5. The phylogenetic tree of 38 different conoinsulin sequences and one insulin-like peptide from 18 species of Conus. The 39 peptide sequences were obtained from UniProt and ConoServer databases (www.uniprot.org/; http://conoserver.org/). The 39 peptide sequences were aligned using MEGA 7.0.14 software. A phylogenetic tree was established using a neighbor-Joining approach (bootstrap method 1000 and pairwise deletion 50%). The color in the inner circle of the figure indicates that insulin comes from three different dietary habits of Conus, while the color in the outer circle indicates different species of Conus.
Figure 5. The phylogenetic tree of 38 different conoinsulin sequences and one insulin-like peptide from 18 species of Conus. The 39 peptide sequences were obtained from UniProt and ConoServer databases (www.uniprot.org/; http://conoserver.org/). The 39 peptide sequences were aligned using MEGA 7.0.14 software. A phylogenetic tree was established using a neighbor-Joining approach (bootstrap method 1000 and pairwise deletion 50%). The color in the inner circle of the figure indicates that insulin comes from three different dietary habits of Conus, while the color in the outer circle indicates different species of Conus.
Toxins 16 00034 g005
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

Guo, Q.; Huang, M.; Li, M.; Chen, J.; Cheng, S.; Ma, L.; Gao, B. Diversity and Evolutionary Analysis of Venom Insulin Derived from Cone Snails. Toxins 2024, 16, 34. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16010034

AMA Style

Guo Q, Huang M, Li M, Chen J, Cheng S, Ma L, Gao B. Diversity and Evolutionary Analysis of Venom Insulin Derived from Cone Snails. Toxins. 2024; 16(1):34. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16010034

Chicago/Turabian Style

Guo, Qiqi, Meiling Huang, Ming Li, Jiao Chen, Shuanghuai Cheng, Linlin Ma, and Bingmiao Gao. 2024. "Diversity and Evolutionary Analysis of Venom Insulin Derived from Cone Snails" Toxins 16, no. 1: 34. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16010034

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