Next Article in Journal
Are Post-Care Recommendations Following Upper-Face Botulinum Toxin Treatment Scientifically Necessary? A Retrospective Study Based on 5000 Patients
Previous Article in Journal
Heatwave-Induced Thermal Stratification Shaping Microbial-Algal Communities Under Different Climate Scenarios as Revealed by Long-Read Sequencing and Imaging Flow Cytometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Animal Venom in Modern Medicine: A Review of Therapeutic Applications

1
College of Veterinary Medicine, Gyeongsang National University, Jinju 52828, Republic of Korea
2
Institute of Animal Medicine, Gyeongsang National University, Jinju 52828, Republic of Korea
3
Pathology Laboratory, Medical Faculty, University of Jember, Jember 68126, Indonesia
4
Department of Pharmacology, Bushehr University of Medical Sciences, Bushehr 14174, Iran
*
Author to whom correspondence should be addressed.
Toxins 2025, 17(8), 371; https://doi.org/10.3390/toxins17080371
Submission received: 12 June 2025 / Revised: 11 July 2025 / Accepted: 22 July 2025 / Published: 28 July 2025
(This article belongs to the Section Animal Venoms)

Abstract

Animal venoms are complex biochemical secretions rich in highly potent and selective bioactive molecules, including peptides, enzymes, and small organic compounds. Once associated primarily with toxicity, these venoms are now recognized as a promising source of therapeutic agents for a wide range of medical conditions. This review provides a comprehensive analysis of the pharmacological potential of venom-derived compounds, highlighting their mechanisms of action, such as ion channel modulation, receptor targeting, and enzyme inhibition. Successful venom-derived drugs like captopril and ziconotide exemplify the translational potential of this biological arsenal. We discuss therapeutic applications in cardiovascular diseases, chronic pain, cancer, thrombosis, and infectious diseases, as well as emerging peptide candidates in clinical development. Technological advancements in omics, structural biology, and synthetic peptide engineering have significantly enhanced the discovery and optimization of venom-based therapeutics. Despite challenges related to stability, immunogenicity, and ecological sustainability, the integration of AI-driven drug discovery and personalized medicine is expected to accelerate progress in this field. By synthesizing current findings and future directions, this review underscores the transformative potential of animal venoms in modern pharmacotherapy and drug development. We also discuss current therapeutic limitations and how venom-derived compounds may address unmet needs in specific disorders.
Key Contribution: This review explores the therapeutic potential of animal venoms, focusing on venom-derived compounds with clinically proven and emerging applications. It highlights key mechanisms such as ion channel modulation and enzyme inhibition, with examples like captopril and ziconotide. Advances in omics, structural biology, and AI are also discussed as drivers accelerating venom-based drug discovery for conditions with unmet medical needs.

1. Introduction

Animal venoms are sophisticated biochemical arsenals composed of a diverse array of bioactive molecules, including peptides, enzymes, and non-proteinaceous compounds, that have evolved for predation and defense. While these venoms have long been associated with toxicity and lethality, recent decades have seen a growing recognition of their potential as a rich source of pharmacologically active agents [1,2]. Venom-derived compounds are often highly selective and potent, characteristics that are desirable in therapeutic development [3]. The translation of venom components into clinically useful drugs has already yielded notable successes, such as captopril, derived from the venom of the Brazilian pit viper Bothrops jararaca, which laid the foundation for modern antihypertensive therapies [4,5].
The modern resurgence of interest in animal venoms is driven not only by the unique pharmacological properties of venom components but also by advancements in omics technologies, peptide synthesis, and structural biology. These tools have enabled researchers to isolate, characterize, and optimize venom-derived compounds with greater precision, facilitating their incorporation into drug development pipelines [6,7]. The diversity of venomous species—ranging from snakes and scorpions to cone snails, spiders, and jellyfish—has further broadened the landscape of potential drug candidates [1,8,9].
Animal venom-derived peptides and proteins exhibit a wide range of biological activities, including analgesic, anticoagulant, antihypertensive, anticancer, antimicrobial, and neuroprotective effects [10,11,12,13]. Their mechanisms of action often involve modulation of specific ion channels, receptors, or enzymes, enabling highly targeted therapeutic interventions [3,14]. As such, venom-based drug discovery represents a promising frontier in biomedical research, particularly for diseases that remain challenging to treat with conventional small molecules [3,10].
This review aims to provide a comprehensive overview of animal venoms as a source of novel therapeutics, with a focus on their biochemical composition, mechanisms of action, and current and emerging applications in modern medicine. By integrating findings from molecular pharmacology, structural biology, and clinical studies, this work highlights the translational potential of venom-derived compounds and the innovative methodologies driving their development.

2. Biochemical Composition of Animal Venoms

Animal venoms are complex secretions comprising a heterogeneous mixture of biologically active molecules, including peptides, proteins, enzymes, nucleotides, lipids, and small organic compounds [9,15]. The composition of venom varies significantly across species, reflecting evolutionary adaptations to different ecological niches and prey types [16]. In general, venom components can be broadly categorized into enzymatic and non-enzymatic constituents. Enzymatic components such as phospholipase A2 enzymes (PLA2s), metalloproteinases, serine proteases, hyaluronidases, and L-amino acid oxidases play roles in tissue degradation, hemorrhage, inflammation, and disruption of homeostasis [17,18,19]. Non-enzymatic components, including neurotoxins, cardiotoxins, cytotoxins, and ion channel blockers, typically target specific cellular receptors and channels with remarkable potency and specificity [20,21].
Recent proteomic and transcriptomic studies have revealed even greater complexity within venom compositions, uncovering novel low-abundance components with potential pharmacological activities that were previously overlooked [22]. Such components may include unique disulfide-rich peptides, natriuretic peptides, and bradykinin-potentiating peptides, which expand the therapeutic landscape of venoms for conditions such as hypertension and chronic pain [23].
Snake venoms are among the most extensively studied and serve as prototypical models for venom pharmacology. Elapid venoms (e.g., from cobras and kraits) are predominantly neurotoxic, containing three-finger toxins (3FTxs) and dendrotoxins that block or modulate ion channels and neurotransmitter receptors [24,25]. In contrast, viperid venoms are rich in hemotoxic and cytotoxic components such as snake venom metalloproteinases (SVMPs), serine proteases, and disintegrins, which impair coagulation, damage vasculature, and disrupt extracellular matrix integrity [26,27].
Beyond snakes, cone snails (Conus spp.) produce conotoxins—small, disulfide-rich peptides with exquisite specificity for voltage-gated ion channels, nicotinic acetylcholine receptors, and G-protein-coupled receptors. These peptides have become valuable pharmacological tools and therapeutic leads, exemplified by ziconotide, an FDA-approved analgesic for chronic pain management [28,29]. Similarly, arachnid venoms (e.g., from spiders and scorpions) contain neuropeptides that modulate sodium, potassium, and calcium channels, offering potential for treating neurological disorders and pain [3,30].
Marine venoms, such as those from jellyfish and sea anemones, contain unique bioactive peptides and proteins with hemolytic, cytolytic, and neurotoxic properties [31,32]. These venoms are relatively underexplored compared to terrestrial venoms but represent an expanding frontier in venom-based drug discovery [33]. Emerging research suggests that marine-derived toxins may provide novel scaffolds for ion channel modulation and cancer therapy, further broadening the application of venom-derived compounds in modern medicine [23].
Additionally, pore-forming toxins (PFTs), including bee melittin, sticholysins from sea anemones, and potent porins from cubozoan jellyfish, represent a critical component of many venoms [34]. PFTs function by forming transmembrane pores, leading to cell lysis or modulation of cellular ion gradients. While their direct therapeutic application may be limited due to cytotoxicity, PFTs are increasingly being investigated as molecular tools for targeted drug delivery and cancer therapy, and as models for synthetic pore-forming agents.
Overall, the structural and functional diversity of venom components across taxa underscores their immense potential as molecular scaffolds for the development of novel therapeutics [10,30].

3. Mechanisms of Action of Venom Components

The biological activity of venom-derived compounds is primarily mediated through their specific interactions with cellular targets, including ion channels, membrane receptors, enzymes, and components of the extracellular matrix. These interactions are often characterized by high affinity and selectivity, enabling venom components to serve as powerful modulators of physiological pathways [35,36] (Figure 1).
Neurotoxins are a prominent class of venom components that target voltage-gated sodium (NaV), potassium (KV), and calcium (CaV) channels. For instance, α-scorpion toxins and sea anemone peptides delay the inactivation of NaV channels, resulting in prolonged neuronal depolarization and paralysis [37,38]. Conversely, spider toxins such as ω-agatoxins and µ-agatoxins can inhibit CaV and NaV channels, disrupting synaptic transmission [39]. These highly specific modulators have been extensively used in neurophysiological research and are being explored as templates for drug development in epilepsy, pain, and cardiac arrhythmias [40,41].
Another well-characterized mechanism is the inhibition of angiotensin-converting enzyme (ACE) by bradykinin-potentiating peptides (BPPs) from viper venoms. BPPs inhibit the conversion of angiotensin I to angiotensin II, thereby reducing vasoconstriction and lowering blood pressure—a mechanism that led to the development of captopril and other ACE inhibitors [42,43]. Similarly, conotoxins from cone snail venom act on nicotinic acetylcholine receptors and N-type CaV channels, offering therapeutic value in neuropathic pain and spasticity [44,45].
The schematic in Figure 1 illustrates the translational pipeline for venom-derived compounds into modern therapeutics. The process includes the following: (i) toxin sourcing; (ii) determination of biological activity (e.g., ion channels, receptors, enzymes); (iii) molecular structure identification (peptides, enzymes, toxins); (iv) toxin production; (v) pharmacodynamic evaluation; and (vi) clinical applications (e.g., pain, hypertension, cancer, thrombosis).
Venom metalloproteinases and serine proteases from viperid snakes degrade components of the extracellular matrix, disrupt basement membranes, and interfere with blood coagulation pathways. These enzymes exert both hemorrhagic and anticoagulant effects, with potential applications in thrombolytic therapy and cancer metastasis inhibition [15,18,26,46]. Disintegrins, another class of venom-derived peptides, bind integrins on cell surfaces and inhibit platelet aggregation, offering antithrombotic potential [47,48]. Disintegrins are a class of low-molecular-weight, cysteine-rich peptides derived from viperid snake venoms. They are characterized by the presence of specific amino acid motifs, such as Arg-Gly-Asp (RGD) or Lys-Gly-Asp (KGD), which enable them to bind selectively to integrins on cell surfaces. By targeting integrins like αIIbβ3 on platelets, disintegrins inhibit fibrinogen binding, thereby preventing platelet aggregation and offering potential as antithrombotic agents. The therapeutic potential of disintegrins has been harnessed in the development of antiplatelet drugs [47]. For instance, tirofiban (Aggrastat) is a nonpeptide glycoprotein IIb/IIIa inhibitor inspired by the RGD motif found in snake venom disintegrins. Tirofiban competes with fibrinogen for binding to the αIIbβ3 integrin, effectively inhibiting platelet aggregation and reducing the risk of thrombotic cardiovascular events.
Ion channel modulation is also a key mechanism underlying analgesic and neuroprotective activities. For example, the conotoxin MVIIA (ziconotide) selectively blocks N-type CaV channels, reducing neurotransmitter release and thereby alleviating chronic pain [45,49]. Likewise, spider venom peptides that target acid-sensing ion channels (ASICs) and transient receptor potential (TRP) channels are under investigation for treating pain and inflammatory conditions [50,51].
Overall, the specificity and potency of venom components in modulating key physiological pathways underscore their therapeutic relevance. By acting on defined molecular targets, venom-derived molecules provide invaluable templates for drug discovery and a deeper understanding of pathophysiological mechanisms [52].

4. Therapeutic Applications of Venom-Derived Compounds

Venom-derived compounds have shown substantial promise in the development of novel therapeutics, owing to their unique ability to modulate specific molecular targets with high affinity and selectivity [53]. Several peptide- and protein-based drugs originating from animal venoms have been approved for clinical use, while many others are in various stages of preclinical or clinical development (Table 1) [30,54,55]. These therapeutics span a broad range of medical applications, including cardiovascular diseases, chronic pain, diabetes, and cancer.
One of the most successful examples is captopril, the first angiotensin-converting enzyme (ACE) inhibitor approved for the treatment of hypertension and heart failure. It was developed based on bradykinin-potentiating peptides derived from the venom of Bothrops jararaca [4,56]. This breakthrough not only validated venom-derived peptides as drug candidates but also opened the door for further exploration of cardiovascular applications. Other notable ACE inhibitors, such as enalapril and lisinopril, were developed as synthetic analogs of captopril, demonstrating how venom-based scaffolds can serve as starting points for rational drug design [57,58].
In the area of pain management, ziconotide (Prialt®) represents a landmark achievement. This synthetic analog of ω-conotoxin MVIIA, derived from Conus magus, selectively inhibits N-type calcium channels and is approved for the treatment of severe chronic pain in patients who are refractory to opioids [29,49]. Unlike opioids, ziconotide does not act via opioid receptors and is not associated with respiratory depression or dependence, making it a valuable alternative in pain pharmacotherapy [59,60].
Venoms have also contributed to the development of antiplatelet agents. Tirofiban and eptifibatide, which are based on disintegrins found in viper venoms, act as glycoprotein IIb/IIIa inhibitors that prevent platelet aggregation, and are used to manage acute coronary syndromes and during percutaneous coronary interventions [61,62]. These agents exemplify how venom proteins that naturally interfere with hemostasis can be adapted for cardiovascular therapeutics.
Exenatide (Byetta®), a synthetic form of exendin-4 originally derived from the venom of the Gila monster (Heloderma suspectum), is a GLP-1 receptor agonist approved for the treatment of type 2 diabetes mellitus. It improves glycemic control by enhancing glucose-dependent insulin secretion, suppressing inappropriate glucagon release, and delaying gastric emptying. Additionally, exenatide has been shown to promote modest weight loss due to its appetite-suppressing effects, contributing to its clinical utility in patients with type 2 diabetes who are overweight or obese [63]. Tirzepatide (Mounjaro™) is a novel synthetic dual agonist of the GLP-1 and glucose-dependent insulinotropic polypeptide (GIP) receptors, often referred to as a “twincretin.” It has been recently approved for the treatment of type 2 diabetes and has demonstrated significant efficacy in improving glycemic control while inducing substantial weight loss in clinical trials. Tirzepatide enhances glucose-dependent insulin secretion, reduces glucagon levels, delays gastric emptying, and improves insulin sensitivity, making it a promising therapeutic option for managing both hyperglycemia and obesity in patients with type 2 diabetes. These venom-derived and venom-inspired therapeutic agents highlight the translational potential of peptide toxins and their analogs in addressing metabolic disorders, demonstrating how bioactive components from animal venoms can inform the development of effective treatments for chronic diseases [64].
Beyond cardiovascular and pain indications, venom-derived peptides are being actively investigated for oncological applications. For instance, chlorotoxin, a peptide from scorpion venom, has demonstrated the ability to bind selectively to matrix metalloproteinase-2 (MMP-2), a protein overexpressed in glioma cells. This specificity has enabled its use in tumor imaging and as a delivery vehicle for targeted therapies [65,66,67]. Likewise, melittin, the principal peptide in bee venom, exhibits anticancer properties through induction of apoptosis, inhibition of angiogenesis, and disruption of cancer cell membranes, although its non-specific cytotoxicity remains a limitation [68,69]. Additionally, melittin has been evaluated for pain management in rheumatoid arthritis, with clinical studies indicating reductions in pain and inflammation [70]. In fact, bee venom has long been used for this purpose in Oriental medicine, particularly in Korean traditional medicine as a form of pharmacopuncture called bee venom acupuncture (BVA).
Recent studies have also explored the antimicrobial and antiparasitic potential of venom peptides. Certain defensin-like peptides from spider and scorpion venoms exhibit broad-spectrum antibacterial and antifungal activity, possibly by disrupting microbial membranes [71,72]. Additionally, peptides from cone snail and wasp venoms have demonstrated inhibitory effects against protozoan parasites such as Plasmodium and Trypanosoma, suggesting utility in tropical infectious diseases [73,74].
The therapeutic versatility of venom-derived compounds is being expanded further through synthetic biology and peptidomimetic design, enabling the enhancement of their pharmacokinetic properties and specificity [30,75]. These efforts are critical for overcoming challenges related to peptide stability, immunogenicity, and delivery, thereby increasing the clinical viability of venom-based drug candidates. These examples underscore the translational potential of venom-derived compounds from bench to bedside, highlighting their impact in addressing conditions such as chronic pain, hypertension, diabetes, cancer pain, and inflammatory disorders.
Table 1. Peptidomimetic venom drugs and candidates in clinical use or development.
Table 1. Peptidomimetic venom drugs and candidates in clinical use or development.
Drug NameSource ToxinOrigin AnimalTarget/MechanismIndicationClinical StatusRefs.
CaptoprilBradykinin-potentiating peptideBothrops jararaca (Viper)ACE inhibitorHypertension, heart failureFDA-Approved[4,56,57,58,76]
Ziconotide (Prialt®)ω-Conotoxin MVIIAConus magus (Cone snail)N-type Ca2+ channel blockerChronic neuropathic painFDA-Approved[45,59,60]
Eptifibatide (Integrilin®)Barbourin analog (Disintegrin)Sistrurus miliarius (Viper)GPIIb/IIIa inhibitorACS, PCIFDA-Approved[61,62]
Tirofiban (Aggrastat®)Echistatin analogEchis carinatus (Saw-scaled viper)GPIIb/IIIa antagonistCoronary artery diseaseFDA-Approved[77]
Exenatide (Byetta®)Exendin-4Heloderma suspectum (Gila monster)GLP-1 receptor AgonistType 2 diabetes,
weight loss
FDA-Approved
(tirzepatide)
[63,64]
DesmoteplaseDSPA α1Desmodus rotundus (Vampire bat)Clot-specific plasminogen activatorIschemic stroke Completed Phase II[78]
Chlorotoxin-based imaging agents (e.g., BLZ-100)ChlorotoxinLeiurus quinquestriatus (Scorpion)MMP-2 bindingGlioma imagingPhase I/II[65,66,67]
Dalazatide (ShK-186)ShK toxinStichodactyla helianthus (Sea anemone)Kv1.3 channel blockerPsoriasis, MS Completed Phase I[79]
Heparin mimetics (e.g., Ancrod)AncrodCalloselasma rhodostoma (Pit viper)Defibrinogenating agentStroke, thrombosisWithdrawn after Phase III[80]
Huwentoxin-IV analogsHuwentoxin-IVOrnithoctonus huwena (Spider)Nav1.7 sodium channel blockerAnalgesiaPreclinical[81]
D-Melittin–nanoparticle conjugatesMelittinApis mellifera (Bee)Membrane disruptionCancerPreclinical[68,82]
Apimol/ApitoxinBee venom extract containing melittinApis mellifera (Bee)Anti-inflammatory and analgesic pathwayRheumatoid arthritis painPhase II (clinical trials, bee venom extract containing melittin)[70]

5. Technological Advances Facilitating Venom-Based Drug Discovery

The successful translation of venom components into therapeutic agents has been significantly accelerated by recent technological advancements across various scientific disciplines. Innovations in high-throughput screening, transcriptomics, proteomics, and structural biology have enabled comprehensive profiling of venom composition and function, thereby transforming venom research from a descriptive to a mechanistic science [7,83,84].
High-throughput sequencing technologies, particularly next-generation sequencing (NGS), have facilitated the generation of venom gland transcriptomes from a wide array of venomous species. For example, a recent study on Bothrops asper and Bothrops jararaca performed de novo transcriptome assembly and bioinformatic analysis [85], identifying not only highly expressed toxin families (e.g., SVMPs, SVSPs, PLA2s, CTLs) but also novel low-abundance proteins such as arylsulfatases and glycerophosphodiester phosphodiesterases, which may contribute to venom variability and present opportunities for antivenom development and drug discovery. This approach has allowed for the rapid identification of novel toxin-coding genes and isoforms, greatly expanding the catalog of potential bioactive peptides [86,87]. Complementary proteomic methods such as liquid chromatography–mass spectrometry (LC-MS/MS) provide detailed information on the molecular weight, sequence, and post-translational modifications of venom proteins, enabling the characterization of mature venom components with functional relevance [84].
Advances in bioinformatics and molecular modeling have also played a pivotal role in venom-based drug discovery. In silico techniques such as molecular docking, molecular dynamics (MD) simulations, and quantitative structure–activity relationship (QSAR) modeling are routinely employed to predict binding affinities, optimize pharmacophores, and identify lead compounds prior to synthesis and in vitro validation [88,89,90]. These tools not only reduce the cost and time associated with experimental assays but also provide mechanistic insights that inform rational drug design. For example, a recent pharmacoinformatic study on plant-derived phytochemicals identified potential inhibitors of the venom metalloproteinase Atrolysin by employing high-throughput virtual screening, molecular docking, MD simulations, and density functional theory (DFT) calculations [83]. This integrative approach enabled the identification of compounds with strong binding affinities (−8.7 to −10.2 kcal/mol) and stable dynamic interactions with the target, confirmed by low RMSD values and consistent hydrogen bond formations throughout 100 ns MD simulations. The study further employed DFT calculations to analyze HOMO-LUMO energy gaps, indicating favorable electronic properties for protein interactions, and MMPBSA analysis to validate binding energies, providing a robust pipeline that can guide venom-targeted drug development while reducing the need for early-stage experimental screening.
Synthetic biology and peptide engineering technologies have further enhanced the therapeutic potential of venom-derived compounds. Through recombinant expression systems, including bacterial, yeast, insect, and mammalian cells, researchers can produce venom peptides in large quantities with defined structural fidelity [91]. Chemical synthesis and cyclization strategies allow for the generation of peptide analogs with improved stability, bioavailability, and target selectivity. Moreover, site-directed mutagenesis and incorporation of non-natural amino acids have facilitated structure–activity relationship studies and the development of peptidomimetics with desirable pharmacokinetic properties [92,93].
Structural biology techniques, including X-ray crystallography, cryo-electron microscopy (cryo-EM), and nuclear magnetic resonance (NMR) spectroscopy, have provided atomic-level insights into toxin–target interactions. These approaches enable the elucidation of binding modes and conformational dynamics, which are critical for the rational design of therapeutic analogs [94,95]. Coupled with data from functional assays and pharmacological profiling, these technologies collectively support the identification and optimization of venom-derived drug candidates.
The integration of these advanced methodologies has ushered in a new era of venom research, allowing for systematic and targeted exploration of the vast chemical diversity encoded within animal venoms. As a result, the discovery and development of venom-based therapeutics are becoming increasingly efficient, scalable, and precise, paving the way for next-generation biologics and peptide drugs [96,97].

6. Venom-Derived Therapeutics for Unmet Medical Needs

Although there have been significant advancements in modern therapeutics, current treatment options for conditions such as chronic pain, hypertension, and certain cancers often face limitations, including insufficient efficacy, development of resistance, and adverse side effects that reduce patient compliance [98]. For instance, opioid-based analgesics, while effective, are associated with the risk of dependence and tolerance, while conventional antihypertensives may be inadequate in refractory cases [99]. In this context, venom-derived compounds offer unique advantages, such as high potency at low doses, novel mechanisms of action targeting specific ion channels and receptors, and the potential to overcome limitations associated with current therapeutic agents [30]. Thus, exploring animal venoms as a source for developing new therapeutics can address unmet medical needs in specific disorders where current treatments are suboptimal (Table 2).

7. Challenges and Future Directions

Despite the promising therapeutic potential of venom-derived compounds, several challenges hinder their seamless translation into clinically approved drugs. One of the foremost obstacles is the complexity and variability of venom composition. Venoms are often composed of hundreds of peptides and proteins, the expression profiles of which can vary significantly depending on species, age, diet, geographical distribution, and even individual physiological states [100,101]. This complexity not only complicates the standardization of venom-derived products but also presents difficulties in reproducibility and scalability for pharmaceutical development.
A related issue is the difficulty in isolating and characterizing low-abundance bioactive components from natural venoms. Even with advanced omics tools, distinguishing pharmacologically relevant molecules from a background of enzymatic or toxic proteins remains a substantial task. Additionally, venom peptides often suffer from poor stability, a short half-life, and low oral bioavailability, limiting their practical application unless chemically modified or delivered via advanced drug delivery systems [30,102].
The issue of immunogenicity is also non-negligible. As exogenous proteins and peptides, venom components can elicit unwanted immune responses, including hypersensitivity and anaphylaxis, particularly upon repeated administration [103,104]. Strategies such as pegylation, cyclization, and nanoparticle encapsulation are being explored to mitigate such risks, but these modifications may alter pharmacodynamics and require extensive validation [105].
Furthermore, the ethical and ecological considerations associated with venom harvesting pose additional constraints. Many venomous species are endangered or difficult to culture in captivity, raising concerns about sustainability and biodiversity loss [106]. Recombinant expression systems and synthetic biology approaches offer alternatives, but they require significant optimization to yield functionally active peptides that mimic natural counterparts in structure and activity [107].
From a regulatory perspective, venom-derived compounds often fall into a gray zone between biologics and small molecules, complicating approval processes. The need for rigorous preclinical and clinical studies, combined with high production costs and uncertain intellectual property pathways, may deter commercial investment despite scientific promise [8].
Looking ahead, future directions in venom-based drug discovery will likely be shaped by the integration of artificial intelligence (AI), machine learning, and systems biology. These approaches can accelerate candidate identification, predict toxicity, and guide structural optimization by mining large-scale datasets of venom sequences, structures, and activities [108]. Moreover, personalized medicine frameworks may allow for the tailoring of venom-based therapies to individual genetic and physiological profiles, particularly in complex diseases such as cancer or autoimmune disorders [109].
Collaborative efforts that bridge academic research, biotechnology, and clinical practice will be essential in advancing venom-derived compounds from bench to bedside. The continued expansion of venom databases, combined with robust validation platforms and translational infrastructure, will determine the success of this promising but still emerging therapeutic frontier.

8. Conclusions

Animal venoms represent a vast and largely untapped resource for the discovery of novel therapeutic agents. With their remarkable biochemical diversity and target specificity, venom-derived peptides and proteins have already led to the development of clinically important drugs such as captopril, ziconotide, and tirofiban. Recent advancements in omics technologies, bioinformatics, structural biology, and synthetic biology have transformed venom research into a systematic and translationally oriented field. Despite ongoing challenges related to bioavailability, immunogenicity, and ecological sustainability, continuous innovation in delivery systems and peptide engineering holds promise for overcoming these barriers. Looking forward, the integration of artificial intelligence and personalized medicine strategies may further unlock the therapeutic potential of venom-based compounds. As the scientific community continues to explore and harness this biological arsenal, venom-based drug discovery is poised to contribute significantly to the future of precision medicine. Future research on venom-based therapeutics should focus on addressing the limitations of existing treatments by leveraging the unique mechanisms and high specificity of venom-derived compounds for specific clinical applications.

Author Contributions

Conceptualization, E.K.; original draft preparation, D.H.H., R.L.M.P., R.D.A., C.K. and E.K. writing—review and editing, D.H.H., R.L.M.P., R.D.A., H.L., Y.H., A.M., R.S., C.K. and E.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2021R1I1A306005711 and RS-2023-00276023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to acknowledge the grant of the National Research Foundation of the Korean government.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Schendel, V.; Rash, L.D.; Jenner, R.A.; Undheim, E.A.B. The Diversity of Venom: The Importance of Behavior and Venom System Morphology in Understanding Its Ecology and Evolution. Toxins 2019, 11, 666. [Google Scholar] [CrossRef]
  2. Chen, N.; Xu, S.; Zhang, Y.; Wang, F. Animal protein toxins: Origins and therapeutic applications. Biophys. Rep. 2018, 4, 233–242. [Google Scholar] [CrossRef] [PubMed]
  3. Lewis, R.J.; Garcia, M.L. Therapeutic potential of venom peptides. Nat. Rev. Drug Discov. 2003, 2, 790–802. [Google Scholar] [CrossRef]
  4. Smith, C.G.; Vane, J.R. The discovery of captopril. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2003, 17, 788–789. [Google Scholar] [CrossRef]
  5. Cushman, D.W.; Ondetti, M.A. Design of angiotensin converting enzyme inhibitors. Nat. Med. 1999, 5, 1110–1113. [Google Scholar] [CrossRef] [PubMed]
  6. Utkin, Y.N. Modern trends in animal venom research—Omics and nanomaterials. World J. Biol. Chem. 2017, 8, 4–12. [Google Scholar] [CrossRef] [PubMed]
  7. Modahl, C.M.; Brahma, R.K.; Koh, C.Y.; Shioi, N.; Kini, R.M. Omics Technologies for Profiling Toxin Diversity and Evolution in Snake Venom: Impacts on the Discovery of Therapeutic and Diagnostic Agents. Annu. Rev. Anim. Biosci. 2020, 8, 91–116. [Google Scholar] [CrossRef] [PubMed]
  8. Bordon, K.C.F.; Cologna, C.T.; Fornari-Baldo, E.C.; Pinheiro-Júnior, E.L.; Cerni, F.A.; Amorim, F.G.; Anjolette, F.A.P.; Cordeiro, F.A.; Wiezel, G.A.; Cardoso, I.A.; et al. From Animal Poisons and Venoms to Medicines: Achievements, Challenges, and Perspectives in Drug Discovery. Front. Pharmacol. 2020, 11, 1132. [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. Vidya, V.; Achar, R.R.; Himathi, M.U.; Akshita, N.; Kameshwar, V.H.; Byrappa, K.; Ramadas, D. Venom peptides—A comprehensive translational perspective in pain management. Curr. Res. Toxicol. 2021, 2, 329–340. [Google Scholar] [CrossRef]
  11. Ageitos, L.; Torres, M.D.; de la Fuente-Nunez, C. Biologically Active Peptides from Venoms: Applications in Antibiotic Resistance, Cancer, and Beyond. Int. J. Mol. Sci. 2022, 23, 15437. [Google Scholar] [CrossRef]
  12. Moga, M.A.; Dimienescu, O.G.; Arvătescu, C.A.; Ifteni, P.; Pleş, L. Anticancer Activity of Toxins from Bee and Snake Venom—An Overview on Ovarian Cancer. Molecules 2018, 23, 692. [Google Scholar] [CrossRef]
  13. 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]
  14. Kalia, J.; Milescu, M.; Salvatierra, J.; Wagner, J.; Klint, J.K.; King, G.F.; Olivera, B.M.; Bosmans, F. From foe to friend: Using animal toxins to investigate ion channel function. J. Mol. Biol. 2015, 427, 158–175. [Google Scholar] [CrossRef]
  15. Oliveira, A.L.; Viegas, M.F.; da Silva, S.L.; Soares, A.M.; Ramos, M.J.; Fernandes, P.A. The chemistry of snake venom and its medicinal potential. Nat. Rev. Chem. 2022, 6, 451–469. [Google Scholar] [CrossRef] [PubMed]
  16. Michálek, O.; King, G.F.; Pekár, S. Prey specificity of predatory venoms. Biol. Rev. Camb. Philos. Soc. 2024, 99, 2253–2273. [Google Scholar] [CrossRef]
  17. Gutiérrez, J.M.; Escalante, T.; Rucavado, A. Experimental Pathophysiology of Systemic Alterations Induced by Bothrops asper Snake Venom. Toxicon 2009, 54, 976–987. [Google Scholar] [CrossRef]
  18. Kini, R.M.; Koh, C.Y. Metalloproteases Affecting Blood Coagulation, Fibrinolysis and Platelet Aggregation from Snake Venoms: Definition and Nomenclature of Interaction Sites. Toxins 2016, 8, 284. [Google Scholar] [CrossRef]
  19. Serrano, S.M.T. The Long Road of Research on Snake Venom Serine Proteinases. Toxicon 2013, 62, 19–26. [Google Scholar] [CrossRef]
  20. AlShammari, A.K.; Abd El-Aziz, T.M.; Al-Sabi, A. Snake Venom: A Promising Source of Neurotoxins Targeting Voltage-Gated Potassium Channels. Toxins 2023, 16, 12. [Google Scholar] [CrossRef]
  21. Kalita, B.; Utkin, Y.N.; Mukherjee, A.K. Current Insights in the Mechanisms of Cobra Venom Cytotoxins and Their Complexes in Inducing Toxicity: Implications in Antivenom Therapy. Toxins 2022, 14, 839. [Google Scholar] [CrossRef]
  22. Aird, S.D.; Aggarwal, S.; Villar-Briones, A.; Tin, M.M.; Terada, K.; Mikheyev, A.S. Snake venoms are integrated systems, but abundant venom proteins evolve more rapidly. BMC Genom. 2015, 16, 647. [Google Scholar] [CrossRef]
  23. Modahl, C.M.; Mrinalini Frietze, S.; Mackessy, S.P. Adaptive evolution of distinct prey-specific toxin genes in rear-fanged snake venom. Proc. Biol. Sci. 2018, 285, 20181003. [Google Scholar] [CrossRef]
  24. Utkin, Y.N. Three-finger toxins, a deadly weapon of elapid venom—Milestones of discovery. Toxicon 2013, 62, 50–55. [Google Scholar] [CrossRef]
  25. Harvey, A.L. Twenty years of dendrotoxins. Toxicon 2001, 39, 15–26. [Google Scholar] [CrossRef]
  26. Gutiérrez, J.M.; Escalante, T.; Rucavado, A.; Herrera, C. Hemorrhage caused by snake venom metalloproteinases: A journey of discovery and understanding. Toxins 2016, 8, 93. [Google Scholar] [CrossRef]
  27. Kini, R.M. Serine proteases affecting blood coagulation and fibrinolysis from snake venoms. Pathophysiol. Haemost. Thromb. 2005, 34, 200–204. [Google Scholar] [CrossRef]
  28. Lebbe, E.K.M.; Tytgat, J. In the picture: Disulfide-poor conopeptides, a class of pharmacologically interesting compounds. J. Venom. Anim. Toxins Incl. Trop. Dis. 2016, 22, 30. [Google Scholar] [CrossRef]
  29. McGivern, J.G. Ziconotide: A review of its pharmacology and use in the treatment of pain. Neuropsychiatr. Dis. Treat. 2007, 3, 69–85. [Google Scholar] [CrossRef]
  30. King, G.F. Venoms as a platform for human drugs: Translating toxins into therapeutics. Expert Opin. Biol. Ther. 2011, 11, 1469–1484. [Google Scholar] [CrossRef]
  31. Mariottini, G.L.; Pane, L. Cytotoxic and Cytolytic Cnidarian Venoms. A Review on Health Implications and Possible Therapeutic Applications. Toxins 2013, 6, 108–151. [Google Scholar] [CrossRef]
  32. Lee, H.; Jung, E.-S.; Kang, C.; Yoon, W.D.; Kim, J.-S.; Kim, E. Scyphozoan jellyfish venom metalloproteinases and their role in the cytotoxicity. Toxicon 2011, 58, 277–284. [Google Scholar] [CrossRef]
  33. Xie, B.; Huang, Y.; Baumann, K.; Fry, B.G.; Shi, Q. From Marine Venoms to Drugs: Efficiently Supported by a Combination of Transcriptomics and Proteomics. Mar. Drugs 2017, 15, 103. [Google Scholar] [CrossRef]
  34. Anderluh, G.; Macek, P. Dissecting the actinoporin pore-forming mechanism. Structure 2003, 11, 1312–1313. [Google Scholar] [CrossRef]
  35. Bittenbinder, M.A.; van Thiel, J.; Cardoso, F.C.; Casewell, N.R.; Gutiérrez, J.M.; Kool, J.; Vonk, F.J. Tissue damaging toxins in snake venoms: Mechanisms of action, pathophysiology and treatment strategies. Commun. Biol. 2024, 7, 358. [Google Scholar] [CrossRef]
  36. Utkin, Y.N. Animal Venoms and Their Components: Molecular Mechanisms of Action. Toxins 2021, 13, 415. [Google Scholar] [CrossRef]
  37. Bosmans, F.; Tytgat, J. Voltage-gated sodium channel modulation by scorpion α-toxins. Toxicon 2007, 49, 142–158. [Google Scholar] [CrossRef]
  38. Moran, Y.; Gordon, D.; Gurevitz, M. Sea anemone toxins affecting voltage-gated sodium channels—Molecular and evolutionary features. Toxicon 2009, 54, 1089–1101. [Google Scholar] [CrossRef]
  39. Adams, M.E. Agatoxins: Ion channel specific toxins from the American funnel web spider, Agelenopsis aperta. Toxicon 2004, 43, 509–525. [Google Scholar] [CrossRef]
  40. Chow, C.Y.; Absalom, N.; Biggs, K.; King, G.F.; Ma, L. Venom-derived modulators of epilepsy-related ion channels. Biochem. Pharmacol. 2020, 181, 114043. [Google Scholar] [CrossRef]
  41. Souza, A.C.N.; Binda, N.S.; Almeida, H.Y.; de Castro Jünior, C.J.; Gomez, M.V.; Ribeiro, F.M.; Da Silva, J.F. Ion Channels-related Neuroprotection and Analgesia Mediated by Spider Venom Peptides. Curr. Protein Pept. Sci. 2023, 24, 365–379. [Google Scholar] [CrossRef]
  42. Ferreira, S.H. A bradykinin-potentiating factor (BPF) present in the venom of Bothrops jararaca. Br. J. Pharmacol. Chemother. 1965, 24, 163–169. [Google Scholar] [CrossRef]
  43. Ondetti, M.A.; Rubin, B.; Cushman, D.W. Design of specific inhibitors of angiotensin-converting enzyme: New class of orally active antihypertensive agents. Science 1977, 196, 441–444. [Google Scholar] [CrossRef]
  44. Satkunanathan, N.; Livett, B.; Gayler, K.; Sandall, D.; Down, J.; Khalil, Z. Alpha-conotoxin Vc1.1 alleviates neuropathic pain and accelerates functional recovery of injured neurons. Brain Res. 2005, 1059, 149–158. [Google Scholar] [CrossRef]
  45. Schmidtko, A.; Lötsch, J.; Freynhagen, R.; Geisslinger, G. Ziconotide for treatment of severe chronic pain. Lancet 2010, 375, 1569–1577. [Google Scholar] [CrossRef]
  46. Calvete, J.J. Snake venomics: From the inventory of toxins to biology. Toxicon 2013, 75, 44–62. [Google Scholar] [CrossRef]
  47. Lazarovici, P.; Marcinkiewicz, C.; Lelkes, P.I. From Snake Venom’s Disintegrins and C-Type Lectins to Anti-Platelet Drugs. Toxins 2019, 11, 303. [Google Scholar] [CrossRef]
  48. McLane, M.A.; Sanchez, E.E.; Wong, A.; Paquette-Straub, C.; Perez, J.C. Disintegrins. Current Drug Targets. Cardiovasc. Hematol. Disord. 2004, 4, 327–355. [Google Scholar] [CrossRef]
  49. Williams, J.A.; Day, M.; Heavner, J.E. Ziconotide: An update and review. Expert Opin. Pharmacother. 2008, 9, 1575–1583. [Google Scholar] [CrossRef]
  50. Cardoso, F.C.; Lewis, R.J. Structure-Function and Therapeutic Potential of Spider Venom-Derived Cysteine Knot Peptides Targeting Sodium Channels. Front. Pharmacol. 2019, 10, 366. [Google Scholar] [CrossRef]
  51. King, G.F.; Vetter, I. No gain, no pain: NaV1.7 as an analgesic target. ACS Chem. Neurosci. 2014, 5, 749–751. [Google Scholar] [CrossRef]
  52. Harvey, A.L. Toxins and drug discovery. Toxicon 2014, 92, 193–200. [Google Scholar] [CrossRef]
  53. Kini, R.M.; Koh, C.Y. Snake venom three-finger toxins and their potential in drug development targeting cardiovascular diseases. Biochem. Pharmacol. 2020, 181, 114105. [Google Scholar] [CrossRef]
  54. Munawar, A.; Ali, S.A.; Akrem, A.; Betzel, C. Snake Venom Peptides: Tools of Biodiscovery. Toxins 2018, 10, 474. [Google Scholar] [CrossRef]
  55. Sadeghi, M.; McArthur, J.R.; Finol-Urdaneta, R.K.; Adams, D.J. Analgesic conopeptides targeting G protein-coupled receptors reduce excitability of sensory neurons. Neuropharmacology 2017, 127, 116–123. [Google Scholar] [CrossRef]
  56. Cushman, D.W.; Ondetti, M.A. History of the design of captopril and related inhibitors of angiotensin converting enzyme. Hypertension 1991, 17, 589–592. [Google Scholar] [CrossRef]
  57. Bader, M.; Ganten, D. Update on tissue renin-angiotensin systems. J. Mol. Med. 2008, 86, 615–621. [Google Scholar] [CrossRef]
  58. Rella, M.; Rushworth, C.A.; Guy, J.L.; Turner, A.J.; Langer, T.; Jackson, R.M. Structure-based pharmacophore design and virtual screening for novel angiotensin converting enzyme 2 inhibitors. J. Chem. Inf. Model. 2006, 46, 708–716. [Google Scholar] [CrossRef]
  59. Staats, P.S.; Yearwood, T.; Charapata, S.G.; Presley, R.W.; Wallace, M.S.; Byas-Smith, M.; Fisher, R.; Bryce, D.A.; Mangieri, E.A.; Luther, R.R.; et al. Intrathecal ziconotide in the treatment of refractory pain in patients with cancer or AIDS: A randomized controlled trial. JAMA 2004, 291, 63–70. [Google Scholar] [CrossRef]
  60. Wallace, M.S.; Rauck, R.; Fisher, R.; Charapata, S.G.; Ellis, D.; Dissanayake, S.; Ziconotide 98-022 Study Group. Intrathecal ziconotide for severe chronic pain: Safety and tolerability results of an open-label, long-term trial. Anesth. Analg. 2008, 106, 628–637. [Google Scholar] [CrossRef]
  61. Scarborough, R.M.; Naughton, M.A.; Teng, W.; Rose, J.W.; Phillips, D.R.; Nannizzi, L.; Arfsten, A.; Campbell, A.M.; Charo, I.F. Design of potent and specific integrin antagonists: Peptide antagonists with high specificity for glycoprotein IIb/IIIa. J. Biol. Chem. 1993, 268, 1066–1073. [Google Scholar] [CrossRef]
  62. Estevez, B.; Shen, B.; Du, X. Targeting integrin and integrin signaling in treating thrombosis. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 24–29. [Google Scholar] [CrossRef]
  63. DeMarsilis, A.; Reddy, N.; Boutari, C.; Filippaios, A.; Sternthal, E.; Katsiki, N.; Mantzoros, C. Pharmacotherapy of type 2 diabetes: An update and future directions. Metabolism 2022, 137, 155332. [Google Scholar] [CrossRef]
  64. Dahl, D.; Onishi, Y.; Norwood, P.; Huh, R.; Bray, R.; Patel, H.; Rodríguez, Á. Effect of Subcutaneous Tirzepatide vs Placebo Added to Titrated Insulin Glargine on Glycemic Control in Patients with Type 2 Diabetes. JAMA 2022, 327, 534–545. [Google Scholar] [CrossRef] [PubMed]
  65. 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] [PubMed]
  66. 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]
  67. Wiranowska, M. Advances in the use of chitosan and chlorotoxin—Functionalized chitosan polymers in drug delivery and detection of glioma—A review. Carbohydr. Polym. Technol. Appl. 2024, 7, 100427. [Google Scholar] [CrossRef]
  68. Yu, X.; Jia, S.; Yu, S.; Chen, Y.; Zhang, C.; Chen, H.; Dai, Y. Recent advances in melittin-based nanoparticles for antitumor treatment: From mechanisms to targeted delivery strategies. J. Nanobiotechnol. 2023, 21, 454. [Google Scholar] [CrossRef]
  69. Oršolić, N. Bee venom in cancer therapy. Cancer Metastasis Rev. 2012, 31, 173–194. [Google Scholar] [CrossRef]
  70. Son, D.J.; Lee, J.W.; Lee, Y.H.; Song, H.S.; Lee, C.K.; Hong, J.T. Therapeutic application of anti-arthritis, pain-releasing, and anti-cancer effects of bee venom and its constituent compounds. Pharmacol. Ther. 2007, 115, 246–270. [Google Scholar] [CrossRef]
  71. Ayroza, G.; Ferreira, I.L.C.; Sayegh, R.S.R.; Tashima, A.K.; da Silva Junior, P.I. Juruin: An antifungal peptide from the venom of the Amazonian Pink Toe spider, Avicularia juruensis, which contains the inhibitory cystine knot motif. Front. Microbiol. 2012, 3, 324. [Google Scholar] [CrossRef] [PubMed]
  72. Cesa-Luna, C.; Muñoz-Rojas, J.; Saab-Rincon, G.; Baez, A. Structural characterization of scorpion peptides and their bactericidal activity against clinical isolates of multidrug-resistant bacteria. PLoS ONE 2019, 14, e0222438. [Google Scholar] [CrossRef] [PubMed]
  73. Vinhote, J.F.C.; Lima, D.B.; Mello, C.P.; de Souza, B.M.; Havt, A.; Palma, M.S.; Martins, A.M.C. Trypanocidal activity of mastoparan from Polybia paulista wasp venom by interaction with TcGAPDH. Toxicon 2017, 137, 168–172. [Google Scholar] [CrossRef]
  74. Akondi, K.B.; Muttenthaler, M.; Dutertre, S.; Kaas, Q.; Craik, D.J.; Lewis, R.J.; Alewood, P.F. Discovery, synthesis and development of structure-activity relationships of Conotoxins. Chem. Rev. 2014, 114, 5815–5847. [Google Scholar] [CrossRef]
  75. Akcan, M.; Craik, D.J. Chapter 11: Engineering Venom Peptides to Improve Their Stability and Bioavailability. In Venoms to Drugs; Series: Drug Discovery Series; Royal Society of Chemistry: Cambridge, UK, 2015; pp. 275–289. [Google Scholar] [CrossRef]
  76. Messadi, E. Snake Venom Components as Therapeutic Drugs in Ischemic Heart Disease. Biomolecules 2023, 13, 1539. [Google Scholar] [CrossRef] [PubMed]
  77. The RESTORE Investigators. Effects of platelet glycoprotein IIb/IIIa blockade with tirofiban on adverse cardiac events in patients with unstable angina or acute myocardial infarction undergoing coronary angioplasty. Circulation 1997, 96, 1445–1453. [Google Scholar] [CrossRef]
  78. Hacke, W.; Albers, G.; Al-Rawi, Y.; Bogousslavsky, J.; Davalos, A.; Eliasziw, M.; Fischer, M.; Furlan, A.; Kaste, M.; Lees, K.R.; et al. The Desmoteplase in Acute Ischemic Stroke Trial (DIAS): A phase II MRI-based 9-hour window acute stroke thrombolysis trial with intravenous desmoteplase. Stroke 2005, 36, 66–73. [Google Scholar] [CrossRef]
  79. Tarcha, E.J.; Olsen, C.M.; Probst, P.; Peckham, D.; Muñoz-Elías, E.J.; Kruger, J.G.; Iadonato, S.P. Safety and pharmacodynamics of dalazatide, a Kv1.3 channel inhibitor, in the treatment of plaque psoriasis: A randomized phase 1b trial. PLoS ONE 2017, 12, e0180762. [Google Scholar] [CrossRef]
  80. Sherman, D.G.; Atkinson, R.P.; Chippendale, T.; Levin, K.A.; Ng, K.; Futrell, N.; Hsu, C.Y.; Levy, D.E. Intravenous ancrod for treatment of acute ischemic stroke: The STAT study: A randomized controlled trial. Stroke Treatment with Ancrod Trial. JAMA 2000, 283, 2395–2403. [Google Scholar] [CrossRef]
  81. Lopez, L.; Montnach, J.; Oliveira-Mendes, B.; Khakh, K.; Thomas, B.; Lin, S.; Caumes, C.; Wesolowski, S.; Nicolas, S.; Servent, D.; et al. Synthetic Analogues of Huwentoxin-IV Spider Peptide with Altered Human NaV1.7/NaV1.6 Selectivity Ratios. Front. Cell Dev. Biol. 2021, 9, 798588. [Google Scholar] [CrossRef]
  82. Lv, S.; Sylvestre, M.; Song, K.; Pun, S.H. Development of D-melittin polymeric nanoparticles for anti-cancer treatment. Biomaterials 2021, 277, 121076. [Google Scholar]
  83. King, G.F. Venoms to Drugs: Venom as a Source for the Development of Human Therapeutics; Royal Society of Chemistry: Cambridge, UK, 2015; ISBN 978-1-84973-663-3. [Google Scholar] [CrossRef]
  84. Slagboom, J.; Derks, R.J.; Sadighi, R.; Somsen, G.W.; Ulens, C.; Casewell, N.R.; Kool, J. High-Throughput Venomics. J. Proteome Res. 2023, 22, 1734–1746. [Google Scholar] [CrossRef]
  85. Espín-Angulo, J.; Vela, D. Exploring the venom gland transcriptome of Bothrops asper and Bothrops jararaca: De novo assembly and analysis of novel toxic proteins. Toxins 2024, 16, 511. [Google Scholar] [CrossRef]
  86. Suryamohan, K.; Krishnankutty, S.P.; Guillory, J.; Jevit, M.; Schröder, 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]
  87. Escoubas, P.; Sollod, B.; King, G.F. Venom landscapes: Mining the complexity of spider venoms via a combined cDNA and mass spectrometric approach. Toxicon 2006, 47, 650–663. [Google Scholar] [CrossRef]
  88. Yang, C.; Chen, E.A.; Zhang, Y. Protein-Ligand Docking in the Machine-Learning Era. Molecules 2022, 27, 4568. [Google Scholar] [CrossRef] [PubMed]
  89. Li, R.; Hasan, M.M.; Wang, D. In Silico Conotoxin Studies: Progress and Prospects. Molecules 2024, 29, 6061. [Google Scholar] [CrossRef] [PubMed]
  90. Ravi, D.A.; Hwang, D.H.; Mohan Prakash, R.L.; Kang, C.; Kim, E. Indian Medicinal Plant-Derived Phytochemicals as Potential Antidotes for Snakebite: A Pharmacoinformatic Study of Atrolysin Inhibitors. Int. J. Mol. Sci. 2024, 25, 12675. [Google Scholar] [CrossRef]
  91. Clement, H.; Flores, V.; Diego-Garcia, E.; Corrales-Garcia, L.; Villegas, E.; Corzo, G. A comparison between the recombinant expression and chemical synthesis of a short cysteine-rich insecticidal spider peptide. J. Venom. Anim. Toxins Incl. Trop. Dis. 2015, 21, 19. [Google Scholar] [CrossRef] [PubMed]
  92. Budisa, N. Expanded genetic code for the engineering of ribosomally and post-translationally modified peptide natural products (RiPPs). Curr. Opin. Biotechnol. 2013, 24, 591–598. [Google Scholar] [CrossRef]
  93. Gordon, D.; Chen, R.; Chung, S.-H. Computational Methods of Studying the Binding of Toxins from Venomous Animals to Biological Ion Channels: Theory and Applications. Physiol. Rev. 2013, 93, 767–802. [Google Scholar] [CrossRef]
  94. Giribaldi, J.; Haufe, Y.; Evans, E.R.J.; Amar, M.; Durner, A.; Schmidt, C.; Faucherre, A.; Moha Ou Maati, H.; Enjalbal, C.; Molgó, J.; et al. Backbone Cyclization Turns a Venom Peptide into a Stable and Equipotent Ligand at Both Muscle and Neuronal Nicotinic Receptors. J. Med. Chem. 2020, 63, 12682–12692. [Google Scholar] [CrossRef]
  95. Sunagar, K.; Undheim, E.A.B.; Scheib, H.; Gren, E.C.; Cochran, C.; Person, C.E.; Koludarov, I.; Kelln, W.; Hayes, W.K.; King, G.F.; et al. Intraspecific venom variation in the medically significant Southern Pacific Rattlesnake (Crotalus oreganus helleri): Biodiscovery, clinical and evolutionary implications. J. Proteom. 2014, 99, 68–83. [Google Scholar] [CrossRef] [PubMed]
  96. Zancolli, G.; Casewell, N.R. Venom systems as models for studying the origin and regulation of evolutionary novelties. Mol. Biol. Evol. 2020, 37, 2777–2790. [Google Scholar] [CrossRef] [PubMed]
  97. Robinson, S.D.; Li, Q.; Bandyopadhyay, P.K.; Gajewiak, J.; Yandell, M.; Papenfuss, A.T.; Purcell, A.W.; Norton, R.S.; Olivera, B.M. Diversity of conotoxin gene superfamilies in the venomous snail Conus victoriae. PLoS ONE 2017, 9, e87648. [Google Scholar] [CrossRef]
  98. Dowell, D.; Tamara, M.; Chou, H.R. CDC Guideline for Prescribing Opioids for Chronic Pain—United States, 2016. JAMA 2016, 315, 1624–1645. [Google Scholar] [CrossRef]
  99. Carey, R.M.; Calhoun, D.A.; Bakris, G.L.; Brook, R.D.; Daugherty, S.L.; Dennison-Himmelfarb, C.R.; Egan, B.M.; Flack, J.M.; Gidding, S.S.; Judd, E.; et al. Resistant Hypertension: Detection, Evaluation, and Management: A Scientific Statement from the American Heart Association. Hypertension 2018, 72, e53–e90. [Google Scholar] [CrossRef] [PubMed]
  100. Tasoulis, T.; Isbister, G.K. A review and database of snake venom proteomes. Toxins 2017, 9, 290. [Google Scholar] [CrossRef]
  101. Casewell, N.R.; Wüster, W.; Vonk, F.J.; Harrison, R.A.; Fry, B.G. Complex cocktails: The evolutionary novelty of venoms. Trends Ecol. Evol. 2013, 28, 219–229. [Google Scholar] [CrossRef]
  102. Dutertre, S.; Jin, A.H.; Vetter, I.; Hamilton, B.; Sunagar, K.; Lavergne, V.; Lewis, R.J. Evolution of separate predation- and defence-evoked venoms in carnivorous cone snails. Nat. Commun. 2014, 5, 3521. [Google Scholar] [CrossRef]
  103. Golden, D.B.K.; Demain, J.G.; Freeman, T.M.; Graft, D.F.; Tankersley, M.S.; Tracy, J.M.; Blessing-Moore, J.; Bernstein, D.; Dinakar, C.; Greenhawt, M.; et al. Stinging insect hypersensitivity: A practice parameter update 2016. Ann. Allergy Asthma Immunol. 2017, 118, 28–54. [Google Scholar] [CrossRef]
  104. Stevens, W.W.; Kraft, M.; Eisenbarth, S.C. Recent Insights into the Mechanisms of Anaphylaxis. Curr. Opin. Immunol. 2023, 81, 102288. [Google Scholar] [CrossRef] [PubMed]
  105. Harris, J.M.; Chess, R.B. Effect of pegylation on pharmaceuticals. Nat. Rev. Drug Discov. 2003, 2, 214–221. [Google Scholar] [CrossRef]
  106. Zamani, A.; Sääksjärvi, E.; Prendini, L. Venom-extraction and exotic pet trade may hasten the extinction of scorpions. Arachnol. Mitteilungen Arachnol. Lett. 2021, 61, 20–23. [Google Scholar] [CrossRef]
  107. Zhang, Z.X.; Nong, F.T.; Wang, Y.Z.; Yan, C.X.; Gu, Y.; Song, P.; Sun, X.M. Strategies for efficient production of recombinant proteins in Escherichia coli: Alleviating the host burden and enhancing protein activity. Microb. Cell Fact. 2022, 21, 191. [Google Scholar] [CrossRef] [PubMed]
  108. Guan, C.; Torres, M.D.T.; Li, S.; de la Fuente-Nunez, C. Venomics AI: A computational exploration of global venoms for antibiotic discovery. bioRxiv 2024. bioRxiv:2024.12.17.628923. [Google Scholar] [CrossRef]
  109. Ma, R.; Mahadevappa, R.; Kwok, H.F. Venom-based peptide therapy: Insights into anti-cancer mechanism. Oncotarget 2017, 8, 100908–100930. [Google Scholar] [CrossRef]
Figure 1. Overview of venom-based drug discovery pipeline.
Figure 1. Overview of venom-based drug discovery pipeline.
Toxins 17 00371 g001
Table 2. Limitations of current therapies and potential advantages of venom-derived therapeutics.
Table 2. Limitations of current therapies and potential advantages of venom-derived therapeutics.
Therapeutic
Area
Limitations of Current TreatmentsAdvantages of Venom-Derived Therapeutics
Chronic PainOpioid tolerance, dependence, limited efficacy in neuropathic painNovel ion channel blockers (e.g., ziconotide) with high specificity
HypertensionRefractory cases, side effects with polytherapyACE inhibitors derived from snake venom (e.g., captopril)
Cancer PainLimited efficacy, opioid-related issuesCrotoxin showing analgesic and potential antitumor effects
DiabetesSuboptimal glycemic control, hypoglycemia riskGLP-1 receptor agonists from Gila monster venom (e.g., exenatide)
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

Kim, E.; Hwang, D.H.; Mohan Prakash, R.L.; Asirvatham, R.D.; Lee, H.; Heo, Y.; Munawir, A.; Seyedian, R.; Kang, C. Animal Venom in Modern Medicine: A Review of Therapeutic Applications. Toxins 2025, 17, 371. https://doi.org/10.3390/toxins17080371

AMA Style

Kim E, Hwang DH, Mohan Prakash RL, Asirvatham RD, Lee H, Heo Y, Munawir A, Seyedian R, Kang C. Animal Venom in Modern Medicine: A Review of Therapeutic Applications. Toxins. 2025; 17(8):371. https://doi.org/10.3390/toxins17080371

Chicago/Turabian Style

Kim, Euikyung, Du Hyeon Hwang, Ramachandran Loganathan Mohan Prakash, Ravi Deva Asirvatham, Hyunkyoung Lee, Yunwi Heo, Al Munawir, Ramin Seyedian, and Changkeun Kang. 2025. "Animal Venom in Modern Medicine: A Review of Therapeutic Applications" Toxins 17, no. 8: 371. https://doi.org/10.3390/toxins17080371

APA Style

Kim, E., Hwang, D. H., Mohan Prakash, R. L., Asirvatham, R. D., Lee, H., Heo, Y., Munawir, A., Seyedian, R., & Kang, C. (2025). Animal Venom in Modern Medicine: A Review of Therapeutic Applications. Toxins, 17(8), 371. https://doi.org/10.3390/toxins17080371

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