Next Article in Journal
Macroalgae-Inspired Brominated Chalcones as Cosmetic Ingredients with the Potential to Target Skin Inflammaging
Previous Article in Journal
Anthelmintic Potential of Agelasine Alkaloids from the Australian Marine Sponge Agelas axifera
Previous Article in Special Issue
A Comparative Study of the In Vitro Intestinal Permeability of Pinnatoxins and Portimine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Saxitoxin: A Comprehensive Review of Its History, Structure, Toxicology, Biosynthesis, Detection, and Preventive Implications

1
Department of Student team, College of Basic Medical Sciences, Naval Medical University, Shanghai 200433, China
2
Department of Biochemistry and Molecular Biology, College of Basic Medical Sciences, Naval Medical University, Shanghai 200433, China
3
College of Chemistry and Materials Science, Fujian Normal University, Fuzhou 350007, China
4
Key Laboratory of Biosafety Defense (Naval Medical University), Ministry of Education, Shanghai 200433, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2025, 23(7), 277; https://doi.org/10.3390/md23070277
Submission received: 20 May 2025 / Revised: 28 June 2025 / Accepted: 30 June 2025 / Published: 2 July 2025
(This article belongs to the Special Issue Marine Biotoxins 3.0)

Abstract

Saxitoxin (STX) is a potent toxin produced by marine dinoflagellates and freshwater or brackish water cyanobacteria, and is a member of the paralytic shellfish toxins (PSTs). As a highly specific blocker of voltage-gated sodium channels (NaVs), STX blocks sodium ion influx, thereby inhibiting nerve impulse transmission and leading to systemic physiological dysfunctions in the nervous, respiratory, cardiovascular, and digestive systems. Severe exposure can lead to paralysis, respiratory failure, and mortality. STX primarily enters the human body through the consumption of contaminated shellfish, posing a significant public health risk as the causative agent of paralytic shellfish poisoning (PSP). Beyond its acute toxicity, STX exerts cascading impacts on food safety, marine ecosystem integrity, and economic stability, particularly in regions affected by harmful algal blooms (HABs). Moreover, the complex molecular structure of STX—tricyclic skeleton and biguanide group—and its diverse analogs (more than 50 derivatives) have made it the focus of research on natural toxins. In this review, we traced the discovery history, chemical structure, molecular biosynthesis, biological enrichment mechanisms, and toxicological actions of STX. Moreover, we highlighted recent advancements in the potential for detection and treatment strategies of STX. By integrating multidisciplinary insights, this review aims to provide a holistic understanding of STX and to guide future research directions for its prevention, management, and potential applications.

1. Introduction

The ocean, as a rich reservoir of natural products, exhibits immense potential for drug research. Saxitoxin (STX) is a remarkable yet perilous neurotoxin, and as a pivotal component of the paralytic shellfish toxins (PSTs), STX is also one of the primary initiators of paralytic shellfish poisoning (PSP) [1,2,3], posing a substantial threat to both human health and the marine ecosystem.
STX emanates from a plethora of sources. Research indicates that STX can be produced both by dinoflagellates in marine environments and by cyanobacteria in freshwater environments. The former includes Alexandrium (comprising 10 species), Pyrodinium bahamense, Gymnodinium catenatum, and Centrodinium punctatum, which have long been considered the primary contributors to PSP [4,5,6,7,8]. Freshwater cyanobacteria include Raphidiopsis brookii, Anabaena circinalis, Aphanizomenon sp., Raphidiopsis raciborskii, and Microseira wollei, among others. It has been confirmed that 15 species of freshwater cyanobacteria possess the capability for STX synthesis [9,10,11,12,13]. Through the intricate food chain, STX accumulates in organisms such as shellfish and fish. Once humans consume these contaminated seafood items, they are at risk of intoxication. The symptoms of STX poisoning are diverse and severe, affecting the nervous system and including nausea, vomiting, oral paresthesia, visual disturbances, muscle paralysis, and myalgia. These symptoms may appear within 0.5 to 2 h after ingestion, and in severe cases, it can lead to respiratory paralysis and death [14,15]. Annually, there are approximately 2000 PSP cases worldwide, with a mortality rate as high as 15% in a comprehensive review [16], highlighting the urgent necessity for in-depth research on STX.
Internationally, the World Health Organization (WHO) and the European Food Safety Authority (EFSA) have set the limit for STX in 100 g of shellfish soft tissue at 80 μg [17]. The United States Environmental Protection Agency (EPA) has placed it on the candidate pollutant list 4 (CCL4) and list 5 (CCL5), and has established health advisories for drinking water regarding microcystin and cylindrospermopsin. Additionally, due to its high toxicity, stability, and potential use as a biological warfare agent, STX has received considerable attention in fields such as food safety, neuroscience, and ecotoxicology [18,19,20,21,22], and is also listed as a controlled substance under the Chemical Weapons Convention; related experiments are subject to strict safety approval.
Although numerous studies have demonstrated that the toxicity of STX is fundamentally attributed to its potent blockade of voltage-gated sodium channels (NaVs) as a specific site-1 sodium channel blocker (S1SCB) acting on the α-subunit [23,24], recent investigations have uncovered that β subunits may also exert a specific modulatory role in mediating the interaction between STX and sodium channels [25,26]. Moreover, different subtypes of NaVs in the human body exhibit varying binding capacities for STX [27,28,29]. Notably, apart from its primary action on NaVs, STX may also affect non-NaV targets to exert more complex toxic effects [30,31]. These aspects collectively constitute the intricate toxicity mechanism of STX. STX has a minimum lethal dose of 1–4 mg/kg and a median lethal dose (LD50) of merely 10 μg/kg (intraperitoneal injection in mice) [32,33], rendering it extremely toxic.
Over the years, extensive studies on STX have been undertaken. By systematically integrating the full-chain evidence spanning from “discovery” to “mechanism” and “application”, the present review innovatively constructs an interdisciplinary research framework, which comprehensively covers STX’s basic toxicology, biosynthesis pathways, and advanced detection technologies. It provides a systematic theoretical foundation for STX prevention and control, precise toxin monitoring, and rational drug design, thereby bridging the gap between fundamental research and practical applications in the field of marine toxins.

2. Historical Traceability and Structural Analysis

2.1. First Detection and Process

The discovery of STX can be traced back to 1957, when Hermann Sommer’s team first isolated the toxin from the Alaska butter clam (Saxidomus gigantea) and named it after its host [34]. In the early stages, research primarily focused on its toxic characteristics and found that it is closely related to PSP. The structural analysis of STX has been hindered due to the lack of high-purity samples and advanced characterization techniques.
In 1975, its chemical structure was completely analyzed; Schantz et al. further purified it and revealed its three-dimensional chemical structure through X-ray diffraction and nuclear magnetic resonance (NMR) techniques, confirming it as a tetrahydropurine derivative and unveiling its unique biguanide and tricyclic skeletal characteristics, while also affirming its role as the primary toxic component of PSP and its unique biguanide, tricyclic skeleton and hydrated ketone structure, laying the foundation for subsequent toxicity mechanism and synthesis research [35,36,37].
In 1977, the Tanino H. team successfully completed the total synthesis of STX racemate (d,l-STX) for the first time, marking the beginning of the artificial synthesis of STX [38]. Subsequently, the researchers developed various synthetic strategies for its complex triazole guanidine structure. For example, the Fleming team reported asymmetric synthesis methods in the 2000s, where natural STX possessed a specific optical purity (+)-configuration; the team achieved the stereoselective synthesis of (+)-STX through 17 reaction steps [39,40]. Iwamoto et al. optimized the synthesis route using the cyanation intermediate, which improved the yield [41,42]. These synthetic breakthroughs not only advanced the development of STX analogs (such as neoSTX and deoxynivalenol), but also provided tools for researching their structure–activity relationships.

2.2. Molecular Structure Characteristics

STX belongs to the class of pyrrolidine and purine alkaloids characterized by high polarity and hydrophilicity. Its molecular formula is C10H17N7O4 (molecular weight = 299.286 g/mol) and it has a CAS number of 35523-89-8. It belongs to pyrrolopurine alkaloids and tetrahydropurine derivatives, with its molecular framework consisting of a tricyclic fused system.
The active sites are concentrated in the guanidine groups at positions 7, 8, and 9, as well as in the C12 hydroxyl group. Additionally, while the formyl side chain is not a core active group, its spatial orientation can influence the binding specificity of the toxin to different subtypes of sodium channels (such as the differences between skeletal muscle and cardiac channels) [43,44,45]. Through reductive preparation, two stereoisomers can be derived: (12R)-dihydrosaxitoxin (12α-OH) and (12S)-dihydrosaxitoxin (12β-OH). The inactive dihydrosaxitoxin features a 12α-OH configuration, with its side-chain conformation resembling that of the ketone form. In contrast, the side-chain conformation of the 12β-OH isomer is similar to the hydrated form of the original toxin. Notably, the stereoelectronic effects of the 12β-OH group exert a significant influence on the side-chain conformation, which is critical for toxicity [46]. Notably, it has been shown that zetekitoxin AB (ZTX), as the only macrocyclic member of the STX family, has a C6-C11 isoxazolidine macrocyclic structure that is speculated to be the key to enhanced toxicity (LD50 is lower than STX). The macrocyclic backbone of ZTX enhances the binding affinity with NaVs by limiting the hydration conformation of C12 (ΔGhγ = 4.29 kcal/mol). This finding challenges the conventional wisdom that C12-hydrated ketones are necessary for the high toxicity of STX [47].
The STX pure product is a white crystalline solid, readily soluble in water, slightly soluble in methanol and ethanol, and insoluble in non-polar solvents. Analysis through the X-ray single-crystal diffraction of its aqueous solution crystallization reveals that the molecules form a strong hydrogen bond network through amino and sulfonic acid groups, and that the crystal structure contains ordered layers of water molecules [35,36]. It is readily soluble in water, sparingly soluble in methanol and ethanol, and insoluble in non-polar solvents. This solubility profile arises from the high polarity of the guanidino groups in the STX molecule, which facilitate the formation of extensive hydrogen-bonding networks, particularly in protic solvents; STX is thermally stable and exhibits high stability under acidic conditions (pH < 4) and low temperatures (−20 °C). It can be stored long-term in diluted hydrochloric acid without losing activity. Carbamate hydrolysis occurs only in highly concentrated acid solutions. Due to its heat and acid resistance, conventional cooking and processing cannot destroy STX, posing a significant threat to food hygiene and safety. However, it is prone to oxidation and inactivation in alkaline environments, with its toxicity rapidly diminishing [48,49]. Specifically, at a neutral pH, STX molecules exist in equilibrium in a divalent cationic hydrate, monovalent cationic hydrate, and monovalent ketone form. With the increase in pH, the C8 guanidine group donates hydronium ions, leading to a decrease in charge value, Additionally, the equilibrium between the ketone and hydrate forms affects the charge state of the molecule, potentially altering its interaction with sodium channels [46,50]. Saxitoxin with a deprotonated C8 guanidino group is susceptible to oxidation, so STX exhibits a more degradable characteristic when the pH value is >8 [51]. Furthermore, STX can be stored stably for a long time under light-proof conditions, while light exposure may accelerate its oxidative degradation. Solid STX remains stable in dry air, but is prone to oxidation when it absorbs moisture or is in solution. Therefore, it should be stored in a sealed and light-proof container [52].
The STX family includes various natural analogs. To date, 57 types of STX analogs have been described, all sharing a tetrahydropurine ring skeleton that can be substituted at positions C11, N1, and C13. Their structural differences primarily arise from the substitution or modification of functional groups. The saxitoxins themselves are further divided into classes such as C toxins, G toxins, and LW toxins depending on their chemical structures and substitutions [18]. They can be classified into several categories, such as non-sulfated, mono-sulfated, di-sulfated, and hydrophobic analogs, each with varying levels of toxicity [53,54]. The new stonehouse clam toxin (neo-STX), based on STX, has an added hydroxyl group at the N1 position, enhancing the hydrogen bond interaction with the sodium channel Tyr401, resulting in a higher affinity than STX [55]. In decarbamoylsaxitoxin (dc-STX), the C12 trancarbamyl group is missing, and the toxicity is reduced, but some channel inhibitory activity is retained [56]. Gonyautoxin (GTX) is generated through mono-sulfation at the N21 or O22 position of STX, which can then be di-sulfated to produce C-toxins [57]. The research on STX analogs is crucial for reducing PST toxicity and chemical or enzymatic degradation, in order to develop detoxification methods, while also exploring the unique medicinal potential of these toxins [58].

3. Molecular Biosynthesis

3.1. Biosynthetic Gene Clusters and Synthetic Pathways

The biosynthetic gene cluster of STX (Sxt gene cluster) was first identified in cyanobacteria, and genomic research showed that this gene cluster contained approximately 36 genes that encoded various functional enzymes, such as methyltransferases and aminotransferases [1].
The biosynthesis of STX is a highly complex process (Figure 1), commencing with the polyketide synthase-like enzyme SxtA, and this enzyme contains multiple catalytic domains (sxtA1-A4). Among these, the sxtA4 domain (functions as condensation domain) is considered a critical marker for toxin synthesis. SxtA is responsible for initiating the synthesis of the STX core structure using arginine and acetyl-coenzyme A (acetyl-CoA) as starting substrates. This step generates the intermediate product 4-amino-3-oxoadipate, which provides the structural framework for subsequent reactions. The subsequent step involves the participation of SxtG and SxtB in the guanylation reaction, catalyzed by guanylyltransferase, which introduces the guanidine derived from arginine into the molecular framework, forming the characteristic guanidine ring structure of the toxin. SxtS is a 2-oxoglutarate-dependent (2OG) dioxygenase that executes the consecutive epoxidation of new double bonds and opens the epoxide to form an aldehyde, accompanied by bicyclization. The dehydrogenase SxtU then reduces the terminal aldehyde of the STX precursor. Following this, SxtH and SxtT (the terminal monooxygenase subunit and the associated ring-hydroxylating dioxygenase encoded by the respective genes) catalyze the consecutive hydroxylation at C-12, resulting in dcSTX. Subsequently, SxtI combines with SxtJ and SxtK to catalyze the formation of STX, with the entire process encompassing approximately 30 biosynthetic steps [1,59] (Figure 1).
On the basis of STX, STX analogs are regenerated through specific reactions, and this process and the transport of toxins are involved in the transport of sxtL, sxtN, sxtO, sxtR, sxtX, sxtW, sxtZ, sxtPER, and other proteins encoded by the sxtACT gene [1,60,61].
The conservation analysis of the gene clusters indicates that the Sxt gene clusters of different toxin-producing cyanobacteria (such as Raphidiopsis brookii and Cylindrospermopsis raciborskii) have high homology, but that there are species-specific modifying enzymes that lead to the diversity of toxin analogs [62,63,64].
Additionally, research utilizing the Alexandrium minutum strain indicates a significant correlation between the toxin content of the strain and the expression of the fourth domain in the sxtA gene (sxtA4), as well as the presence of genetic subtypes. The quantitative polymerase chain reaction (qPCR) determination based on sxtA4 has been applied in the study of toxic algal blooms associated with PST in marine environments, demonstrating broad application prospects [60,65]. There are also newly designed sxtB TaqMan probes for the targeted detection of the STX’s biosynthetic gene sxtB from Alexandrium catenella and Alexandrium Pacificum [66]. Notably, dinoflagellates, as the primary producers of STX, have posed challenges in identifying their biosynthetic genes. A recent comparative transcriptomic study of Gymnodinium catenatum identified over 1000 homologs of STX biosynthesis genes in this species. Among these, sxtA and sxtG exhibited significantly higher expression levels than other genes, suggesting their central roles in the toxin biosynthesis pathway. Intriguingly, the study revealed the absence of sxtB. This finding implies that Gymnodinium catenatum may employ alternative metabolic routes for STX biosynthesis or rely on functionally redundant genes to compensate for the loss of sxtB [67].

3.2. Environmental Regulation of Toxin Synthesis

The biosynthesis of STX is coordinately regulated by various environmental factors, including nutritional conditions, light, temperature, salinity, and others [68,69,70].
Different toxin-producing algae have their own suitable temperature ranges for growth and STX synthesis. Within the appropriate temperature range, enzymatic activity is high, and reactions related to the STX synthesis pathway can proceed smoothly. Extreme temperatures, whether too high or too low, can affect enzyme activity and thus inhibit STX synthesis. Additionally, temperature can influence the expression of core genes sxtA and sxtG, thereby affecting STX content, particularly evidenced by the increased expression of sxtA and sxtG under low-temperature stress [71].
The intensity and quality of light can affect the photosynthesis and related metabolic pathways of toxin-producing algae, thereby influencing the synthesis of STX [72,73]. Appropriate light intensity can promote the growth and metabolism of algae, facilitating the synthesis of STX, while excessive or insufficient light can inhibit STX production. Research indicated that an increase in temperature combined with adequate light can enhance the biomass of Raphidiopsis raciborskii. However, excessively high temperatures can lead to a decrease in its toxicity [74].
Toxic cyanobacteria (such as Raphidiopsis raciborskii) and dinoflagellates (such as the genus Alexandrium) perceive environmental changes and dynamically adjust the expression of the Sxt gene cluster to adapt to ecological pressures and maintain competitive advantages [75,76]. Specifically, changes in the nitrogen and phosphorus content in nutrients lead to nutritional stress, with nitrogen stress inhibiting the growth of Alexandrium minutum and reducing the biosynthesis of STX, while phosphorus deficiency can lead to an increase in the biosynthesis of saxitoxin [77,78]. The NO3 concentration also affects the expression of sxt core genes and affects the production of STXs, suggesting that N is directly involved in the biosynthetic pathway of STXs [79].
Apart from the natural environment, the biological environment can also affect the content of STX. For instance, under the chemical induction of the zooplankton Daphnia gessneri, the sxt genes of Raphidiopsis raciborskii (Cyanobacteria) are upregulated [80]. Other studies have shown that Jannaschia cystaugens (representing Alexandrium cyst formation-promoting bacteria, Alex-CFPB) significantly influences the life cycle and toxin production of Alexandrium pacificum. Co-culturing with Jannaschia cystaugens induces stress that promotes the synthesis of highly toxic PST, leading to increased intracellular toxin levels, specifically involving oxidative stress, nutrient stress, and quorum sensing signals, all of which contribute to an increase in toxin synthesis [81].

3.3. Latest Research Developments

Although significant achievements have been made in the biosynthesis research of STX, many details still require refinement.
In exploring the details of the STX biosynthetic pathway, multi-omics technologies such as metabolomics and proteomics have played a crucial role. Through metabolomics, it is possible to comprehensively analyze the metabolite profiles of toxic algae at different growth stages and under various environmental conditions, and by utilizing proteomics, it is possible to identify differentially expressed proteins in toxic algae under varying environments, thus providing a pathway for the in-depth exploration of the dynamic changes in the STX biosynthetic pathway under environmental influences [2].
In addition, biotechnological techniques provide new strategies for the efficient production of STX, and researchers are attempting to reconstruct the STX biosynthetic pathway in heterologous hosts (such as Escherichia coli and Synechocystis) to achieve the extracellular synthesis of PST [82,83]. Jiao Y et al. established an asymmetric synthesis strategy featuring an intramolecular [2 + 2] photocycloaddition of an alkenylboronate ester equipped with a new chiral auxiliary, leveraging chiral additives to stabilize the transition state through hydrogen bonding, which achieved high stereoselectivity in constructing tricyclic frameworks and offered a novel paradigm for synthesizing complex toxins [84]. Furthermore, Greenhough H et al. scaled up the cultivation of Alexandrium pacificum using a 1250 L photobioreactor (PBR), thereby surpassing the production limitations of traditional natural sources for GTX-1,4. This advancement provides a gram-scale supply of toxins for preclinical research and offers a groundbreaking solution for the production of STX and its analogs [85].

4. Bioenrichment of STX: Pathways, Influencing Factors, and Ecological Impacts

The traditional route by which STX causes human poisoning involves the production of toxic dinoflagellates (such as the genus Alexandrium), their concentration through filter-feeding bivalves (mussels, oysters, and clams), and subsequent ingestion by humans, representing a common pathway in reported poisoning incidents. This pathway is closely associated with the species composition, geographic distribution, and toxin-producing capacity of the involved dinoflagellate populations. These factors are further influenced by environmental conditions, including salinity, temperature, and eutrophication levels [86,87,88].
Different traditional media have variations in the speed of toxin removal, retention time, and the retention of toxic components. The toxicity produced by dinoflagellate species also varies, resulting in different toxic effects on humans. Many countries have implemented monitoring programs for STX in toxic microalgae and shellfish, which help minimize public health risks [89].
Non-traditional media include non-bivalve invertebrates (gastropods and cephalopods), crustaceans, fish, annelid worms such as Eudistylia sp., and echinoderms like Asterias amurensis, Astropecten scoparius, Astropecten polyacanthus, and Pisaster ochraceus. Among these, although gastropods and cephalopods accumulate STX, they do not cause significant negative effects, while the yellow crab is the most commonly known non-filter-feeding non-molluscan species containing STXs. The mechanism by which STX is found in fish involves the transport of STXs through the food chain and the dissemination and accumulation of toxins through zooplankton [90,91]. Research indicates that the internal organs are the primary sites for PST accumulation. Simultaneously, these organs exhibit a higher STX clearance rate and excrete via the kidneys. Among STX and its analogs, those with higher solubility have a greater excretion rate [77,92]. The regulation and control of these non-traditional media is of significant importance for preventing STX poisoning incidents.
In the field of ecotoxicology, the accumulation of STX at the top of the food chain can lead to population cascading effects. For instance, high concentrations of STX in shellfish can cause foodborne poisoning, including PSP [93,94]. Toxins’ metabolic transformation significantly affects their toxicity level. The latest studies show that when mussels are exposed to dynamic algal cell densities (simulating natural blooms), the concentration of PSTs (paralytic shellfish toxins) in their bodies increases by 1.8 times compared with exposure to fixed densities. Moreover, low-toxicity sulfocarbamoyl toxins (such as GTX5) are converted into highly toxic derivatives (e.g., dcSTX) through decarbamoylation reactions, a process associated with the inhibition of sulfotransferase activity in the hepatopancreas [95]. The consumption of contaminated fish may lead to the mass death of marine predators that feed on fish (such as seabirds and cetaceans) [96,97,98]. Exposure to STX during the development of zebrafish can lead to changes in the expression of genes related to axon growth and affect the functionality of NaVs [99]. Studies on Saxidomus gigantea (Alaska butter clams) have shown that when humans consume 200 g of tissue containing 900 μg of saxitoxin (STX), the probability of severe symptoms in average males reaches 11%, and the mortality risk is 0.27%. This confirms the potential threats of STX to the ecological chain and human health [100].
The acute guideline value for STXs proposed by the WHO is 0.003 mg/L (3 μg/L) for the total STXs (including congeners, free forms, and cell-bound forms). Exposure routes include drinking contaminated surface water, consuming contaminated marine shellfish, and having contact with high-concentration water bodies during recreational activities. In terms of protection, avoiding contact and accidental ingestion is fundamental. So far, there is no specifically approved antidote, and treatment after poisoning mainly focuses on symptomatic and supportive care [101,102]. Health risks should be addressed through water source protection and water treatment process control, while paying attention to their impacts on the sensory quality of drinking water. Controlling water eutrophication can reduce the risk of cyanobacterial blooms. Additionally, water contaminated with STXs can be treated to mitigate poisoning risks. Extracellular STXs can be physically removed via activated carbon adsorption or membrane filtration [103]. Alternatively, chemical inactivation methods such as potassium permanganate, ozonation, and chlorination, etc., can disrupt the toxin structure. However, the risks of disinfection by-products should be noted [104,105]. In addition, H2O2 may not be able to effectively degrade saxitoxin [106]. Biological activity-based removal or degradation methods, such as adsorption by lactic acid bacteria and degradation by bacteria isolated from the digestive tract of blue mussels (Mytilus edulis), have been reported [107]. To remove intracellular toxins, complete cyanobacterial cells need to be eliminated, which can be achieved through traditional coagulation and sand filtration processes [108]. During the removal process, care must be taken to prevent cell lysis and the release of toxins, as the review [109] showed.

5. Toxicity Mechanisms of STX

STX, as a guanidine alkaloid, features a core structure with two guanidine groups, both belonging to the guanidinium toxin family with TTX. These guanidine groups exhibit an extremely high affinity with and ion flux blocking ability towards NaVs [25,110], which endows STX with high biological activity and extreme toxicity.

5.1. Blocking Mechanism of STX on NaVs

Toxins can be roughly divided into two categories: pore blockers that physically close the pore of the channel and “gating modifier toxins” (GMTs) that act on the voltage-sensing domain of the channel to cause abnormal activation (opening) or inactivation (closing) of the sodium ion channel.
Mammalian NaVs consist of a pore-forming α subunit and one or more β subunits. The α subunit is the main part that forms the ion channel pore and has voltage-sensing and gating functions [111,112]. The α subunit is folded from a single polypeptide chain containing four homologous repeat sequences. Each repeat sequence contains six transmembrane helix segments (S1–S6). These segments are connected by extracellular and intracellular loops. S1–S4 form the voltage sensor, and S5 (outer helix), S6 (inner helix), and the P loop that re-enters the extracellular membrane between S5 and S6 together form the pore domain [113,114]. Site 1 on the α subunit is the common receptor-binding site for these types of toxins, such as STX, tetrodotoxin (TTX), and polypeptide toxins like μ-conotoxin [25,115]. The active sites of STX are the guanidyl groups at positions 7, 8, and 9. They have a high affinity with the amino acid residues at site 1 of the voltage-gated Na⁺ channel on the excitable cell membrane, thereby physically blocking the pore [29]. Through the analysis of the structure of the sodium channel–STX complex via X-ray crystallography and cryo-electron microscopy, it was observed that after STX binding, the selective filter region of the channel contracted, narrowing the ion permeation path and preventing sodium ions from passing through smoothly [116].
STX blocks the influx of sodium ions, inhibiting the establishment and collapse of the transmembrane electrochemical charge gradient in animal nerve, muscle, and endocrine cells. This ion imbalance hinders the initiation of action potentials [117,118], leading to serious consequences. The symptoms of STX poisoning usually appear within 30 min after ingestion, including oral numbness, limb paralysis, and breathing difficulties, etc. In severe cases, it can cause death within a few hours [119,120] (Figure 2).

5.2. Association of β Subunits in the Action of STX

Furthermore, the process of STX blocking sodium channels also involves interactions with β subunits. The β subunit is not just a simple accessory to the pore-forming α subunit. Under normal circumstances, β subunits play an important role in the localization of sodium channels, function regulation, and coupling with other cell signaling pathways [115,121], allowing for the precise control of excitability in a cell type-specific manner. At the same time, the regulatory effect of the β subunit on voltage-gated potassium channels affects neuronal excitability. Abnormal neuronal excitability is related to many nervous system diseases. Therefore, this regulatory effect may play a role in the occurrence and development of nervous system-related diseases [26].
Messner, D.J. and Catterall, W.A. detected the functional integrity of the resulting protein complexes by selectively removing β1 or β2 subunits from the detergent-solubilized channels in rat brains. They found that selectively removing the β1 subunit from the detergent-solubilized channels in rat brains led to a complete loss of 3H-labeled STX binding activity in sodium channels, confirming that the presence of the β1 subunit is crucial for maintaining the high-affinity conformation of the STX binding site on sodium channels [122].
Nonetheless, currently, there are very few studies on the specific role of β subunits in the process of STX toxicity, and further exploration is needed.

5.3. Affinity Difference in Different Subtypes of NaVs to STX

In the human body, there are nine subtypes of sodium channels, such as Nav1.1–Nav1.9. They are specifically distributed in different tissues. The amino acid sequences, spatial conformations, and accessory subunits of different subtypes of sodium channels vary [27,28,123], resulting in different binding abilities of STX to each subtype.
However, STX binds to the amino acid residues at site 1, which is located on the outer surface of the excitable cell membrane and closer to the outer opening of the sodium channel [29,43]. Sodium channels in nerve tissues have a high affinity for STX. Ritchie and Rogart conducted binding experiments using radioactively labeled STX with rabbit vagus nerves, lobster walking leg nerves, and gar olfactory nerve fibers and found that STX can tightly bind to the NaVs on nerve cell membranes and effectively block nerve impulse conduction at extremely low concentrations [124].
Nav1.4 is mainly distributed in skeletal muscle. The high affinity of Nav1.4 in skeletal muscle for STX may be determined by the charge complementarity of acidic residues in the pore region, the cooperative effect of domains, and subtype-specific conformations [113,125], which provides a key molecular target for the development of toxin antagonists targeting skeletal muscle sodium channels. In contrast, the sodium channels in cardiac muscle tissue have a relatively low affinity for STX [126]. Catterall and Coopersmith used radioactive ligand binding experiments to detect STX receptor sites in vertebrate heart tissues. However, the binding ability of heart tissues to STX is weaker than that of nerve tissues, and a higher concentration of STX is required to produce a significant inhibitory effect [127].

5.4. Potential Toxicity Targets of STX

Interestingly, although the toxicity of STX mainly stems from the inhibition of specific NaVs, STX can also affect other channels, enzymes, and proteins, thereby exerting toxic effects [128]. Studies have found that STX also affects L-type calcium channels [30] and HERG potassium channels [31], interfering with the ion balance and electrophysiological activities within cells. In addition, STX may also act on neuronal nitric oxide synthase, STX-metabolizing enzymes, a transferrin-like family of proteins, and a unique protein found in the blood of pufferfish. Although the specific mechanisms and degrees of influence remain unclear, these additional effects may further exacerbate the toxic effects of STX on organisms [120,128].

5.5. Synergistic Effect of STX with TTX

In addition to structural analogs, the combined exposure effects of STX with other toxins have garnered increasing attention in recent years. Despite their pronounced chemical, structural, and affinity disparities, both STX and TTX are categorized as S1SCBs, and their neurotoxicity may exhibit additive effects [129,130].
Furthermore, investigations into the combined toxic effects of STX and TTX have made significant strides recently. Concurrently, Andrea Boente-Juncal et al. validated through chronic toxicity experiments that combined exposure to TTX and STX elicits additive toxic effects, particularly when the STX dosage exceeds the maximum exposure threshold recommended by the EFSA, leading to significantly elevated mortality and biochemical derangements. They demonstrated via intragastric gavage that co-exposure to STX and TTX in mice induced significantly more aberrant changes in blood biochemical parameters and higher mortality rates compared to single-toxin exposure groups, thereby confirming their combined toxic effects [131]. Nevertheless, the intragastric gavage methodology harbors methodological limitations, as it may lead to inadvertent pulmonary aspiration of the administered toxins—potentially causing rapid lethality—and mechanical reflux induced by gastric tube insertion [132,133]. Nevertheless, Finch et al. systematically demonstrated the additive toxicity of STX and TTX through LD50 determinations via multiple routes (intraperitoneal injection and feeding), a comparison of predictive models with experimental data, and consistency analysis of action mechanisms and clinical symptoms [134]. Despite the lack of in-depth explanations of the mechanisms of the combined effects of toxins, these findings provide a scientific basis for incorporating TTX into the regulatory system of PSTs.

6. Self-Resistance Mechanisms to STX Enlighten Treatment Development

Currently, there is no specific antidote for STX poisoning. Treatment mainly involves symptomatic supportive therapies such as artificial ventilation and gastric lavage, as the review summarized [135]. However, research on the strategies of organisms in nature to cope with STX exposure has inspired the research on STX treatment [136,137]. In recent years, relevant research achievements in exploring STX antagonists and optimizing toxin delivery systems have also provided clues for the development of drug treatments.

6.1. Target Protein Resistance Mutations

Target protein resistance mutations play a crucial role in the research on the mechanisms of biological toxicity resistance and have been investigated in depth and systematically [137,138]. Soft-shell clams (Mya arenaria) from areas affected by “red tides” have been found to be more resistant to PSTs. This is due to a natural mutation of a single amino acid residue, which reduces the affinity of the STX binding site in the sodium channel pore of resistant soft-shell clams (but not sensitive ones) to one-thousandth of the original level [139]. When cysteine (Cys) at position 374 of the cardiac sodium channel α subunit (RHI) is replaced by tyrosine (Tyr) (similar to the situation of the μl subtype), the sensitivity of RHI to STX changes significantly [140], which may significantly affect the STX resistance of organisms. When single-gene mutations were studied in the SS2 and adjacent regions of the four internal repeat sequences in sodium channel II, it was found that mutations involving a negatively charged residue clustered at specific positions significantly changed the STX sensitivity, and that these mutations may affect the extracellular mouth and/or pore wall structure of the sodium channel [125]. All these mutations have an impact on the STX resistance of organisms.

6.2. Toxin Sequestration

Soluble STX-binding proteins found in some organisms such as frogs, pufferfish, cockles, and crabs constitute a special mechanism for organisms to cope with toxins—toxin sequestration [141]. Among the strategies of organisms to cope with toxins, the resistance mechanism of frogs to STX has attracted much attention [142]. It has been found that saxiphilin (Sxph) in bullfrogs can bind to STX with a high affinity (Kd is approximately 0.2 nM) [143,144,145]. In the study of the binding site of STX to NaVs, it was found that Sxph and Navs use similar molecular logic to recognize STX. Both use side-chain carboxylates to interact with the five-membered and six-membered guanidyl groups of STX. Moreover, Sxph can bind STX through a specific binding pocket in the C1 domain, thus acting as a high-affinity “toxin sponge” protein to prevent STX poisoning. In addition, a similar sequence of Sxph, SxphNP, was found in the Himalayan frog (Nanorana parkeri), which also has the key characteristics of recognizing STX [128,146]. These findings expand the research direction of developing effective STX antidotes based on biological resistance.

6.3. Biotransformation and Metabolism

Some reports have indicated that the digestive glands of bivalve molluscs possess a high capacity for PST transformation [147,148], suggesting the presence of microorganisms and/or toxin-transforming enzymes within them, as demonstrated in the review [149]. Vasama Mshou et al. were the first to confirm that Lactobacillus rhamnosus GG and LC-705 strains (both in viable and non-viable forms) can remove STX and neoSTX from neutral solutions with a pH of 7.3 and acidic solutions with a pH of 2, achieving removal rates of as high as 77–97.2% [150]. Pseudoalteromonas bacteria are capable of degrading PSTs, and STX is one of them. These bacteria have the potential to remove STX. Such bacteria are relatively common in the marine environment. They may degrade toxins through special enzymes secreted by themselves or metabolic pathways. Marine bacterial isolates cultured from the digestive tracts of blue mussels (Mytilus edulis) contaminated with PST can reduce the overall toxicity of algal extracts to less than 10% of the original within 3 days [151]. These bacteria and similar ones may assist marine bivalves in the natural metabolism and elimination of PSTs and other marine biotoxins, and also promote research on the biological treatment of STX poisoning.
Meanwhile, some newly reported C-11 hydroxyl analogs of PSTs in shellfish are considered novel derivatives of PSTs and are called “M-toxins” [152,153,154,155]. The research team led by Li Aifeng confirmed that the lower the pH and temperature under experimental conditions, the more stable the M-toxins. Compared with common components of PSTs, the stability of M-toxins is poor. They also confirmed the chemical transformation pathways of “M-toxins” as M1→M3→M5, M2→M4→M6, and NEO→M10 [156]. This is of great significance for understanding the detoxification mechanism of shellfish and achieving rapid detoxification by converting highly toxic PSTs into weakly toxic M-toxins. It also provides new ideas for studying the detoxification mechanism of organisms to STX.

7. Detection

In the context of the high risk of STX, countries around the world have listed it as a mandatory item for aquatic product safety inspection, and the detection technology of STX and its derivatives is the key to its bioaccumulation and toxicological mechanism study. At present, there are various detection methods, mainly including biological detection methods and instrument detection methods, etc.
Biological detection methods, based on the biological activity of toxins, were the main means for early STX detection. The Mouse Bioassay (MBA) is a classic method for early STX detection. It involves the intraperitoneal injection of sample extracts and semi-quantitative analysis of STX toxicity based on mouse death time. Since it was confirmed as the statutory detection method by the Association of Official Analytical Chemists (AOAC) in 1958 (AOAC 959.08), it has been used as the reference standard for other detection techniques [157]. However, it has limitations such as an inability to distinguish toxin analogs, low sensitivity, and animal ethics controversies [158]. Additionally, the Daphnia acute toxicity test utilizes the sensitivity of Daphnia pulex (limit of detection (LOD): 0.3–10 ng equivalent to saxitoxin (Eq.STX)/L) and Moina micrura (LOD: 2.7–10.7 ng Eq.STX/L) to STXs, serving as a rapid screening tool for environmental samples [159]. Similarly, the speckled cockroach (Nauphoeta cinerea) bioassay can act as a supplementary field screening tool for STXs, with its core value lying in cost and speed advantages, though specificity remains low [160].
Cell-level detection, achieved through toxicity effects mediated by NaVs, can replace traditional mouse bioassays. For example, the Neuro-2a cell-based assay operates on the principle that veratridine, in the presence of ouabain, increases sodium influx in the mouse neuroblastoma cell line Neuro-2a (ATCC, CCL131), leading to cell swelling and death [159]. STX specifically blocks NaVs, antagonizing the toxic effect of veratridine. The toxicity of STX is indirectly reflected by detecting functional changes in intracellular ion channels or cytotoxic effects (such as MTT metabolism and lactate dehydrogenase release). Similarly, this method can also be used for the detection of sodium channel-blocking toxins such as TTX [161,162]. However, this method requires a long analysis time. The analysis time can be shortened by using modified forms of short toxins or the calcium agonist maitotoxin [163,164]. In addition, it also has high requirements for reagents, professional knowledge, and equipment.
With the acquisition of anti-STX antibodies, immunoassay techniques have been developed. The enzyme-linked immunosorbent assay (ELISA) method is the most commonly used, relying on the color reaction between enzyme-labeled antibodies and substrates for detection, including direct and indirect ELISA [165,166]. Immunoassay can achieve the rapid screening of PSP, but its cross-reactivity to different toxins is low, making it difficult to fully reflect the composition of toxic components in samples. Therefore, when detecting STX and its analogs, additional approaches are required to identify and quantify specific toxins, which highly depends on laboratory equipment [165]. The subsequently developed MIST Alert™ rapid detection method can be used for the determination of the total PSP toxicity and related toxin toxicity in acidic crude toxin extracts. It achieves a 100% detection rate for toxic extracts containing at least 80 μg Eq.STX/100 g, with a LOD of approximately 40 μg Eq.STX/100 g (varying with toxin profiles). This method features high sensitivity, rapid detection, and does not require specialized tissue culture equipment or expertise. However, it is unable to distinguish individual toxin components [167,168]. Rabbit monoclonal antibodies (r-mAbs) developed by Li et al. achieved the ultra-sensitive recognition of STX through a single-cell sorting and cross-screening strategy. The sensitivity is 100 times higher than that of traditional mouse antibodies, which solves the problem of the poor stability of traditional antibodies in complex matrices [169].
The receptor binding assay (AOAC 2011.27) utilizes the principle that PSP toxins can competitively bind to NaVs with [3H]-STX diHCl. This method applies isotope labeling to detect PSP toxins, featuring rapidity, reliability, and accuracy [170]. The substitution of Na+ channel proteins with saxiphilin, which specifically binds to STX, has been studied to be specific for STX identification and can be distinguished from TTX identification [171,172].
In chemical instrumental analysis, high-performance liquid chromatography (HPLC) was the first instrumental analysis technique applied to the detection of PSTs. It offers advantages such as high sensitivity, strong specificity, a low LOD (0.1 μg/kg), and fast analysis speed. HPLC can simultaneously provide qualitative and quantitative information for multiple toxins, making it the core method for toxin detection currently. Due to the weak chromophoric group of STX, pre-column oxidation (AOAC 2005.06) or post-column oxidation (AOAC 2011.02) is required before detection to generate fluorescent chromophores (e.g., dimethylaminobenzoic acid derivatives), which are then detected using a fluorescence detector. Among them, post-column oxidation HPLC is noted for its operational simplicity, strong matrix tolerance, and high stability [173,174].
Subsequently, with the advancement of mass spectrometry technology, the combination of liquid chromatography and mass spectrometry (LC/MS) has emerged as one of the preferred analytical methods for the highly selective and sensitive quantification of toxins. In a study, dansyl chloride was used for the chemical derivatization of STX, followed by separation via ultra-high-performance liquid chromatography (UHPLC) coupled with heated electrospray ionization (HESI), and quantification using a Q-Exactive mass spectrometer. This approach achieved a LOD as low as 0.01 μg/L [175,176]. LC-MS/MS serves as the gold standard for detecting guanidine toxins. Its chromatographic separation module is compatible with both HPLC and UHPLC systems [177]. Liquid chromatography separates STX via a chromatographic column by leveraging the partitioning difference between the stationary phase and mobile phase. Meanwhile, mass spectrometry ionizes STX through techniques such as electrospray ionization (ESI), generates characteristic product ions via collision-induced dissociation (CID), and enables the simultaneous qualitative and quantitative analysis of STX and more than 50 of its analogs using the multiple reaction monitoring (MRM) mode [178]. Subsequent improvements in the separation and extraction methods for STX have been made. For example, the combination of isotope dilution liquid chromatography tandem mass spectrometry (HILIC) and the isotope dilution method has enabled the simultaneous detection of multiple toxins and improved sensitivity [179,180]. The extraction and purification of STX are performed using immunoaffinity column (IAC) combined with solid-phase extraction (SPE) technology. STX is extracted with phosphate-buffered saline (PBS), separated on a TSK-GEL Amide column, and detected via LC-MS/MS. This method completely eliminates the matrix effects and does not require matrix matching, meeting the requirements for the detection of trace STX in bivalve aquatic products [181]. Additionally, extraction is performed by acidifying plasma samples and precipitating with acetonitrile, followed by LC-MS-MS analysis. The entire process takes ≤30 min, with a LOD of 2.8 ng/mL and a lower limit of quantitation (LOQ) of 5.0 ng/mL [182]. This technique utilizes the characteristic fragmentation of the guanidino group in STX molecules, combined with the high sensitivity of electrospray ionization (ESI) to distinguish structurally similar toxins (such as neoSTX and dcSTX) and achieve accurate confirmation in complex matrices (e.g., biological tissues and seawater). For instance, Dahlmann et al. used LC-ESI-MS/MS to simultaneously detect STX and microcystins in phytoplankton, confirming the tricyclic skeleton structure of STX through MS/MS fragment analysis [183]. LC-MS/MS offers significant advantages in STX detection, enabling simultaneous multi-component analysis without the need for derivatization, thus ensuring high operational efficiency. However, it also has limitations, such as expensive instrumentation and complex maintenance.
Thin-layer chromatography (TLC) is one of the traditional methods for detecting PSP toxins, which is currently used less frequently [184]. Research has been conducted on the chromatographic column for the separation of STX and its congeners using thin-layer chromatography in a rod shape, with detection by a flame thermionic detector (FTID). The detection limit for STX was 5 ng, while the detection limits for the other congeners also reached the ng level [185]. In addition, traditional techniques for detecting STX toxins include gas chromatography, ion-exchange column chromatography, and electrophoresis [184,186,187,188]. However, these traditional methods are currently less used due to their low resolution and poor sensitivity.
In recent years, new detection technologies have been continuously emerging. For example, some research is dedicated to developing detection methods based on biosensors, utilizing the specific interactions between biological recognition elements and STX, achieving the rapid and sensitive detection of STX through signal transduction, thereby enhancing selectivity and sensitivity [189,190]. These methods hold promise for overcoming some of the limitations of the traditional techniques, such as expensive instrumentation and complex operation, featuring advantages of portability, rapidity, high sensitivity, and low cost. They provide new possibilities for on-site detection and real-time monitoring. However, they also suffer from limitations, including complicated fabrication, short operational lifespan, complex manipulation, high professional requirements, and significant susceptibility to matrix effects [191,192,193]. In addition, microfluidic chip technology has also begun to be applied in the field of STX detection. The microfluidic chip can integrate multiple steps such as sample processing, reaction, and detection onto a small chip, offering advantages such as fast analysis speed, the low consumption of samples and reagents, and the ability to achieve high-throughput detection, thereby providing direction for the miniaturization, automation, and integration of STX detection. However, its fabrication is complicated, and each detection causes the consumption of the detection chip, resulting in a short operational lifespan [194,195].
Based on the existing research methods and their characteristics, combined with the scenario requirements for STX detection (on-site screening, laboratory confirmation, and research-grade analysis), the following feasible detection system of “rapid screening–precise quantification–mechanism verification” is proposed. By integrating the sensitivity, specificity, and operational complexity of different techniques, this system enables efficient detection. Quality control is essential before and during the detection process. The first stage involves on-site rapid screening, applicable to critical scenarios including aquatic product harvesting sites, the emergency monitoring of environmental water bodies, and initial food safety screening. The Daphnia acute toxicity bioassay and MIST Alert™ rapid test are employed here, leveraging their combined advantages of operational simplicity, cost-effectiveness, and rapid response. These techniques enable the preliminary triage of non-toxic samples, thereby streamlining the subsequent analytical workload. The second stage is the laboratory confirmation and quantification stage (applicable scenarios: regulatory sample re-inspection, market access testing, and acquisition of basic data for toxicological research). Firstly, the indirect ELISA method is adopted, which uses the specific binding of enzyme-labeled antibodies to STX for color development, rapidly quantifies the total STX equivalents in samples, and screens positive samples. The second step is precise analysis via HPLC (post-column oxidation method) or LC-MS/MS, with LC-MS/MS being the preferred choice. The third stage is the research-grade analysis and mechanism verification stage (applicable scenarios: STX toxicological mechanism research, identification of novel toxin analogs, and drug development target screening), involving the utilization of cell-based assays (Neuro-2a cell-based assay) and receptor-binding assays. By integrating LC-MS/MS for structural characterization with cell function analysis, this stage deepens STX research from the multi-dimensional perspectives of “toxicity effect–molecular mechanism–structural characterization” (Table 1).

8. Conclusions

STX, as a potent neurotoxin derived from marine dinoflagellates and freshwater or brackish water cyanobacteria, has been the subject of extensive research in recent years. This review has comprehensively summarized the current state of knowledge regarding STX, encompassing its discovery history, chemical structure, molecular biosynthesis, bioenrichment, toxicity mechanisms, self-resistance mechanisms, and detection methods.
The discovery and structural elucidation of STX were pivotal milestones in understanding its properties and biological activities. The identification of its unique tricyclic skeleton and biguanide group provided a foundation for subsequent studies on its interactions with biological targets, particularly NaVs.
In the field of toxicity mechanism research, the traditional mechanism of STX as a specific S1SCB of NaVs has been systematically elucidated. However, recent studies have continuously broken through the boundaries of cognition, not only discovering potential novel action targets of STX beyond NaVs, but also achieving key advancements in the exploration of biological natural resistance mechanisms. These innovative discoveries provide entirely new ideas and theoretical support for the development of STX-related drugs. Scholars such as Mingxuan Kai et al. have developed Neuron-MOF/SxtN-NPs (neuronal membrane-coated MOF nanoparticles loaded with SxtN enzyme), a dual-mode cellular nanoparticle formulation. They enable broad-spectrum neurotoxin neutralization via membrane mimicry and continuous detoxification through enzymatic activity. In STX-intoxicated mice, the nanoparticles show enhanced survival, no acute toxicity, and validate a biomimetic nanomedicine platform for neurotoxin countermeasures [199]. Dali Wang et al. have pioneered the design of an aptamer-based small-molecule therapeutic sustained-release system, applying STX to nerve blocks [200], which has inspired the progress of “seeking drugs from the ocean”.
As a highly toxic marine neurotoxin, STX as a single component has posed a threat to the marine ecological balance and human health. In natural environments, STX often forms complex contamination systems with multiple toxicants, and more than 50 natural structural analogs of STX have been identified. In this context, research on the combined action of toxins is gradually becoming a frontier in the field of STX toxicology. This research not only helps to reveal the real toxic risks of natural contamination systems, but also provides key scientific evidence for establishing the safety thresholds of marine toxins based on mixture effects and developing multi-target detoxification strategies.
In terms of detection technology, although numerous methods have been developed and applied to STX detection, and LC-MS/MS has become the most reliable method for detecting STX, systematically comparing the advantages and disadvantages of the existing detection methods and scientifically classifying their applicable scenarios still holds significant practical significance for achieving the rapid and precise detection of STX.

Author Contributions

Writing—original draft preparation, H.D. and X.S.; writing—review and editing, M.S. and N.H.; supervision, H.Z.; investigation, L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (NSFC) under Grant No. 82173732 and the National Key Research and Development Program of China under Grant No. 2019YFC0312603. Lianghua Wang received financial support from the above grants and acknowledges their contribution.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were generated in this review.

Acknowledgments

The contributions of all reviewers and editors are deeply valued by the authors. The authors extend sincere gratitude to the peer reviewers for their meticulous and constructive suggestions on the sections of toxicological mechanisms, biosynthesis process, and detection methods significantly enhanced the scientific rigor and clarity of the draft. Special appreciation is also due to the editors for their continuous guidance and reminders throughout the review process, which ensured the manuscript’s compliance with academic standards and publication requirements. The authors extend heartfelt gratitude to Andrew Turner for his generous provision of references about the detection of paralytic shellfish toxins.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kellmann, R.; Mihali, T.K.; Jeon, Y.J.; Pickford, R.; Pomati, F.; Neilan, B.A. Biosynthetic Intermediate Analysis and Functional Homology Reveal a Saxitoxin Gene Cluster in Cyanobacteria. Appl. Environ. Microbiol. 2008, 74, 4044–4053. [Google Scholar] [CrossRef]
  2. Akbar, M.A.; Mohd, Y.N.; Tahir, N.I.; Ahmad, A.; Usup, G.; Sahrani, F.K.; Bunawan, H. Biosynthesis of Saxitoxin in Marine Dinoflagellates: An Omics Perspective. Mar. Drugs 2020, 18, 103. [Google Scholar] [CrossRef] [PubMed]
  3. Oshima, Y.; Hasegawa, M.; Yasumoto, T.; Hallegraeff, G.; Blackburn, S. Dinoflagellate Gymnodinium catenatum as the Source of Paralytic Shellfish Toxins in Tasmanian Shellfish. Toxicon 1987, 25, 1105–1111. [Google Scholar] [CrossRef]
  4. Anderson, D.M.; Alpermann, T.J.; Cembella, A.D.; Collos, Y.; Masseret, E.; Montresor, M. The Globally Distributed Genus Alexandrium: Multifaceted Roles in Marine Ecosystems and Impacts on Human Health. Harmful Algae 2012, 14, 10–35. [Google Scholar] [CrossRef] [PubMed]
  5. Bates, H.A.; Kostriken, R.; Rapoport, H. The Occurrence of Saxitoxin and Other Toxins in Various Dinoflagellates. Toxicon 1978, 16, 595–601. [Google Scholar] [CrossRef]
  6. Ciminiello, P.; Fattorusso, E.; Forino, M.; Montresor, M. Saxitoxin and Neosaxitoxin as Toxic Principles of Alexandrium andersoni (Dinophyceae) from the Gulf of Naples, Italy. Toxicon 2000, 38, 1871–1877. [Google Scholar] [CrossRef] [PubMed]
  7. Landsberg, J.H.; Hall, S.; Johannessen, J.N.; White, K.D.; Conrad, S.M.; Abbott, J.P.; Flewelling, L.J.; Richardson, R.W.; Dickey, R.W.; Jester, E.L.; et al. Saxitoxin Puffer Fish Poisoning in the United States, with the First Report of Pyrodinium bahamense as the Putative Toxin Source. Environ. Health Perspect. 2006, 114, 1502–1507. [Google Scholar] [CrossRef]
  8. Shin, H.H.; Li, Z.; Reveillon, D.; Rovillon, G.A.; Mertens, K.N.; Hess, P.; Kim, H.J.; Lee, J.; Lee, K.W.; Kim, D.; et al. Centrodinium punctatum (Dinophyceae) Produces Significant Levels of Saxitoxin and Related Analogs. Harmful Algae 2020, 100, 101923. [Google Scholar] [CrossRef]
  9. Llewellyn, L.E.; Negri, A.P.; Doyle, J.; Baker, P.D.; Beltran, E.C.; Neilan, B.A. Radioreceptor Assays for Sensitive Detection and Quantitation of Saxitoxin and Its Analogues from Strains of the Freshwater Cyanobacterium, Anabaena circinalis. Environ. Sci. Technol. 2001, 35, 1445–1451. [Google Scholar] [CrossRef]
  10. Ramos, T.K.; Costa, L.; Yunes, J.S.; Resgalla, C.J.; Barufi, J.B.; Bastos, E.O.; Horta, P.J.; Rörig, L.R. Saxitoxins from the Freshwater Cyanobacterium Raphidiopsis raciborskii Can Contaminate Marine Mussels. Harmful Algae 2021, 103, 102004. [Google Scholar] [CrossRef]
  11. Ballot, A.; Fastner, J.; Wiedner, C. Paralytic Shellfish Poisoning Toxin-Producing Cyanobacterium Aphanizomenon gracile in Northeast Germany. Appl. Environ. Microbiol. 2010, 76, 1173–1180. [Google Scholar] [CrossRef] [PubMed]
  12. Carmichael, W.W.; Evans, W.R.; Yin, Q.Q.; Bell, P.; Moczydlowski, E. Evidence for Paralytic Shellfish Poisons in the Freshwater Cyanobacterium Lyngbya wollei (Farlow Ex Gomont) Comb. Nov. Appl. Environ. Microbiol. 1997, 63, 3104–3110. [Google Scholar] [CrossRef] [PubMed]
  13. Christensen, V.G.; Khan, E. Freshwater Neurotoxins and Concerns for Human, Animal, and Ecosystem Health: A Review of Anatoxin-a and Saxitoxin. Sci. Total Environ. 2020, 736, 139515. [Google Scholar] [CrossRef]
  14. Pedrosa, C.; Souza, L.; Gomes, T.A.; de Lima, C.; Ledur, P.F.; Karmirian, K.; Barbeito-Andres, J.; Costa, M.; Higa, L.M.; Rossi, A.D.; et al. The Cyanobacterial Saxitoxin Exacerbates Neural Cell Death and Brain Malformations Induced by Zika Virus. PLoS Negl. Trop. Dis. 2020, 14, e0008060. [Google Scholar] [CrossRef] [PubMed]
  15. Farabegoli, F.; Blanco, L.; Rodriguez, L.P.; Vieites, J.M.; Cabado, A.G. Phycotoxins in Marine Shellfish: Origin, Occurrence and Effects on Humans. Mar. Drugs 2018, 16, 188. [Google Scholar] [CrossRef]
  16. Guillotin, S.; Delcourt, N. Marine Neurotoxins’ Effects on Environmental and Human Health: An OMICS Overview. Mar. Drugs 2021, 20, 18. [Google Scholar] [CrossRef]
  17. Watanabe, R.; Takayanagi, Y.; Chiba, O.; Itoda, S.; Ishizuka, H.; Odagi, M.; Ozawa, M.; Uchida, H.; Matsushima, R.; Konoki, K.; et al. Nontoxic Enantiomeric Reference Materials for Saxitoxins. Anal. Chem. 2022, 94, 11144–11150. [Google Scholar] [CrossRef]
  18. Morris, Z.J.; Stommel, E.W.; Metcalf, J.S. Airborne Cyanobacterial Toxins and Their Links to Neurodegenerative Diseases. Molecules 2025, 30, 2320. [Google Scholar] [CrossRef]
  19. Stern, D.B.; Raborn, R.T.; Lovett, S.P.; Boise, N.R.; Carrasquilla, L.; Enke, S.; Radune, D.; Woodruff, D.L.; Wahl, K.L.; Rosovitz, M.J. Novel Toxin Biosynthetic Gene Cluster in Harmful Algal Bloom-Causing Heteroscytonema crispum: Insights into the Origins of Paralytic Shellfish Toxins. Genome Biol. Evol. 2025, 17, evae248. [Google Scholar] [CrossRef]
  20. Zoffoli, G.; Pezzolesi, L.; Simonazzi, M.; Guerrini, F.; Vanucci, S.; Calfapietra, A.; Dall’Ara, S.; Servadei, I.; Pistocchi, R. A Decade-Long Study on Harmful Dinoflagellate Blooms and Biotoxin Contamination in Mussels from the North-Western Adriatic Sea (Mediterranean Sea). Harmful Algae 2025, 146, 102870. [Google Scholar] [CrossRef]
  21. Fachon, E.; Pickart, R.S.; Sheffield, G.; Pate, E.; Pathare, M.; Brosnahan, M.L.; Muhlbach, E.; Horn, K.; Spada, N.N.; Rajagopalan, A.; et al. Tracking a Large-Scale and Highly Toxic Arctic Algal Bloom: Rapid Detection and Risk Communication. Limnol. Oceanogr. Lett. 2025, 10, 62–72. [Google Scholar] [CrossRef]
  22. Diez-Chiappe, A.; Cirés, S.; Muñoz-Martín, M.Á.; Justel, A.; Quesada, A.; Perona, E. Unexpected Cyanobacterial Communities in Highly Heterogeneous Toxic Blooms from a Mediterranean Protected Area. Environ. Res. 2025, 270, 120953. [Google Scholar] [CrossRef]
  23. Villegas, R.; Barnola, F.V. Ionic Channels and Nerve Membrane Constituents. Tetrodotoxin-like Interaction of Saxitoxin with Cholesterol Monolayers. J. Gen. Physiol. 1972, 59, 33–46. [Google Scholar] [CrossRef]
  24. Barchi, R.L.; Weigele, J.B. Characteristics of Saxitoxin Binding to the Sodium Channel of Sarcolemma Isolated from Rat Skeletal Muscle. J. Physiol. 1979, 295, 383–396. [Google Scholar] [CrossRef]
  25. Duran-Riveroll, L.M.; Cembella, A.D. Guanidinium Toxins and Their Interactions with Voltage-Gated Sodium Ion Channels. Mar. Drugs 2017, 15, 303. [Google Scholar] [CrossRef]
  26. Hull, J.M.; Isom, L.L. Voltage-Gated Sodium Channel Beta Subunits: The Power Outside the Pore in Brain Development and Disease. Neuropharmacology 2018, 132, 43–57. [Google Scholar] [CrossRef]
  27. Yan, Z.; Zhou, Q.; Wang, L.; Wu, J.; Zhao, Y.; Huang, G.; Peng, W.; Shen, H.; Lei, J.; Yan, N. Structure of the Na(v)1.4-Beta1 Complex from Electric Eel. Cell 2017, 170, 470–482.e11. [Google Scholar] [CrossRef]
  28. Ghovanloo, M.R.; Ruben, P.C. Cannabidiol and Sodium Channel Pharmacology: General Overview, Mechanism, and Clinical Implications. Neuroscientist 2022, 28, 318–334. [Google Scholar] [CrossRef]
  29. Catterall, W.A.; Cestele, S.; Yarov-Yarovoy, V.; Yu, F.H.; Konoki, K.; Scheuer, T. Voltage-Gated Ion Channels and Gating Modifier Toxins. Toxicon 2007, 49, 124–141. [Google Scholar] [CrossRef]
  30. Su, Z.; Sheets, M.; Ishida, H.; Li, F.; Barry, W.H. Saxitoxin Blocks L-Type ICa. J. Pharmacol. Exp. Ther. 2004, 308, 324–329. [Google Scholar] [CrossRef]
  31. Wang, J.; Salata, J.J.; Bennett, P.B. Saxitoxin Is a Gating Modifier of HERG K+ Channels. J. Gen. Physiol. 2003, 121, 583–598. [Google Scholar] [CrossRef]
  32. Carnicero, B.; Fuentes, R.; Sanhueza, N.; Mattos, H.; Aguirre-Campos, C.; Contreras, D.; Troncoso, E.; Henriquez, J.P.; Boltana, S. Sublethal Neurotoxicity of Saxitoxin in Early Zebrafish Development: Impact on Sensorimotor Function and Neurotransmission Systems. Heliyon 2024, 10, e27874. [Google Scholar] [CrossRef]
  33. Sun, Q.; Chen, X.; Liu, W.; Li, S.; Zhou, Y.; Yang, X.; Liu, J. Effects of Long-Term Low Dose Saxitoxin Exposure on Nerve Damage in Mice. Aging 2021, 13, 17211–17226. [Google Scholar] [CrossRef]
  34. Schantz, E.J.; Mold, J.D.; Stanger, D.W.; Shavel, J.; Riel, F.J.; Bowden, J.P.; Lynch, J.M.; Wyler, R.S.; Riegel, B.; Sommer, H. Paralytic Shellfish Poison. VI. A Procedure for the Isolation and Purification of the Poison from Toxic Clam and Mussel Tissues. J. Am. Chem. Soc. 1957, 79, 5230–5235. [Google Scholar] [CrossRef]
  35. Schantz, E.J.; Ghazarossian, V.E.; Schnoes, H.K.; Strong, F.M.; Springer, J.P.; Pezzanite, J.O.; Clardy, J. Letter: The Structure of Saxitoxin. J. Am. Chem. Soc. 1975, 97, 1238. [Google Scholar] [CrossRef]
  36. Bordner, J.; Thiessen, W.E.; Bates, H.A.; Rapoport, H. The Structure of a Crystalline Derivative of Saxitoxin. Struct. Saxitoxin. J. Am. Chem. Soc. 1975, 97, 6008–6012. [Google Scholar] [CrossRef]
  37. Whitefleet-Smith, J.L.; Divan, C.L.; Schantz, E.J.; Schnoes, H.K. Distribution of Paralytic Toxins in California Shellfish. Toxicon 1985, 23, 346–349. [Google Scholar] [CrossRef]
  38. Tanino, H.; Nakata, T.; Kaneko, T.; Kishi, Y. A Stereospecific Total Synthesis of d,l-Saxitoxin. J. Am. Chem. Soc. 1977, 99, 2818–2819. [Google Scholar] [CrossRef]
  39. Fleming, J.J.; McReynolds, M.D.; Du Bois, J. (+)-Saxitoxin: A First and Second Generation Stereoselective Synthesis. J. Am. Chem. Soc. 2007, 129, 9964–9975. [Google Scholar] [CrossRef]
  40. Fleming, J.J.; Du Bois, J. A Synthesis of (+)-Saxitoxin. J. Am. Chem. Soc. 2006, 128, 3926–3927. [Google Scholar] [CrossRef]
  41. Iwamoto, O.; Shinohara, R.; Nagasawa, K. Total Synthesis of (-)- and (+)-Decarbamoyloxysaxitoxin and (+)-Saxitoxin. Chem. Asian J. 2009, 4, 277–285. [Google Scholar] [CrossRef]
  42. Iwamoto, O.; Nagasawa, K. Total Synthesis of (+)-Decarbamoylsaxitoxin and (+)-Gonyautoxin 3. Org. Lett. 2010, 12, 2150–2153. [Google Scholar] [CrossRef]
  43. Kao, C.Y.; Walker, S.E. Active Groups of Saxitoxin and Tetrodotoxin as Deduced from Actions of Saxitoxin Analogues on Frog Muscle and Squid Axon. J. Physiol. 1982, 323, 619–637. [Google Scholar] [CrossRef]
  44. Yang, L.; Kao, C.Y. Actions of Chiriquitoxin on Frog Skeletal Muscle Fibers and Implications for the Tetrodotoxin/Saxitoxin Receptor. J. Gen. Physiol. 1992, 100, 609–622. [Google Scholar] [CrossRef]
  45. Yang, L.; Kao, C.Y.; Oshima, Y. Actions of Decarbamoyloxysaxitoxin and Decarbamoylneosaxitoxin on the Frog Skeletal Muscle Fiber. Toxicon 1992, 30, 645–652. [Google Scholar] [CrossRef]
  46. Shimizu, Y.; Hsu, C.-P.; Genenah, A. Structure of Saxitoxin in Solutions and Stereochemistry of Dihydrosaxitoxins. J. Am. Chem. Soc. 1981, 103, 605–609. [Google Scholar] [CrossRef]
  47. Li, W.; Paladugu, S.R.; Liles, J.P.; Karthikeyan, M.; Chase, K.; Raghuraman, S.; Sigman, M.S.; Looper, R.E. Macrocycles of Saxitoxin: Insights into the Structure of Zetekitoxin AB. Chembiochem Eur. J. Chem. Biol. 2025, 26, e202500170. [Google Scholar] [CrossRef]
  48. Vicentini, M.; Calado, S.; Pessati, J.; Perussolo, M.C.; Lirola, J.R.; Marcondes, F.R.; Nascimento, N.D.; Beghetto, C.L.; Vilar, M.; Mela, M.; et al. Temperature Rise and Its Influence on the Toxic Effects Caused by Cyanotoxins in a Neotropical Catfish. Environ. Pollut. 2024, 353, 124166. [Google Scholar] [CrossRef]
  49. Leal, J.F.; Cristiano, M. Revisiting the HPLC-FLD Method to Quantify Paralytic Shellfish Toxins: C3,4 Quantification and the First Steps towards Validation. Toxins 2022, 14, 179. [Google Scholar] [CrossRef]
  50. Rogers, R.S.; Rapoport, H. The pKa’s of Saxitoxin. J. Am. Chem. Soc. 1980, 102, 7335–7339. [Google Scholar] [CrossRef]
  51. Nicholson, B.C.; Shaw, G.R.; Morrall, J.; Senogles, P.J.; Woods, T.A.; Papageorgiou, J.; Kapralos, C.; Wickramasinghe, W.; Davis, B.C.; Eaglesham, G.K.; et al. Chlorination for Degrading Saxitoxins (Paralytic Shellfish Poisons) in Water. Environ. Technol. 2003, 24, 1341–1348. [Google Scholar] [CrossRef]
  52. Bowers, E.K.; Stimmelmayr, R.; Hendrix, A.; Lefebvre, K.A. Stability of Saxitoxin in 50% Methanol Fecal Extracts and Raw Feces from Bowhead Whales (Balaena mysticetus). Mar. Drugs 2022, 20, 547. [Google Scholar] [CrossRef]
  53. Alkassar, M.; Tudo, A.; Rambla-Alegre, M.; Ferreres, L.; Diogene, J.; Sureda, F.X.; Campas, M. First Record of Paralytic Shellfish Toxins in Marine Pufferfish from the Spanish Mediterranean Coast Using Cell-Based Assay, Automated Patch Clamp and HPLC-FLD. Chemosphere 2024, 364, 143053. [Google Scholar] [CrossRef]
  54. Wiese, M.; D’Agostino, P.M.; Mihali, T.K.; Moffitt, M.C.; Neilan, B.A. Neurotoxic Alkaloids: Saxitoxin and Its Analogs. Mar. Drugs 2010, 8, 2185–2211. [Google Scholar] [CrossRef]
  55. Strichartz, G. Structural Determinants of the Affinity of Saxitoxin for Neuronal Sodium Channels. Electrophysiol. Stud. Frog Peripher. Nerve. J. Gen. Physiol. 1984, 84, 281–305. [Google Scholar] [CrossRef]
  56. Viallon, J.; Chinain, M.; Darius, H.T. Revisiting the Neuroblastoma Cell-Based Assay (CBA-N2a) for the Improved Detection of Marine Toxins Active on Voltage Gated Sodium Channels (VGSCs). Toxins 2020, 12, 281. [Google Scholar] [CrossRef]
  57. Yoshida, T.; Sako, Y.; Uchida, A.; Kakutani, T.; Arakawa, O.; Noguchi, T.; Ishida, Y. Purification and Characterization of Sulfotransferase Specific to O-22 of 11-Hydroxy Saxitoxin from the Toxic Dinoflagellate Gymnodinium catenatum (Dinophyceae). Fish. Sci. 2002, 68, 634–642. [Google Scholar] [CrossRef]
  58. Smith, E.A.; Grant, F.; Ferguson, C.M.; Gallacher, S. Biotransformations of Paralytic Shellfish Toxins by Bacteria Isolated from Bivalve Molluscs. Appl. Environ. Microbiol. 2001, 67, 2345–2353. [Google Scholar] [CrossRef]
  59. Mihali, T.K.; Kellmann, R.; Neilan, B.A. Characterisation of the Paralytic Shellfish Toxin Biosynthesis Gene Clusters in Anabaena circinalis AWQC131C and Aphanizomenon Sp. NH-5. BMC Biochem. 2009, 10, 8. [Google Scholar] [CrossRef]
  60. Geffroy, S.; Lechat, M.M.; Le Gac, M.; Rovillon, G.A.; Marie, D.; Bigeard, E.; Malo, F.; Amzil, Z.; Guillou, L.; Caruana, A. From the sxtA4 Gene to Saxitoxin Production: What Controls the Variability Among Alexandrium minutum and Alexandrium pacificum Strains? Front. Microbiol. 2021, 12, 613199. [Google Scholar] [CrossRef]
  61. Hackett, J.D.; Wisecaver, J.H.; Brosnahan, M.L.; Kulis, D.M.; Anderson, D.M.; Bhattacharya, D.; Plumley, F.G.; Erdner, D.L. Evolution of Saxitoxin Synthesis in Cyanobacteria and Dinoflagellates. Mol. Biol. Evol. 2013, 30, 70–78. [Google Scholar] [CrossRef]
  62. Murray, S.A.; Mihali, T.K.; Neilan, B.A. Extraordinary Conservation, Gene Loss, and Positive Selection in the Evolution of an Ancient Neurotoxin. Mol. Biol. Evol. 2011, 28, 1173–1182. [Google Scholar] [CrossRef]
  63. Vico, P.; Bonilla, S.; Cremella, B.; Aubriot, L.; Iriarte, A.; Piccini, C. Biogeography of the Cyanobacterium Raphidiopsis (Cylindrospermopsis) raciborskii: Integrating Genomics, Phylogenetic and Toxicity Data. Mol. Phylogenet. Evol. 2020, 148, 106824. [Google Scholar] [CrossRef]
  64. Ballot, A.; Cerasino, L.; Hostyeva, V.; Cires, S. Variability in the Sxt Gene Clusters of PSP Toxin Producing Aphanizomenon gracile Strains from Norway, Spain, Germany and North America. PLoS ONE 2016, 11, e0167552. [Google Scholar] [CrossRef]
  65. Hyung, J.-H.; Moon, S.J.; Kim, E.J.; Kim, D.W.; Park, J. Quantification of Alexandrium catenella (Group I) Using sxtA4-Based Digital PCR for Screening of Paralytic Shellfish Toxins in Jinhae-Masan Bay, Korea. Mar. Pollut. Bull. 2024, 200, 116048. [Google Scholar] [CrossRef]
  66. Kim, H.-S.; Kim, T.; Park, J.; Park, T.G.; Ki, J.-S. Development of Saxitoxin Biosynthesis Gene sxtB-Targeted qPCR Assay for the Quantification of Toxic Dinoflagellates Alexandrium catenella (Group I) and A. pacificum (Group IV) Occurring in the Korean Coast. Harmful Algae 2024, 134, 102603. [Google Scholar] [CrossRef]
  67. Muhammad, B.L.; Kim, H.S.; Bui, Q.; Ki, J.S. Transcriptomic Comparison Unveils Saxitoxin Biosynthesis Genes in the Marine Dinoflagellate Gymnodinium catenatum. Harmful Algae 2025, 147, 102872. [Google Scholar] [CrossRef]
  68. Bui, Q.; Kim, H.; Park, H.; Ki, J.S. Salinity Affects Saxitoxins (STXs) Toxicity in the Dinoflagellate Alexandrium pacificum, with Low Transcription of SXT-Biosynthesis Genes sxtA4 and sxtG. Toxins 2021, 13, 733. [Google Scholar] [CrossRef]
  69. Cires, S.; Delgado, A.; Gonzalez-Pleiter, M.; Quesada, A. Temperature Influences the Production and Transport of Saxitoxin and the Expression of Sxt Genes in the Cyanobacterium Aphanizomenon gracile. Toxins 2017, 9, 322. [Google Scholar] [CrossRef]
  70. Alonso-Rodríguez, R.; Pichardo-Velarde, J.G. Effects of Temperature and Nutrients on Growth and Toxicity of Alexandrium affine from Southeastern Gulf of California. Mar. Pollut. Bull. 2024, 203, 116464. [Google Scholar] [CrossRef]
  71. Wang, H.; Kim, H.; Park, H.; Ki, J.S. Temperature Influences the Content and Biosynthesis Gene Expression of Saxitoxins (STXs) in the Toxigenic Dinoflagellate Alexandrium pacificum. Sci. Total Environ. 2022, 802, 149801. [Google Scholar] [CrossRef]
  72. Oh, S.J.; Kim, D.I.; Sajima, T.; Shimasaki, Y.; Matsuyama, Y.; Oshima, Y.; Honjo, T.; Yang, H.S. Effects of Irradiance of Various Wavelengths from Light-Emitting Diodes on the Growth of the Harmful Dinoflagellate Heterocapsa circularisquama and the Diatom Skeletonema costatum. Fish. Sci. 2008, 74, 137–145. [Google Scholar] [CrossRef]
  73. Rivkin, R.B. Influence of Irradiance and Spectral Quality on the Carbon Metabolism of Phytoplankton I. Photosynthesis, Chemical Composition and Growth. Mar. Ecol. Prog. 1989, 55, 291–304. [Google Scholar] [CrossRef]
  74. Mesquita, M.; Lurling, M.; Dorr, F.; Pinto, E.; Marinho, M.M. Combined Effect of Light and Temperature on the Production of Saxitoxins in Cylindrospermopsis raciborskii Strains. Toxins 2019, 11, 38. [Google Scholar] [CrossRef]
  75. Kim, H.; Park, H.; Wang, H.; Yoo, H.Y.; Park, J.; Ki, J.S. Low Temperature and Cold Stress Significantly Increase Saxitoxins (STXs) and Expression of STX Biosynthesis Genes sxtA4 and sxtG in the Dinoflagellate Alexandrium catenella. Mar. Drugs 2021, 19, 291. [Google Scholar] [CrossRef]
  76. Eckford-Soper, L.K.; Bresnan, E.; Lacaze, J.P.; Green, D.H.; Davidson, K. The Competitive Dynamics of Toxic Alexandrium fundyense and Non-Toxic Alexandrium tamarense: The Role of Temperature. Harmful Algae 2016, 53, 135–144. [Google Scholar] [CrossRef]
  77. Lim, P.T.; Leaw, C.P.; Kobiyama, A.; Ogata, T. Growth and Toxin Production of Tropical Alexandrium minutum Halim (Dinophyceae) under Various Nitrogen to Phosphorus Ratios. J. Appl. Phycol. 2010, 22, 203–210. [Google Scholar] [CrossRef]
  78. Akbar, M.A.; Mohd, Y.N.; Usup, G.; Ahmad, A.; Baharum, S.N.; Bunawan, H. Nutrient Deficiencies Impact on the Cellular and Metabolic Responses of Saxitoxin Producing Alexandrium minutum: A Transcriptomic Perspective. Mar. Drugs 2023, 21, 497. [Google Scholar] [CrossRef]
  79. Abassi, S.; Kim, H.-S.; Bui, Q.T.N.; Ki, J.-S. Effects of Nitrate on the Saxitoxins Biosynthesis Revealed by Sxt Genes in the Toxic Dinoflagellate Alexandrium pacificum (Group IV). Harmful Algae 2023, 127, 102473. [Google Scholar] [CrossRef]
  80. Vilar, M.; Rodrigues, T.; Silva, L.O.; Pacheco, A.; Ferrao-Filho, A.S.; Azevedo, S. Ecophysiological Aspects and Sxt Genes Expression Underlying Induced Chemical Defense in STX-Producing Raphidiopsis raciborskii (Cyanobacteria) against the Zooplankter Daphnia Gessneri. Toxins 2021, 13, 406. [Google Scholar] [CrossRef]
  81. Jiang, Y.; Ramanan, R.; Yoon, S.; Lee, B.M.; Kang, Y.H.; Li, Z. Toxin Production in Bloom-Forming, Harmful Alga Alexandrium pacificum (Group IV) Is Regulated by Cyst Formation-Promoting Bacteria Jannaschia cystaugens NBRC 100362(T). Water Res. 2025, 272, 122930. [Google Scholar] [CrossRef]
  82. Soeriyadi, A.H.; Mazmouz, R.; Pickford, R.; Al-Sinawi, B.; Kellmann, R.; Pearson, L.A.; Neilan, B.A. Heterologous Expression of an Unusual Ketosynthase, SxtA, Leads to Production of Saxitoxin Intermediates in Escherichia Coli. Chembiochem Eur. J. Chem. Biol. 2021, 22, 845–849. [Google Scholar] [CrossRef]
  83. D’Agostino, P.M.; Al-Sinawi, B.; Mazmouz, R.; Muenchhoff, J.; Neilan, B.A.; Moffitt, M.C. Identification of Promoter Elements in the Dolichospermum circinale AWQC131C Saxitoxin Gene Cluster and the Experimental Analysis of Their Use for Heterologous Expression. BMC Microbiol. 2020, 20, 35. [Google Scholar] [CrossRef]
  84. Jiao, Y.; Liu, J.; Mao, W.; Fang, R.; Xia, T.; Lang, Q.; Luo, T. Synthesis of (+)-Saxitoxin Facilitated by a Chiral Auxiliary for Photocycloadditions Involving Alkenylboronate Esters. J. Am. Chem. Soc. 2025, 147, 9091–9097. [Google Scholar] [CrossRef]
  85. Greenhough, H.; Waugh, C.; van Ginkel, R.; Bowater, J.; Kaur, G.; Oakly, J.; Plouviez, M.; Ingebrigtsen, R.A.; Svenson, J.; Selwood, A.I. Mass Cultivation of the Dinoflagellate Alexandrium pacificum for Gonyautoxin-1,4 Production. Sci. Rep. 2025, 15, 7430. [Google Scholar] [CrossRef]
  86. Pućko, M.; Rourke, W.; Hussherr, R.; Archambault, P.; Eert, J.; Majewski, A.R.; Niemi, A.; Reist, J.; Michel, C. Phycotoxins in Bivalves from the Western Canadian Arctic: The First Evidence of Toxigenicity. Harmful Algae 2023, 127, 102474. [Google Scholar] [CrossRef]
  87. Fraga, M.; Churro, C.; Leão-Martins, J.; Rudnitskaya, A.; Botelho, M.J. Cyanotoxins on the Move—Freshwater Origins with Marine Consequences: A Systematic Review of Global Changes and Emerging Trends. Mar. Pollut. Bull. 2025, 216, 118017. [Google Scholar] [CrossRef] [PubMed]
  88. Aboualaalaa, H.; Rijal Leblad, B.; Elkbiach, M.L.; Ibghi, M.; Boutaib, R.; Maamour, N.; Savar, V.; Masseret, E.; Abadie, E.; Rolland, J.L.; et al. Effect of Temperature, Salinity and Nutrients on the Growth and Toxin Content of the Dinoflagellate Gymnodinium catenatum from the Southwestern Mediterranean. Sci. Total Environ. 2024, 945, 174094. [Google Scholar] [CrossRef]
  89. Vianney, M.H.; Sébastien, O.K.; Kafoumba, B.; Dangui, D.; Olivier, K. Assessment of the Health Risks Associated with the Consumption of Bivalve Mollusks Potentially Contaminated with Phycotoxins from the Coastal Ecosystem of the Ebrié Lagoon, Côte d’Ivoire. Environ. Monit. Assess. 2024, 196, 259. [Google Scholar] [CrossRef]
  90. Da, S.C.; Oba, E.T.; Ramsdorf, W.A.; Magalhaes, V.F.; Cestari, M.M.; Oliveira, R.C.; Silva, D.A.H. First Report about Saxitoxins in Freshwater Fish Hoplias malabaricus through Trophic Exposure. Toxicon 2011, 57, 141–147. [Google Scholar] [CrossRef]
  91. Niu, Y.; Wei, H.; Zhang, Y.; Su, J. Transcriptome Response of a Marine Copepod in Response to Environmentally-Relevant Concentrations of Saxitoxin. Mar. Pollut. Bull. 2024, 205, 116546. [Google Scholar] [CrossRef] [PubMed]
  92. Barbosa, V.; Santos, M.; Anacleto, P.; Maulvault, A.L.; Pousao-Ferreira, P.; Costa, P.R.; Marques, A. Paralytic Shellfish Toxins and Ocean Warming: Bioaccumulation and Ecotoxicological Responses in Juvenile Gilthead Seabream (Sparus aurata). Toxins 2019, 11, 408. [Google Scholar] [CrossRef]
  93. Sinno-Tellier, S.; Abadie, E.; Haro, L.D.; Paret, N.; Langrand, J.; Roux, G.L.; Labadie, M.; Boels, D.; Bloch, J.; Delcourt, N. Human Poisonings by Neurotoxic Phycotoxins Related to the Consumption of Shellfish: Study of Cases Registered by the French Poison Control Centres from 2012 to 2019. Clin. Toxicol. 2022, 60, 759–767. [Google Scholar] [CrossRef] [PubMed]
  94. Cadaillon, A.M.; Mattera, B.; Albizzi, A.; Montoya, N.; Maldonado, S.; Raya Rey, A.; Riccialdelli, L.; Almandoz, G.O.; Schloss, I.R. Multispecies Mass Mortality in the Beagle Channel Associated with Paralytic Shellfish Toxins. Harmful Algae 2024, 132, 102581. [Google Scholar] [CrossRef]
  95. Zheng, G.; Che, H.; Wu, H.; Deng, Y.; Guo, M.; Peng, J.; Geng, Q.; Tan, Z. Metabolic Transformation of Paralytic Shellfish Toxins in the Mussel Mytilus galloprovincialis under Different Exposure Modes. Harmful Algae 2025, 141, 102771. [Google Scholar] [CrossRef] [PubMed]
  96. Hemert, C.V.; Schoen, S.K.; Litaker, R.W.; Smith, M.M.; Pearce, J.M. Algal Toxins in Alaskan Seabirds: Evaluating the Role of Saxitoxin and Domoic Acid in a Large-Scale Die-off of Common Murres. Harmful Algae 2020, 92, 101730. [Google Scholar] [CrossRef]
  97. Shearn-Bochsler, V.; Lance, E.W.; Corcoran, R.; Piatt, J.; Bodenstein, B.; Frame, E.; Lawonn, J. Fatal Paralytic Shellfish Poisoning in Kittlitz’s Murrelet (Brachyramphus brevirostris) Nestlings, Alaska, USA. J. Wildl. Dis. 2014, 50, 933–937. [Google Scholar] [CrossRef]
  98. Hall, A.J.; Kershaw, J.L.; Fraser, S.; Davidson, K.; Rowland-Pilgrim, S.; Turner, A.D.; McConnell, B. Estimating the Risks of Exposure to Harmful Algal Toxins among Scottish Harbour Seals. Harmful Algae 2024, 136, 102653. [Google Scholar] [CrossRef]
  99. Aluru, N.; Chapman, D.P.; Becker, K.W.; Van Mooy, B.; Karchner, S.I.; Stegeman, J.J.; Hahn, M.E. Developmental Exposure of Zebrafish to Saxitoxin Causes Altered Expression of Genes Associated with Axonal Growth. Neurotoxicology 2024, 105, 303–312. [Google Scholar] [CrossRef]
  100. Litaker, R.W.; Matweyou, J.A.; Kibler, S.R.; Hardison, D.R.; Holland, W.C.; Tester, P.A. Paralytic Shellfish Toxins in Alaskan Butter Clams: Does Cleaning Make Them Safe to Eat? Toxins 2025, 17, 271. [Google Scholar] [CrossRef]
  101. Coleman, R.M.; Ojeda-Torres, G.; Bragg, W.; Fearey, D.; McKinney, P.; Castrodale, L.; Verbrugge, D.; Stryker, K.; DeHart, E.; Cooper, M.; et al. Saxitoxin Exposure Confirmed by Human Urine and Food Analysis. J. Anal. Toxicol. 2018, 42, e61–e64. [Google Scholar] [CrossRef] [PubMed]
  102. Vilarino, N.; Louzao, M.C.; Abal, P.; Cagide, E.; Carrera, C.; Vieytes, M.R.; Botana, L.M. Human Poisoning from Marine Toxins: Unknowns for Optimal Consumer Protection. Toxins 2018, 10, 324. [Google Scholar] [CrossRef]
  103. Lebad, M.S.; Ballot, A.; Vogelsang, C.; Abdessemed, D.; Krzeminski, P. Removal of a Cyanotoxins Mixture by Loose Nanofiltration Membranes Applied in Drinking Water Production. J. Water Process Eng. 2024, 57, 104694. [Google Scholar] [CrossRef]
  104. Ho, L.; Tanis-Plant, P.; Kayal, N.; Slyman, N.; Newcombe, G. Optimising Water Treatment Practices for the Removal of Anabaena circinalis and Its Associated Metabolites, Geosmin and Saxitoxins. J. Water Health 2009, 7, 544–556. [Google Scholar] [CrossRef]
  105. Orr, P.T.; Jones, G.J.; Hamilton, G.R. Removal of Saxitoxins from Drinking Water by Granular Activated Carbon, Ozone and Hydrogen Peroxide—Implications for Compliance with the Australian Drinking Water Guidelines. Water Res. 2004, 38, 4455–4461. [Google Scholar] [CrossRef] [PubMed]
  106. Gobler, C.J.; Kramer, B.J.; Lusty, M.W.; Thraen, J.; McTague, S. The Ability of Hydrogen Peroxide (H2O2) to Degrade Saxitoxin-, Microcystin-, Anatoxin-, and Non-Toxin-Producing Strains of the Harmful Cyanobacterium, Dolichospermum. J. Environ. Manage. 2025, 387, 125696. [Google Scholar] [CrossRef]
  107. Donovan, C.J.; Ku, J.C.; Quilliam, M.A.; Gill, T.A. Bacterial Degradation of Paralytic Shellfish Toxins. Toxicon 2008, 52, 91–100. [Google Scholar] [CrossRef]
  108. Chambers, C.; Grimes, S.; Smith, R.C.; Weil, A.; Reza, M.T. Investigation of Adsorption Parameters of Saxitoxin onto Loblolly Pine-Derived Biochar Synthesized at Various Pyrolysis Temperature. Chemosphere 2025, 370, 143965. [Google Scholar] [CrossRef] [PubMed]
  109. Gawankar, S.; Masten, S.J.; Lahr, R.H. Review of the Occurrence, Treatment Technologies, and Detection Methods for Saxitoxins in Freshwaters. J. Water Health 2024, 22, 1472–1490. [Google Scholar] [CrossRef]
  110. Heggeness, S.T.; Starkus, J.G. Saxitoxin and Tetrodotoxin. Electrostatic Effects on Sodium Channel Gating Current in Crayfish Axons. Biophys. J. 1986, 49, 629–643. [Google Scholar] [CrossRef]
  111. Catterall, W.A. Cellular and Molecular Biology of Voltage-Gated Sodium Channels. Physiol. Rev. 1992, 72, S15-48. [Google Scholar] [CrossRef] [PubMed]
  112. Hartshorne, R.P.; Catterall, W.A. The Sodium Channel from Rat Brain. Purification and Subunit Composition. J. Biol. Chem. 1984, 259, 1667–1675. [Google Scholar] [CrossRef] [PubMed]
  113. Tikhonov, D.B.; Zhorov, B.S. Predicting Structural Details of the Sodium Channel Pore Basing on Animal Toxin Studies. Front. Pharmacol. 2018, 9, 880. [Google Scholar] [CrossRef] [PubMed]
  114. Pan, X.; Li, Z.; Zhou, Q.; Shen, H.; Wu, K.; Huang, X.; Chen, J.; Zhang, J.; Zhu, X.; Lei, J.; et al. Structure of the Human Voltage-Gated Sodium Channel Na(v)1.4 in Complex with Beta1. Science 2018, 362, eaau2486. [Google Scholar] [CrossRef]
  115. Yanagawa, Y.; Abe, T.; Satake, M. Mu-Conotoxins Share a Common Binding Site with Tetrodotoxin/Saxitoxin on Eel Electroplax Na Channels. J. Neurosci. 1987, 7, 1498–1502. [Google Scholar] [CrossRef]
  116. Shen, H.; Li, Z.; Jiang, Y.; Pan, X.; Wu, J.; Cristofori-Armstrong, B.; Smith, J.J.; Chin, Y.; Lei, J.; Zhou, Q.; et al. Structural Basis for the Modulation of Voltage-Gated Sodium Channels by Animal Toxins. Science 2018, 362, eaau2596. [Google Scholar] [CrossRef]
  117. Catterall, W.A. From Ionic Currents to Molecular Mechanisms: The Structure and Function of Voltage-Gated Sodium Channels. Neuron 2000, 26, 13–25. [Google Scholar] [CrossRef]
  118. Shcherbatko, A.; Ono, F.; Mandel, G.; Brehm, P. Voltage-Dependent Sodium Channel Function Is Regulated through Membrane Mechanics. Biophys. J. 1999, 77, 1945–1959. [Google Scholar] [CrossRef]
  119. Kao, C.Y.; Suzuki, T.; Kleinhaus, A.L.; Siegman, M.J. Vasomotor and Respiratory Depressant Actions of Tetrodotoxin and Saxitoxin. Arch. Int. Pharmacodyn. Ther. 1967, 165, 438–450. [Google Scholar]
  120. Llewellyn, L.E. Saxitoxin, a Toxic Marine Natural Product That Targets a Multitude of Receptors. Nat. Prod. Rep. 2006, 23, 200–222. [Google Scholar] [CrossRef]
  121. Coward, K.; Jowett, A.; Plumpton, C.; Powell, A.; Birch, R.; Tate, S.; Bountra, C.; Anand, P. Sodium Channel Beta1 and Beta2 Subunits Parallel SNS/PN3 Alpha-Subunit Changes in Injured Human Sensory Neurons. Neuroreport 2001, 12, 483–488. [Google Scholar] [CrossRef] [PubMed]
  122. Messner, D.J.; Catterall, W.A. The Sodium Channel from Rat Brain. Role of the Beta 1 and Beta 2 Subunits in Saxitoxin Binding. J. Biol. Chem. 1986, 261, 211–215. [Google Scholar] [CrossRef] [PubMed]
  123. Goldfarb, M. Voltage-Gated Sodium Channel-Associated Proteins and Alternative Mechanisms of Inactivation and Block. Cell Mol. Life Sci. 2012, 69, 1067–1076. [Google Scholar] [CrossRef]
  124. Ritchie, J.M.; Rogart, R.B. The Binding of Saxitoxin and Tetrodotoxin to Excitable Tissue. Rev. Physiol. Biochem. Pharmacol. 1977, 79, 1–50. [Google Scholar] [CrossRef]
  125. Terlau, H.; Heinemann, S.H.; Stuhmer, W.; Pusch, M.; Conti, F.; Imoto, K.; Numa, S. Mapping the Site of Block by Tetrodotoxin and Saxitoxin of Sodium Channel II. FEBS Lett. 1991, 293, 93–96. [Google Scholar] [CrossRef]
  126. Moczydlowski, E.; Olivera, B.M.; Gray, W.R.; Strichartz, G.R. Discrimination of Muscle and Neuronal Na-Channel Subtypes by Binding Competition between [3H]Saxitoxin and Mu-Conotoxins. Proc. Natl. Acad. Sci. USA 1986, 83, 5321–5325. [Google Scholar] [CrossRef] [PubMed]
  127. Catterall, W.A.; Coppersmith, J. High-Affinity Saxitoxin Receptor Sites in Vertebrate Heart. Evidence for Sites Associated with Autonomic Nerve Endings. Mol. Pharmacol. 1981, 20, 526–532. [Google Scholar] [CrossRef]
  128. Yen, T.J.; Lolicato, M.; Thomas-Tran, R.; Du Bois, J.; Minor, D.J. Structure of the Saxiphilin:Saxitoxin (STX) Complex Reveals a Convergent Molecular Recognition Strategy for Paralytic Toxins. Sci. Adv. 2019, 5, eaax2650. [Google Scholar] [CrossRef]
  129. Biessy, L.; Boundy, M.J.; Smith, K.F.; Harwood, D.T.; Hawes, I.; Wood, S.A. Tetrodotoxin in Marine Bivalves and Edible Gastropods: A Mini-Review. Chemosphere 2019, 236, 124404. [Google Scholar] [CrossRef]
  130. Walker, J.R.; Novick, P.A.; Parsons, W.H.; McGregor, M.; Zablocki, J.; Pande, V.S.; Du Bois, J. Marked Difference in Saxitoxin and Tetrodotoxin Affinity for the Human Nociceptive Voltage-Gated Sodium Channel (Nav1.7) [Corrected]. Proc. Natl. Acad. Sci. USA 2012, 109, 18102–18107. [Google Scholar] [CrossRef]
  131. Boente-Juncal, A.; Otero, P.; Rodriguez, I.; Camina, M.; Rodriguez-Vieytes, M.; Vale, C.; Botana, L.M. Oral Chronic Toxicity of the Safe Tetrodotoxin Dose Proposed by the European Food Safety Authority and Its Additive Effect with Saxitoxin. Toxins 2020, 12, 312. [Google Scholar] [CrossRef] [PubMed]
  132. Damsch, S.; Eichenbaum, G.; Tonelli, A.; Lammens, L.; Van den Bulck, K.; Feyen, B.; Vandenberghe, J.; Megens, A.; Knight, E.; Kelley, M. Gavage-Related Reflux in Rats: Identification, Pathogenesis, and Toxicological Implications (Review). Toxicol. Pathol. 2011, 39, 348–360. [Google Scholar] [CrossRef] [PubMed]
  133. Finch, S.C.; Webb, N.G.; Boundy, M.J.; Harwood, D.T.; Munday, J.S.; Sprosen, J.M.; Somchit, C.; Broadhurst, R.B. A Sub-Acute Dosing Study of Saxitoxin and Tetrodotoxin Mixtures in Mice Suggests That the Current Paralytic Shellfish Toxin Regulatory Limit Is Fit for Purpose. Toxins 2023, 15, 437. [Google Scholar] [CrossRef]
  134. Finch, S.C.; Boundy, M.J.; Harwood, D.T. The Acute Toxicity of Tetrodotoxin and Tetrodotoxin(-)Saxitoxin Mixtures to Mice by Various Routes of Administration. Toxins 2018, 10, 423. [Google Scholar] [CrossRef]
  135. Donaghy, M. Neurologists and the Threat of Bioterrorism. J. Neurol. Sci. 2006, 249, 55–62. [Google Scholar] [CrossRef]
  136. Almabruk, K.H.; Dinh, L.K.; Philmus, B. Self-Resistance of Natural Product Producers: Past, Present, and Future Focusing on Self-Resistant Protein Variants. ACS Chem. Biol. 2018, 13, 1426–1437. [Google Scholar] [CrossRef] [PubMed]
  137. Hunter, P. Do Not Poison Thyself: Mechanisms to Avoid Self-Toxicity Could Inspire Novel Compounds and Pathways for Synthetic Biology and Applications for Agriculture. EMBO Rep. 2018, 19, e46756. [Google Scholar] [CrossRef]
  138. Tarvin, R.D.; Borghese, C.M.; Sachs, W.; Santos, J.C.; Lu, Y.; O’Connell, L.A.; Cannatella, D.C.; Harris, R.A.; Zakon, H.H. Interacting Amino Acid Replacements Allow Poison Frogs to Evolve Epibatidine Resistance. Science 2017, 357, 1261–1266. [Google Scholar] [CrossRef]
  139. Bricelj, V.M.; Connell, L.; Konoki, K.; Macquarrie, S.P.; Scheuer, T.; Catterall, W.A.; Trainer, V.L. Sodium Channel Mutation Leading to Saxitoxin Resistance in Clams Increases Risk of PSP. Nature 2005, 434, 763–767. [Google Scholar] [CrossRef]
  140. Satin, J.; Kyle, J.W.; Chen, M.; Bell, P.; Cribbs, L.L.; Fozzard, H.A.; Rogart, R.B. A Mutant of TTX-Resistant Cardiac Sodium Channels with TTX-Sensitive Properties. Science 1992, 256, 1202–1205. [Google Scholar] [CrossRef]
  141. Mahar, J.; Lukacs, G.L.; Li, Y.; Hall, S.; Moczydlowski, E. Pharmacological and Biochemical Properties of Saxiphilin, a Soluble Saxitoxin-Binding Protein from the Bullfrog (Rana catesbeiana). Toxicon 1991, 29, 53–71. [Google Scholar] [CrossRef] [PubMed]
  142. Kao, C.Y.; Fuhrman, F.A. Differentiation of the Actions of Tetrodotoxin and Saxitoxin. Toxicon 1967, 5, 25–34. [Google Scholar] [CrossRef] [PubMed]
  143. Doyle, D.D.; Wong, M.; Tanaka, J.; Barr, L. Saxitoxin Binding Sites in Frog-Myocardial Cytosol. Science 1982, 215, 1117–1119. [Google Scholar] [CrossRef] [PubMed]
  144. Morabito, M.A.; Moczydlowski, E. Molecular Cloning of Bullfrog Saxiphilin: A Unique Relative of the Transferrin Family That Binds Saxitoxin. Proc. Natl. Acad. Sci. USA 1995, 92, 6651. [Google Scholar] [CrossRef]
  145. Llewellyn, L.E.; Moczydlowski, E.G. Characterization of Saxitoxin Binding to Saxiphilin, a Relative of the Transferrin Family That Displays pH-Dependent Ligand Binding. Biochemistry 1994, 33, 12312–12322. [Google Scholar] [CrossRef]
  146. Chen, Z.; Zakrzewska, S.; Hajare, H.S.; Alvarez-Buylla, A.; Abderemane-Ali, F.; Bogan, M.; Ramirez, D.; O’Connell, L.A.; Du Bois, J.; Minor, D.J. Definition of a Saxitoxin (STX) Binding Code Enables Discovery and Characterization of the Anuran Saxiphilin Family. Proc. Natl. Acad. Sci. USA 2022, 119, e2210114119. [Google Scholar] [CrossRef]
  147. Shimizu, Y.; Yoshioka, M. Transformation of Paralytic Shellfish Toxins as Demonstrated in Scallop Homogenates. Science 1981, 212, 547–549. [Google Scholar] [CrossRef]
  148. Sullivan, J.J.; Iwaoka, W.T.; Liston, J. Enzymatic Transformation of PSP Toxins in the Littleneck Clam (Protothaca staminea). Biochem. Biophys. Res. Commun. 1983, 114, 465–472. [Google Scholar] [CrossRef]
  149. Raposo, M.I.C.; Gomes, M.T.S.R.; Botelho, M.J.; Rudnitskaya, A. Paralytic Shellfish Toxins (PST)-Transforming Enzymes: A Review. Toxins 2020, 12, 344. [Google Scholar] [CrossRef]
  150. Vasama, M.; Kumar, H.; Salminen, S.; Haskard, C.A. Removal of Paralytic Shellfish Toxins by Probiotic Lactic Acid Bacteria. Toxins 2014, 6, 2127–2136. [Google Scholar] [CrossRef]
  151. Donovan, C.J.; Garduno, R.A.; Kalmokoff, M.; Ku, J.C.; Quilliam, M.A.; Gill, T.A. Pseudoalteromonas Bacteria Are Capable of Degrading Paralytic Shellfish Toxins. Appl. Environ. Microbiol. 2009, 75, 6919–6923. [Google Scholar] [CrossRef] [PubMed]
  152. Dell’Aversano, C.; Walter, J.A.; Burton, I.W.; Stirling, D.J.; Fattorusso, E.; Quilliam, M.A. Isolation and Structure Elucidation of New and Unusual Saxitoxin Analogues from Mussels. J. Nat. Prod. 2008, 71, 1518–1523. [Google Scholar] [CrossRef] [PubMed]
  153. Ding, L.; Qiu, J.; Li, A. Proposed Biotransformation Pathways for New Metabolites of Paralytic Shellfish Toxins Based on Field and Experimental Mussel Samples. J. Agric. Food Chem. 2017, 65, 5494–5502. [Google Scholar] [CrossRef]
  154. Qiu, J.; Meng, F.; Ding, L.; Che, Y.; McCarron, P.; Beach, D.G.; Li, A. Dynamics of Paralytic Shellfish Toxins and Their Metabolites during Timecourse Exposure of Scallops Chlamys Farreri and Mussels Mytilus galloprovincialis to Alexandrium pacificum. Aquat. Toxicol. 2018, 200, 233–240. [Google Scholar] [CrossRef]
  155. Choudhary, G.; Yotsu-Yamashita, M.; Shang, L.; Yasumoto, T.; Dudley, S.J. Interactions of the C-11 Hydroxyl of Tetrodotoxin with the Sodium Channel Outer Vestibule. Biophys. J. 2003, 84, 287–294. [Google Scholar] [CrossRef] [PubMed]
  156. Che, Y.; Ding, L.; Qiu, J.; Ji, Y.; Li, A. Conversion and Stability of New Metabolites of Paralytic Shellfish Toxins under Different Temperature and pH Conditions. J. Agric. Food Chem. 2020, 68, 1427–1435. [Google Scholar] [CrossRef]
  157. Turner, A.D.; Hatfield, R.G.; Rapkova, M.; Higman, W.; Algoet, M.; Suarez-Isla, B.A.; Cordova, M.; Caceres, C.; van de Riet, J.; Gibbs, R.; et al. Comparison of AOAC 2005.06 LC Official Method with Other Methodologies for the Quantitation of Paralytic Shellfish Poisoning Toxins in UK Shellfish Species. Anal. Bioanal. Chem. 2011, 399, 1257–1270. [Google Scholar] [CrossRef]
  158. Ferrao-Filho, A.S.; Soares, M.C.; de Magalhaes, V.F.; Azevedo, S.M. A Rapid Bioassay for Detecting Saxitoxins Using a Daphnia Acute Toxicity Test. Environ. Pollut. 2010, 158, 2084–2093. [Google Scholar] [CrossRef]
  159. Aballay-Gonzalez, A.; Gallardo-Rodriguez, J.J.; Silva-Higuera, M.; Rivera, A.; Ulloa, V.; Delgado-Rivera, L.; Rivera-Belmar, A.; Astuya, A. Neuro-2a Cell-Based Assay for Toxicity Equivalency Factor—Proposal and Evaluation in Chilean Contaminated Shellfish Samples. Food Addit. Contam. Part. Chem. Anal. Control Expo. Risk Assess. 2020, 37, 162–173. [Google Scholar] [CrossRef]
  160. Ruebhart, D.R.; Radcliffe, W.L.; Eaglesham, G.K. Alternative Bioassay for the Detection of Saxitoxin Using the Speckled Cockroach (Nauphoeta cinerea). J. Toxicol. Environ. Health A 2011, 74, 621–637. [Google Scholar] [CrossRef]
  161. Jellett, J.F.; Stewart, J.E.; Laycock, M.V. Toxicological Evaluation of Saxitoxin, Neosaxitoxin, Gonyautoxin II, Gonyautoxin II plus III and Decarbamoylsaxitoxin with the Mouse Neuroblastoma Cell Bioassay. Toxicol. Vitr. 1995, 9, 57–65. [Google Scholar] [CrossRef] [PubMed]
  162. Manger, R.L.; Leja, L.S.; Lee, S.Y.; Hungerford, J.M.; Hokama, Y.; Dickey, R.W.; Granade, H.R.; Lewis, R.; Yasumoto, T.; Wekell, M.M. Detection of Sodium Channel Toxins: Directed Cytotoxicity Assays of Purified Ciguatoxins, Brevetoxins, Saxitoxins, and Seafood Extracts. J. AOAC Int. 1995, 78, 521–527. [Google Scholar] [CrossRef]
  163. Manger, R.L.; Leja, L.S.; Lee, S.Y.; Hungerford, J.M.; Kirkpatrick, M.A.; Yasumoto, T.; Wekell, M.M. Detection of Paralytic Shellfish Poison by Rapid Cell Bioassay: Antagonism of Voltage-Gated Sodium Channel Active Toxins in Vitro. J. AOAC Int. 2003, 86, 540–543. [Google Scholar] [CrossRef] [PubMed]
  164. Okumura, M.; Tsuzuki, H.; Tomita, B. A Rapid Detection Method for Paralytic Shellfish Poisoning Toxins by Cell Bioassay. Toxicon 2005, 46, 93–98. [Google Scholar] [CrossRef]
  165. Charapata, P.; Bowers, E.K.; Hardison, D.R.; Kibler, S.; Anderson, D.M.; Fachon, E.; Lefebvre, K.A. Paralytic Shellfish Toxin Concentrations Measured in Alaskan Arctic Clams Using ELISA and HPLC Methods. Toxins 2025, 17, 60. [Google Scholar] [CrossRef]
  166. Chu, F.S.; Fan, T.S. Indirect Enzyme-Linked Immunosorbent Assay for Saxitoxin in Shellfish. J. Assoc. Anal. Chem. 1985, 68, 13–16. [Google Scholar] [CrossRef]
  167. Mackintosh, F.H.; Gallacher, S.; Shanks, A.M.; Smith, E.A. Assessment of MIST AlertTM, a Commercial Qualitative Assay for Detection of Paralytic Shellfish Poisoning Toxins in Bivalve Molluscs. J. AOAC Int. 2002, 85, 632–641. [Google Scholar] [CrossRef] [PubMed]
  168. Jellett, J.F.; Roberts, R.L.; Laycock, M.V.; Quilliam, M.A.; Barrett, R.E. Detection of Paralytic Shellfish Poisoning (PSP) Toxins in Shellfish Tissue Using MIST Alert, a New Rapid Test, in Parallel with the Regulatory AOAC Mouse Bioassay. Toxicon 2002, 40, 1407–1425. [Google Scholar] [CrossRef] [PubMed]
  169. Wei, L.-N.; Luo, L.; Li, B.; Yin, Q.-C.; Lei, H.; Wang, B.-Z.; Guan, T.; Xu, Z.-L. Ultrasensitive “Hunter of Target”: Rabbit Monoclonal Antibody-Based Competitive Lateral Flow Immunoassay for Saxitoxin. Anal. Chem. 2025, 97, 1410–1418. [Google Scholar] [CrossRef]
  170. Zeng, W.; Tang, X.; Wu, T.; Han, B.; Wu, L. Development of a Highly Sensitive Aptamer-Based Electrochemical Sensor for Detecting Saxitoxin Based on K3Fe(CN)6 Regulated Silver Nanoparticles. Anal. Chim. Acta 2024, 1287, 342134. [Google Scholar] [CrossRef]
  171. Llewellyn, L.E.; Doyle, J.; Negri, A.P. A High-Throughput, Microtiter Plate Assay for Paralytic Shellfish Poisons Using the Saxitoxin-Specific Receptor, Saxiphilin. Anal. Biochem. 1998, 261, 51–56. [Google Scholar] [CrossRef] [PubMed]
  172. Negri, A.; Llewellyn, L. Comparative Analyses by HPLC and the Sodium Channel and Saxiphilin 3H-Saxitoxin Receptor Assays for Paralytic Shellfish Toxins in Crustaceans and Molluscs from Tropical North West Australia. Toxicon 1998, 36, 283–298. [Google Scholar] [CrossRef]
  173. Rey, V.; Botana, A.M.; Botana, L.M. Quantification of PSP Toxins in Toxic Shellfish Matrices Using Post-Column Oxidation Liquid Chromatography and Pre-Column Oxidation Liquid Chromatography Methods Suggests Post-Column Oxidation Liquid Chromatography as a Good Monitoring Method of Choice. Toxicon 2017, 129, 28–35. [Google Scholar] [CrossRef] [PubMed]
  174. Vale, P. Time Dependence of the Pre-Chromatographic Oxidation Method for Bivalves Contaminated with Paralytic Shellfish Poisoning Toxins. Food Addit. Contam. Part. Chem. Anal. Control Expo. Risk Assess. 2025, 42, 479–490. [Google Scholar] [CrossRef] [PubMed]
  175. Roy-Lachapelle, A.; Solliec, M.; Sauve, S. Determination of BMAA and Three Alkaloid Cyanotoxins in Lake Water Using Dansyl Chloride Derivatization and High-Resolution Mass Spectrometry. Anal. Bioanal. Chem. 2015, 407, 5487–5501. [Google Scholar] [CrossRef]
  176. Ross, A.R.S.; Ip, B.; Mueller, M.; Surridge, B.; Hartmann, H.; Hundal, N.; Matthews, N.; Shannon, H.; Hennekes, M.; Sastri, A.; et al. Seasonal Monitoring of Dissolved and Particulate Algal Biotoxins in the Northern Salish Sea Using High Performance Liquid Chromatography and Tandem Mass Spectrometry. Harmful Algae 2025, 145, 102854. [Google Scholar] [CrossRef]
  177. Gallo, P.; Lambiase, S.; Duro, I.; Esposito, M.; Pepe, A. Paralytic Shellfish Poisoning (PSP) Toxins in Bivalve Molluscs from Southern Italy Analysed by Liquid Chromatography Coupled with High-Resolution Mass Spectrometry (UHPLC-HRMS/MS). Toxins 2024, 16, 502. [Google Scholar] [CrossRef]
  178. Turner, A.D.; McNabb, P.S.; Harwood, D.T.; Selwood, A.I.; Boundy, M.J. Single-Laboratory Validation of a Multitoxin Ultra-Performance LC-Hydrophilic Interaction LC-MS/MS Method for Quantitation of Paralytic Shellfish Toxins in Bivalve Shellfish. J. AOAC Int. 2015, 98, 609–621. [Google Scholar] [CrossRef]
  179. Haddad, S.P.; Bobbitt, J.M.; Taylor, R.B.; Lovin, L.M.; Conkle, J.L.; Chambliss, C.K.; Brooks, B.W. Determination of Microcystins, Nodularin, Anatoxin-a, Cylindrospermopsin, and Saxitoxin in Water and Fish Tissue Using Isotope Dilution Liquid Chromatography Tandem Mass Spectrometry. J. Chromatogr. A 2019, 1599, 66–74. [Google Scholar] [CrossRef]
  180. Numano, S.; Watanabe, R.; Ozawa, M.; Uchida, H.; Matsushima, R.; Suzuki, T. Validation Study of LC-MS/MS for Paralytic Shellfish Toxins and Tetrodotoxin in Scallops. Shokuhin Eiseigaku Zasshi 2024, 65, 129–135. [Google Scholar] [CrossRef]
  181. Yue, Y.; Zhu, B.; Lun, L.; Xu, N. Quantifications of Saxitoxin Concentrations in Bivalves by High Performance Liquid Chromatography-Tandem Mass Spectrometry with the Purification of Immunoaffinity Column. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2020, 1147, 122133. [Google Scholar] [CrossRef] [PubMed]
  182. Odehnalova, K.; Pribilova, P.; Marsalek, B.; Babica, P. A Fast and Reliable LC-MS-MS Method for the Quantification of Saxitoxin in Blood Plasma Samples. J. Anal. Toxicol. 2024, 48, 119–125. [Google Scholar] [CrossRef]
  183. Dahlmann, J.; Budakowski, W.R.; Luckas, B. Liquid Chromatography-Electrospray Ionisation-Mass Spectrometry Based Method for the Simultaneous Determination of Algal and Cyanobacterial Toxins in Phytoplankton from Marine Waters and Lakes Followed by Tentative Structural Elucidation of Microcystins. J. Chromatogr. A 2003, 994, 45–57. [Google Scholar] [CrossRef] [PubMed]
  184. Kodama, M.; Ogata, T.; Noguchi, T.; Maruyama, J.; Hashimoto, K. Occurrence of Saxitoxin and Other Toxins in the Liver of the Pufferfish Takifugu Pardalis. Toxicon 1983, 21, 897–900. [Google Scholar] [CrossRef]
  185. Indrasena, W.M.; Ackman, R.G.; Gill, T.A. Separation of Paralytic Shellfish Poisoning Toxins on Chromarods-SIII by Thin-Layer Chromatography with the Iatroscan (Mark 5) and Flame Thermionic Detection. J. Chromatogr. A 1999, 855, 657–668. [Google Scholar] [CrossRef]
  186. Gulland, F.; Danil, K.; Bolton, J.; Ylitalo, G.; Rojas-Bracho, L. Vaquitas (Phocoena sinus) Continue to Die from Bycatch Not Pollutants. Vet. Rec. 2020, 187, e51. [Google Scholar] [CrossRef] [PubMed]
  187. Barchi, R.L.; Cohen, S.A.; Murphy, L.E. Purification from Rat Sarcolemma of the Saxitoxin-Binding Component of the Excitable Membrane Sodium Channel. Proc. Natl. Acad. Sci. USA 1980, 77, 1306–1310. [Google Scholar] [CrossRef]
  188. Wu, Y.; Ho, A.Y.; Qian, P.Y.; Leung, K.S.; Cai, Z.; Lin, J.M. Determination of Paralytic Shellfish Toxins in Dinoflagellate Alexandrium tamarense by Using Isotachophoresis/Capillary Electrophoresis. J. Sep. Sci. 2006, 29, 399–404. [Google Scholar] [CrossRef]
  189. Chwan, C.C.J.; Akbar, M.A.; Mohd, Y.N.; Pike, A.; Goh, C.T.; Mustapha, S.; Tan, L.L. Enhancing Early Warning: A DNA Biosensor with Polyaniline/Graphene Nanocomposite for Label-Free Voltammetric Detection of Saxitoxin-Producing Harmful Algae. Chemosphere 2024, 364, 143114. [Google Scholar] [CrossRef]
  190. Wei, L.N.; Luo, L.; Lei, H.T.; Guan, T.; Jiang, C.; Yin, Q.C.; Xu, Z.L.; Li, C. Nanoflower Microreactor Based Versatile Enhancer for Recognition Cofactor-Dependent Enzyme Biocatalysis toward Saxitoxin Detection. ACS Appl. Mater. Interfaces 2024, 16, 46495–46505. [Google Scholar] [CrossRef]
  191. Liu, X.; Bai, X.; Song, Z.; Zeng, J.; Wang, W.; Li, X. Luxuriant Foliage: High-Sensitivity Detection of Saxitoxin Using an Enzyme-Assisted Double-Cycling Amplification SERS Aptamer Sensor. Anal. Chim. Acta 2025, 1364, 344213. [Google Scholar] [CrossRef] [PubMed]
  192. Zhu, L.; Zeng, W.; Li, Y.; Han, Y.; Wei, J.; Wu, L. Development of Magnetic Fluorescence Aptasensor for Sensitive Detection of Saxitoxin Based on Fe3O4@Au-Pt Nanozymes. Sci. Total Environ. 2024, 921, 171236. [Google Scholar] [CrossRef]
  193. Zeng, W.; Fu, X.; Tang, X.; Liu, X.; Pan, F.; Wu, L. Novel Clean-Background Coffee Ring SERS Platform for Trace Toxin Analysis. Anal. Chem. 2025, 97, 7429–7438. [Google Scholar] [CrossRef] [PubMed]
  194. Celio, L.; Ottaviani, M.; Cancelliere, R.; Di Tinno, A.; Panjan, P.; Sesay, A.M.; Micheli, L. Microfluidic Flow Injection Immunoassay System for Algal Toxins Determination: A Case of Study. Front. Chem. 2021, 9, 626630. [Google Scholar] [CrossRef]
  195. Hyung, J.-H.; Hwang, J.; Moon, S.-J.; Kim, E.-J.; Kim, D.-W.; Park, J. Development of a Method for Detecting Alexandrium pacificum Based on the Quantification of sxtA4 by Chip-Based Digital PCR. Toxins 2022, 14, 111. [Google Scholar] [CrossRef]
  196. Petropoulos, K.; Bodini, S.F.; Fabiani, L.; Micheli, L.; Palleschi, G. Re-Modeling ELISA Kits Embedded in an Automated System Suitable for On-Line Detection of Algal Toxins in Seawater. Sens. Actuators B Chem. 2019, 283, 865–872. [Google Scholar] [CrossRef]
  197. Manger, R.L.; Leja, L.S.; Lee, S.Y.; Hungerford, J.M.; Wekell, M.M. Tetrazolium-Based Cell Bioassay for Neurotoxins Active on Voltage-Sensitive Sodium Channels: Semiautomated Assay for Saxitoxins, Brevetoxins, and Ciguatoxins. Anal. Biochem. 1993, 214, 190–194. [Google Scholar] [CrossRef] [PubMed]
  198. Van Dolah, F.M.; Fire, S.E.; Leighfield, T.A.; Mikulski, C.M.; Doucette, G.J. Determination of Paralytic Shellfish Toxins in Shellfish by Receptor Binding Assay: Collaborative Study. J. AOAC Int. 2012, 95, 795–812. [Google Scholar] [CrossRef]
  199. Kai, M.; Shen, W.-T.; Yu, Y.; Wang, D.; Zhang, J.A.; Wang, S.; Fang, R.H.; Gao, W.; Zhang, L. Dual-Modal Cellular Nanoparticles for Continuous Neurotoxin Detoxification. Nano Lett. 2024. Online ahead of print. [Google Scholar] [CrossRef]
  200. Wang, D.; Li, Y.; Deng, X.; Torre, M.; Zhang, Z.; Li, X.; Zhang, W.; Cullion, K.; Kohane, D.S.; Weldon, C.B. An Aptamer-Based Depot System for Sustained Release of Small Molecule Therapeutics. Nat. Commun. 2023, 14, 2444. [Google Scholar] [CrossRef]
Figure 1. Biosynthesis pathway of STX [1,59]: Steps 1–3, Claisen condensation: Acetyl-CoA transfers an acetate group to the SxtA complex (subunits: MTF, ACP, ACTP, and AONS); methyl donation from SAM generates propionyl-SxtA, which undergoes Claisen condensation with arginine. Step 4, amidino group transfer: SxtG transfers an amidino group from arginine to the SxtA-derived intermediate, forming 4,7-diguanidino-3-oxoheptane. Steps 5–7, cyclization and double bond formation: Step 5, SxtB mediates cyclization of the amidino intermediate, forming the first heterocycle; Step 6, SxtD introduces a C1–C5 double bond; Step 7, SxtS catalyzes epoxidation of the new double bond, and epoxide ring opening generates aldehydes, triggering bicyclization. Steps 8–9, reduction and hydroxylation: Step 8, SxtU reduces the terminal aldehyde to an alcohol; Step9, SxtH/SxtT catalyze C12 dihydroxylation, with O2 and succinate as co-substrates. Step 10, carbamoyl transfer: the SxtI/SxtJ/SxtK complex transfers a carbamoyl group to the dihydroxylated intermediate (dsSTX), yielding saxitoxin (STX).
Figure 1. Biosynthesis pathway of STX [1,59]: Steps 1–3, Claisen condensation: Acetyl-CoA transfers an acetate group to the SxtA complex (subunits: MTF, ACP, ACTP, and AONS); methyl donation from SAM generates propionyl-SxtA, which undergoes Claisen condensation with arginine. Step 4, amidino group transfer: SxtG transfers an amidino group from arginine to the SxtA-derived intermediate, forming 4,7-diguanidino-3-oxoheptane. Steps 5–7, cyclization and double bond formation: Step 5, SxtB mediates cyclization of the amidino intermediate, forming the first heterocycle; Step 6, SxtD introduces a C1–C5 double bond; Step 7, SxtS catalyzes epoxidation of the new double bond, and epoxide ring opening generates aldehydes, triggering bicyclization. Steps 8–9, reduction and hydroxylation: Step 8, SxtU reduces the terminal aldehyde to an alcohol; Step9, SxtH/SxtT catalyze C12 dihydroxylation, with O2 and succinate as co-substrates. Step 10, carbamoyl transfer: the SxtI/SxtJ/SxtK complex transfers a carbamoyl group to the dihydroxylated intermediate (dsSTX), yielding saxitoxin (STX).
Marinedrugs 23 00277 g001
Figure 2. Schematic of STX Blockade Mechanism on NaVs: (A) Structural organization of NaV complex: The NaV complex is composed of a pore-forming α subunit and one or more β subunits [113]. The α subunit, a single polypeptide chain with four homologous repeat domains (I–IV), constitutes the ion-conducting pore and voltage-sensing apparatus. Each domain contains six transmembrane helices (S1–S6)—S1–S4 form the voltage sensor, while S5–S6 form the pore domain [114]; (B) STX binding and functional disruption: Upon binding to the outer vestibule of the α-subunit pore via site 1, STX induces conformational contraction of the selective filter, restricting Na⁺ influx [24]. This α-subunit-mediated disruption abrogates neuronal action potential generation (Vmax reduction ≥ 80%) and propagation, leading to clinical conditions: facial paresthesia, limb paralysis, and respiratory arrest, etc. [23,30]. (Some drawing elements are from www.figdraw.com).
Figure 2. Schematic of STX Blockade Mechanism on NaVs: (A) Structural organization of NaV complex: The NaV complex is composed of a pore-forming α subunit and one or more β subunits [113]. The α subunit, a single polypeptide chain with four homologous repeat domains (I–IV), constitutes the ion-conducting pore and voltage-sensing apparatus. Each domain contains six transmembrane helices (S1–S6)—S1–S4 form the voltage sensor, while S5–S6 form the pore domain [114]; (B) STX binding and functional disruption: Upon binding to the outer vestibule of the α-subunit pore via site 1, STX induces conformational contraction of the selective filter, restricting Na⁺ influx [24]. This α-subunit-mediated disruption abrogates neuronal action potential generation (Vmax reduction ≥ 80%) and propagation, leading to clinical conditions: facial paresthesia, limb paralysis, and respiratory arrest, etc. [23,30]. (Some drawing elements are from www.figdraw.com).
Marinedrugs 23 00277 g002
Table 1. Established detection system for saxitoxin (STX).
Table 1. Established detection system for saxitoxin (STX).
StageTechnical MethodLOD *Application ScenariosKey FeaturesRef
On-site rapid screeningDaphnia acute toxicity bioassay0.3–10.7 ng Eq.STX/LAquatic product preliminary screeningLow cost and simple operation[158]
MIST Alert™ rapid test40 μg Eq.STX/100 gAquatic product preliminary screeningEquipment-free and adapted for on-site rapid turnaround[168]
Laboratory confirmation and quantificationELISA0.015 ng/mLPositive sample screening and preliminary quantificationHigh stability and applicable for rapid triage of large-scale samples[196]
HPLC (post-column oxidation method)0.1 μg/kgRoutine sample quantificationHigh specificity[181]
LC-MS/MS0.1 μg/kgComplex matrices and multitoxin simultaneous analysisHigh specificity and capable of structural confirmation[178]
Research-grade analysis and mechanism verificationNeuro-2a cell-based assay2 ng/LToxicological mechanisms and low-dose effect studiesDirect correlation between STX ion channel blockade toxicity and biological effect simulation[197]
Receptor-binding assay2 ug Eq.STX/LToxin–receptor interaction mechanismReal-time reflection of binding kinetics and high sensitivity[198]
*: The limit of detection (LOD) may be affected by various factors such as experimental conditions, instrument performance, and sample matrix, which may vary in actual detection.
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

Deng, H.; Shang, X.; Zhu, H.; Huang, N.; Wang, L.; Sun, M. Saxitoxin: A Comprehensive Review of Its History, Structure, Toxicology, Biosynthesis, Detection, and Preventive Implications. Mar. Drugs 2025, 23, 277. https://doi.org/10.3390/md23070277

AMA Style

Deng H, Shang X, Zhu H, Huang N, Wang L, Sun M. Saxitoxin: A Comprehensive Review of Its History, Structure, Toxicology, Biosynthesis, Detection, and Preventive Implications. Marine Drugs. 2025; 23(7):277. https://doi.org/10.3390/md23070277

Chicago/Turabian Style

Deng, Huiyun, Xinrui Shang, Hu Zhu, Ning Huang, Lianghua Wang, and Mingjuan Sun. 2025. "Saxitoxin: A Comprehensive Review of Its History, Structure, Toxicology, Biosynthesis, Detection, and Preventive Implications" Marine Drugs 23, no. 7: 277. https://doi.org/10.3390/md23070277

APA Style

Deng, H., Shang, X., Zhu, H., Huang, N., Wang, L., & Sun, M. (2025). Saxitoxin: A Comprehensive Review of Its History, Structure, Toxicology, Biosynthesis, Detection, and Preventive Implications. Marine Drugs, 23(7), 277. https://doi.org/10.3390/md23070277

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