Next Article in Journal
Fast and Sensitive Analysis of Short- and Long-Chain Perfluoroalkyl Substances in Foods of Animal Origin
Previous Article in Journal
Essential Oils and COVID-19
Previous Article in Special Issue
Current Sample Preparation Methodologies for Determination of Catecholamines and Their Metabolites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of In Situ Methods—Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) for the Collection and Concentration of Marine Biotoxins and Pharmaceuticals in Environmental Waters

1
Mass Spectrometry Group, Department Physical Sciences, Munster Technological University (MTU), Rossa Avenue, Bishopstown, T12 P928 Cork, Ireland
2
HALPIN Centre for Research & Innovation, National Maritime College of Ireland (NMCI), Munster Technological University (MTU), P43 XV65 Ringaskiddy, Ireland
3
CREATE (Centre for Research in Advanced Therapeutic Engineering) and BioExplore, Munster Technological University (MTU), Rossa Avenue, Bishopstown, T12 P928 Cork, Ireland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 7898; https://doi.org/10.3390/molecules27227898
Submission received: 28 June 2022 / Revised: 14 September 2022 / Accepted: 20 September 2022 / Published: 15 November 2022
(This article belongs to the Special Issue Recent Advances and Future Trends in Sample Preparation)

Abstract

:
Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) are in situ methods that have been applied to pre-concentrate a range of marine toxins, pesticides and pharmaceutical compounds that occur at low levels in marine and environmental waters. Recent research has identified the widespread distribution of biotoxins and pharmaceuticals in environmental waters (marine, brackish and freshwater) highlighting the need for the development of effective techniques to generate accurate quantitative water system profiles. In this manuscript, we reviewed in situ methods known as Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) for the collection and concentration of marine biotoxins, freshwater cyanotoxins and pharmaceuticals in environmental waters since the 1980s to present. Twelve different adsorption substrates in SPATT and 18 different sorbents in POCIS were reviewed for their ability to absorb a range of lipophilic and hydrophilic marine biotoxins, pharmaceuticals, pesticides, antibiotics and microcystins in marine water, freshwater and wastewater. This review suggests the gaps in reported studies, outlines future research possibilities and guides researchers who wish to work on water contaminates using Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) technologies.

1. Introduction

Determining ‘Quality of the Environment’ is a high priority global challenge. Consistent monitoring and surveillance programmes are required to highlight harmful environmental trends, to detect consequences of pollution on human life, and to initiate remedial action when regulatory limits are exceeded. Water-borne pollutants and biotoxins are key anthropogenic and natural environmental agents which must be reliably, accurately and rapidly detected in both the marine and freshwater ecosystems [1].
Marine biotoxins are naturally occurring metabolites of planktonic and microalgae that enter the food chain so that when ingested by marine organisms and accumulated in their tissues, they pose dangers to humans who consume contaminated seafood [2]. Various toxicological symptoms result from their consumption including gastrointestinal illness such as diarrhoea, nausea or vomiting or neurological effects causing in some cases unconsciousness and death [3]. The aquaculture industry, which is responsible for delivering environmentally sustainable produce devoid of anthropogenic or natural pollutants, maintains consumer safety by the routine monitoring and examination of water and aquaculture products [4]. For example, to counter the worldwide problem of shellfish contamination, produce is subject to biotoxin concentration limits based on strict internationally agreed permissible levels of biotoxins in edible shellfish [5] and consumers are protected by routine monitoring (on a weekly basis) during harvesting periods [6].
Passive sampling is an environmental monitoring technique developed for measuring organic and inorganic compounds in different sample matrices [7]. The concept has proven useful for monitoring pollutants in air, soil and water with applications dating back to 1873 when the technique was first used to estimate atmospheric ozone. In 1927 passive sampling was applied to determine carbon monoxide levels in the atmosphere [8] and in 1973 it was used for the quantitative analysis of the molecular diffusion of gaseous pollutants in the atmosphere [9]. In the 1970s passive sampling was extended to study aqueous environments such as water monitoring system for organic contaminants [1] and to determine pesticides that accumulate in soil and traverse into freshwater environments [10].
Periodic spot sampling regimes cannot quantify average exposure of aquatic organisms to contaminants, nor reveal long-term background level fluctuations and pre-treatment of sample material is necessary to isolate bioavailable fractions of contaminants prior to analysis [11]. Whereas time-integrative continuous in situ sampling, which arose from advances in material science (and in particular molecular imprinting technology) can be used to selectively accumulate and concentrate contaminants over periods of days or weeks. This technique not only serves to pre-concentrate low concentrations of analyte, but also provides a means to observe pollution impacts by mimicking bioaccumulation [11].
In recent decades passive sampling techniques and devices have been developed to monitor organic and inorganic contaminants in aquatic environments. Those technologies have been incorporated into European Commission guidance on surface water chemical monitoring. The aim of the NORMAN network [12] is to investigate and promote passive sampling to complement data gathered by spot sampling, this deployment indicates a future role for passive sampling in the regulatory domain [13].
Passive sampling receiving systems comprise either membrane materials, sorbent phases that act alone, or sorbents that act in combination with partition materials, that accumulate and concentrate contaminant species when exposed to analyte-laden matrices over extended periods. Analyte up-take in samplers is subject to chemical behaviour of the receiving system, characterised by the balance between analyte adsorption and desorption. The general form of this relationship fits Equation (1) [14] shown in Figure 1 where k1 is the uptake constant, k2 the elimination constant, Cs analyte concentration in the sampler and Cw analyte compound concentration in water, which is assumed to remain constant. During the integrative sampling phase indicated in Figure 1, when an approximately linear relationship between concentrations of analyte in the water and sorbent exists, thermodynamics are modelled by Equation (2), a simplification of Equation (1).
C s = C w k 1 k 2 [ 1 e x p ( k 2 × t ) ]
C s = R s × t × C w m s
When first exposed, negligible desorption gives way to a linear rise of analyte concentration. Integrative sampling periods last between several days to weeks and give rise to time-weighted average values for water concentration. Polar Organic Chemical Integrative Sampler (POCIS) and other commercially available passive sampling systems (those commonly applied for freshwater and wastewater quality monitoring), typically operate in this thermodynamic regime. As time passes and available adsorption sites fill, desorption offsets adsorption until the analyte concentration in the receiving material achieves thermodynamic equilibrium, affinity of receiver material for compounds of interest dictate maximum value of Cs achieved. Solid Phase Adsorption Toxin Tracking (SPATT) and generally qualitative sampling systems operate via this regime (although Cw is estimable provided the sampler-water partition coefficient is known) [11]. Regardless of the thermodynamic mode, accumulation within receiving systems selectively concentrates compounds of interest compared to the environmental matrix, thus simplifying pre-treatment of samples [15]. The pre-concentration capacity of passive sampler devices also makes them highly practical for obtaining ultra-trace quantities of chemicals [16]. Performance reference compounds (PRCs) can improve accuracy of quantitative sampling attempts. If the analyte accumulation rate (Rs) for each pollutant captured by the sampler is known, then the time-weighted average concentrations of these compounds in environmental water can be estimated. However, Rs values measured in laboratory conditions differ from those encountered in the field due to local environmental factors (pH, water flow rate, and temperature) therefore Cw estimates will be uncertain. By exposing sorbents to known concentrations of PRCs (compounds not present in the field environment for which sorbent characteristics are known, whose Rs values are affected by environmental factors identical to compounds-of-interest sample rates, and whose desorption follows first-order kinetics analogous to uptake) prior to deployment, the subsequent decay of these PRC levels during deployment indicate in-field Rs values for target analytes.
Proprietary passive samplers most used for water-borne hydrophobic (nonpolar) compounds are Semi-Permeable Membrane Devices (SPMDs), silicone rubber strips and sheets, and low-density polyethylene (LDPE) membranes but ‘Membrane Enclosed Sorptive Sampler’ (MEMSO) devices, non-polar disks for the ‘Chemcatcher’ sampler, and naked chromatography disks (e.g., C18, C8) may also be used. For hydrophilic (polar) compounds, including biotoxins, POCIS and Chemcatcher disks products are available and suitable naked chromatography disks or ion-exchange resins may also be used. Diffusive gel discs and sheets can be used to sample metals and cations, oxyanions and polar organic compounds [17] however, evidence of extant research [18] indicates that the application of this technology to study marine toxins remains in its infancy.

2. Marine Biotoxins

Absorbent substrates employed in passive sampling tools are matched with classes of biotoxins (Table 1 and Table 2) for which they show affinity. These can be categorised according to lipophilic (lipid soluble contaminants) properties, hydrophilic (water soluble contaminants) [19] properties, or for affinity with compounds falling in the category of novel emerging aquatic biotoxins [3].

2.1. Lipophilic Biotoxins

Lipophilic biotoxins are naturally occurring compounds that can be found individually or grouped as biotoxin mixtures and are categorised into four groups: (i) okadaic acid (OA) and derivative dinophysistoxins (DTXs), (ii) pectenotoxins (PTXs), (iii) yessotoxins (YTXs) and (iv) azaspiracids (AZAs), based on chemical structure and bioavailability [2] (Table 2 and Supplementary Figure S1). Consumption of lipophilic toxins predominantly causes gastrointestinal symptoms (vomiting, nausea and diarrhoea) [20].
OA, DTXs and PTXs (Supplementary Figure S1) are classified as diarrhetic shellfish poisoning (DSP) biotoxins and were initially detected in the Netherlands in the 1960s through a spate of gastrointestinal incidents caused by the consumption of contaminated mussels [21]. DSPs are heat resistant so remain stable after cooking in some species of shellfish. This group of biotoxins are produced by dinoflagellates (single-celled organisms possessing two flagella, a group that includes the species Dinophysis spp. [22]) that occur in large numbers in communities of marine and fresh water plankton [23]. Many dinoflagellates are known to be photosynthetic, but a large fraction of these are in fact mixotrophic, combining photosynthesis with ingestion of prey (phagotrophy [feeding by engulfing a food cell or particle and ingesting it in a phagocytic vacuole] and myzocytosis [“cellular vampirism” as the predatory cell pierces the cell wall and/or cell membrane of the prey cell with a feeding tube, the conoid, and sucks out the cellular content and digests it]) [23,24]. The population of Dinophysis spp. can proliferate quickly and spread throughout bays through strong winds and currents [21]. A variety of nutritional and physical parameters can affect the reproduction and survival of microalgae. Nitrogen and phosphate are two essential nutrients: nitrogen plays an important role in the synthesis of protein and chlorophyll; phosphorus is the main constituent in membrane lipids, adenosine triphosphates (ATPs) and DNA [25]. Limitation of nitrogen and phosphorous concentrations would affect the amount of DSP toxin production [26]. Other than nutrition, physical parameters such as salinity, turbidity, temperature and light also affect algae bloom growth and DSP toxin production [26]. The limit value of DSP toxin in bivalves determined by the European Commission is 0.16 µg OA equivalent per kg in test shellfish tissue (Table 1) [27,28]. (The amount of toxins expressed as the amount of okadaic acid that gives the same toxic response followed intraperitoneal administration to mice. This applies similarly for the group of yessotoxins and azaspiracids, respectively [27]).
Okadaic acid (OA), a polycyclic ether, is commonly found in shellfish across Europe, Asia and South America from Spring to late Autumn [29]. Okadaic acid as a lipophilic polyether toxin (Supplementary Figure S1a) that accumulates in the hepatopancreas of shellfish causing illness (e.g., nausea, vomiting, diarrhoea) to consumers of the contaminated shellfish [30,31]. Four metabolites, 19-epi-okadaic acid, belizeanic acid, 11-oxo okadaic acid, 11S-hydroxy and 11R-hydroxy okadaic acid are reported for okadaic acid [32]. These metabolites are produced by the human recombinant cytochrome CYP3A4. These metabolites are similar to okadaic acid and are protein phosphatase PP2A inhibitors [33]. In addition, OA has been studied for the last few decades not only as a diarrhetic toxin [29] but also as a DNA and cellular disruptor that can affect the immune and nervous systems [34]. Furthermore, OA has been studied as a potential carcinogenic agent [35]. OA is an inhibitor of serine/threonine protein phosphatase PP1 and PP2A that increases phosphorylation of some proteins such as myosin. It also affects sodium secretion of intestinal cells resulting in gastro-intestinal fluid accumulation and gastric cramp [31]. OA also reduces activity of PP2A that affects the brain of patients suffering with Alzheimer’s disease [36]. In summary, OA has been shown to be a tumour promoter, genotoxic [31] and hepatotoxic [36].
Liver cells are one of the major targets of OA, causing an interruption to intercellular communication, and precipitating apoptosis in hepatocytes, OA can change the cell morphology of human liver tumour cells and can change the cell proliferation [37]. In addition, to the damage to human cells, OA can induce cytotoxicity and cell death in cervical cancer cell lines [38], affect human kidney cell lines by increasing oxidative DNA damage [39], morphologically change human osteosarcoma cells and cause DNA fragmentation [40]. Availability of two major target of OA, PP1 and PP2A, mixed with OA structures, could be a new design of drug targeting cell. This complex can target the cells by protein phosphatase inhibition [41].
Studies of animal cell lines have reported negative effects on monkey kidney epithelial cells [30]; it decreased l viability of Chinese hamster cells [42] and increased death in Chinese hamster ovary cells [43]. In addition, in vivo studies of okadaic acid toxicology, such as direct injection of OA (≥75 ng/larva) into larva (Galleria mellonella) resulting in reduced larval survival (>65%) [44]; studies which show OA promoting tumour growth in Hawaiian green turtle’s fibropapillomatosis [45], and studies demonstrating diarrhetic effects in mouse small intestine via inhibition of the serine/threonine protein phosphatases [46], provide further evidence of OA toxicity.
DTXs, (such as DTX1, DTX2 and their derivative DTX3) (Supplementary Figure S1p–r), belong to the OA structural class of toxins [47] and are known as the diarrhetic lipophilic marine toxins, that during blooms of Dinophysis dinoflagellates can be accumulated in shellfish. The 7-O-acyl fatty acid ester of OA in combination with DTX1 and DTX2 are collectively designed as DTX3 [29]. DTX1 acts as protein phosphate inhibitor (PP2A) that acts similarly to OA; DTX2 has reduced protein inhibitor activity [48]. DTX2 typically is known as the late summer toxin occurring during blooms of Dinophysis acuta [49].

2.1.1. Yessotoxins (YTXs)

Yessotoxins (YTXs), are a group of lipophilic toxins having 11 contiguous ether rings and one saturated side chain (Supplementary Figure S1b), that show chemical and structural similarities to brevotoxins and ciguatoxins [50]. YTX (Table 2) was first detected in the digestive gland of a bivalve scalene called Patinopecten yessoensis (Japanese scallop), in 1986 in Japan [51]. However, more recently YTX was identified in Spain, Italy, Norway, Chile, Norway and New Zealand [52]. YTXs mainly accumulate in the digestive glands of bivalves and can also accumulate in hepatopancreas and muscle tissue [53]. Some YTXs are produced by dinoflagellates directly or are the result of shellfish metabolism, for example YTX and homoYTX, which are produced by dinoflagellates, whilst 45-OH-YTX separates from shellfish as a shellfish metabolite [52].
Yessotoxins (YTX), homoYTX, 45 OH-YTX and 45 OH-homoYTX, are polyether toxins constructed by Protoceratium reticulatum, Lingulodinium polyedrum and Gonyaulax spinifera [54]. The variety of the toxin profile depends on the origin of the strain. The vast majority of toxin detected in Protoceratium reticulatum is YTX. In 2006 Gonyaulax spinifera was also verified as an YTX producer [55]. Homo YTX is reported to be produced by Lingulodinium polyedrum [56] and has also been detected in a strain of P. reticulatum in Spain [54], China [57] and Japan [58].
YTX, is a ladder-shaped polycyclic polyether compound containing two sulphur groups with a high polarity end that allows it to interact with plasma membrane components [50] (Supplementary Figure S1b). The main target of YTXs are scallops and mussels (filter feeders) species which can siphon water in from one side and expel water from the other side while extracting oxygen and foods in between, this filter feeding mechanism in mussels and scallops enables them to accumulate high amounts of YTXs (4.5–65 pg/cell). The concentration of Homo-YTX in this study reported between zero to 0.80 pg/cell in Lingulodinium polyedrum and 33.4 pg/cell in Gonyaulax spinifera [56].
The individual injection of 100 µg/kg YTX and homoYTX are reported as a lethal concentration in mice [59]. In addition, the neuronal brain damage, liver and pancreas failure and cardiovascular toxicity of YTXs indicate the possible health impact of YTXs on human health [60,61]. The mouse bioassay test shows YTXs produce toxicological symptoms similar to diarrheic shellfish poisoning (DSP) [62]. However, coexisting YTXs and OA changes the toxicological functionality as protein phosphatases 2A (PP2A) inhibits or causing diarrhoea, therefore it hindered YTX to be categorised in the DSP toxin group, although YTX still remains listed as a DSP toxin by accumulating in shellfish predominantly [62,63]. YTX have been recognized as a potent cytotoxins, affecting the neural networks, [64] neuronal death in rats was found following exposure to ≥25 nM YTX for around 48 h [60]. Therefore, the authorized level of yessotoxins has increased to 3.75 mg equivalent per kg shellfish meat (i.e., this is the YTX concentration limit for food as recommended by the European Commission (Table 1) [65].

2.1.2. Azaspiracids (AZAs)

Azaspiracid (Supplementary Figure S1c) was first discovered after the ingestion of contaminated mussels originally from Ireland in 1995 caused human illness (Table 2). Extensive chemical analysis revealed the culprit was a new class of toxin [66,67]. Since then, more than 60 analogues of AZA have been reported [68,69,70] and almost 21 different type of AZA characterized [71]. AZA1 was first reported in 1998 in blue mussels, AZA2 with an extra methyl group on C8, AZA3 with a less methyl group in C22, AZA4 and AZA5 showed an extra OH group in C3 and C23, respectively, were also observed [72,73,74]. AZA6 is reported to be like AZA2 and AZA3 having an extra methyl group similar to AZA2 with the lack of a methyl group in the same position as AZA3. Examination of AZAs in shellfish shows that the main AZAs detected in shellfish are AZA1 and AZA2, while AZA3 is detected at low concentrations [66,67]. AZA4 up to AZA23 (except AZA11) have been shown to be shellfish metabolites, as they have not been detected in planktonic samples [75].
Azaspiracids toxins are fat-soluble polycyclic ethers that are non-neurotoxic in the lipophilic toxins categories. AZAs toxins are heat resistance, acid stable and have a long-term stability under refrigeration [71]. Although AZAs induce some symptoms such as stomach cramp, diarrhoea and nausea they are not categorised as diarrhetic shellfish poisoning (DSP) [76]. In vivo studies, injecting AZA into mice show a neurological action causing paralysis and spasm [70]. Abal et al. studied the effect of AZAs on Caco-2 cell lines, by passing AZA1 through the cell monolayer and cell alterations [76]. In addition, human breast cancer cell line (MCF-7) was exposed to AZA1. The results indicated that AZA1 induced selective cell death [77].
The lethal dose of AZAs injected into mice is reported as 200 µg/kg for AZA1, 110 µg/kg for AZA2, 140 µg/kg for AZA3, 470 µg/kg for AZA4 and 1000 µg/kg for AZA5. However, the lethal dose of oral fed AZA1 in mice is reported as >700 µg/kg (Table 1) [78]. AZAs are produced by Azadinium spinosum and Amphidoma languida [71] and they have a regulatory testing limit of 160 µg azaspiracids per kg (bivalve tissue) reported by the European Commission due to their multi-organ toxicity [28]. Calcium, adenosine 3,5-cyclic monophosphate (cAMP), Protein kinase C (PKC), phosphodiesterases (PDEs) and mitochondria play roles in mechanism of action of AZAs [52]. Table 1 shows the highest permitted level of marine biotoxins in total quantities (measured in the whole bivalve body or any separate edible part) by the European Commission [28].
Table 1. Maximum permitted level of marine biotoxins in total quantities. “Total quantity means measured in the whole bivalve molluscs body or each edible parts separately” [28].
Table 1. Maximum permitted level of marine biotoxins in total quantities. “Total quantity means measured in the whole bivalve molluscs body or each edible parts separately” [28].
BiotoxinLimited Level
Paralytic Shellfish Poison (PSP)<800 microgram/kg
Amnesic Shellfish Poison (ASP)<20 milligrams of domoic acid/kg
okadaic acid (OA), dinophysistoxins and pectenotoxins together<160 micrograms of OA equivalents/kg
Yessotoxins (YTXs)<3.75 milligram of YTX equivalent/kg
Saxitoxin (STXs)≤800 µg STX.2HCL equivalent/kg
Azaspiracids (AZAs)<160 micrograms of AZA equivalents/kg
Domoic acid (DA)<20 mg domoic acid/kg
Brevetoxin<200 mouse units or equivalent
Ciguatoxins<0.1 µg/kg fish

2.2. Hydrophilic Biotoxins

Hydrophilic biotoxins (domoic acids and saxitoxin) are naturally occurring toxins that can be found in phytoplankton and in the shellfish that consume them. They are classified into two groups according to the type of illness associated with ingestion: amnesic shellfish poisoning (ASP) and paralytic shellfish poison (PSP) respectively [79].

2.2.1. Domoic Acid

Amnesic shellfish poisoning (ASP) was first detected in 1987 in Canada, the toxin responsible was domoic acid (DA), an amino acid of kainoid class produced by phytoplankton. DA can bioaccumulate in shellfish and thus can be introduced into the human food chain (Table 2) [80]. DA can contaminate shellfish such as mussels, crabs, scallops, razor clams and cockles in which it is primarily located in digestive glands [81]. Domoic acid is a heterocyclic amino acid that includes a proline ring and imino group in the structure. In addition, DA consists of three carboxylic acids (Supplementary Figure S1d) [82]. Epi-domoic acid (epi-DA) and isodomoic acids A-H (iso-Das) are isomers of DA that can co-occur in shellfish [27]. These toxins cause gastrointestinal and neurological symptoms after the consumption of contaminated seafood. Although DA is heat resistant and stable at cooking temperatures, there is the possibility of concentration reduction due to the hydrophilicity (water solubility) of the toxin [27]. In vivo studies showed that DA absorption is very low through the gut, the detected amount of DA in animals fed DA showed that rats and cynomolgus monkeys expelled 2% and 4–7% DA, respectively, in their urine after 24 h [83]. Further, DA has a LD50 of 2.4 mg domoic acid/kg b.w., the LD50 was established by injecting DA extracted from mussels into mice, this induced neurological symptoms including memory loss. Abdominal cramp was also observed [84]. The injection of 1 mg/kg b.w. domoic acid into rodents shows that DA stimulates the thyroid hormones [85]. The mechanism of neural dysfunctionality refers to the tendency of DA binding to glutamate and kainate receptors in the brain. This strong association is due to the similarity of DA to glutamic acid and kainic acid [86]. According to the regulations, the maximum concentration of DA in shellfish should not exceed 20 µg/g (20 mg/kg), otherwise gastrointestinal and neurological symptoms such as vomiting, diarrhoea abdominal cramps, confusion and seizures may result (Table 1) [20].

2.2.2. Saxitoxin (STX)

Saxitoxin (STX) and gonyautoxins are two potent biotoxins produced by Harmful Algae Blooms (HABs) such as red tides, these toxins are found in algae spices, mostly in Alexandrium dinoflagellates (Table 2) [87]. STX is primarily related to marine dinoflagellates (eukaryotes) and freshwater cyanobacteria (prokaryotes) [88]. A trait of Alexandrium spp. is that they can be found in the same place for several years as their cysts can reside and reproduce in marine sludge [87]. STX has three rings in the structure (Supplementary Figure S1e) and can be described as a 3,4,6-trialkyl tetrahydropurine (Supplementary Figure S1d) with a ring containing a hydrated ketone created through the 3,4 positions of the purine ring. In addition, STX has two guanidine moieties that form by NH2 group in 2,8 position of the purine ring [89,90].
STX and its derivatives are often called paralytic shellfish toxins (PST) and cause paralytic shellfish poisoning (PSP) [27]. Neurological symptoms of these toxins can appear shortly after contaminated shellfish is consumed which can cause muscular paralysis and respiratory failure by obstructing voltage-dependent sodium channels resulting in death, a few hours after the first symptoms is observed [91,92]. The lethal dose of STX based on mouse bioassay is 10 µg/g b.w.; however, the oral dose for human is 7 µg/kg b.w. (Table 1) [93].
The first STX was reported in 1798 in Canada where those who consumed mussels became ill [94]. STX including 3,4-propinoperhydropurine tricyclic system and two guanidine groups are highly polar [92] amorphous compounds having a pKa value of 8.24 and 11.60 [95]. STX is a water soluble contaminant, however, it is also soluble in methanol and ethanol [92]. The stability of STX is based on the environmental conditions such as pH and temperature. STX is more stable in acidic environments at a pH around 2–4, however at higher pH there is a decrease in the stability [93]. In addition, suitable temperatures for STX stability is reported at around 20 °C, therefore a pH (2–7) and a temperature of 20–25 °C, are the most suitable environmental conditions for STX and its derivatives’ to remain stable [96].

2.3. Emerging Biotoxins

This category of toxins have distinctive differences to the known toxins, as they might be found in the water based on changes caused to the environment such as dinoflagellate cysts discharges, ballast water from ship discharge, red tides, climate change and changes to ocean temperature currents [3,97].

2.3.1. Brevetoxins

Neurotoxic Shellfish Poisoning (NSP) is a gastrointestinal and neurological disease caused by molluscs and shellfish contaminated with brevetoxins (PbTxs; Table 2) produced by the dinoflagellate, Karenia brevis [98]. Blooms [20] mainly occur in the Gulf of Mexico and off New Zealand coasts [97]. Brevetoxins are potent cyclic polyether compounds and have nine analogues (Supplementary Figure S1f). Depending on the ether rings there are two type of backbone structure for brevetoxins including type-A 10 and type-B 11 trans-fused ether rings. Among brevetoxin derivatives, PbTX1 is the most potent and PbTX2 is the most naturally abundant in algal blooms [99,100], PbTX3 is the most abundant in beachside marine aerosols [101]. PbTX1, PbTX2, PbTX3 (Supplementary Figure S1f–h) and PbTX7 have an A-type backbone (open lactone ring) and show greater polarity, compared with close-ring brevetoxin derivatives [101]. Brevetoxins are fat-soluble neurotoxins that can cause extensive fatality in marine mammalians and fish [99]. Brevetoxins are highly stable under conditions such as dry state and over the pH scale, ranging between pH 2 and pH 10. Brevetoxins are stable in solvents such as DMSO, alcohol, acetonitrile and acetone [100] but its degradation by incubating at 500 °C for 10–15 min has been reported [102]. Vomiting, diarrhoea, hypotension, arthralgia, myalgia, hyporeflexia are symptoms of NSP [20] additionally, NSP can cause neurological symptoms, respiratory irritation and chest pain. Symptoms of neurological illness persist for longer periods in patients compared to gastrointestinal symptoms [97]. Brevetoxins can cause damage in nerve membranes by influencing and activating voltage-gated sodium channels (VGSC), depolarizing the nerve membranes causing neuro-excitation that leads to neurological symptoms [97]. Moreover, studies have revealed that muscle depolarization and nerve depolarization can happen at the same time [103]. In vivo studies injecting brevetoxin in rats showed that due to the lipophilic nature of NSP, it the toxin easily penetrates cell membranes [104]. The lethal oral dose in mice is 0.520 mg/kg body weight and 0.094 mg/kg body weight (Table 1) after injecting PbTx-3 [97].

2.3.2. Ciguatoxins

Ciguatera fish poisoning (CFP) is a fat-soluble toxin accumulated in a range of fish and marine species (Table 2) [3]. CFP is the most frequently detected toxin in seafood affecting human health [105]. This type of toxin can be found mainly in tropical areas however, these biotoxins can remain in frozen food and may be spread to other regions by the importation of fresh or frozen contaminated seafood [106]. These toxins are heat resistance and stable under moderate acidic or basic environment. Ciguatoxins (Supplementary Figure S1i–m) are polyether compounds including 13–14 aromatic rings joined by ethers that create a ladder-like structure [107]. Ciguatoxin (CTX) was found in 1606 in the South Pacific Islands [100]. These toxins are of a class of compounds with 24 associated structures [100]. They can be produced by micro algae such as dinoflagellate that live in shallow tropical waters that attach to dead coral or seaweeds. Therefore, fish grazing on such corals and seaweeds are at risk of contamination with CTX. In addition, accidental consumption of marine in shore or coastal water containing CFPs can cause illness in humans [105]. The CTX mechanism of action is similar to that of brevetoxin, bonding to sodium-voltage channels that are closed during the resting membrane potential and causing neurological symptoms by depolarization of the membrane of the nerve [105,108]. CTX mostly accumulates in vivo in fish heads, fish livers or fish gonads rather than fish meat in which much lower concentrations are found. More serious poisoning therefore happens after consumption of contaminated fish heads or fish organs [105]. Ciguatera fish poisoning shows neurological [109], gastrointestinal as well as cardiovascular symptoms [107]. Cardiovascular symptoms usually happen in parallel with gastrointestinal or neurological symptoms and patients require immediate medical care [110]. In vivo studies on CTXs shows that these toxins can quickly absorb via the gastrointestinal tract and pass around the body [105]. Symptoms of contamination with CTX include tingling, vomiting, diarrhoea and some neuropsychiatric symptoms such as depression and memory loss [106]. However, neurological symptoms such as confusion, depression [111], loss of memory and anxiety [112] may be observed from one day to weeks after consumption. The lethal dose (LD50) is 0.25 µg/kg for P-CTX1, 2.3 µg/kg for P-CTX2 and 0.9 µg/kg for P-CTX3 in mice (Table 1) [113]. Among ciguatoxins, P-CTX1 is the most polar and most toxic form. Indeed, increasing polarity of ciguatoxins in vivo is due to their oxidative metabolism as they climbs up the food chain [107].
Table 2. Marine biotoxin classification.
Table 2. Marine biotoxin classification.
ToxinFormulaMW (g/mol)Chemical ClassSyndrome CategorySolubilityOriginPolarityRefs.
Okadaic acid (OA)C10H17N7 O4804Polyether, spiro-keto ring assemblyDSPLipophilicHalichondria
okadaii
Low polarity[30,41]
Yessotoxin (YTX)C55H82O21S21143.4 Sulfur bearing polyetherGastrointestinal, NeurologicalLipophilicProtoceratium reticulatum, Lingulodinium polyedrum and Gonyaulax spiniferaHighly polar[50]
Azaspiracids (AZA)C47H71NO12842.1Polyether, second amine, 3-spiro-ring assemblyDSPLipophilicA. spinosumLow polarity[66,114]
Domoic acid (DA)C15H21NO6311Cyclic amino acid, 3 carboxyilic acid groupsASPHydrophilicPhytoplanktonHighly polar[115]
Saxitoxin (STX)C10H17N7O4299.3Tetrahydro-purine alkaloidPSPHydrophilicPhytoplanktonHigh polarity[89]
Brevetoxin (PbTxs)C49H70O13867.1Polyether with contiguously fused ringsNSPLipophilicDinoflagellatesPolar[99,101]
Ciguatoxins (CTX)C60H86O191111.313Polyether Gastrointestinal, Cardiovascular, NeurologicalLipophilicDinoflagellatePolar to moderate polarity[105]

3. Solid Phase Absorption Toxin Tracking

In countries that host indigenous shellfish industries, regular monitoring of algae blooms and biotoxin levels in marine waters are essential public health measures; warnings of imminent or actual contamination events may minimize consumer exposure to potentially harmful shellfish products [116]. Occurrences of harmful algae blooms have dramatically increased in recent decades due to increasing eutrophication and other factors such as the spread of toxins in ballast water discharges [117,118], climate change [119,120], ocean acidification [121], and global shellfish product trading [119]. Alterations to the physical, chemical or biological properties of water environments may cause socioeconomic impacts across multiple sectors (e.g., public health, tourism, recreation and commercial fishing, etc.) in addition to incurring monitoring and management costs [122]. Historically, human health threats posed by biotoxins were identified by direct food product chemical monitoring and by animal testing [75]. Thus far, food tracking programmes are limited to known toxin categories and have not linked toxin occurrences to all HAB populations that give rise to them, so these programmes have no preventative value [19]. Detection techniques based on toxin ingestion by live animals give rise to irreconcilable ethical issues [123] and have very low selectivity therefore limit their effectiveness for protecting consumers [124]. Nevertheless, the Mouse Bioassay (MBA) method developed in the 1970s to detect Lipophilic Shellfish Toxins (LSTs) has until recently remained the reference technique for biotoxin detection, a role now filled by non-animal testing methods especially LC-MS/MS. Nevertheless, animal test regimes can be advantageous: MBA is applicable for new or other type of toxins rather than the known regular biotoxins [75] and cyclic imines that are not categorized in the EU regulations may be detected using the MBA [75]. Rat bioassay (RBA) is specifically used for the detection of OA, DTXs and AZAs and is not commonly used as a monitoring method [75].
Mouse bioassay methods only present qualitative (negative or positive) results and combination with and interference from other compounds during sample preparation gives rise to high risk of false positive or negative results [125]. Furthermore, the use of live animals for bioassays raises ethical issues [123]. These test methods are insensitive therefore their capability to protect consumers is limited [124]. Animal testing has been superseded by liquid chromatography–mass spectrometry (LC-MS/MS) [47,67,126,127,128] and high resolution mass spectrometry methods that are able to detect both qualitatively and quantitatively the different toxin classes OA/DTXs, PTXs, YTXs, AZAs and emerging biotoxins either individually or in combination (Scheme 1) [75,129].
Functional assays, methods of detecting toxins based on observable or measurable responses of cells within cell cultures to biotoxins in sample water have been reported by Rossini et al. [130] for neuro-2a neuroblastoma system cell line for STXs and Jellett et al. [131] for a BE(2)-M17 neuroblastoma cell line for STXs; Nicholson et al. [132] for cell-free cell line for okadaic acid and Vieytes et al. [133] for a MCF-7 cell line for YTX. However, it must be noted that cell-based methods are limited to specific toxin groups [19].
Biochemical methods are based on the bonding of marine biotoxins to antibodies taken from animals or cell cultures. The advantage of this method is that they are suitable for the detection of very low concentrations of specific toxins, including AZAs [75], the cyclic imines group [134], domoic acid [135], okadaic acid [136], PTXs [137], STXs [131] and YTXs [138]. The commonest chemical analysis methods employed for detecting marine biotoxins are based on liquid chromatography (LC), in tandem with other physicochemical methods such LC-MS, LC-UV and LC-FL [75]. LC-MS/MS is the official method for lipophilic toxin detection [139], HPLC-UV for domoic acid detection [140], HPLC-FLD for PST toxin detection. Emerging toxins and other toxins are primarily analysed using LC-MS/MS and with biological functional methods [141,142,143,144,145]. The application of LC with fluorescent detection (FL) [146,147], LC-MS [139,148,149] and LC-MS/MS [148] has been widely applied and evaluated for detection of the DSP group of toxins; LC-MS [149] and LC-MS/MS [150] for PTXs; LC-MS/MS [151] and LC-FL [152] for STXs and LC-FL [153], LC-MS [149] and LC-MS/MS [151] for detection of YTXs (Scheme 1). Applying analytical methods for the detection of marine toxins is expensive [154], requires extensive validation and highly trained staff for instrument operation but analytical methods are the most dependable and reliable analytical approaches for protecting human health [75,155].
The foregoing discussion suggests a need for cost-effective but efficient methods to replace discrete grab (non-continuous) sampling traditionally employed to monitor marine biotoxins. However, concentration of toxins present is an important requirement in these sampling methods and sampling might not capture spatial concentrations due to vertical migration, hydrologic or circulation effects, therefore results might be inherently biased [156,157].
The concept of SPATT (small packages of adsorbent phase deployed for extended durations at various depths in the water column) was pioneered by McKenzie et al., an idea that first arose from observations of dissolved polar and non-polar biotoxins in seawater in 1998 and from culture studies conducted in 2003. Passive adsorption followed with LC-MS detection gives rise to a fast but simple and sensitive biotoxin monitoring system that McKenzie et al. concluded could provide accurate predictions of net toxin accumulation by bivalves (Supplementary Figures S2 and S3). The technique was first field tested in 2002 and 2003, a period that included DSP and YTX bloom events (Supplementary Figure S4) [116,158].
SPATT technology, in common with other continuous monitoring techniques, can be directly deployed in the marine and freshwater environment, an advantage over traditional lab-based methods such as shellfish tissue testing or phytoplankton microscopic cell counting. In addition, SPATT technology is a simple, cost effective technique for recovering targeted analytes from a aqueous environment [159]. However, SPATT will only detect dissolved biotoxins (extracellular biotoxins) which is one of its limitations, another drawback is the lack of calibration and validation for this technique [19]. Furthermore, the SPATT technique is not able to detect at the ng/g of toxin concentration required by health advisory authorities [159]. However, it is still a useful predictive and preventative tool for toxin monitoring when used in conjunction with analytical instrumentation for the screening of seafood.

3.1. SPATT Bag Construction and Adsorbent Phase Activation

SPATT bag dimensions are typically 60 mm × 60 mm, bags being fabricated from 95 µm polyester mesh (Polymon PES 95/27). Edges are closed by polyester thread stitching [116] or by heat-sealing. Supplementary Figure S2 illustrates construction of SPATT bags [160].
To form a thin layer of resin Rundberget et al., placed the resin between two layers of nylon mesh that was clamped tightly into a frame (Supplementary Figure S2) [161].
Another example (Supplementary Figure S2) illustrates SPATT bag and associated mounting system construction. For this river deployment, bags were secured to a large aluminium alloy mounting tube using 4 mm diameter clamping screws [162].

3.2. SPATT Sorbents

A number of synthetic adsorbent materials show affinity for biotoxins. In common with naturally occurring counterparts (e.g., Zeolytes and activated carbon), human-made sorbents are characterized by spherical particles with large surface areas and high porosities that create three-dimensional crystalline lattices, including pores, in which water molecules are held loosely. Adsorption results from interactions between atoms at the sorbent surface [163].
While sorbent particles are in contact with a solution, smaller solutes can diffuse into pores using physical interactions between the surface and the species. The precise nature and combination of bonds include non-polar hydrophobic interactions, polar interactions (h-bonding, pi–pi bonds, dipole–dipole or induced dipole), electrostatic attraction; and depend on analyte species. Adsorption and desorption occur simultaneously according to differential concentrations of analyte in solution and in the sorbent, giving rise to characteristic analyte-specific behaviour indicated in Figure 1 in Section 1. The complex nature of interactions necessitates an experimental approach to determining mathematical models that best fit observed isotherms. Studies such as Shin and Kim [163] demonstrate this principle; however, searches for evidence of similar research into SPATT sorbent biotoxin affinities at the time this review was prepared were unsuccessful. Larger molecules, those that are bigger than pore size, will remain unabsorbed, a phenomenon known as the ‘sieving’ effect.
Synthetic adsorbents have a wide range of applications in the separation of valuable compounds from plant extracts, fermentation products, food additives and in pharmaceutical applications. Synthetic sorbents are designed to remain stable in acidic and alkaline environments and are resistance to organic solvents. Analyte extraction from synthetic sorbents is safer and reduces solvent usage compared to other solvent-based extraction techniques.
Synthetic sorbents can be categorised into three different groups based on their chemical structures:

3.2.1. Aromatic Adsorbents

Aromatic adsorbents are characterized by a crossed linked polymeric matrix that are suitable for the separation of peptides, antibiotics and food additives [164], for example, the type of commercial sorbents of this type are DIAION HP20, HP21 SEPABEADS SP825L, SP850, SP70 and SP700. DIAION HP20 is the most popular non-polar aromatic-based synthetic sorbent with particle density of 1.01 g/mL, specific surface area of 600 m2/g and large pore size (200–300 Angstrom), which enables adsorption of large natural products and organic compounds with molecular mass > 1000 Da, such as peptides, proteins and phenols [164]. This sorbent can adsorb many lipophilic marine toxins and some hydrophilic toxins (see Section 3.4) and has exhibited the reliable adsorption and recovery of the freshwater hepatotoxins, microcystin (MC) MC-LR, MC-YR, MC-LA, and MC-RR [165]. SEPABEADS SP825L and SP850 aromatic sorbents are similar to HP20 having high porosity however, they have higher surface areas and a more uniform pore-size distribution compared to HP20 (Supplementary Figure S3a) [164].
SEPABEADS SP700 (Supplementary Figure S3b) has also been reported as an effective lipophilic shellfish toxin detection sorbent. SP700 pore size is smaller than that of HP20 with pore radius of 90 Å [166]. SP700 provide greater surface area (1100 m2/g), with similar particle density (1.02 g/mL) (Supplementary Figure S3b) [167].
SF700 has applications in the separation of food additives and in food purification (refining impurities and unwanted products from food ingredients) of chemicals. In addition, SP700 are able to adsorb vitamins, antibiotics, enzyme and steroids from aqueous environment [167].

3.2.2. Modified Aromatic Adsorbents

Modified aromatic adsorbents are designed to give the highest hydrophobicity (water repellent). Modification by brominating the aromatic region enables this phase to adsorb dissolved components or those components existing in a low concentration in an aqueous solution. An example of a modified resins is SEPABEAD SP207 (Supplementary Figure S3c), which has a 1.18 g/mL particle density, particle size > 250 µm, and it is suitable for the adsorption of organic compounds at very low concentrations. Furthermore, being a hydrophobic sorbent with high selectivity for non-polar molecules, this material is suitable both in applications with upward flow (fixed bed processes) and in batch processes [166].

3.2.3. Methacrylic Adsorbents

Methacrylic series are methacrylic ester copolymers with a highly hydrophilic nature, suitable for adsorption of polyphenols and surfactants. DIAION HP2MG belongs to this group of sorbents. DIAION HP2MG is synthesised from methacrylate and has no aromatic compounds in its chemical structure (Supplementary Figure S3d). This polymer resin is of intermediate polarity and has more hydrophilic specifications in its matrix, this characteristic enables it to be used in desalting and in adsorption of high polarity organic compounds [166].

3.2.4. Other SPATT Adsorbents

Other SPATT adsorbents include AG 50W-X4 (strong Cation exchange) and Amberlite IRP-64 (weak Cation exchange) phases which both have been shown to adsorb the polar cyanotoxin anatoxin-a [168]. HLB (Hydrophilic-lipophilic polymeric resin) show a high adsorption capability for freshwater hepatotoxins microcystins (MC), MC-RR and MC-LR [6].

3.2.5. SPATT Sorbent Comparisons

Since SPATT bags were first developed, numerous studies have assessed the efficiency of adsorbing a range of aquatic toxins (i.e., amount recovered compared to concentration in environmental water) [116]. Table 5 shows an overview of the main results from in vitro and in situ studies conducted on SPATT technology and different sorbent substrate types for monitoring microalgae and cyanobacteria in marine and freshwater environments.
DIAION HP20 is inexpensive (approximately EUR 200/kg), widely used and has been tested for collecting a variety of lipophilic and hydrophilic toxins. Zendong et al. [154] illustrated HP20 resins use in SPATT bags and determined a strong correlation between toxins extracted from SPATT devices and toxins levels in phytoplankton rich aqueous solutions. This study also investigated the amounts of OA, DTX1, PTX2 accumulating according to resin mass (0.3, 3 and 10 g), concluding that 3.0 g of sorbent had better extraction results compared to 0.3 g but the larger 10 g quantity was not recommended because of the risk of clogging.
Caillaud et al. [169] reported 89% and 66.2% recovery efficiencies of ciguatoxins (CTX1B) and maitotoxins (MTX), respectively, from 10 g of wet DIAION HP20 resin after 72-hrs exposure. Amounts of 4, 2.8, 0.3, 0.4, 0.5, 0.2 ng/g HP20 resin were detected for OA, YTX, AZA1, PTX2, GYM and SPX-13-DesMe-C, respectively.
A study conducted by Fux et al. [5] off the Irish coast reported 36% and 62% recovery efficiency of OA, DTX1, PTX2 and YTX from DIAION SP700 and HP2MG, respectively. PTX2, OA and YTX showed 50%, 60% and 80% extraction ability, respectively, from the HP20 resin [116].
Lane et al. [170] applied HP20, SP700, SP207, SP207SS to detect DA and PST. All resin sorbents were successful in extracting DA although the HP20 was reported as the most effective. For lipophobic biotoxins, the efficiency order was reported as: HP20 > SP700 > SP207 > SP207SS.
Oasis HLB, Strata-X and HP20 resins were tested alongside low-density polyethylene strips and silicone rubber strips for the detection of OA, AZAs, PnTX-G, SPX1, PITXs in laboratory and field tests [2]. Polymeric strips yielded significantly lower recoveries than sorbents, and the toxins of all groups were absorbed more slowly by HP20 when compared to Strata-X and Oasis HLB. All sorbents showed different efficiencies and accumulation speeds with Strata-X and Oasis HLB proving to be more suitable for short-term deployments or in-field evaluation of toxin presence, but HP20 was more appropriate for exposure periods of greater than 5 days.
SPATT bags containing PAC and Strata-X were deployed in a river containing toxic benthic cyanobacterial mats by Wood et al. [162]. Powdered activated carbon (PAC), a cheap, widely available sorbent that slightly outperformed Strata-X in tests, with anatoxin-a (ATX) and homoanatoxin-a (HTX) toxin substrates, giving 45% desorption recovery compared to 42% recovery from Strata-X.
SP700 and HP20 have been compared in several studies [171]. SP700 showed more rapid short-term accumulation but no significant difference after 72 h exposure. However, Fux et al. [172] found HP20 did not reach equilibrium in that time frame (72 h), but in long-term field trials HP20 showed significantly greater adsorption potential than other sorbent resins such as SP700. Other researchers, taking desorption into account, found that the lower recovery efficiency of SP700 (78% and 72% for MC-LR and [Dha7] MC-LR freshwater toxins, respectively, compared to 91% and 89% for HP20) offset speed of adsorption and recommended HP20 for toxin tracking [173]. Adsorption data analysed by Li et al. [174] indicated SP700 has a lower capacity for OA and DSP (1088 and 1872 g/g) toxins than HP20 (1639 and 2934 g/g), respectively, and that adsorption capacity is determined primarily by pore size distribution and analyte polarity, rather than sorbent specific surface area [174].

3.3. SPATT Bag Preparation and Analyte Extraction

SPATT bags are typically prepared in the following steps: (i) cleaning, then rinsing, with deionized water, (ii) activation by exposure to 100% MeOH, (iii) dispersal of specific volumes into plankton net or nylon mesh and sealing. The activation and dispersal sequence varies in different studies, for instance dry resin can be dispersed through the mesh before the activation process [170].
Synthetic sorbents are spherical cross-linked polymers, that due to this specific structure, show noticeable swelling in the presence of polar solvents such as methanol and water [175]. Extraction of toxin substrates from a resin sorbent involves several steps [159], and with different eluates as outlined in Table 3. The choice of eluates varies based on the type of targeted toxin. For example, with toxins PTX, PTX-2 SA, PTX-11, PTX-11 SA, OA, OA-ester or YTX, the optimised elution solvent is 100% MeOH [2,156]. In addition, DA and CTXs elution was reported as requiring three different elutes in each step starting with MeOH in water, followed up with ammonium acetate in MeOH for the last two steps [176] (Table 3 and Table 4).

3.4. SPATT Bag Storage and Stability

Methods for SPATT bag storage reported in various studies typically focus on stability issues. It is necessary for bags to remain hydrated before and after deployment, therefore researchers commonly report storing them in deionized water at 4–6 °C prior to deployment [159]. In addition, after deployment, the SPATT bags are required to be soaked in the elution solvent (elute) immediately or be kept in storage at −20 °C [159]. SPATT bags stored in the freezer (−20 °C) were stable with no loss of biotoxins for up to three months; to-date no longer freezer (−20 °C) storage duration has been assessed. In addition, under the same conditions there are no sign of degradation of extracts up to 12 months [6,165]. The advantages of the SPATT monitoring include:
(i)
simplicity, low cost, ease of application, transport and storage [160];
(ii)
allows sampling throughout the water column where no shellfish exist naturally [19];
(iii)
targeting toxin substances directly [19,160];
(iv)
impervious to biotransformation with no sign of degradation when stored in −20 °C [160];
(v)
a sufficient pre-concentration technique to ensure adequate adsorption and analytical detection;
(vi)
can be used as an early warning system for bloom events when coupled with appropriate analysis (e.g., ELISA, LC-MS) [160];
(vii)
reveals unique information on toxins such as origin, environmental durability, and variations in the specific toxicity [6,160];
(viii)
profiles the water for toxins generated by HABs prior to their biochemical transformations within shellfish tissues that leads to a variety of toxin derivatives and
(ix)
assesses biotoxin frequency, and the duration of algae blooms in a specific region [160].
Phytoplankton monitoring disadvantages include:
(i)
difficult to detect spatially and temporally integrated water samples [116];
(ii)
monitoring only shows the evidence of a possible shellfish contamination [116];
(iii)
phytoplankton monitoring is intensive, difficult to identify and needs specifically skilled observers [116].
The difficulties of routine testing of shellfish include:
(i)
high cost of instrumentation with training requirements and complex sample preparation and clean-up optimisation and validation processes;
(ii)
biochemical transformations within shellfish tissues leads to a variety of toxin derivatives, a more complex toxin profile than what originated from the HABs and
(iii)
matrix problems from biological samples makes the extraction and analysis slow, and sometimes analytically challenging [2,5].

3.5. SPATT Applications

3.5.1. Application of SPATT to Marine and Freshwater Toxins

Solid Phase Adsorbent and Toxin Tracking (SPATT) is a powerful, reliable and efficient biotoxin monitoring tool applicable to both marine and freshwater environments. Harmful Algae Blooms (HAB) can produce toxins that cause illness or fatalities following consumption of contaminated shellfish and seafood [47], however, the amount of toxin accumulated in shellfish varies based on toxin production and growth stage so the presence of toxin does not necessarily mean that shellfish are a contamination risk [165]. HAB proliferation has resulted from eutrophication, unregulated ballast water discharges and climate change. Total toxin regulatory limits for marine biotoxin in seafood has been set by EC No 853/2004 regulations [28]. The SPATT tool coupled with a laboratory based analytical approach such as LC-MS/MS provides a selective and highly sensitive tool for detecting biotoxins. Table 5 (Marine and Fresh water applications) reviews the use of SPATT tool in monitoring toxin-producing algal, results indicate that SPATT performs well for monitoring dissolved polar (domoic acid) and non-polar (polyether) biotoxins species and SPATT can also be used as an early warning, forecasting system [116].
The first SPATT bag was designed by MacKenzie et al. in 2004 to detect dissolved biotoxins during diarrhetic shellfish poisoning (DSP) and yessotoxin (YTX) blooms caused by Dinophysis acuminate and Protoceratium reticulatum, respectively. Such an early warning method may predict the net accumulation of polyether marine biotoxins in mussels. HP20 with a 3 g dry weight (12 mL in solution) sorbent was placed and sealed in the 95 µm polyester bags and conditioned with MeOH and rinsed with MQ-water in advance of deployment and bags deployed at different depths to track cumulative adsorption within the water column. The extraction process performed after initially washing the retrieved SPATT with ultrapure water (Milli-Q water) to remove salt residues, involved removing the phase from the SPATT bag and transferring the phase to an empty solid phase extraction (SPE) tube with a glass filter. The adsorbent phase is eluted with 100% MeOH. Recoveries were determined by LC-MS/MS [116].
Fux et al. studied the uptake and extraction behaviour of OA and DTX1 applying five different sorbents, DIAION HP-20, SP850, Sepabeads1 SP825L, Amberlite1 XAD4, Dowex1 Optipore1 L-493. Results demonstrated that HP20, SP850 and SP825L are similar in adsorbing the lipophilic marine biotoxins with recoveries of 99%, 97% and 98%, respectively [172]. In addition, Fux et al. applied HP20 to the adsorption of numerous lipophilic toxins and demonstrated that SPATT disks are sensitive tools for toxin profile investigations and suggested that the OA toxin groups in the absence of toxic phytoplankton did not result in shellfish contamination. It means that feeding bivalves on toxic phytoplankton enhance the accumulation of toxins such as the OA or DTX group [177]. Furthermore, SEPABEADS SP825L, SP850 & SP700 were applied in the adsorption studies of OA, PTX2, AZA and YTX [160]. Recoveries of SP700 for OA (61%), PTX2 (22%), AZA (41%), YTX (47%) was reported by MacKenzie [160]; HP20, SP825 and SP850 showed similar recovery results. HP20 having a larger pore size could not reach the equilibrium within 72 hrs as achieved for sorbents SP850 and SP825L. Fux et, al. related this to the diffusion concept film diffusion, that is the migration of molecules to the surface of the particles and internal diffusion that is the migration from the surface to the internal part of the resin particles [172]. SPATT was used to determine recovery characteristics of MC-LR, MC-YR, MC-LA and MC-RR in several studies [178]. The recovery results showed the suitability of SPATT for monitoring microcystin (MC) toxins in freshwater and marine environments [157,158,179]. Stata-X polymeric resin was successfully used in SPATT bags for monitoring cyanobacteria, ATX and HTX, in river waters [162]. A comparison study on individual HP20, Strata-X, BundEly C18 and Oasis HLB SPATT bags showed that the adsorption rate on Oasis HLB and Strata-X were higher than on HP20, although HP20 and Strata-X gave a higher recovery after 24 h of exposure [2].
Table 5. Marine and Freshwater applications.
Table 5. Marine and Freshwater applications.
SPATT ResinsToxins
Detected
YearCountryEluteApplicationDeployment
Condition
Adsorbent
Quantity
Analyte
% Recoveries
Ref.
DIAION HP-20
(Bags)
PTX, PTX2 SA, PTX11, PT11 SA, OA, OA-ester & YTX.2004NZMeOH
Ace
MeOH > Ace
MarineDeployed at selected depth12 mL = 3 g dry weightAve. = 62%[116]
DIAION SP-207
(Bags)
OA, DTX1, PTX2, YTX, 36% less
recovery than
DIAION HP-20
2004NZMeOH
Ace
MeOH > Ace
MarineDeployed at
selected depth
12 mL = 3 g dry weightAve. = 36%[116]
DIAION HP-20 (Large scale pumping)OA DTX-2 PTX-2 PTX-2SA2007NorwayMeOHMarineSeawater0.5
kg/column
DTX-2: 73%
OA: 78%
Accumulation: 2.7 mg OA,
1.3 mg DTX-2 and 1.8 mg PTX-2 during an 18-h
[161]
DIAION HP-20, SP850, Sepabeads1 SP825L, Amberlite1 XAD4, Dowex1 Optipore1 L-493 (Bags & Disks)OA, DTX2, PTX2, AZA1, -2 and -32008IrelandMeOHMarineDeployed3 gOA and DTX1 were determined in positive ionisation mode[172]
Membrane (Polycarbonate, polyethersulfone, polyester, nylon) and POCIS Oasis HLBMC-RR, MC- LR2008Czech
Republic
90:10 v/v
MeOH/water Acidified with 0.1% TFA
Fresh water Exposed
in a natural
reservoir
Membrane
exchanging area: 47.5 cm2, Oasis HLB: 2.75, 5.55, 11.1 mg/cm2
MC-RR: 0.022 L/day, MC-LR: 0.017 L/day[168]
POCIS Oasis HLB29 organic chemicals: Antibiotics, fungicides, herbicides, biocides2009SpainMeOH acidified in three
different levels
MarineFish farm200 mgThe
detected conc’s do not have impact on aquatic
organism
[179]
Diaion HP20
(disk)
D. acuta bloom2009SpainMeOHMarineDeployed at different depth3 gPlankton: PTX2 ranged from 19–73 pg/cell[180]
DIAION HP-20 (Disks) OA, PTX, YTX & AZA group 2009IrelandMeOHMarineDeployed at different depth3 gAccumulation rate of toxins in the mussels and SPATT discs correlated [177]
DIAION HP-20
(Disks)
20-methylSPX-G, AZA-1, AZA-2, OA, DTX-1, DTX-2, PTX-2, PTX-12 & YTX.2009NorwayMeOHMarine and
Fresh water
Deployed attaching to a fixed point (1 m depth)3 gPTX-2: 5–40 ng/disk
20-methylSPX-G: 706.5 ng/disk
SPX-C: 164.2 ng/disk
[181]
SEPABEADS SP825L, SP850 & SP700
(Bags)
OA, PTX2, AZA, YTX2010NZMeOHMarineDeployed-OA: 61%
PTX2: 22%
AZA: 41%
YTX: 47%
[160]
DIAION HP-20
(Disks)
OA, PTX, YTX and AZA group 2010IrelandMeOHMarine
(Deployed in four different depth)
Deployed Different depth3 g Recovery discussed based on
period and depth
[5]
HP20, SP700, SP207, SP207SSDomoic acid and
saxitoxin
2010USAMeOHCoastal Deployed3 gSP700:
69–72%
HP20: 99%
[170]
DIAON HP20MCY-RR and -LR, 2010USAMeOHFreshwater and Lake waterDeployed3 g2.9 million ppb[157]
Strata-X
(Bags)
ATX, HTX,
Dihydroanatoxin-a, Dihydrohomotoxin-a
2011NZMeOHFreshwaterRiver
1.2 m3/s
1 g7% of water loading[162]
PAC G-60
(Bags)
ATX, HTX, Dihydroanatoxin-a, Dihydrohomotoxin-a2011NZ5% FA and 70% MeOHFreshwaterRiver
1.2 m3/s
1 g4–12% of water
loading
[162]
Diaion HP20CTX, MTX2011SpainMeOHMarine
G. Pacificus
culture
In vitro experiment10 g wetCTX1B: 85.5–90.9%
MTX: 66.2%
[169]
SP700PSP toxins2010SpainMeOHMarinePSP and LSTs producing culture1 gGTX2,3: 406.02 ± 13.30 ng/g resin
STX: 219.02 ± 37.71 ng/g resin
[182]
DIAION HP-20
(Disks)
Spirolide C,
iso-spirolide C,13-desmethylspirolide C, 20-methylspirolide G
2011NorwayMeOHMarineDeployed3 gSpirolide C: 69%,
iso-spirolide C: 13%,
13-desmethylspirolide C: 22%
20-methylspirolide
G: 77%
13,19-didesmethylspirolide
C: 33%
[183]
SEPABEADS SP700Toxic Alexandrium
okadaic acid, 13-desmethyl SPX C, 20-methyl SPX G
2011IrelandMeOHHarbour Deployed
in water
5 gOA, DTX-2 and PTX: 2.5 ng/g [184]
HP20Pinnatoxin (PnTx), analogues PnTx-E, PnTx-F, okadaic acid (OA) and its esters2011NZ MeOHMarineDeployed over two summers4 gOA: 14%
PTXs: 50%
OA-esters: 10%
[185]
DIAION HP20MC-LR, -YR, -LA, and -RR2011USAMeOHFreshwaterDeployed for 16 months3 gMC-LR: 66.4 ng/L, 18,400 ng/g resin equal parts MC-RR, MC-YR, MC-LR[156]
DIAION HP20Chlorophyll-a,
Secchi depth, total phosphorus and
total nitrogen
2012NZMeOHLakeDeployed3 g CYN82/91, CYN83/87/95) and the Calothrix sp. (CYN100) had low similarities (<94%) to GenBank
sequences
[186]
HP20 and SP700Cyanobacterial
cultures
2013ChinaMeOHFreshwater
M. aeruginosa cultures
Deployed 2 gHP20 better result than SP700[173]
HP20OA, DTX2, PTX2)2013SpainMeOHMarineDeployed at 3, 7 and 12 m depths2.5 gOA/DTX2:
1.5–6.0 ng
PTX2:
1.8–7.0 ng
PTX2SA:
0.5–3.0 ng
[187]
Diaion HP20, Strata-X, Oasis HLB, BondElut C18SPX1, AZA1,
PnTX-G, ovTX-a
2014FranceMeOHSeawater
Agilent reservoir
Conditioning method the same as
Shea et al. 2006
Oasis (30 mg), Strata-X (200 mg), HP20 (200 mg), Bond Elut C18 (200 mg)SPX1: 14 ng, AZA1:
19 ng, PnTX-G:
238 ng, ovTX-a:
359 ng
[2]
HP20
XAD761
DSP toxins2014IrelandMeOH: Water
80:20
MarineDeployed at different depth5 g of each separatelyHP20
XAD761
[165]
DIAON HP20Cyclic imines
(SPXs, PnTXs, GYMs)
2014Spain MeOHMarineDeployed10 g wetDIAON HP20[188]
DIAION HP-20 resinMicrocystin-LR, RR, YR, LA2014USMeOHFreshwater Deployed blow surface water3 gDIAION HP-20 resin[189]
HP20A range of lipophilic toxins2015FranceMeOHMarineDeployed300 mgHP20[190]
Amberlite XAD761, HP20OA, PTX2, DTX22015Ireland80:20
MeOH: Water
MarineDeployed in different depth5 g dry weightAmberlite XAD761, HP20[6]
Diaion HP20OA, DTX1, DTX2, PTX22016SpainMeOHMarineDeployed in semi enclosed river at 3 m depth10 gOA 17.75 pg/cell
PTX2 13.2 pg/cell
DTX1 trace amount
[191]
HP20PTX2, D. fortii, D. acuminate, P. rotundatum, OA, DTX12016ChinaMeOHMarineDeployed at 8 m depth3 gD. fortii (0.28 pg/cell), D. acuminata complex (0.08 pg/cell) & P. rotundatum (D. rotundata) (0.02 pg/cell). PTX2 (nd~5.7 mg/kg), OA (nd~2.8 mg/kg) and DTX1 (nd~1.6 mg/kg)[192]
HP20OA, DTX1, PTX2,
PTX2sa, 13-desMe-C, PnTX-G
2016FranceMeOHMarineDeployed during
summer
0.3, 3, 10 gThe higher amount of resin captured more toxins [154]
HP20DSTs, AZA2017USAMeOHMarineDeployed3 gConc’s during different month are discussed.[193]
SPATT
HP20
Microcystins2017USAMeOHWater reservoir and lakeDeployed at different sites3 gMC-LR: ~88%, MC-YR: ~100%, MC-LA: ~100%[178]
HP20Gambierdiscus
toxins (CTXs)
2018FranceMeOHMarineDeployed 20 g55 ng P-CTX-3C equiv./g resin[194]
DIAION HP20DSTs2018USAMeOH & then
MeOH: ammonium acetate
MarineDeployed3 gFour toxins were identified in 37% of mussels. one toxin in 99% of
mussels
[176]
Strata-XToxic Microcoleus autumnalis (Basionym Phormidium autumnale)-dominated2018NZMeOH acidified with FAStream waterDeployed in River 10, 20, 40 m1 g0.91 ng mL−1 and 95 ng g−1 of strata-x hr−1[195]
HP20 (Diaion) and XAD-2 (Amberlite) OA, STX, DTX1, PTX2, PTX2 isomers2018USAACN acidified with FAMarineDeployed 3 gFor both
resins: OA: 53%
DTX1: 20%
Esterified OA: 19%
Esterified DTX1: 8%
PTX2: 88%
PTX2 isomers: 5%
PTX11: 4%
secoPTX2: 3%
[196]
Diaion HP20PCTXs, MTXs2018NZDCM and aqueous MeOHMarineIn vitro2.5–10 gPCTX- 3C (70%)
P-CTX-1B (92%).
MTX3 not possible to detect
[197]
HP20Domoic acid (DA), saxitoxin (STX), okadaic acid (OA)2019USAExtract 1 50% MeOH (v/v) and Extract 2 and Extract 3 with 1 M C2H7NO2 in 50% MeOH MarineDeployed DA from 9.2 to 37 ng/g
STX from 1.3 to 5.3 ng/g
[198]
HP20Phycotoxin
pectenotoxin-2 (PTX2)
2020AntarcticaMeOHMarineDeployed in cove10 gVery low background conc. [199]
Abbreviation: MeOH (Methanol), Formic acid (FA), DCM (Dichlorometane), Ace (Acetone), ACN (Acetonitrile), Ammonia Acetate (C2H7NO2), New Zealand (NZ).

3.5.2. SPATT Sorbents and Biotoxin Harvesting

Rundberget et al. exploited the affinity shown by synthetic adsorbents for biotoxins by developing a large-scale toxin harvesting system, conceived to overcome the problems encountered by toxicology researchers and scientists developing analytical methods in obtaining sufficient quantities of pure toxins and their metabolites [161]. The system comprised novel filter configurations to release toxins from cells while eliminating debris and unwanted insoluble compounds, and a solid phase extraction column of fixed-bed design was used to gather analytes as shown in (Supplementary Figure S4). The system included a submersible pump (A), a pre-filter with 100 µm plankton net (B), 50 mm filter (C), flow distributor (D), columns contain adsorbent resin. On the right-hand side picture of the pre-concentration device, as the second pump has shown, the system passes the water through a 40 µm filter to increase the algal cell concentration (10-fold), the second pump passes the concentrated water to the pumping system. With this apparatus, quantities of okadaic acid and dinophysistoxin-2 were obtained from algal blooms in Spain and Norway, with results indicating that the method was also applicable for collecting toxins such as azaspiracids, pectenotoxins, spirolides and microcystins from in-field sources or microalgal cultures [161].

4. Polar Organic Chemical Integrative Sampler (POCIS)

The ‘Polar Organic Chemical Integrative Sampler’ device is a semi-quantitative time integrative sampler or qualitative sampling tool for separating hydrophilic or polar organic compounds with different physicochemical properties from water [200]. POCIS has been using in screening, monitoring and in the determination of dissolved bioavailable contaminants, pharmaceutical residues [201], endocrine disrupting substances [202], pesticides [203], per-fluorinated compounds and mixtures of chemical compounds [204].
POCIS systems comprise one or more analyte receptor capsules directly exposed to environmental water or mounted on a metal frame held in a perforated canister through which water passes freely. Capsules are made up of a sorbent phase component, chosen for its affinity for targeted compounds, trapped between biofouling-resistant microporous polyethersulfone membranes (pore size 0.1 µm). Capsule assemblies are circular or rectangular, secured by stainless steel rings (Supplementary Figure S5). The water matrix, which can include a mixture of dissolved compounds such as pesticides, personal care products and pharmaceuticals [205] can pass through the membrane while chemicals of interest are trapped in the sorbent [206,207]. Total membrane areas are in the order of 41–46 cm2 with sorbent masses of 228 mg [205,208], give surface area to mass ratios of around 180 cm2/g (membrane diameter 3.3 cm giving surface area 17 cm2). Deployment durations reported in different studies range between one week to one month (see Section 1)
Accumulation takes place during three successive steps simultaneously: (i) diffusion of the dissolved compound crossing the water boundary layer, in which the thickness of this boundary layer, varies based on the water flow velocity and turbulence around the sample [205]. The water boundary layer is a result of friction between water and sampler that causes a higher viscosity on the surface of the sampler [209]; (ii) the next step is compound transportation. The transportation can occur through two routes, either through the water-filled pores of the PES membrane or through the membrane polymer itself; (iii) the final step is adsorption, that is penetration of the analyte into sorbent resin [205]. Indeed, the challenge of applying POCIS and understanding the accumulation of different compounds and sampling rates is because of the last two steps. Advantages of this multi-transport method include the ability to measure time-weighted average concentrations of analytes, the ability to detect ultra-trace micropollutants, the capacity to use different membranes, its simple construction and the omission of a high-power demand [210].
Analyte components with log values between (0 to 5) 0 ≤ Kow ≤ 5, in which Kow refers to the ratio (without a unit) of concentration of solute to determine the hydrophobicity (the lower the log Kow, the higher solubility of solute in the water) [208], are usually targeted by POCIS sorbents, for example the copolymer poly(divinylbenzene)-co-N-vinylpyrrolidone, [206,207]. Proprietary sorbent mixtures are available for target applications, for example triphasic sorbent admixture included in ‘pest-POCIS’ (10 mg sorbent per POCIS disk) for pesticide screening or ‘pharm-POCIS’ which is adaptable for both pharmaceuticals and pesticides monitoring and shows a higher efficiency for sampling most analytes (Supplementary Figure S5) [207].

4.1. POCIS Applications in the Marine Environment

Bioactive pharmaceutical residues, as metabolites or original compounds, that end up in the environment through effluent rejection after treatment from wastewater treatment plants, can accumulate and negatively affect the environment by interacting with marine organisms and exacerbating current environmental stresses such as climate change or eutrophication [211]. The use of Polar Organic Chemical Integrative Sampling (POCIS) as a qualitative and semi-quantitative tool for the analysis and monitoring of pharmaceutical compounds in marine environments has been reported [211,212,213]. The determination of the uptake rate or sampling rate (Rs) of the compounds onto the resin sorbent is important in all POCIS studies. The uptake rate varies for different compounds and metabolites. The parameters that can affect the sampling rate include water flow, amount of resin sorbent, temperature, dissolved organic compounds (DOC) and conductivity [212]. Table 6 shows different sorbents used in POCIS for the adsorption of a variety of compounds such as pharmaceuticals, pesticides and personal care products in the marine environment.
Carbamazepine as a highly prescribed pharmaceutical ingredient is the most frequently detected pharmaceutical compound found in marine environment, it is highly resistance to biodegradation and elimination after water treatment. Buento et al. studied carbamazepine and oxacarbamazapine and their related metabolites in marine environment in the Mediterranean Sea using 200 mg Oasis HLB sorbent, at 21 °C sea water temperature. Average recoveries of all the compounds reported was >94%, excluding oxacarmazapine which was <69% [212]. However, modified POCIS using 600 mg Strata-X resin sorbent and Chemcatcher (SDB-RPS or SDB-XC) at pH 6.5 was successful for the detection of caffeine, carbamazapine, capsone, DEET, hydrochlorothiazide and troclosan [214]. However, Chemcatcher SDB-RPS showed a higher accumulation in comparison with Chemcatcher SDB-XC. Chemcatcher with SDB-RPS, due to having the larger particle size showed less linear accumulation profile however, most neutral pesticides and personal care and pharmaceutical products produce linear accumulation on both Chemcatcher and Strata-X, although, in addition codeine at pH 6.5 showed a linear profile [214]. Furthermore, POCIS was used in the marine environment and 93 pharmaceuticals in 43 different sites were discovered in the Baltic Sea and in the Skagerrac strait, Jutland peninsula of Denmark using Oasis HLB as the POCIS absorbent. In this study, 200 mg HLB sorbent was sandwiched between a polyethersulphone (PES) membrane. Carbamazepine was the most frequently obtained compound present in 37 out of 43 samples and reported to be between 0.57–3.2 ng/L in different locations around the Baltic Sea [215].
Table 6. The application of POCIS in the marine environment.
Table 6. The application of POCIS in the marine environment.
POCIS ResinsAnalyteYearCountryEluteApplicationDeployment ConditionAdsorbent QuantityAnalyte
%Recoveries
Ref.
Modified POCIS Strata-X and Chemcatcher™ (SDB-RPS)Caffeine, Carbamazapine, Dapsone, DEET,
Hydrochlorothiazide,
Troclosan
2014AustraliaMeOH, ACN & AcePharmaceuticalsMarine and freshwater Environments (grab sampling & passive sampler)600 mg PES membranes (47 mm diameter) with 147 mm thickness and a pore size of 0.2 mm (used on Chemcatcher with SDB-RPS) or 140 mm thickness and a pore size of 0.45 mmCaf: 102%
CBZ: 104%
Dap: 74%
DEET: 77%
Hydro: 99%
Tro: 84
[214]
POCIS Oasis HLBCarbamazapine (CBZ), Oxacarbamazapine (Ox), and their related metabolites2014Mediterranean SeaMeOHPharmaceuticals one POCIS disk was placed into glass aquaria containing 1.5 L of filtered spiked seawater at 5 μg/LMarine
environment (lab
experiment)
200 mgCBZ: 110 ± 4 (5 ng), 95 ± 11 (10 ng) & OX: 58 ± 7 (5 ng), 69 ± 3
(10 ng)
[212]
POCIS HLB
PES membrane
93 pharmaceuticals2018SwedenDCM/ACN (8/2, v/v), & DCMPharmaceuticalsMarine (grab sampling) 200 mgConc’s ranging between 0.01 & 80 ng/L[215]
Abbreviation: MeOH (Methanol), EtOAc (Ethyl acetate), DCM (Dichlorometane), Ace (Acetone), ACN (Acetonitrile).

4.2. POCIS and Wastewater Monitoring

The application of POCIS in sampling wastewater treatment discharge to detect compounds such as pharmaceuticals, beta-blockers, pesticides and personal care products over traditional grab sampling methods has been well developed (Table 7, Table 8, Table 9 and Table 10). Six drugs, azithromycin, fluoxetine, omeprazole, levothyroxine were the most frequent prescribed drugs in the United States, and two illicit drugs, methamphetamine and 3,4-Methylenedioxymethamphetamine (MDMA), as polar components were monitored in wastewater effluent using 200 mg Oasis HLB. The recovery values, azithromycin 15–66 ng/L, methamphetamine, 2 ng/L and methylenedioxymethamphetamine 0.5 ng/L showed that although the concentration of these substances in the wastewater is not high there is the concern that accumulation in wastewater can cause chronic effects on human health as well as on native biota, that are continuously exposed to those substances [216].
Oasis HLB is the most popular sorbent used in passive sampling and POCIS in wastewater studies. Studies indicate 200 mg of sorbent is sufficient sorbent mass for sampling analytes from sewage however, there are some other studies that showed 220 mg HLB sorbent to be the optimum mass to sample a range of pharmaceuticals such as neutral pharmaceutical ingredients, antibiotics and illicit drugs [217]. In addition to the amount of sorbent used, other parameters such as sampling rate, temperature and turbulence had an impact on the uptake of these compounds by the sorbent. Furthermore, artificial sweeteners such as sucralose, acesulfame and beta-blockers showed a high affinity to POCIS sorbent. Sucralose (128–213 ng/L) and acesulfame (4–33 ng/L) was reported in treated wastewater [217,218]. In contrast, applying POCIS using 30 mg HLB sorbent for five perfluoroalkyl substances in surface water and hospital wastewater, showed that all five substances could accumulate in POCIS with a concentration range of 6.5 × 10−3 to 3.6 × 10−1 nmol/L in the influent and accumulate in POCIS at a concentration of 1.3 × 10−2–2.2 × 10−1 nmol/L in the effluent [136]. Although, the efficiency of HLB in POCIS has been discussed in different studies, the efficiency of some other modified POCIS using different type of sorbents in sampling of biocides, pharmaceutical ingredients, organophosphates, beta-blockers and chlorinated pollutants have also been studied [219]. Octadecyl-functionalized silica gel (C18) and triphasic sorbents were applied in a POCIS device for the extraction and monitoring of alkylphenols (APs), hormones, bisphenol-A (BPA), synthetic musk fragrances and herbicides (trifluralin (Tri) and alachlor (Ala)) in wastewater treatment plant effluent. Determination of the sampling rate and uptake behaviour of those compounds demonstrated that non-polar C18 sorbent shows acceptable results for analysing these compounds which have moderate polarities such as 4t-OP (octylphenol) alkylphenols with 59% recovery, however for other alkylphosphenols compounds with hydrophobic natures such as NPs (nonylphenols) with 47% recovery, 4n-OP alkylphenols with 38% recovery and MeEE2 with 15% recovery, non-polar C18 is not a satisfactory sorbent [219].
POCIS is also known and accepted as a powerful and reliable tool in monitoring a wide range of non-pharmaceuticals, Challis et al. assessed the application of POCIS in the evaluation of non-targeted analytes for high resolution mass spectrometry of wastewater. They presented this study in comparison with the application of organic-diffusive gradients in thin-films (o-DGT) for the determination of both targeted and non-targeted compounds. The efficiency of using 200 mg Oasis HLB in POCIS for the evaluation of neonicotinoids, organophosphates, triazines, antibiotics, b-blockers, SSRI’s and sodium channel blockers was compared to o-DGT samplers which were constructed using two layered gels made of 1.5% agarose. This study suggested o-DGT was more reliable for TWA (Time-Weighted Average) determination contaminant concentrations during the deployment period, this is due to the o-DGT feature, having a diffusive hydrogel membrane consisting of 96–98% water together with the control of the viscosity of water, which reduces the water boundary layer effects [201].
Table 7. Application of POCIS on wastewater.
Table 7. Application of POCIS on wastewater.
POCIS ResinsAnalyteYearCountryEluteApplicationDeployment ConditionAdsorbent
Quantity
Analyte
%Recoveries
Ref
Oasis HLBOmeprazole, fluoxetine, azithromycin, levothyroxine, methamphetamine, and methylenedioxymethamphetamine2004USAMeOHPharmaceuticalsWW River water Deployed200 mgAzithromycin 15–66 ng/L, methamphetamine, 2 ng/L, methylenedioxymethamphetamine 0.5 ng/L[216]
POCIS Oasis HLB25 pharmaceuticals and personal care products2007CanadaMeOHPharmaceuticals Uptake rates were 0.040 to 2.462 L/d in uptake rates between 0.016 and 0.223 L/dWW and SW200 mgRS values for 13 of the 25 analytes could be
determined
[220]
Oasis HLB
PES membrane
A range of Pharmaceuticals, personal care products, endocrine disrupting2010SwitzerlandMeOHPharmaceuticals,
personal care
products, endocrine disrupting
Treated WW Flow rate 2.6 and 37 cm/s200 mgDifferent recoveries based on flow rate is reported[221]
Oasis HLB (pharmaceutical & pesticide), PES membraneA range of pharmaceuticals, hormones and pesticide are reported 2011USMeOHPharmaceuticals,
steroid hormones,
pesticides
WW200 mgRecoveries as
reported in the
paper
[222]
Oasis HLB
PES membrane
Beta-blockers and hormones2012FranceMeOH containing 5% NH4OH.Pharmaceuticals, hormonesWW200 mgHormones low conc’s prevented determination of reliable sampling rates. Suitable for b-blockers[223]
POCIS Oasis HLBAtenolol, Prednisolone, Methylprednisolone, Sulfamethoxazole, Ofloxacin, Ketoprofen2013FranceMeOHPharmaceuticals Uptake profiles (25 °C, flow velocity 0.16 m/s), an automatic sampler, taking 100 mL every 15 min between 7:00 a.m. and 9:00 p.m. and 100 mL every 30 min during the night.Hospital sewage200 mg64–95%[224]
SBSE
Oasis HLB
19 moderately hydrophobic to hydrophobic pesticides2013FranceMeOH then, MeOH/EtOAc (75:25, v/v)PesticideAgricultural
WW
200 mgSBSE was able to
integrate a
concentration peak triggered by a quick flood
[225]
POCIS HLB
PES membrane
Naproxen, Ibuprofen, triclosan2014South AfricaACN, then ACN: MeOH, 50:50 (v/v),PharmaceuticalsWW200 mgNaproxen 92%,
Ibuprofen 108%, triclosan 75%
[226]
POCIS HLBAcetaminophen, caffeine, 1,7-dimethylxanthine, cotinine, dextroamphetamine, diethyltoluamide (DEET), diphenhydramine, ibuprofen, methamphetamine, carbamazepine, azithromycin, erthyromycin, lincomycin, monensin, sulfachloropyridazine, sulfamethazine, sulfadimethoxine, sulfamethazole, sulfamethoxazole, sulfamerazine, sulfathiazole, thiabendazole, tiamulin, tylosin, and ractopamine2015NebraskaMeOHPharmaceuticalsWW
effluents
200 mgResults available for two different
locations
[227]
POCIS HLB, PES membrane17-β-estradiol, 17-α-estradiol, 17-αethinylestradiol, estrone and estriol2016Czech
republic
DCM, Ace, MeOHHormonesWW200 mg3.4 8 to 3.66 at nominal steroid concentration in water 100 ng/L[228]
POCIS HLBA range of Pharmaceuticals, artificial sweeteners, food additives, antibiotics personal care product, fragrances, sugar substitutes, steroid hormone2016CanadaMeOHPharmaceuticals, artificial sweeteners, food additives, antibiotics personal care product, fragrances, sugar substitutes, steroid hormoneWW220 mgDetected
concentrations
discussed
[217]
POCIS HLBCiprofloxacin2016FranceACNPharmaceuticalHospital
effluents
200 mgIndicate a potential ecotoxicological risk[229]
POCIS HLB
PES membrane
Atrazine, carbendazim, desethylatrazine, desethylterbutylazine, diuron, S-metolachlor, terbutylazine, alprazolam, atenolol, carbamazepine, diazepam, diclofenac, ibuprofen, naproxen, 17-alpha-estradiol, 17-alpha-ethinylestradiol, 17-beta-estradiol, estriol, estrone, (BDE 28, BDE 47, BDE 99, BDE 100, BDE 153, BDE 154, PFOA, PFOS, Bisphenol A, triclosan2016Germany-Pesticides, pharmaceuticals, hormones, fluorinated surfactants, bisphenol A, triclosanTreated WW 200 mgRecovery details discussed[204]
C18 & triphasicAlkylphenols (APs),
several hormones, bisphenol-A (BPA), synthetic musk fragrances and herbicides, e.g., trifluralin (Tri) & alachlor (Ala), DES hormones
2016GermanyDCM/EtOAc/MeOH (4:4:2, v/v)Herbicides, alkyphenols, hormonesWW treatment plant200 mgRecovery
percentages vary between 6% for DES to 96% for Tri
[219]
POCIS OASIS HLBA range of Pharmaceuticals and illicit drugs2017Norway5% NH4OH in MeOH, and 5% HOAc in MeOHPharmaceuticals and illicit drugsWW220 mgResults discussed in the paper[218]
POCIS HLB
PES membrane
PFHxA, PFOA, PFHxS, PFDoDA, PFOS2017ChinaMeOH containing 5% NH4OHPerfluorinated compoundsWW30 mgConcentration shown on diagram[136]
POCIS HLB12 pharmaceuticals2017Ukraine MeOHPharmaceuticalsSurface
water
Hospital WW
200 mgRemoval patterns of pharmaceuticals were discussed based on specific physical chemical properties of
molecules
[230]
POCIS HLB
MIP membrane
BTEX, chlorinated
pollutants and
pharmaceuticals
2017Czech republicMeOH, & MeOH/DCM (1:1, v/v), then MeOHPharmaceuticalsWater
remediation
200 mg(POCIS) for the pharmaceuticals and in situ soil
microcosms for
microbial
community analysis, was proven
[231]
POCIS HLB
PES membrane
Carbamazepine (CBZ) and sucralose (SCR)2018BrazilMeOH/water (1:2, v/v),PharmaceuticalsSewage 200 mgCBZ: <LOD 3.6 ng/g, SCR: <LOD 139.9 ng/g[232]
POCISClarithromycin, metoprolol, propranolol, carbamazepine, sulfamethoxazole, Atenolol2018Canada-PharmaceuticalsWW-Recoveries
compared over three years
[233]
Modified- POCIS Strata-X PES membrane8 organophosphate flame retardants (OPFRs)2018China MeOHPharmaceuticals and their metabolitesWW200 mgResults discussed[234]
Oasis HLB PES membraneBiocides, carbamazepine, diclofenac, terbutryne, diuron, carbendazime 2020LuxembourgDCM/ACN (1:1, v/v),Pesticides, pharmaceuticalsWW200 mgResults discussed in different flooding time[203]
ODGT and POCIS Oasis HLB PES membraneNeonicotinoids, organophosphates, triazines, antibiotics, b-blockers, SSRI’s, and sodium channel blockers2020Canada-PharmaceuticalsWW200 mgQuantitative comparison of o-DGT, POCIS is discussed[201]
Abbreviation: MeOH (Methanol), EtOAc (Ethyl acetate), DCM (Dichlorometane), Ace (Acetone), ACN (Acetonitrile), HOAc (Acetic acid), NH4OH (Ammonium hydroxide), Wastewater (WW), Surface water (SW).

4.3. Application of POCIS to the Detection of Pollutants in Freshwater, Rivers, Lakes and Drinking Water Sources

Residues of personal care products including cosmetics, UV-filters, pesticides, pharmaceuticals including beta-blockers, herbicides, and pesticides in freshwater lakes and rivers is a matter of public concern and monitoring of lakes and reservoirs is necessary to evaluate the amount of pharmaceutical and personal care products (PPCPs) in drinking water pre and post treatment [235,236].
A range of pharmaceutical components from different biological classification systems (BCS) [237] and household materials along with personal care products, were monitored and identified as present in several studies [220,238]. These compounds can be introduced to surface and ground water through different routes such as household waste, clinical wastewater and landfill leachate [235].
Pharm-POCIS has been used to sample a range of pharmaceuticals, pesticides, microcystins, plasticisers and UV filters used in cosmetic personal care products from wastewater, groundwater, river and lakes [207].
Hydrophilic-Lipophilic Balanced sorbent (Oasis HLB) and Triphasic sorbent mixture [207] are commonly used sorbent phases, Oasis HLB dominates studies as the sorbent resin with the highest efficiency, although the triphasic admixture, Strata-X, Strata-XW, also demonstrate reliable results. Oasis HLB can be used to detect compounds with hydrophilic tendencies (such as benzene) and those exhibiting hydrophobic behaviour, e.g., aliphatic chains and pyrrolidone, and is suitable for the separation of low molecular weight and polar components. However, water samples containing high polarity compounds such as atrazine, desisopropylatrazine (DIA), diuron and dicamba-d3 with an acidity level of pKa < 4.5 reach thermodynamic equilibrium level with the HLB membrane quickly, thus preventing and reducing accumulation on the sorbent and limiting its use for quantitative studies [239]. However, an investigation that involved increasing the amount of sorbent from 200 mg to 600 mg showed a higher efficiency in the accumulation of highly acidic (bentazon, dicamba, mesotrione, and metsulfuron) and polar (atrazine and diuron) compounds from water samples [239]. In contrast, there are other studies that show a lower concentration of Oasis HLB sorbent such as 2.75, 5.55 and 11.10 mg/cm2 having lower mass to surface area ratio which results in a higher efficiency by creating a thin layer of sorbent that increases the water speed into the sorbent and thus increased accumulation of compounds [168]. In addition, Muller demonstrated the bioavailability of endocrine disrupting chemicals (EDS) from suspended sediments (that occurs due to flood events) on HLB sorbent. In the bioavailability screening of estrogenic compounds using 54.5 ± 0.5 mg Oasis HLB, this phase showed reliable analytical results for the following list of compounds (nonylphenol detected at 18 mg/L concentration, estrone (E1) detected at 14 ng/L concentration, 17β-estradiol (E2) detected at 0.2 ng/L concentration and 17β-ethinylestradiol EE2 detected at 0.5 ng/L concentration) [240].
An application of POCIS using 200 mg Oasis HLB for in situ studies on the most frequent prescribed drugs examined the efficiency of POCIS for monitoring azithromycin, fluoxetine, levothyroxine, omeprazole and sulfamethazole and recorded 100%, 95%, 86%, 95% and ~100% recoveries, respectively [206,241,242,243].
Other pharmaceutical recovery studies report reasonable recovery rates for carbamazepine, caffeine, aspirin, diazepam, naproxen, theophylline, amitriptyline at 71%, 77%, 66%, 105%, 61%, 81% and 97%, respectively. Other compounds detected reported recovery vales of 103%, 99%, 94%, 98% for alph-ethynylestradiol, beta-estradiol, levonorgestrel and progestron, respectively. In addition, in the same study the recovery ratios of pesticides were measured and included the following compounds diazinon, diuron, dimethacholor, atrazine, CDPU, CDPMU with 117%, 94%, 148%, 87%, 73% and 71% recoveries, respectively [244]. A separate study on 21 pharmaceuticals using 200 mg Oasis HLB accumulated from river water showed the highest recovery values for alprazolam, clenbuterol, bromazepam, ibuprofen and theophylline with 97%, 92%, 93%, 90% and 90% recoveries, respectively. In addition, the lowest recovery value reported was for aspirin at 51% [245]. Estragon steroid hormones were also evaluated using POCIS apparatus with 230 mg Oasis HLB used for monitoring oestrogens, oestrone (E1) and 17β-oestradiol (E2), 17α-ethynylestradiol (EE2) and diclofenac. The limit of detection (LOD) reported was 0.001 mg/L for pharmaceutical and endocrine disrupting chemicals (EDC) [246].
Eight classes of pesticides, carbamates, chloroacetanilides, dicarboximides, morpholines, organophosphorous, phenylureas, strobilurines and triazines were studied using POCIS and three different amounts of sorbent (60 mg, 150 mg, 500 mg) by Lissalde et al. [16]. Reasonable recoveries of 75% was reported for the test analytes using 60 mg of sorbent however, the recoveries using 150 mg and 500 mg of sorbent were much lower, reported at 22.8% and 33.6%, respectively. The examination of polar pesticides in ground water in France, using a pharma-POCIS with 450 mg Oasis HLB resin showed the enhanced ability of applying passive samplers over traditional manual sampling methods in detecting trace substances in ground water [247].
Other resin sorbents studies for detecting pharmaceutical and pesticides in river and ground water, have used Strata-X using a nylon Membrane and mixed Polymer Sampler (MPS) [248,249,250,251]. A copolymer of poly(divinylbenzene)-N-vinylpyrrolidone was used as resin sorbent for monitoring 46 pesticides, 17 pharmaceuticals and some artificial sweeteners [252]. Strata-X CW was chosen for monitoring pollutants such as pharmaceuticals, pesticides and corrosion inhibitors in river water, and this resin sorbent was compared with Oasis HLB [253]. The recovery results for Strata-X CW and Oasis HLB were 18–75% and 64–97%, respectively, which shows that although the same amount of sorbent, 200 mg, was used in both studies, the hydrophilic-lipophilic balance sorbent (Oasis HLB) had better results than Strata-X in the up-take of compounds. However, in this study metformin was an exceptional compound showing only 1% recovery [253]. In addition, a modified POCIS device using 600 mg Strata-XAW and a polyrthersulfone (PES) membrane to assess the effects of water velocity on the concentration of the analyte, was used for monitoring prefluorinated chemicals (PFCs) in river water. A comparison between PFCs with a molecular weight ≤ 464 and PFCs with molecular weight ≥ 500 showed that by increasing water flow rate, the sampling rate improved for those PFCs with molecular weight ≤ 464 [254].
Several lake water pollution studies demonstrated the capability of POCIS with Oasis HLB to detect pesticides and herbicides. A 200 mg device was used to evaluate pesticides such as neonicotinoid insecticides (NNIs), thiamethoxam, clothianidin, imidacloprid, [255] atrazine, azadirachtin, carbofuran, chlorpyrifos, cypermethrin, dieldrin, imidacloprid, and profenofos (0.1 μg/L at 30.8 ± 1.3 °C) [256]. Additionally, atrazine, diurom, 2,4-D, mecoprope, fluconazole, climbazole and chlorothalonil were evaluated in the western lake of Ontario [257]. The concentrations of endocrine disruptions such as bisphenol A, 17β-estradiol, estrone and 4-nonylphenol in fish in lakes were detected in ppm concentration levels [258]. In addition, monitoring pharmaceuticals in lake water has been reported for 35 active pharmaceutical ingredients (APIs) with the recoveries of 26 compounds detected between 80–120% in which it appeared that 60 mg HLB is appropriate for the evaluation of 1 L sample quantities [257,259]. Brophy et al. assessed microcystin-LR (MC-LR) absorption using POCIS with 220 mg Oasis HLB resin in lake water, the sampling rate showed 0.045 and 0.041 L/day for concentrations of 0.5 and 1 µg/L, respectively. These values detected from POCIS are higher than the MC-LR concentrations detected from the traditional grab sampling method. The grab sampling method had a value of 0.3 µg/L MC-LR for its detection limit and for the POCIS the detection limit was 1 ng/L, per day [260].
Studies on drinking water confirm Oasis HLB to be the most used sorbent for monitoring clothianidin at 300 µg/L concentration, imidaclopid at 500 µg/L concentration and thiamethoxam at 5 µg/L concentration with POCIS [261,262]. Oasis HLB and DOWEX sorbents were compared for the sampling of pesticides in drinking water. Recovery averages for Dowex Optipore L-493 was 90% (range between 66–127%), and for HLB was 91% (range between 66–135%) [263,264]. A modified POCIS sampler using two different sorbents, Chromabond HR-X and Oasis MAX were compared for monitoring acidic herbicides in drinking water. This comparison showed that Oasis MAX had >85% elute recovery, a higher sampling rate compared with HR-X which had <20% recovery [265]. Further studies reported the detection of pharmaceuticals, alkylphenols, hormones, UV filters [266] personal care products [263,264] endocrine disrupting and drugs of abuse [262] and prefluorinated compounds [267] in drinking water. Microcystin levels in drinking water reservoir during two vegetation seasons were determined, using 90 mg Oasis HLB. The results after 14 days deployment showed a concentration of 1–12 ng/L for the toxin, this detected concentration level was not a risk to human health [268].
Table 8. Application of POCIS to river water studies.
Table 8. Application of POCIS to river water studies.
POCIS ResinsAnalyteYear CountryEluteApplicationDeployment ConditionAdsorbent QuantityAnalyte %RecoveriesRef
Oasis HLBAzithromycin; Fluoxetine; Levothyroxine; Omeprazole 2004UKMeOHPharmaceuticalsRiver,
deployment
at 8 sites
200 mgRanging
between
86–100%
[241]
Oasis Tribasic admixtureAtrazine, Diazinon, Diuron,
17a-Ethynylestradiol, Isoproturon
2004UK MeOHPharmaceuticalsRiver,
deployment
at 8 sites
200 mgRanging between 88–99%[241]
Oasis HLB
PES, PE membrane
Estrone, 17-estradiol, 17-ethynylestradiol, bisphenol A, propranolol, sulfamethoxazole, meberverine, thioridazine, carbamazepine, tamoxifen, indomethacine, diclofenac and meclofenamic acid in sewage effluent and river water.2008UKMeOHPharmaceuticals, endocrine disrupting compounds, personal care productsRiver100 mgPropranolol, sulfamethoxazole, carbamazepine, indomethacine & diclofenac, varied between 3.0 & 45.6 ng L−1, <LOD & 17.6 ng L−1, 16.6 and 539 ng L−1, 0.4 & 7.2 ng L−1 & 2.4 & 65.2 ng L−1, respectively; applying POCIS, conc’s were between 2.8 & 40.5 ng L−1, <LOD &18.2 ng L−1, 26.3 & 427 ng L−1, 0.5 & 11.9 ng L−1 & 4.4 & 165 ng L−1,
respectively.
[242]
Oasis HLB, PES membraneDiuron, (1-(3,4-dichlorophenyl) urea (DCPU), 1-(3,4-dichlorophenyl)-3-methylurea (DCPMU)2010FranceMeOH,
&
75% MeOH/25%, EtOAc (v/v),
HerbicidesRiver500 mgConc’s of
diuron and its transformation
products in
microcosm
[269]
Oasis HLB, PES membrane20 pesticide analytes2010USAMeOHPesticides, polycyclic, aromatic hydrocarbonsGround water cave200 mgVary during different month[270]
Strata-X, PES membraneE1, E2, EE22010UKEthyl acetate, & MeOH & ultrapure waterEndocrine disrupting substancesRiver300 mgRanging
between
0.9–2.2 ng/L
[250]
Strata-X, PES membranePrometryn2011GermanyMeOHPrometrynRiver300 mg0.01–0.07 mg/L[251]
Oasis HLB, PES membraneDesethylatrazine, Deisopropylatrazine, Simazine, Desethylterbuthylazine, Atrazine, Metolachlor, Terbuthylazine2011FranceMeOH, 75% & MeOH/25% ethyl acetate (v/v)PesticideRiver
deployment
200 mgAfter 24 h terbuthylazine > atrazine > simazine > metolachlor. Daily conc. varies [10]
Oasis HLB, PES membraneRange of substances in different pH reported2011CanadaMeOHPharmaceuticals, personal care products, disrupting substancesRiver & Tap water (Water chamber in lab)200 mgRecoveries in different pH are reported[174]
Oasis HLB, PES membraneA wide range of pollutants Pharmaceuticals, Polycyclic aromatic hydrocarbons, Hormones, Pesticides, Phenols2011FranceDCM/MeOH (50:50 v/v),Pharmaceuticals, polycyclic aromatic hydrocarbons, hormones, phenolsRiver
deployment
200 mgRecovery
tables are
reported in the paper
[244]
Oasis HLB33 Pesticides 2011FranceMeOH & MeOH/ethyl acetate, 75/25 (v/v)PesticideRiver 60 mg,
150 mg,
500 mg
Dimetomorph: 14.8 ng/L, linuron: 5.1 ng/L, metazachlor: 11.3 ng/L, terbuthylazine: 4.8 ng/L[16]
Oasis HLB, PES membraneAlkylphenols, Phenolated polymer, Oestrogenic hormones, Antidepressants, Anti-inflammatory, b-Blockers, Hypolipidemic agent2012France-Alkylphenols, phenolated polymers, hormones, pharmaceuticalsRiver and wastewater treatment plant200 mgResults shows the diagnostic capacity of POCIS tools[271]
Oasis HLB, PES membraneSulfamethoxazole2012Czech
Republic
MeOH:water (9:1 v/v acidified with 0.1% TFA)Pharmaceuticals Sulphonamides in streamRiver200 mg20 up to 736 ng/L[243]
Oasis HLB, PES membranePerfluorinated alkylcarboxylates2012Australia0.1% (v/v) ammonia solution in MeOH, then MeOHPerfluorinated chemicalsHarbour200 mg0.1−12 ng/L[272]
Oasis HLB, PES membraneRange of pharmaceuticals2012FranceMeOH, MeOH/DCM mixture (50:50), & DCMPharmaceuticalsRiver200 mgRanging
between
51–97%
[245]
Oasis HLB
PES membrane
A range of 14 different pesticides2012France-PesticideRiver200 mgConcs.
discussed
[206]
Oasis HLB
PES membrane
Chloro, Propic a, Propic b, Hex, Phos2012USAMeOHPesticide Synthesized river water (Effect of flow velocity was assessed)200 mgLevels of
organic carbon (<0.1–5 mg/L)
[273]
Oasis HLB
PES membrane
21 pharmaceuticals, 6 alkylphenols and 27 hydrophilic pesticides and biocides2012FranceMeOH & MeOH/DCM (v/v: 50/50), & DCMPharmaceuticals, alkylphenols and pesticidesSurface water200 mgRanging
between
2.5–33 ng/L
[274]
Pharma-POCIS Oasis HLBPolar pesticides 2013FranceMeOHPesticideGround water Deployed in 15 m depth and drinking water450 mgPOCIS could be tested on groundwater sites which present temporal variations in concentrations for studying its integrative
capacity
[247]
Strata XAW, PES membraneRange of prefluorinated chemicals2013Australia0.1% (v/v) ammonia sol in MeOH & MeOHPerfluorinated chemicalsRiver600 mgRanging
between
71–92%
[254]
POCIS HLB
PES membrane
Terbuthylazine, diuron and linuron2014SwitzerlandMeOHHerbicidesRiver200 mgTerbuthylazine: 220 ng/L, diuron: 70 ng/L linuron:
50 ng/L
[275]
POCIS HLBRange of pesticides 2014FranceMeOHPesticidesRiver200 mgRanging between 138–1080 ng/L[276]
POCIS HLB, PES membrane23 polar pesticides and 8 metabolites2014FranceMeOH, & MeOH/ethyl acetate, 75:25 (v/v)PesticidesRiver200 mgDetails discussed in the paper[277]
POCIS HLB, PES membraneAtrazine2014CanadaMeOHPesticidesRiver 200 mgAtrazine conc. in 24 streams discussed[278]
POCIS HLB
PES membrane
Atrazine, propazine, terbutylazine, diclofenac, ibuprofen, ketoprofen, perfluorooctanoic acid and perfluorooctanesulfonate2014ItalyAcetonePerfluorinated chemicals, pharmaceuticals, pesticidesRiver (linear velocity of 2.0, 5.1, 10.2 and 15.3 cm/s)200 mgSpiked conc. in different flow velocities is shown[279]
POCIS HLB
PES membrane
Diuron2014FranceMeOH, & MeOH/DCM (v/v: 50/50), then DCMPesticidesCoastal water200 mgOysters were exposed to
diuron integrated conc’s as low as 0.2 and 0.3 g/L
[280]
POCIS HLB39 pesticides and metabolites2015FranceMeOH, & MeOH: EtOAc, 75:25 (v/v)Pesticides and metabolitesRiver200 mgFrequency of detection and concentration reported [281]
POCIS HLB12 veterinary antibiotics2015USAMeOHPharmaceuticalsRiver 200 mgConc’s ranging from 0.0003 ng/L to 68 ng/L[282]
POCIS HLB, PES membraneAPIs and pesticides2015Portugal MeOH & DCM/MeOH (50:50; v/v), & DCMPharmaceuticals, pesticidesRiver 200 mgCaffeine: 804 ± 209 ng/L, theophylline:184 ± 44 ng/L, Carbendazim: 45 ± 18 ng/L, atrazine, diuron, Isoproturon and simazine levels were below the Environmental Quality Stds[283]
POCIS-Pest and POCISPharm, PES membraneOrgan halogen herbicides, organophosphorous pesticides, carbamate, triazine, urea, pharmaceuticals, phenols, and industrial chemicals2016GreeceHexane/DCM Additionally, DCM/EtOAc (50:50 v/v)Pesticides, carbamate, triazine, urea, pharmaceuticals, phenols, and industrial chemicalsRiver200 mgMost compounds showed recoveries ranged from 60 to 110%. The coefficient of variation (CV) ranged from 0.84 to 23.8%. LOD and LOQ ranged from 6.4 to 40.1 ng/L and from 21.5 to 134 ng/L, respectively. [284]
POCIS sorbents, HLB and Strata X-CW, PES membraneBenzotriazole, methylbenzotriazole, atrazine, diuron, isoproturon, linuron, metolachlor, penconazole, terbuthylazine, carbamazepine, diclofenac, metformin, sulfamethoxazole2016SwitzerlandMeOHCorrosion inhibitors, pesticides, pharmaceuticalsRiver200 mgRs: Benzotriazole: 0.134, methylbenzotriazole: 0.148, atrazine: 0.26, diuron: 0.15, isoproturon: 0.254, metolachlor: 0.139, terbuthylazine: 0.197, carbamazepine: 0.231, diclofenac: 0.165, sulfamethoxazole: 0.103[253]
POCIS HLB
PES membrane
Metaldehyde, isoproturon, simazine, chlorotoluron, atrazine, epoxiconazole, chlorpyrifos, cypermethrin and permethrin2016UKEthyl acetatePesticidesRiver200 mgResults compared in three different sites[285]
POCIS HLBIn vitro (i.e., zf liver cell lines stably expressing zfERα, zfERβ1 and zfERβ2 subtypes) and in vivo (i.e., transgenic cyp19a1b-GFP zf embryos)2016FranceMeOH, & DCM/MeOH (50:50; v/v), & DCMEndocrine
disrupting
substances
River 200 mgResults in different sites discussed[202]
POCIS HLB
PES membrane
20 parent compounds (PCs) and 11 characteristic TPs in four 11 wastewater-impacted rivers2016Sweden-PharmaceuticalsRiver200 mgResults of four different rivers discussed[192]
POCIS HLB
PES membrane
CBZ: carbamazepine, CAF: caffeine, BPA: bisphenol A, LIN: lincomycin, SFA: sulfamethazine, SFO: Sulfamethoxazole, ATZ: atrazine, GEM: gemfibrozil2016Singapore -PharmaceuticalRiver200 mgSediment concentrations for carbamazepine (r = 0.79, p b 0.001), caffeine (r = 0.93, p b 0.001) but not BPA (p = 0.16)[286]
POCIS HLB
PES membrane
Complex mixtures of micropollutants, including emerging substances or transformation products2016FranceACN, MeOHRodenticide, hormones, antiparasitic, cardiovascular agent, pharmaceuticals, pesticides and their metabolitesGroundwater200 mgResults of two different sites discussed[287]
POCIS HLB
PES membrane
Tebuconazole, Propiconazole, Carbendazim, Azoxystrobin, Myclobutanil, Iprodione Fluconazole, Ketoconazole, Climbazole, Mecoprop, Agriculture, Dicamba, 2,4-D, Irgarol 1051, Terbutryn, Estrone Natural estrogen, Androstenedione, Ibuprofen, Acetaminophen, Naproxen, Trimethoprim, Sulfamethoxazole, Gemfibrozil, Carbamazepine, Sucralose2016CanadaMeOHPharmaceuticals, steroid hormones, the artificial sweetener, sucralose, fungicides, herbicides, biocidesRiver200 mgResults
compared in
different sites
[288]
Passive sampler copolymer of poly(divinylbenzene)-N-vinylpyrrolidone46 pesticides, 17 pharmaceuticals, 1 stimulant (caffeine) and 1 artificial sweetener (sucralose)2017FranceMeOH, & MeOH/EtOAc (50/50 v/v) and EtOAcPharmaceuticals Average flow of river over ten years 1.0 m3/sThe Marque River because of agricultural activities200 mgAtrazine 0.22 L/day, Cyprodinil 0.22 L/day, Desethylatrazine 0.09 L/day, Desisopropylatrazine 0.09 L/day, Diclofenac 0.08 L/day, Dimethenamid 0.20 L/day, Isoproturon 0.16 L/day, Metolachlor 0.17 L/day, Metalaxyl 0.19 L/day[252]
POCIS HLB, PES membrane13 parent pharmaceuticals and 8 of their transformation products (TPs)2017ChinaMeOHPharmaceuticals and their metabolitesRiver200 mgThe max concentration: 544.0 ng/L (CBZE), and the minimum value was 0.43 ng/L1 (SDZ)[289]
POCIS HLB, PES membrane45 pesticides2017FranceMeOHPesticidesSurface water in vitro200 mgAverage concentrations discussed in the paper[290]
POCIS HLBA range of pesticide, fungicide, herbicide, and insecticide 2017USAMeOHpesticidesSurface water200 mgA total of 141 compounds detected at one or more of the 97 sites sampled[291]
POCIS HLB, PES membrane37 pharmaceuticals and 3 human tracers2018FranceMeOH, & MeOH/EtO Ac, 75:25 v/vPesticides and their Metabolites,
pharmaceuticals
River200 mgFrequency and concentrations in the paper[292]
POCISOrganophosphate flame retardants
(OPFRs)
2018ChinaEthyl acetateEndocrine disrupting substancesRiver200 mgSix sampling locations ranged from 8.99 to 112.45 ng/L with an average concentration of 47.04 ng/L[293]
POCIS HLB, PES membrane17 pharmaceuticals, pesticides, per- and polyfluoroalkyl substances (PFASs)2018USA0.1% (v/v) ammonia solution in MeOH, & MeOHPesticides, pharmaceuticals, and perfluorinated chemicalsRiver200 mgConcentration shown during different month of the year [294]
POCIS HLB, PES membraneAtrazine, thiamethoxam, clothianidin, imidacloprid, 2,4-D and carbamazepine 2018CanadaMeOHPesticides and pharmaceuticalsRiver200 mgRecoveries compared during years[295]
POCIS Oasis HLB37 pharmaceuticals and 3 human traces2018France75:25 v:v MeOH:EthylPharmaceuticals flow rate at each sampling points was calculated proportionally at the size of sampling point watershed (75, 145 and 55 km2) agricultural rural headwater river200 mg23 compound out of 37 detected[238]
POCIS HLB, PES membrane & Mixed Polymer Sampler (MPS)Alachlor, atrazine, cybutryne, diclofenac, diuron, isoproturon, PCP, Simazine, terbutryne 2018GermanyMeOHPharmaceuticals, pesticidesRiver220 mgDissolved concentration of the compounds shown [249]
POCISS-metolachlor2018France-PesticidesRiver200 mgConcentration discussed in different sites[296]
POCIS HLB, PES membraneMalathion, diuron, carbofuran, carbendazim, trifluralin, imidacloprid, metolachlor, and acetamiprid2018BrazilMeOHPesticidesRiver220 mgMalathion 7.7%, diuron 5.1%, carbofuran 35.9%, Carbendazim 12.8%, trifluralin 5.1%, imidacloprid 5.1%, metolachlor 7.7%, acetamiprid 2.6%[297]
POCIS HLB, PES membraneAtrazine, 2,6-dichlorobenzamide, bentazone, chloridazon, isoproturon, and propiconazole2018SwedenEthyl
acetate
pesticidesSurface water220 mgHerbicides 36%–48%, fungicides 36%–21%, metabolites 11%–12%, insecticides 8%–10%, and other or mixed types 8%–10%[298]
POCIS HLB
32 selected herbicides, fungicides, and insecticides (mainly polar)2018GermanyACNpesticideSurface water230 mgDetails discussed in the paper[299]
POCIS HLB, PES membraneA range of pesticides2018JapanMeOHPesticidesRiver220 mgCompared POCIS with grab sampling[300]
Strata XAW & HLB
Nylon membrane
Acetaminophen, atrazine, diuron and norfloxacin hydrochloride, amitriptyline, irbesartan, ketoprofen and progesterone2018Spain2.5%
methanolic
ammonia & MeOH
Pesticide, pharmaceuticalsEstuarine100 mg
each
Feasibility of the simultaneous uptake of hydrophilic, acidic and basic compound[248]
Oasis HLB, PES membraneEstrone (E1), Nonylphenol (NP), 17b-estradiol (E2), ethynylestradiol (EE2),2019Germany-Endocrine disrupting
substances
River water54.5 mgNP 18 mg/L, E1 14 ng/L, E2 0.2 ng/L, EE2 0.5 ng/L[240]
Oasis HLB20 pesticides and 32 point source chemicals2019SpainMeOHPharmaceuticals, pesticides, hormonesRiver 200 mgRecovery % is reported in the paper[301]
POCIS20 pesticides, and 32 point source chemicals, mainly pharmaceuticals2019Spain_Pharmaceuticals, pesticides, hormonesRiver -High recoveries reported[302]
Oasis HLB, PES membraneBiomarkers of estrogenic endocrine disruption in smallmouth bass2019USADCM/MTBE 80:20 (v/v),Endocrine disrupting substancesRiver 200 mgRanging between 28–92%[303]
Oasis HLB, PES membranePharmaceuticals, endocrine disrupting substances, pesticides2019IrelandMeOHPharmaceuticals, endocrine disrupting substances, pesticidesRiver 230 mgConc’s in different years reported[246]
Oasis HLB, PES membrane168 targeted compounds2019SlovakiaMeOH/DCM, (1:1 v/v)Pesticides, pharmaceuticals, hormones, polycyclic aromatic hydrocarbons, polychlorinated biphenylsRiver200 mgRisk assessment of the detected compounds
revealed
[304]
Oasis HLB
PES membrane
A range of pesticide, herbicide, fungicide, metabolite, and insecticide 2019FranceMeOH, & MeOH/ethyl acetate, 75:25 v/vpesticidesRiver200 mg Bags deployed based on the depth >100 m or <100 mResults discussed in the paper[305]
Oasis HLB
Six different membrane
25 pharmaceuticals and personal care products2022USAFormic acid: MeOH
88:12%
Pharmaceuticals and personal care productsVernal pools Results compared to grab sampling in the paper[306]
Abbreviations: MeOH (Methanol), EtOAc (Ethyl acetate), Hx (Hexane), Rs = Sampling Rate.
Table 9. Application of POCIS on Lake water.
Table 9. Application of POCIS on Lake water.
POCIS ResinsAnalyteYear of StudyCountryEluteApplicationDeployment ConditionAdsorbent
Quantity
Analyte
%Recoveries
Ref
Oasis HLB
PES memberane
Range of substances reported2010USAMeOHEndocrine disruptingLake200 mg Recoveries
reported
[258]
Oasis HLB
PES membrane
Ibuprofen, Gemfibrozil, Caffeine, Carbamazepine, Trimethoprim, Venlafaxine, desmethyl-venl, Citalopram, Galaxolide, Tonalide, Triclosan2012CanadaMeOHPharmaceuticals,
antidepressants, personal care products
Lake 200 mgRecoveries discussed in the paper in different
seasons
[307]
POCIS HLB
PES membrane
35 APIs and endocrine disruption2014Singapore -PharmaceuticalsTropical Lake flow ~3–5 cm/S60 mgAtorvastatin and norfluoxetine,
from 52 to 196% (109 ± 32%), the 80–120% range for 26 of the compounds
[259]
POCIS HLB
PES membrane
Pestisides, herbicides, fungicides and pharmaceuticals2016CanadaMeOHFungicides, herbicides,
pharmaceuticals
Lake200 mgAtrazine was detected at all sites, and diuron, 2,4-D, and mecoprop were frequently detected. The fungicides carbendazim and thiophanate-methyl were detected at all sites, & a hydroxy-metabolite of the fungicide chlorothalonil was also widely
detected
[257]
POCIS HLB
PES membrane
25 pesticides 2018Burkina FasoMeOH, & MeOH/EtOAc(1:1, v/v), & EtOAc/Hx (1:4, v/v)pesticidesLake200 mgAtrazine, azadirachtin, carbofuran, chlorpyrifos, cypermethrin, dieldrin, imidacloprid, & profenofos exceeded 0.1 μg/L[256]
Oasis HLB
PES membrane
microcystin-LR (MC-LR)2019CanadaMeOHMicrocystin-LRLake220 mgThe Rs 0.045 (±0.001) and 0.041 (±0.001) L per day for initial concentrations of 0.5 and 1.0 mg/L[260]
Oasis HLB
PES membrane
neonicotinoid insecticides NNIs, thiamethoxam, clothianidin and imidacloprid, Atrazine, 2,4-D, dicamba, carbendazim, thiophanate methyl and several azoles2019CanadaMeOHPesticidesLake 200 mgNNIs, thiamethoxam, clothianidin and imidacloprid 0.23 μg/L, Atrazine, 2,4-D, dicamba, carbendazim, thiophanate methyl and several azole-based fungicides were also widely detected[255]
Oasis HLBPesticides2019TunisiaACNPesticidesLagoon watershed The results in different sites are reported in the paper [308]
Table 10. Application of POCIS on drinking water.
Table 10. Application of POCIS on drinking water.
POCIS ResinsAnalyteYear CountryEluteApplicationDeployment ConditionAdsorbent QuantityAnalyte %RecoveriesRef
Oasis MAX, HRX, HLB, PES membraneAcidic herbicides 2012FranceMeOH, & MeOH/EtOAc 5:5 (v/v)PesticideDrinking
water
-POCIS-MAX showed no influence of nitrates. MAX sorbent >82% recoveries[265]
Oasis HLB, PES membraneEight alkylphenols, nine hormones, 11 pesticides, 27 pharmaceuticals and one UV filter2013FranceMeOH, & MeOH: DCM 50:50Alkylphenols, hormones, pesticides, pharmaceutical, UV filtersTap water
(using external thermostat tank)
200 mgDetails
discussed in the paper
[266]
POCIS HLBCarbamazepine, trimethoprim, sulfamethoxazole, ibuprofen, gemfibrozil, estrone and sucralose2014CanadaMeOHPharmaceuticalsDrinking
water
220 mgAfter 10 days: CBZ 894.7 ± 37.2 ng/L, Ibuprofen: 262.4 ± 80.1 ng/L, Gemfibrozil: 144.5 ± 31.3 ng/L, TPM: 8.7 ± 2.2 ng/L, Sucralose: 186.1± 15.0 ng/L[309]
POCIS HLB, PES membraneAtrazine-d5, caffeine-13C3, cotinine-d3, DIA-d5, fluoranthene-d10, lindane2016USAAcetone, DCMPesticides, polycyclic, aromatic hydrocarbons, personal care productsTap water 200 mg4.51 ± 0.34 g/g (atrazine-d5), 4.62 ± 0.30 g/g (caffeine-13C3), 4.01 ± 0.08 g/g (cotinine-d3), 3.87 ± 0.24 g/g (DIA-d5), 4.42 ± 0.16 g/g (fluoranthene-d10), and 4.65 ± 0.14 g/g (lindane).[264]
POCIS HLB, Additionally, DOWEX, PES membrane34 pesticide, personal care products and hydrocarbons 2016USAAcetone, & DCM,Pesticides, polycyclic aromatic hydrocarbons, personal care productsTap water200 mgRecoveries average: Dowex Optipore L-493: 90% (range: 66–127%), HLB: 91% (range: 66–135%), and Osorb media: 96% (range:63–127%) [263]
POCIS HLB, PES membrane73 compounds2016NorwayMeOHPharmaceuticals, endocrine disrupting substances, pesticides, herbicides, drugs of abuseDrinking
water
200 mgResults for prediction model discussed[262]
POCIS HLB, PES membraneDiclofenac (DIC), ketoprofen (KET), mefenamic acid (MEF), naproxen (NAP), ibuprofen (IBU), ketoprofen-d3 (KET-d3), perfluorooctanoic acid (PFOA), perfluorooctanesulfonate (PFOS) and Caffeine (CAF)2018ItalyAcetonePharmaceuticals, perfluorinated compounds, caffeineDrinking
water
in vitro
200 mgCaffeine: 0.07–0.93 ng/L, perfluorinated compounds: 2.93–13.42 ng/L[267]
POCIS HLB, PES membraneImidacloprid, clothianidin, thiamethoxam, acetamiprid, thiacloprid, a hydroxy metabolite2018CanadaMeOH:
Acetone
60:40 v/v
PesticidesDrinking
water
220 mgClothianidin 300 µg/L imidaclopid 500 µg/L thiamethoxam 5 µg/L[261]
Oasis HLBMicrocystins risk assessment 2019Czech Republic-MicrocystinsDrinking water reservoir Depth: 13, 28, 46 m, flow velocities ranging between 0.01 and 0.15 m/s90 mg20–200 pg/L after 14-d deployment and 1–12 ng/L[268]
POCIS Oasis WAX, PES membrane26 per- and polyfluoroalkyl substances (PFASs)2019SwedenMeOHPer, polyfluoroalkyl substances (PFASs)Drinking water in treatment plant200 mg64–89%[310]

4.4. Application of POCIS, In Vitro Laboratory Studies

The efficiency of POCIS has been assessed in vitro under laboratory simulated conditions for the evaluation of a range of pharmaceuticals, pesticides and beta-blockers using different amount of sorbent (Table 11). Oasis HLB (200 mg) of sorbent was applied to moderate polarity pesticides and their metabolites such as propiconazole (log Kow = 3.72) and tebuconazole (log Kow = 3.7) and showed a low efficiency of POCIS in adsorbing these compounds, which were initially at low concentrations in the aquatic environment of study [311]. Additionally, ten pharmaceutical compounds ranging between log KOW 0.16 to 4.51 (0.16_log KOW_4.51) under laboratory conditions were studied to evaluate the effect of flow velocity on the accuracy of passive sampling [312]. Using the POCIS with PRC (Performance Reference Compounds) approach and o-DGT, these absorption systems were applied under different water velocities (2 < V < 18 cm/s), to evaluate the sampling of atenolol, carbamazepine, diclofenac, fluoxetine, ketoprofen, metoprolol, paroxetine, propranolol, sulfamethoxazole, and trimethoprim. Both samplers were able to limit the flow effects ensuring the accuracy of POCIS performance (within 20% uncertainty). In addition, even though o-DGT was more efficient in quiescent (dormant) situations it has less sensitivity compared to POCIS, at adsorbing these analytes [312].
In the in vitro laboratory scale studies, unlike the other applications for POCIS, the most dominant studies have reported higher amounts of HLB sorbent efficiency (>200 mg). Ibrahim et al. have used 220 mg of Oasis HLB resin for a range of 17 polar pesticides (1.15 ≤ log Kow ≤ 3.71) during a 15-day study where the sampling rate ranged between 67.9 to 279 mL/day increased with increasing hydrophobicity of the pesticides [313]. In addition, Fauvelle et al., presented a study on the comparison of the capacity of different amounts of Oasis HLB (200 and 600 mg) in a pharma-POCIS device on the uptake of acidic (2,4-dichlorophenoxyacetic acid, (ESA), acetochlor oxanilic acid, bentazon, dicamba, mesotrione, and metsulfuron) and polar herbicides such as (atrazine, diuron, and desisopropylatrazine). According to these studies 200 mg Oasis HLB is not efficient in absorbing compounds with high acidity and polarity because of their fast thermodynamic equilibrium with the HLB sorbent. Separately, 600 mg of sorbent showed a sampling rate two times higher compared with 200 mg over 35 days [239].
The effects of water temperature on the accumulation of 48 pesticides, insecticides and fungicides on POCIS using 220 mg of Oasis HLB sealed between hydrophilic microporous polyethersulfone (PES) membranes for 28 days was investigated [314]. The results demonstrated that increasing the water temperature led to an increasing sampling rate, reported at 18, 24, and 30 °C ranged from 0.00676 to 0.262, 0.00603 to 0.312, and 0.00426 to 0.603, respectively [314]. Togola et al. evaluated the application of POCIS for pharmaceutical monitoring under environmental conditions such as salinity, temperature, and pollutants. Applying three conditions for Rs (0 PSU/21 °C, 35 PSU/21 °C and 0 PSU/15 °C), showed average recoveries for caffeine of 1622 ng/L, amitriptyline 355 ng/L, doxepin 253 ng/L, imipramine 377 ng/L, carbamazepine 226 ng/L, diazepam 435 ng/L, nordiazepam 629 ng/L and ibuprofen 1128 ng/L. This study showed the applicability of POCIS in the detection of trace concentration of compounds below the detection limit of discrete (non-continuous) sampling approaches [315]. Five different types of POCIS-SR, POCIS-A, POCIS-B, chemcatcher PRS, chemcatcher C18 were tested for the determination of 124 different pesticides in water under simulated laboratory ambient (20 °C) temperature and 10 cm/s turbulence conditions. POCIS-SR showed a better capability in the up-take of hydrophobic compounds (log Kow > 5.3) whereas the other devices (POCIS-A, POCIS-B, chemcatcher PRS, chemcatcher C18) showed better results for hydrophilic compounds (log Kow < 0.7) [316].
Table 11. Application of POCIS, in vitro laboratory scale.
Table 11. Application of POCIS, in vitro laboratory scale.
POCIS ResinsAnalyteYearCountry EluteApplication Deployment ConditionAdsorbent
Quantity
Analyte
%Recoveries
Ref
Oasis HLB, PES membraneCaffein, Amitripthiline, Doxepine, Imipramine, Carbamazapine, Diazepam, Nordizepam, Ibuprofen, Gemfibrozile, Naproxine, Diclofenac, Ketoprofen2007France EtOAc/Ace 50/50 v/vPharmaceuticals Laboratory
simulation
200 mgAverage recoveries: Caf: 1622 ng/L, Ami: 355 ng/L, Dox: 253 ng/L, Imi: 377 ng/L, Cbz: 226 ng/L, Dzp: 435 ng/L, Ndzp: 629 ng/L, Ibu: 1128 ng/L, Gem: 1744 ng/L, Nap: 673 ng/L, Diclo: 606 ng/L, Keto: 388 ng/L[315]
Oasis HLB, PES membraneRange of substances in different pH reported2011Canada MeOHPharmaceuticals, personal care products, disrupting substances Laboratory scale River and Tap
water (water chamber in lab)
200 mgRecoveries in different pH are reported[174]
Oasis HLB, PES membraneAtrazine, simazine, desethylatrazine (DEA), desisopropylatrazine (DIA), desethylterbuthylazine (DET), terbuthylatrazine, diuron, isoproturon, chlortoluron, linuron, propyzamide, alachlor, metolachlor, acetochlor, metalaxyl, penconazole, and azoxystrobine2013France ACNpesticide Laboratory
calibration
experiment
230 mgSampling rate: 67.9–279 mg/L[313]
Pharma-POCIS Oasis HLB PES membranePolar pesticides and metabolites2013France ACNpesticide Laboratory in situ sampling200 mg169 to 479 mL/g day[311]
POCIS HLB, Nylon membraneA wide range of pharmaceuticals and pesticides2014France MeOH, DCM/MeOH (50:50; v/v), and DCMPesticides, pharmaceuticals Laboratory water samples200 mgResults discussed in the paper [317]
Strata XAW, PES membranePerfluorohexanoate (PFHxA), perfluoroheptanoate (PFHpA), perfluorooctanoate (PFOA), perfluorononanoate (PFNA), perfluorodecanoate (PFDA), perfluoroundecanoate (PFUnDA)2014Australia 0.1% (v/v) ammonia MeOH, & MeOHPerfluorinated chemicals (PFCs) Laboratory water sample600 mgPFPeA 0.078 ± 0.02 L/d, PFHxA 0.118 ± 0.01 L/d, PFNA 0.165 ± 0.004 L/d, PFHxS 0.182 ± 0.01 L/d[214]
POCIS HLB, PES membrane2,4-dichlorophenoxyacetic acid (2,4-D), acetochlor ethanesulfonic acid (ESA), acetochlor oxanilic acid, bentazon, dicamba, esotrione, and metsulfuron, atrazine, diuron, esisopropylatrazine herbicides 2014France MeOH & MeOH/EtOAc 50: 50 (v/v)Herbicides Laboratory water samples200, 600 mgIncreasing sorbent to 600 mg resulted in sampling rates (Rss) twice as high as those observed with 200 mg[239]
POCIS (5 different types) SR, POCIS-A, POCIS-B, Chemcatcher RPS, Chemcatcher C18124 pesticides 2015Sweden MeOH, & DCM/MeOH (8/2, v/v)Pesticides Laboratory condition220 mgResults for different POCIS devices
discussed
[316]
Pharma-POCIS HLB
PES membrane
20 pesticides, insecticides, herbicides2016Japan EtOHPesticides Laboratory
pesticide sample water
220 mgSampling rate increased at 18 °C from 0.00676 to 0.262, 24 °C 0.00603 to 0.312, 30 °C 0.00426 to 0.603.[314]
Carbon nanotubes, PES membraneCarbamazepine, diclofenac, β-estradiol, p-nitrophenol, 3,5-dichlorphenol, sulfapyridine, sulfamethoxazole2017Poland ACN/MeOH/DCM, (40:40:20; v/v),Pharmaceuticals, phenols Laboratory water sample100 mgSulfapyridine: 79.8 ± 0.2%, Sulfamethoxazole: 41.5 ± 0.1% Carbamazepine: 96.6 ± 1.5% p-nitrophenol: 70.5 ± 0.1% 17-β-estradiol: 77.1 ± 0.5% 3,5-dichlorophenol: 103.1 ± 1.8% diclofenac: 76.3 ± 1.4% [318]
Oasis HLB
PES membrane
Atenolol, cabamazapine, Diclofenac, Fluoxetine, Ketoprofen, Metoprolol, Paroxetine, Propaonalol, Sulfamethaxazole, Trimethoprime2019France MeOH, & MeOH/ EtOAc, 75:25 v/vPharmaceuticals Ultrapure water200 mgEffect of flow velocities is assessed (2 < V < 18 cm/s)[312]
POCIS Oasis HLB44 pharmaceuticals2020France MeOH & 75:25 (v/v) MeOH: EtOAcPharmaceuticals
pump (flow rate =
13 m3/h)
laboratory-scale artificial
river
200 mgEconazole, fenbendazole, fenofibrate, metformin, thioridazine, and triclabendazole) were not sampled by POCIS and 12 compounds are not available[319]
POCIS HLB
PES membrane
neonicotinoid pesticides2020Japan MeOH: ACE
2:1
neonicotinoid pesticides Laboratory scale20 mgSuitable for neonicotinoid detection in lower concentration[320]
Abbreviations: EtOAc (Ethyl acetate), Ace (Acetone), MeOH (Methanol), CAN (Acetonitrile), DCM (Dichloromethane), Ammonia MeOH (Methanolic ammonia), EtOH (Ethanol).

5. Conclusions

In this review, scientific literature associated with two in situ methods, Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS), for the collection and concentration of biotoxins and pharmaceuticals in environmental waters, has been investigated. The application of Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) to pre-concentrate a range of marine toxins, pesticides and pharmaceutical compounds that occur at low levels in marine and environmental waters has been critically discussed and summarised in tabular format
A variety of adsorption substrates in SPATT and different sorbents in POCIS were reviewed. Laboratory and field studies demonstrated the efficacy and ability of SPATT technology as reliable in situ methods to absorb a range of lipophilic and hydrophilic marine biotoxins, pharmaceuticals, pesticides, antibiotics and microcystins in marine water, freshwater and wastewater ecosystems.
Furthermore, analytical methods such as TLC, LC-MS and LC-MS/MS that had been used to detect the biotoxins of different toxin classes OA/DTXs, PTXs, YTXs, AZAs DA, have been addressed.
The review shows that previous studies primarily focused on adsorption and desorption efficiencies; this leaves gaps in the knowledge regarding quantitative sampling and isotherm characterization in relation to specific biotoxins. Studies showed that although HP20 is successful in the adsorption of a range of lipophilic and hydrophilic toxins, it has relatively slow uptake, however the efficiency of HP20 in adsorption of toxins is highest compared to other aromatic sorbents due to its large pore size. These studies emphasise the importance of knowledge gathering and experimentation to determine the duration that HP20 remains in the integrative phase.
The maximal capacity of HP20 is relatively higher than SP700 that means pore size plays an important rule is adsorption efficiency along with the polarity of toxins. The average recoveries of lipophilic and some hydrophilic toxins with HP20 is around 90% and SP700 shows 69–72%. Average recoveries of lipophilic toxins for HP20, Strata-X and Oasis HLB after 24 h shows 70, 50 and 40%. POCIS using oasis HLB shows an average recovery between 80–120% recoveries for APIs and endocrine disruption. Finally, this review examines the marine toxin area categorising the main toxin groups, according to toxicity and aetiology, and historically reviews the application of SPATT as an early warming strategy for marine toxin surveillance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27227898/s1: Marin Biotoxins Figure S1: Marine biotoxin structures. (a) okadaic acid, (b) yessotoxin, (c) azaspiracid, (d) saxitoxin, (e) domoic acid, (f) Brevetoxin (PbTX-1) Type-A, (g) PbTX-2 (Type-B), (h) PbTX-3(Type-B), (i) Ciguatoxin-1 (CTX), (j) Ciguatoxin-2, (k) Ciguatoxin-3, (l) Ciguatoxin-4A, (m) Ciguatoxin-4B, (n) Pectenotoxin-1 (PTX-1), (o) Pectenotoxin-2 (PTX-2), (p) Dinophysistoxin 1 (DTX-1), (q) Dinophysistoxin 2 (DTX-2), (r) Dinophysistoxin 3 (DTX-3). Figure S2: (a) SPATT bags and discs with various resins contained within 80 mm polyester mesh; (b) SPATT bags being deployed. However, after this initial design other studies applied the same bags in a different manner to support the SPATT bags in the water flow. Diagram [160]; (c) showing SPATT bags among holding tubes. (A) 100 mm nylon mesh, (B) resin, (C) inner holding ring, (D) outer holding ring, (F) 75 mm diameter embroidery ring and (E) final assembled sampling disk [181]. SPATT bags attached to aluminum alloy [162]. Figure S3: (a) Chemical Structure DIAION HP-20 (Aromatic synthetic adsorbent ion-exchange resin) Styrene-divinylbenzene [164]; (b) SEPABEADS SP700 (Aromatic synthetic adsorbent ion-exchange resin) [167] (c) SEPABEAD SP207 (Modified Aromatic synthetic adsorbent ion-exchange resin) Brominated styrene—divinylbenzene [166] (d) DIAION HP2MG (Methacrylic synthetic adsorbent ion-exchange resin) Polymethacrylate [166]. Figure S4: Schematic diagram of pumping system [161]. Figure S5: Organic Chemical Integrative Sampler (POCIS) device [321]. (a) POCIS or Aquasense-P disk; (b) Polar Organic Chemical Integrative Sampler (POCIS) carrier; (c) carrier on which one to three POCIS can be mounted. (d) schematic diagram of extraction of analyte from POCIS device.

Author Contributions

N.K.: draft preparation, writing the pharmaceutical, Marine Toxins and SPATT sections and editing the review, preparing databases and resources, preparing tables and images, drawing diagrams and chemical structures. F.A.: draft preparation, writing the SPATT section of the review and preparing figures. M.L.: second academic supervisor on the design, text correction and rewriting, editing and reviewing of the manuscript. M.G.: reviewing the literature on phase adsorption and editing of the review. A.F.: Article originator and principal supervisor, preparing the review structure and supervising the manuscript literature searches, drafting, correcting, reviewing, section rewriting, figures/images/tables planning/drafting and preparation and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This review was funded as a deliverable for the STREAM project, which is part-funded by the European Regional Development Fund (ERDF) through the Ireland-Wales Cooperation programme 2014–2020 (Grant #C80943).

Acknowledgments

The authors also acknowledge the funding of a High-Resolution Sector Field ICP-MS (ICP-SFMS) under the Science Foundation Ireland (SFI) Research Infrastructure Call 2012 [Proposal #12/RI/2335 (2)] and the funding from the Higher Education Authority of Ireland—The Programme for Research in Third Level Institutions (HEA-PRTLI 2 & 4) that generated research expertise that led to the funding for the STREAM project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vrana, B.; Smedes, F.; Hilscherová, K. Passive sampling of waterborne contaminants. In In Situ Bioavailability and Toxicity of Organic Chemicals in Aquatic Systems; Springer: New York, NY, USA, 2020. [Google Scholar]
  2. Zendong, Z.; Herrenknecht, C.; Abadie, E.; Brissard, C.; Tixier, C.; Mondeguer, F.; Séchet, V.; Amzil, Z.; Hess, P. Extended evaluation of polymeric and lipophilic sorbents for passive sampling of marine toxins. Toxicon 2014, 91, 57–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. The Occurrence of Marine Biotoxins and Risk of Exposure to Seafood Consumers in Ireland; Food Safety Authority of Ireland: Dublin, Ireland, 2016; p. 64.
  4. Fernandez, R.; Maman, L.; Jaen, D.; Fernandez Fuentes, L.; Ocana, M.A.; Gordillo, M.M. Dinophysis Species and Diarrhetic Shellfish Toxins: 20 Years of Monitoring Program in Andalusia, South of Spain. Toxins 2019, 11, 189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Fux, E.; Gonzalez-Gil, S.; Lunven, M.; Gentien, P.; Hess, P. Production of diarrhetic shellfish poisoning toxins and pectenotoxins at depths within and below the euphotic zone. Toxicon 2010, 56, 1487–1496. [Google Scholar] [CrossRef] [PubMed]
  6. McCarthy, M.; Bane, V.; García-Altares, M.; van Pelt, F.N.; Furey, A.; O’Halloran, J. Assessment of emerging biotoxins (pinnatoxin G and spirolides) at Europe’s first marine reserve: Lough Hyne. Toxicon 2015, 108, 202–209. [Google Scholar] [CrossRef] [PubMed]
  7. Górecki, T.; Namieśnik, J. Passive sampling. TrAC Trends Anal. Chem. 2002, 21, 276–291. [Google Scholar] [CrossRef]
  8. Salim, F.; Gorecki, T. Theory and modelling approaches to passive sampling. Environ. Sci. Process. Impacts 2019, 21, 1618–1641. [Google Scholar] [CrossRef]
  9. Palmes, E.; Gunnison, A.F. Personal monitoring device for gaseous contaminants. Am. Ind. Hyg. Assoc. J. 1973, 34, 78–81. [Google Scholar] [CrossRef]
  10. Pesce, S.; Morin, S.; Lissalde, S.; Montuelle, B.; Mazzella, N. Combining polar organic chemical integrative samplers (POCIS) with toxicity testing to evaluate pesticide mixture effects on natural phototrophic biofilms. Environ. Pollut. 2011, 159, 735–741. [Google Scholar] [CrossRef]
  11. Grodtke, M.; Paschke, A.; Harzdorf, J.; Krauss, M.; Schuurmann, G. Calibration and field application of the Atlantic HLB Disk containing Chemcatcher(R) passive sampler—Quantitative monitoring of herbicides, other pesticides, and transformation products in German streams. J. Hazard. Mater. 2021, 410, 124538. [Google Scholar] [CrossRef]
  12. Network of Reference Laboratories, Research Centres and Related Organisations for Monitoring of Emerging Environmental Substances. Available online: https://www.norman-network.net/ (accessed on 27 June 2022).
  13. Miège, C.; Mazzella, N.; Allan, I.; Dulio, V.; Smedes, F.; Tixier, C.; Vermeirssen, E.; Brant, J.; O’Toole, S.; Budzinski, H.; et al. Position paper on passive sampling techniques for the monitoring of contaminants in the aquatic environment–achievements to date and perspectives. Trends Environ. Anal. Chem. 2015, 8, 20–26. [Google Scholar] [CrossRef]
  14. Vrana, B.; Allan, I.J.; Greenwood, R.; Mills, G.A.; Dominiak, E.; Svensson, K.; Knutsson, J.; Morrison, G. Passive sampling techniques for monitoring pollutants in water. TrAC Trends Anal. Chem. 2005, 24, 845–868. [Google Scholar] [CrossRef]
  15. Gong, X.; Li, K.; Wu, C.; Wang, L.; Sun, H. Passive sampling for monitoring polar organic pollutants in water by three typical samplers. Trends Environ. Anal. Chem. 2018, 17, 23–33. [Google Scholar] [CrossRef]
  16. Lissalde, S.; Mazzella, N.; Fauvelle, V.; Delmas, F.; Mazellier, P.; Legube, B. Liquid chromatography coupled with tandem mass spectrometry method for thirty-three pesticides in natural water and comparison of performance between classical solid phase extraction and passive sampling approaches. J. Chromatogr. A 2011, 1218, 1492–1502. [Google Scholar] [CrossRef] [PubMed]
  17. DGT, Research. DGT Research for Measurements in Waters, Soils and Sediments; DGT, Research: Lancaster, UK; Available online: https://www.dgtresearch.com/what-dgt-does/ (accessed on 27 June 2022).
  18. Howard, M.D.; Kudela, R.; Caron, D.; Smith, J.; Hayashi, K. Standard Operating Procedure for Solid Phase Adsorption Toxin Testing (SPATT) Assemblage and Extraction of HAB Toxins; University of California and University of Southern California: Santa Cruz, CA, USA, 2018. [Google Scholar]
  19. Roué, M.; Darius, H.T.; Chinain, M. Solid phase adsorption toxin tracking (SPATT) technology for the monitoring of aquatic toxins: A review. Toxins 2018, 10, 167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Munday, R.; Reeve, J. Risk assessment of shellfish toxins. Toxins 2013, 5, 2109–2137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Salas, R.; Clarke, D. Review of DSP Toxicity in Ireland: Long-Term Trend Impacts, Biodiversity and Toxin Profiles from a Monitoring Perspective. Toxins 2019, 11, 61. [Google Scholar] [CrossRef] [Green Version]
  22. Young, N.; Robin, C.; Kwiatkowska, R.; Beck, C.; Mellon, D.; Edwards, P.; Turner, J.; Nicholls, P.; Fearby, G.; Lewis, D.; et al. Outbreak of diarrhetic shellfish poisoning associated with consumption of mussels, United Kingdom, May to June 2019. Eurosurveillance 2019, 24, 1900513. [Google Scholar] [CrossRef] [Green Version]
  23. Stoecker, D.K. Mixotrophy among Dinoflagellates 1. J. Eukaryot. Microbiol. 1999, 46, 397–401. [Google Scholar] [CrossRef]
  24. Ulrich Lüttge, F.M.C.; Risueño, M.-C.; Leuschner, C. Progress in Botany; Springer: Berlin, Germany, 2021; Volume 82. [Google Scholar]
  25. Azad, H.S.; Borchardt, J.A. Variations in phosphorus uptake by algae. Environ. Sci. Technol. 1970, 4, 737–743. [Google Scholar] [CrossRef]
  26. Lee, T.C.; Fong, F.L.; Ho, K.C.; Lee, F.W. The Mechanism of Diarrhetic Shellfish Poisoning Toxin Production in Prorocentrum spp.: Physiological and Molecular Perspectives. Toxins 2016, 8, 272. [Google Scholar] [CrossRef]
  27. Chain, E. Scientific Opinion on marine biotoxins in shellfish—Palytoxin group. EFSA J. 2009, 7, 1393. [Google Scholar]
  28. Council, E. Regulation (EC) No 853/2004 of the European parliament and of the council of 29 April 2004 laying down specific hygiene rules for food of animal origin. J. Eur. Union 2004, 139, 55–205. [Google Scholar]
  29. Manita, D.; Alves, R.N.; Braga, A.C.; Fogaca, F.H.; Marques, A.; Costa, P.R. In vitro bioaccessibility of the marine biotoxins okadaic acid, dinophysistoxin-2 and their 7-O-acyl fatty acid ester derivatives in raw and steamed shellfish. Food Chem. Toxicol. 2017, 101, 121–127. [Google Scholar] [CrossRef] [PubMed]
  30. Fu, L.L.; Zhao, X.Y.; Ji, L.D.; Xu, J. Okadaic acid (OA): Toxicity, detection and detoxification. Toxicon 2019, 160, 1–7. [Google Scholar] [CrossRef]
  31. Solter, P.F.; Beasley, V.R. Phycotoxins. In Haschek and Rousseaux’s Handbook of Toxicologic Pathology; Elsevier: Amsterdam, The Netherlands, 2013; pp. 1155–1186. [Google Scholar]
  32. Guo, F.; An, T.; Rein, K.S. The algal hepatoxoxin okadaic acid is a substrate for human cytochromes CYP3A4 and CYP3A5. Toxicon 2010, 55, 325–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liu, L.; Guo, F.; Crain, S.; Quilliam, M.A.; Wang, X.; Rein, K.S. The structures of three metabolites of the algal hepatotoxin okadaic acid produced by oxidation with human cytochrome P450. Bioorganic Med. Chem. 2012, 20, 3742–3745. [Google Scholar] [CrossRef] [Green Version]
  34. Valdiglesias, V.; Prego-Faraldo, M.V.; Pasaro, E.; Mendez, J.; Laffon, B. Okadaic acid: More than a diarrheic toxin. Mar. Drugs 2013, 11, 4328–4349. [Google Scholar] [CrossRef] [Green Version]
  35. Suganuma, M.; Fujiki, H.; Suguri, H.; Yoshizawa, S.; Hirota, M.; Nakayasu, M.; Ojika, M.; Wakamatsu, K.; Yamada, K.; Sugimura, T. Okadaic acid: An additional non-phorbol-12-tetradecanoate-13-acetate-type tumor promoter. Proc. Natl. Acad. Sci. USA 1988, 85, 1768–1771. [Google Scholar] [CrossRef] [Green Version]
  36. Kamat, P.K.; Rai, S.; Nath, C. Okadaic acid induced neurotoxicity: An emerging tool to study Alzheimer’s disease pathology. Neurotoxicology 2013, 37, 163–172. [Google Scholar] [CrossRef]
  37. Jiao, Y.-H.; Dou, M.; Wang, G.; Li, H.-Y.; Liu, J.-S.; Yang, X.; Yang, W.-D. Exposure of okadaic acid alters the angiogenesis in developing chick embryos. Toxicon 2017, 133, 74–81. [Google Scholar] [CrossRef]
  38. Jayaraj, R.; Gupta, N.; Rao, P.L. Multiple signal transduction pathways in okadaic acid induced apoptosis in HeLa cells. Toxicology 2009, 256, 118–127. [Google Scholar] [CrossRef] [PubMed]
  39. Tubaro, A.; Florio, C.; Luxich, E.; Vertua, R.; Yasumoto, T. Suitability of the MTT-based cytotoxicity assay to detect okadaic acid contamination of mussels. Toxicon 1996, 34, 965–974. [Google Scholar] [CrossRef]
  40. Chen, L. Okadaic acid induces apoptosis through the PKR, NF-κB and caspase pathway in human osteoblastic osteosarcoma MG63 cells. Toxicol. Vitr. 2011, 25, 1796–1802. [Google Scholar] [CrossRef] [PubMed]
  41. Cruz, P.G.; Norte, M.; Creus, A.H.; Fernandez, J.J.; Daranas, A.H. Self-association of okadaic acid: Structural and pharmacological significance. Mar. Drugs 2013, 11, 1866–1877. [Google Scholar] [CrossRef] [Green Version]
  42. Rodrigues, S.M.; Vale, P.; Chaveca, T.; Laires, A.; Rueff, J.; Oliveira, N.G. Naturally contaminated shellfish samples: Quantification of diarrhetic shellfish poisoning toxins in unhydrolysed and hydrolysed extracts and cytotoxicity assessment. J. Appl. Toxicol. 2010, 30, 699–707. [Google Scholar] [CrossRef]
  43. Huynh-Delerme, C.; Fessard, V.; Kiefer-Biasizzo, H.; Puiseux-Dao, S. Characteristics of okadaic acid—Induced cytotoxic effects in CHO K1 cells. Environ. Toxicol. Int. J. 2003, 18, 383–394. [Google Scholar] [CrossRef]
  44. Coates, C.J.; Lim, J.; Harman, K.; Rowley, A.F.; Griffiths, D.J.; Emery, H.; Layton, W. The insect, Galleria mellonella, is a compatible model for evaluating the toxicology of okadaic acid. Cell Biol. Toxicol. 2019, 35, 219–232. [Google Scholar] [CrossRef] [Green Version]
  45. Landsberg, J.H.; Balazs, G.H.; Steidinger, K.A.; Baden, D.G.; Work, T.M.; Russell, D.J. The potential role of natural tumor promoters in marine turtle fibropapillomatosis. J. Aquat. Anim. Health 1999, 11, 199–210. [Google Scholar] [CrossRef]
  46. Wang, J.; Wang, Y.Y.; Lin, L.; Gao, Y.; Hong, H.S.; Wang, D.Z. Quantitative proteomic analysis of okadaic acid treated mouse small intestines reveals differentially expressed proteins involved in diarrhetic shellfish poisoning. J. Proteom. 2012, 75, 2038–2052. [Google Scholar] [CrossRef]
  47. Visciano, P.; Schirone, M.; Berti, M.; Milandri, A.; Tofalo, R.; Suzzi, G. Marine Biotoxins: Occurrence, Toxicity, Regulatory Limits and Reference Methods. Front. Microbiol. 2016, 7, 1051. [Google Scholar] [CrossRef] [Green Version]
  48. Bialojan, C.; Takai, A. Inhibitory effect of a marine-sponge toxin, okadaic acid, on protein phosphatases. Specificity and kinetics. Biochem. J. 1988, 256, 283–290. [Google Scholar] [CrossRef] [PubMed]
  49. Moita, M.T.; Pazos, Y.; Rocha, C.; Nolasco, R.; Oliveira, P.B. Toward predicting Dinophysis blooms off NW Iberia: A decade of events. Harmful Algae 2016, 53, 17–32. [Google Scholar] [CrossRef] [PubMed]
  50. Alfonso, A.; de la Rosa, L.; Vieytes, M.R.; Yasumoto, T.; Botana, L.M. Yessotoxin, a novel phycotoxin, activates phosphodiesterase activity. Effect of yessotoxin on cAMP levels in human lymphocytes. Biochem. Pharmacol. 2003, 65, 193–208. [Google Scholar] [CrossRef]
  51. Murata, M.; Kumagai, M.; Lee, J.S.; Yasumoto, T. Isolation and structure of yessotoxin, a novel polyether compound implicated in diarrhetic shellfish poisoning. Tetrahedron Lett. 1987, 28, 5869–5872. [Google Scholar] [CrossRef]
  52. Alfonso, A.; Vieytes, M.R.; Botana, L.M. Yessotoxin, a Promising Therapeutic Tool. Mar. Drugs 2016, 14, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ciminiello, P.; Fattorusso, E.; Forino, M.; Magno, S.; Poletti, R.; Satake, M.; Viviani, R.; Yasumoto, T. Yessotoxin in mussels of the northern Adriatic Sea. Toxicon 1997, 35, 177–183. [Google Scholar] [CrossRef]
  54. Paz, B.; Riobo, P.; Ramilo, I.; Franco, J.M. Yessotoxins profile in strains of Protoceratium reticulatum from Spain and USA. Toxicon 2007, 50, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Rhodes, L.; McNabb, P.; De Salas, M.; Briggs, L.; Beuzenberg, V.; Gladstone, M. Yessotoxin production by Gonyaulax spinifera. Harmful Algae 2006, 5, 148–155. [Google Scholar] [CrossRef]
  56. Pistocchi, R.; Guerrini, F.; Pezzolesi, L.; Riccardi, M.; Vanucci, S.; Ciminiello, P.; Dell’Aversano, C.; Forino, M.; Fattorusso, E.; Tartaglione, L.; et al. Toxin levels and profiles in microalgae from the North-Western Adriatic Sea—15 years of studies on cultured species. Mar. Drugs 2012, 10, 140–162. [Google Scholar] [CrossRef] [Green Version]
  57. Liu, Y.; Yu, R.-C.; Kong, F.-Z.; Li, C.; Dai, L.; Chen, Z.-F.; Geng, H.-X.; Zhou, M.-J. Contamination status of lipophilic marine toxins in shellfish samples from the Bohai Sea, China. Environ. Pollut. 2019, 249, 171–180. [Google Scholar] [CrossRef]
  58. Suzuki, T.; Horie, Y.; Koike, K.; Satake, M.; Oshima, Y.; Iwataki, M.; Yoshimatsu, S. Yessotoxin analogues in several strains of Protoceratium reticulatum in Japan determined by liquid chromatography-hybrid triple quadrupole/linear ion trap mass spectrometry. J. Chromatogr. A 2007, 1142, 172–177. [Google Scholar] [CrossRef] [PubMed]
  59. Ogino, H.; Kumagai, M.; Yasumoto, T. Toxicologic evaluation of yessotoxin. Nat. Toxins 1997, 5, 255–259. [Google Scholar] [CrossRef]
  60. Pérez-Gómez, A.; Ferrero-Gutierrez, A.; Novelli, A.; Franco, J.M.; Paz, B.; Fernández-Sánchez, M.T. Potent neurotoxic action of the shellfish biotoxin yessotoxin on cultured cerebellar neurons. Toxicol. Sci. 2006, 90, 168–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Seymour, B.; Andreosso, A.; Seymour, J. Cardiovascular toxicity from marine envenomation. In Heart and Toxins; Elsevier: Amsterdam, The Netherlands, 2015; pp. 203–223. [Google Scholar]
  62. Paz, B.; Daranas, A.H.; Norte, M.; Riobo, P.; Franco, J.M.; Fernandez, J.J. Yessotoxins, a group of marine polyether toxins: An overview. Mar. Drugs 2008, 6, 73–102. [Google Scholar] [CrossRef]
  63. Rubini, S.; Albonetti, S.; Menotta, S.; Cervo, A.; Callegari, E.; Cangini, M.; Dall’Ara, S.; Baldini, E.; Vertuani, S.; Manfredini, S. New Trends in the Occurrence of Yessotoxins in the Northwestern Adriatic Sea. Toxins 2021, 13, 634. [Google Scholar] [CrossRef]
  64. Bianchi, C.; Fato, R.; Angelin, A.; Trombetti, F.; Ventrella, V.; Borgatti, A.R.; Fattorusso, E.; Ciminiello, P.; Bernardi, P.; Lenaz, G. Yessotoxin, a shellfish biotoxin, is a potent inducer of the permeability transition in isolated mitochondria and intact cells. Biochim. Biophys. Acta (BBA)-Bioenerg. 2004, 1656, 139–147. [Google Scholar] [CrossRef]
  65. Barroso, J.M. Commission Regulation (EU) No 786/2013.of 16 August 2013 amending Annex III to Regulation (EC) No 853/2004 of the European Parliament and of the Council as regards the permitted limits of yessotoxins in live bivalve molluscs. Off. J. Eur. Union 2013, 14, 1. [Google Scholar]
  66. Twiner, M.J.; Rehmann, N.; Hess, P.; Doucette, G.J. Azaspiracid shellfish poisoning: A review on the chemistry, ecology, and toxicology with an emphasis on human health impacts. Mar. Drugs 2008, 6, 39–72. [Google Scholar] [CrossRef]
  67. Furey, A.; O’Doherty, S.; O’Callaghan, K.; Lehane, M.; James, K.J. Azaspiracid poisoning (AZP) toxins in shellfish: Toxicological and health considerations. Toxicon 2010, 56, 173–190. [Google Scholar] [CrossRef]
  68. McGirr, S.; Clarke, D.; Kilcoyne, J.; Salas, R.; Koehler, H.; Silke, J.; Touzet, N. Insights into the discrepancy between Azadinium spp. and azaspiracid toxins near strategically important aquaculture operations in the west and southwest of Ireland. Estuar. Coast. Shelf Sci. 2021, 262, 107622. [Google Scholar] [CrossRef]
  69. Hess, P.; McCarron, P.; Krock, B.; Kilcoyne, J.; Miles, C.O. Azaspiracids: Chemistry, biosynthesis, metabolism, and detection. In Seafood and Freshwater Toxins: Pharmacology, Physiology, and Detection; CRC Press: Boca Raton, FL, USA, 2014; pp. 799–821. [Google Scholar]
  70. Satake, M.; Ofuji, K.; Naoki, H.; James, K.J.; Furey, A.; McMahon, T.; Silke, J.; Yasumoto, T. Azaspiracid, a new marine toxin having unique spiro ring assemblies, isolated from Irish mussels, Mytilus edulis. J. Am. Chem. Soc. 1998, 120, 9967–9968. [Google Scholar] [CrossRef]
  71. Wu, X.; Hou, L.; Lin, X.; Xie, Z. Application of novel nanomaterials for chemo-and biosensing of algal toxins in shellfish and water. In Novel Nanomaterials for Biomedical, Environmental and Energy Applications; Elsevier: Amsterdam, The Netherlands, 2019; pp. 353–414. [Google Scholar]
  72. Lehane, M.; Brana-Magdalena, A.; Moroney, C.; Furey, A.; James, K. Liquid chromatography with electrospray ion trap mass spectrometry for the determination of five azaspiracids in shellfish. J. Chromatogr. A 2002, 950, 139–147. [Google Scholar] [CrossRef]
  73. Diaz Sierra, M.; Furey, A.; Hamilton, B.; Lehane, M.; James, K.J. Elucidation of the fragmentation pathways of azaspiracids, using electrospray ionisation, hydrogen/deuterium exchange, and multiple-stage mass spectrometry. J. Mass Spectrom. 2003, 38, 1178–1186. [Google Scholar] [CrossRef]
  74. Hamilton, B.; Díaz Sierra, M.; Lehane, M.; Furey, A.; James, K.J. The fragmentation pathways of azaspiracids elucidated using positive nanospray hybrid quadrupole time-of-flight (QqTOF) mass spectrometry. Spectroscopy 2004, 18, 355–362. [Google Scholar] [CrossRef]
  75. Hess, P.; Grune, B.; Anderson, D.B.; Aune, T.; Botana, L.M.; Caricato, P.; van Egmond, H.P.; Halder, M.; Hall, S.; Lawrence, J.F.; et al. Three Rs Approaches in Marine Biotoxin Testing: The Report and Recommendations of a joint ECVAM/DG SANCO Workshop (ECVAM Workshop 54). Altern. Lab. Anim. 2006, 34, 193–224. [Google Scholar] [CrossRef] [PubMed]
  76. Abal, P.; Louzao, M.C.; Fraga, M.; Vilarino, N.; Ferreiro, S.; Vieytes, M.R.; Botana, L.M. Absorption and Effect of Azaspiracid-1 Over the Human Intestinal Barrier. Cell. Physiol. Biochem. 2017, 43, 136–146. [Google Scholar] [CrossRef]
  77. Ronzitti, G.; Hess, P.; Rehmann, N.; Rossini, G.P. Azaspiracid-1 alters the E-cadherin pool in epithelial cells. Toxicol. Sci. 2007, 95, 427–435. [Google Scholar] [CrossRef] [Green Version]
  78. Authority, E.F.S. Marine biotoxins in shellfish–Azaspiracid group-Scientific Opinion of the Panel on Contaminants in the Food chain. EFSA J. 2008, 6, 723. [Google Scholar]
  79. Alves, R.N.; Rambla-Alegre, M.; Braga, A.C.; Maulvault, A.L.; Barbosa, V.; Campàs, M.; Reverté, L.; Flores, C.; Caixach, J.; Kilcoyne, J.; et al. Bioaccessibility of lipophilic and hydrophilic marine biotoxins in seafood: An in vitro digestion approach. Food Chem. Toxicol. 2019, 129, 153–161. [Google Scholar] [CrossRef] [Green Version]
  80. Jeffery, B.; Barlow, T.; Moizer, K.; Paul, S.; Boyle, C. Amnesic shellfish poison. Food Chem. Toxicol. 2004, 42, 545–557. [Google Scholar] [CrossRef]
  81. Vale, P.; Sampayo, M.A.M. Domoic acid in Portuguese shellfish and fish. Toxicon 2001, 39, 893–904. [Google Scholar] [CrossRef]
  82. Hambright, K.D.; Zamor, R.M.; Easton, J.D.; Allison, B. Encyclopedia of Toxicology, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2014; Volume 1. [Google Scholar]
  83. Truelove, J.; Mueller, R.; Pulido, O.; Martin, L.; Fernie, S.; Iverson, F. 30-day oral toxicity study of domoic acid in Cynomolgus monkeys: Lack of overt toxicity at doses approaching the acute toxic dose. Nat. Toxins 1997, 5, 111–114. [Google Scholar] [CrossRef] [PubMed]
  84. Grimmelt, B.; Nijjar, M.S.; Brown, J.; Macnair, N.; Wagner, S.; Johnson, G.R.; Amend, J.F. Relationship between domoic acid levels in the blue mussel (Mytilus edulis) and toxicity in mice. Toxicon 1990, 28, 501–508. [Google Scholar] [CrossRef]
  85. Alfonso, M.; Duran, R.; Arufe, M.C. Effect of excitatory amino acids on serum TSH and thyroid hormone levels in freely moving rats. Horm. Res. Paediatr. 2000, 54, 78–83. [Google Scholar] [CrossRef]
  86. Hampson, D.R.; Huang, X.-p.; Wells, J.W.; Walter, J.A.; Wright, J.L. Interaction of domoic acid and several derivatives with kainic acid and AMPA binding sites in rat brain. Eur. J. Pharmacol. 1992, 218, 1–8. [Google Scholar] [CrossRef]
  87. Burns, J.M.; Hall, S.; Ferry, J.L. The adsorption of saxitoxin to clays and sediments in fresh and saline waters. Water Res. 2009, 43, 1899–1904. [Google Scholar] [CrossRef] [PubMed]
  88. 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] [Green Version]
  89. 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]
  90. 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]
  91. Fox, J.W. 114—Venoms and Poisons from Marine Organisms. In Goldman’s Cecil Medicine, 24th ed.; Goldman, L., Schafer, A.I., Eds.; W.B. Saunders: Philadelphia, PA, USA, 2012; pp. 697–700. [Google Scholar] [CrossRef]
  92. Faber, S. Saxitoxin and the induction of paralytic shellfish poisoning. J. Young Investig. 2012, 23, 1–7. [Google Scholar]
  93. 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] [PubMed]
  94. Falconer, I.R. Algal Toxins in Seafood and Drinking Water; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  95. Rogers, R.S.; Rapoport, H. The pKa’s of saxitoxin. J. Am. Chem. Soc. 1980, 102, 7335–7339. [Google Scholar] [CrossRef]
  96. Harland, F.; Wood, S.A.; Broady, P.; Williamson, W.; Gaw, S. Changes in saxitoxin-production through growth phases in the metaphytic cyanobacterium Scytonema cf. crispum. Toxicon 2015, 103, 74–79. [Google Scholar] [CrossRef] [PubMed]
  97. Watkins, S.M.; Reich, A.; Fleming, L.E.; Hammond, R. Neurotoxic shellfish poisoning. Mar. Drugs 2008, 6, 431–455. [Google Scholar] [CrossRef]
  98. Botana, L.M.; Alfonso, A. Phycotoxins: Chemistry and Biochemistry; John Wiley & Sons: New York, NY, USA, 2015. [Google Scholar]
  99. Plakas, S.M.; Dickey, R.W. Advances in monitoring and toxicity assessment of brevetoxins in molluscan shellfish. Toxicon 2010, 56, 137–149. [Google Scholar] [CrossRef]
  100. Turchiano, R. Brief review of natural nonprotein neurotoxins. Biol. Warf. 2015, 89, 16–24. [Google Scholar]
  101. Abraham, A.; Plakas, S.M.; Wang, Z.; Jester, E.L.; El Said, K.R.; Granade, H.R.; Henry, M.S.; Blum, P.C.; Pierce, R.H.; Dickey, R.W. Characterization of polar brevetoxin derivatives isolated from Karenia brevis cultures and natural blooms. Toxicon 2006, 48, 104–115. [Google Scholar] [CrossRef]
  102. Poli, M.A. Laboratory procedures for detoxification of equipment and waste contaminated with brevetoxins PbTx-2 and PbTx-3. J. Assoc. Off. Anal. Chem. 1988, 71, 1000–1002. [Google Scholar] [CrossRef]
  103. Trainer, V.L.; Edwards, R.A.; Szmant, A.M.; Stuart, A.M.; Mende, T.J.; Baden, D.G. Brevetoxins: Unique Activators of Voltage-Sensitive Sodium Channels; ACS Publications: Washington, DC, USA, 1990. [Google Scholar]
  104. Poli, M.A.; Templeton, C.B.; Pace, J.G.; Hines, H.B. Detection, Metabolism, and Pathophysiology of Brevetoxins; ACS Publications: Washington, DC, USA, 1990. [Google Scholar]
  105. Friedman, M.A.; Fernandez, M.; Backer, L.C.; Dickey, R.W.; Bernstein, J.; Schrank, K.; Kibler, S.; Stephan, W.; Gribble, M.O.; Bienfang, P.; et al. An updated review of ciguatera fish poisoning: Clinical, epidemiological, environmental, and public health management. Mar. Drugs 2017, 15, 72. [Google Scholar] [CrossRef] [Green Version]
  106. Friedman, M.A.; Fleming, L.E.; Fernandez, M.; Bienfang, P.; Schrank, K.; Dickey, R.; Bottein, M.-Y.; Backer, L.; Ayyar, R.; Weisman, R. Ciguatera fish poisoning: Treatment, prevention and management. Mar. Drugs 2008, 6, 456–479. [Google Scholar] [CrossRef]
  107. Lehane, L.; Lewis, R.J. Ciguatera: Recent advances but the risk remains. Int. J. Food Microbiol. 2000, 61, 91–125. [Google Scholar] [CrossRef]
  108. Dechraoui, M.Y.; Tiedeken, J.A.; Persad, R.; Wang, Z.; Granade, H.R.; Dickey, R.W.; Ramsdell, J.S. Use of two detection methods to discriminate ciguatoxins from brevetoxins: Application to great barracuda from Florida Keys. Toxicon 2005, 46, 261–270. [Google Scholar] [CrossRef] [PubMed]
  109. Schlaich, C.; Hagelstein, J.G.; Burchard, G.D.; Schmiedel, S. Outbreak of ciguatera fish poisoning on a cargo ship in the port of hamburg. J. Travel Med. 2012, 19, 238–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Chateau-Degat, M.L.; Dewailly, E.; Cerf, N.; Nguyen, N.L.; Huin-Blondey, M.O.; Hubert, B.; Laudon, F.; Chansin, R. Temporal trends and epidemiological aspects of ciguatera in French Polynesia: A 10-year analysis. Trop. Med. Int. Health 2007, 12, 485–492. [Google Scholar] [CrossRef] [PubMed]
  111. Arena, P.; Levin, B.; Fleming, L.; Friedman, M.; Blythe, D. A pilot study of the cognitive and psychological correlates of chronic ciguatera poisoning. Harmful Algae 2004, 3, 51–60. [Google Scholar] [CrossRef]
  112. Friedman, M.A.; Arena, P.; Levin, B.; Fleming, L.; Fernandez, M.; Weisman, R.; Bernstein, J.; Schrank, K.; Blythe, D.; Backer, L.; et al. Neuropsychological study of ciguatera fish poisoning: A longitudinal case-control study. Arch. Clin. Neuropsychol. 2007, 22, 545–553. [Google Scholar] [CrossRef] [Green Version]
  113. Caillaud, A.; de la Iglesia, P.; Darius, H.T.; Pauillac, S.; Aligizaki, K.; Fraga, S.; Chinain, M.; Diogene, J. Update on methodologies available for ciguatoxin determination: Perspectives to confront the onset of ciguatera fish poisoning in Europe. Mar. Drugs 2010, 8, 1838–1907. [Google Scholar] [CrossRef]
  114. James, K.; Lehane, M.; Moroney, C.; Fernandez-Puente, P.; Satake, M.; Yasumoto, T.; Furey, A. Azaspiracid shellfish poisoning: Unusual toxin dynamics in shellfish and the increased risk of acute human intoxications. Food Addit. Contam. 2002, 19, 555–561. [Google Scholar] [CrossRef]
  115. La Barre, S.; Bates, S.S.; Quilliam, M.A. Domoic Acid. In Outstanding Marine Molecules; Wiley-VCH: Weinheim, Germany, 2014; pp. 189–216. [Google Scholar] [CrossRef]
  116. MacKenzie, L.; Beuzenberg, V.; Holland, P.; McNabb, P.; Selwood, A. Solid phase adsorption toxin tracking (SPATT): A new monitoring tool that simulates the biotoxin contamination of filter feeding bivalves. Toxicon 2004, 44, 901–918. [Google Scholar] [CrossRef]
  117. McCarthy, H.P.; Crowder, L.B. An overlooked scale of global transport: Phytoplankton species richness in ships’ ballast water. Biol. Invasions 2000, 2, 321. [Google Scholar] [CrossRef]
  118. Hallegraeff, G.M. Transport of toxic dinoflagellates via ships ballast water: Bioeconomic risk assessment and efficacy of possible ballast water management strategies. Mar. Ecol. Prog. Ser. 1998, 168, 297–309. [Google Scholar] [CrossRef] [Green Version]
  119. van den Bergh, J.C.; Nunes, P.A.; Dotinga, H.M.; Kooistra, W.H.; Vrieling, E.G.; Peperzak, L. Exotic harmful algae in marine ecosystems: An integrated biological–economic–legal analysis of impacts and policies. Mar. Policy 2002, 26, 59–74. [Google Scholar] [CrossRef]
  120. Moore, S.K.; Trainer, V.L.; Mantua, N.J.; Parker, M.S.; Laws, E.A.; Backer, L.C.; Fleming, L.E. Impacts of climate variability and future climate change on harmful algal blooms and human health. In Environmental Health; Springer: Berlin/Heidelberg, Germany, 2008; pp. 1–12. [Google Scholar]
  121. Callaway, R.; Shinn, A.P.; Grenfell, S.E.; Bron, J.E.; Burnell, G.; Cook, E.J.; Crumlish, M.; Culloty, S.; Davidson, K.; Ellis, R.P.; et al. Review of climate change impacts on marine aquaculture in the UK and Ireland. Aquat. Conserv. Mar. Freshw. Ecosyst. 2012, 22, 389–421. [Google Scholar] [CrossRef]
  122. Sanseverino, I.; Conduto, D.; Pozzoli, L.; Dobricic, S.; Lettieri, T. Algal bloom and its economic impact. Eur. Comm. Jt. Res. Cent. Inst. Environ. Sustain. 2016. [CrossRef]
  123. Panda, D.; Dash, B.P.; Manickam, S.; Boczkaj, G. Recent advancements in LC-MS based analysis of biotoxins: Present and future challenges. Mass Spectrom. Rev. 2021, 41, 766–803. [Google Scholar] [CrossRef] [PubMed]
  124. Vilarino, N.; Louzao, M.C.; Fraga, M.; Rodriguez, L.P.; Botana, L.M. Innovative detection methods for aquatic algal toxins and their presence in the food chain. Anal. Bioanal. Chem. 2013, 405, 7719–7732. [Google Scholar] [CrossRef]
  125. Turrell, E.A.; Stobo, L. A comparison of the mouse bioassay with liquid chromatography-mass spectrometry for the detection of lipophilic toxins in shellfish from Scottish waters. Toxicon 2007, 50, 442–447. [Google Scholar] [CrossRef]
  126. Gerssen, A.; Pol-Hofstad, I.E.; Poelman, M.; Mulder, P.P.; Van den Top, H.J.; De Boer, J. Marine toxins: Chemistry, toxicity, occurrence and detection, with special reference to the Dutch situation. Toxins 2010, 2, 878–904. [Google Scholar] [CrossRef] [Green Version]
  127. Puente, P.F.; Sáez, M.J.F.; Hamilton, B.; Lehane, M.; Ramstad, H.; Furey, A.; James, K.J. Rapid determination of polyether marine toxins using liquid chromatography–multiple tandem mass spectrometry. J. Chromatogr. A 2004, 1056, 77–82. [Google Scholar] [CrossRef]
  128. Bane, V.; Brosnan, B.; Barnes, P.; Lehane, M.; Furey, A. High-resolution mass spectrometry analysis of tetrodotoxin (TTX) and its analogues in puffer fish and shellfish. Food Addit. Contam. Part A 2016, 33, 1468–1489. [Google Scholar] [CrossRef]
  129. Bane, V.; Lehane, M.; Dikshit, M.; O’Riordan, A.; Furey, A. Tetrodotoxin: Chemistry, toxicity, source, distribution and detection. Toxins 2014, 6, 693–755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Rossini, G.P. Functional assays in marine biotoxin detection. Toxicology 2005, 207, 451–462. [Google Scholar] [CrossRef] [PubMed]
  131. Jellett, J.F.; Marks, L.J.; Stewart, J.E.; Dorey, M.L.; Watson-Wright, W.; Lawrence, J.F. Paralytic shellfish poison (saxitoxin family) bioassays: Automated endpoint determination and standardization of the in vitro tissue culture bioassay, and comparison with the standard mouse bioassay. Toxicon 1992, 30, 1143–1156. [Google Scholar] [CrossRef]
  132. Nicholson, R.A.; Li, G.H.; Buenaventura, E.; Graham, D. A rapid and sensitive assay for paralytic shellfish poison (PSP) toxins using mouse brain synaptoneurosomes. Toxicon 2002, 40, 831–838. [Google Scholar] [CrossRef]
  133. Vieytes, M.; Fontal, O.; Leira, F.; de Sousa, J.B.; Botana, L. A fluorescent microplate assay for diarrheic shellfish toxins. Anal. Biochem. 1997, 248, 258–264. [Google Scholar] [CrossRef]
  134. Louhimies, S. Directive 86/609/EEC on the protection of animals used for experimental and other scientific purposes. Altern. Lab. Anim. 2002, 30, 217–219. [Google Scholar] [CrossRef]
  135. Garthwaite, I.; Ross, K.M.; Miles, C.O.; Hansen, R.P.; Foster, D.; Wilkins, A.L.; Towers, N.R. Polyclonal antibodies to domoic acid, and their use in immunoassays for domoic acid in sea water and shellfish. Nat. Toxins 1998, 6, 93–104. [Google Scholar] [CrossRef]
  136. Wang, L.; Gong, X.; Wang, R.; Gan, Z.; Lu, Y.; Sun, H. Application of an immobilized ionic liquid for the passive sampling of perfluorinated substances in water. J. Chromatogr. A 2017, 1515, 45–53. [Google Scholar] [CrossRef]
  137. Taniyama, S.; Arakawa, O.; Terada, M.; Nishio, S.; Takatani, T.; Mahmud, Y.; Noguchi, T. Ostreopsis sp., a possible origin of palytoxin (PTX) in parrotfish Scarus ovifrons. Toxicon 2003, 42, 29–33. [Google Scholar] [CrossRef]
  138. He, K.; Zhang, X.; Wang, L.; Du, X.; Wei, D. Production of a soluble single-chain variable fragment antibody against okadaic acid and exploration of its specific binding. Anal. Biochem. 2016, 503, 21–27. [Google Scholar] [CrossRef]
  139. Bodero, M.; Gerssen, A.; Portier, L.; Klijnstra, M.D.; Hoogenboom, R.; Guzman, L.; Hendriksen, P.J.M.; Bovee, T.F.H. A Strategy to Replace the Mouse Bioassay for Detecting and Identifying Lipophilic Marine Biotoxins by Combining the Neuro-2a Bioassay and LC-MS/MS Analysis. Mar. Drugs 2018, 16, 501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Lawrence, J.F.; Charbonneau, C.F.; Ménard, C. Liquid chromatographic determination of domoic acid in mussels, using AOAC paralytic shellfish poison extraction procedure: Collaborative study. J. Assoc. Off. Anal. Chem. 1991, 74, 68–72. [Google Scholar] [CrossRef]
  141. Hess, P.; McGovern, E.; McMahon, T.; Morris, S.; Stobo, L.A.; Brown, N.A.; Gallacher, S.; McEvoy, J.D.G.; Kennedy, G.; Young, P.B.; et al. LC-UV and LC-MS methods for the determination of domoic acid. TrAC Trends Anal. Chem. 2005, 24, 358–367. [Google Scholar] [CrossRef]
  142. Lefebvre, K.A.; Hendrix, A.; Halaska, B.; Duignan, P.; Shum, S.; Isoherranen, N.; Marcinek, D.J.; Gulland, F.M. Domoic acid in California sea lion fetal fluids indicates continuous exposure to a neuroteratogen poses risks to mammals. Harmful Algae 2018, 79, 53–57. [Google Scholar] [CrossRef] [PubMed]
  143. Bogan, Y. Factors Affecting the Concentration of Domoic Acid in Scallop, Pecten maximus. Ph.D. Thesis, Letterkenny Institute of Technology, Letterkenny, Ireland, 2006. [Google Scholar]
  144. López-Rivera, A.; Pinto, M.; Insinilla, A.; Isla, B.S.; Uribe, E.; Alvarez, G.; Lehane, M.; Furey, A.; James, K.J. The occurrence of domoic acid linked to a toxic diatom bloom in a new potential vector: The tunicate Pyura chilensis (piure). Toxicon 2009, 54, 754–762. [Google Scholar] [CrossRef]
  145. Furey, A.; Lehane, M.; Gillman, M.; Fernandez-Puente, P.; James, K.J. Determination of domoic acid in shellfish by liquid chromatography with electrospray ionization and multiple tandem mass spectrometry. J. Chromatogr. A 2001, 938, 167–174. [Google Scholar] [CrossRef]
  146. Lee, J.S.; Yanagi, T.; Kenma, R.; Yasumoto, T. Fluorometric determination of diarrhetic shellfish toxins by high-performance liquid chromatography. Agric. Biol. Chem. 1987, 51, 877–881. [Google Scholar]
  147. Dorne, J.; Bordajandi, L.; Amzal, B.; Ferrari, P.; Verger, P. Combining analytical techniques, exposure assessment and biological effects for risk assessment of chemicals in food. TrAC Trends Anal. Chem. 2009, 28, 695–707. [Google Scholar] [CrossRef]
  148. Kilcoyne, J.; Fux, E. Strategies for the elimination of matrix effects in the liquid chromatography tandem mass spectrometry analysis of the lipophilic toxins okadaic acid and azaspiracid-1 in molluscan shellfish. J. Chromatogr. A 2010, 1217, 7123–7130. [Google Scholar] [CrossRef]
  149. Holland, P.; McNabb, P.; Selwood, A.; Neil, T.; Slattery, D.; Van de Riet, J.; Van Egmond, H.; Van den Topp, H.; Yasumoto, T. A multiresidue LC-MS method for algal toxins in shellfish: Inter-laboratory study. In Proceedings of the HABTech03 Workshop, Nelson, New Zealand, November 2003. [Google Scholar]
  150. Otero, P.; Miguéns, N.; Rodríguez, I.; Botana, L.M. LC–MS/MS analysis of the emerging toxin pinnatoxin-G and high levels of esterified OA group toxins in Galician commercial mussels. Toxins 2019, 11, 394. [Google Scholar] [CrossRef] [Green Version]
  151. Harju, K.; Rapinoja, M.-L.; Avondet, M.-A.; Arnold, W.; Schär, M.; Burrell, S.; Luginbühl, W.; Vanninen, P. Optimization of sample preparation for the identification and quantification of saxitoxin in proficiency test mussel sample using liquid chromatography-tandem mass spectrometry. Toxins 2015, 7, 4868–4880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Lawrence, J.F.; Niedzwiadek, B.; Menard, C. Quantitative determination of paralytic shellfish poisoning toxins in shellfish using prechromatographic oxidation and liquid chromatography with fluorescence detection: Collaborative study. J. AOAC Int. 2005, 88, 1714–1732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Yasumoto, T.; Takizawa, A. Fluorometric measurement of yessotoxins in shellfish by high-pressure liquid chromatography. Biosci. Biotechnol. Biochem. 1997, 61, 1775–1777. [Google Scholar] [CrossRef]
  154. Zendong, Z.; Bertrand, S.; Herrenknecht, C.; Abadie, E.; Jauzein, C.; Lemée, R.; Gouriou, J.; Amzil, Z.; Hess, P. Passive Sampling and High Resolution Mass Spectrometry for Chemical Profiling of French Coastal Areas with a Focus on Marine Biotoxins. Environ. Sci. Technol. 2016, 50, 8522–8529. [Google Scholar] [CrossRef] [PubMed]
  155. Banack, S.; Metcalf, J.; Spáčil, Z.; Downing, T.; Downing, S.; Long, A.; Nunn, P.B.; Cox, P.A. Distinguishing the cyanobacterial neurotoxin β-N-methylamino-L-alanine (BMAA) from other diamino acids. Toxicon 2011, 57, 730–738. [Google Scholar] [CrossRef] [PubMed]
  156. Kudela, R.M. Characterization and deployment of Solid Phase Adsorption Toxin Tracking (SPATT) resin for monitoring of microcystins in fresh and saltwater. Harmful Algae 2011, 11, 117–125. [Google Scholar] [CrossRef]
  157. Miller, M.A.; Kudela, R.M.; Mekebri, A.; Crane, D.; Oates, S.C.; Tinker, M.T.; Staedler, M.; Miller, W.A.; Toy-Choutka, S.; Dominik, C.; et al. Evidence for a novel marine harmful algal bloom: Cyanotoxin (microcystin) transfer from land to sea otters. PLoS ONE 2010, 5, e12576. [Google Scholar] [CrossRef]
  158. Mackenzie, L.; White, D.; Oshima, Y.; Kapa, J. The resting cyst and toxicity of Alexandrium ostenfeldii (Dinophyceae) in New Zealand. Phycologia 1996, 35, 148–155. [Google Scholar] [CrossRef]
  159. Kudela, R.M. Passive sampling for freshwater and marine algal toxins. Compr. Anal. Chem. 2017, 78, 379–409. [Google Scholar]
  160. MacKenzie, L.A. In situ passive solid-phase adsorption of micro-algal biotoxins as a monitoring tool. Curr. Opin. Biotechnol. 2010, 21, 326–331. [Google Scholar] [CrossRef]
  161. Rundberget, T.; Sandvik, M.; Larsen, K.; Pizarro, G.M.; Reguera, B.; Castberg, T.; Gustad, E.; Loader, J.I.; Rise, F.; Wilkins, A.L.; et al. Extraction of microalgal toxins by large-scale pumping of seawater in Spain and Norway, and isolation of okadaic acid and dinophysistoxin-2. Toxicon 2007, 50, 960–970. [Google Scholar] [CrossRef] [PubMed]
  162. Wood, S.A.; Holland, P.T.; MacKenzie, L. Development of solid phase adsorption toxin tracking (SPATT) for monitoring anatoxin-a and homoanatoxin-a in river water. Chemosphere 2011, 82, 888–894. [Google Scholar] [CrossRef]
  163. Shin, H.S.; Kim, J.-H. Isotherm, kinetic and thermodynamic characteristics of adsorption of paclitaxel onto Diaion HP-20. Process Biochem. 2016, 51, 917–924. [Google Scholar] [CrossRef]
  164. Mitsubishi Chemical Corporation. In Diaion HP20 Product Data Sheet; Mitsubishi Chemical Corporation: Tokyo, Japan.
  165. McCarthy, M.; van Pelt, F.N.A.M.; Bane, V.; O’Halloran, J.; Furey, A. Application of passive (SPATT) and active sampling methods in the profiling and monitoring of marine biotoxins. Toxicon 2014, 89, 77–86. [Google Scholar] [CrossRef] [PubMed]
  166. Mitsubishi Chemical Corporation. DIAION Technical Manual; Mitsubishi Chemical Corporation: Tokyo, Japan.
  167. Mitsubishi Chemical Corporation. SEPABEADS SP700 Product Data Sheet; Mitsubishi Chemical Corporation: Tokyo, Japan.
  168. Kohoutek, J.; Babica, P.; Bláha, L.; Maršálek, B. A novel approach for monitoring of cyanobacterial toxins: Development and evaluation of the passive sampler for microcystins. Anal. Bioanal. Chem. 2008, 390, 1167–1172. [Google Scholar] [CrossRef]
  169. Caillaud, A.; de la Iglesia, P.; Barber, E.; Eixarch, H.; Mohammad-Noor, N.; Yasumoto, T.; Diogene, J. Monitoring of dissolved ciguatoxin and maitotoxin using solid-phase adsorption toxin tracking devices: Application to Gambierdiscus pacificus in culture. Harmful Algae 2011, 10, 433–446. [Google Scholar] [CrossRef]
  170. Lane, J.Q.; Roddam, C.M.; Langlois, G.W.; Kudela, R.M. Application of Solid Phase Adsorption Toxin Tracking (SPATT) for field detection of the hydrophilic phycotoxins domoic acid and saxitoxin in coastal California. Limnol. Oceanogr. Methods 2010, 8, 645–660. [Google Scholar] [CrossRef]
  171. Turrell, E.; Stobo, L.; Lacaze, J.-P.; Bresnan, E.; Gowland, D. Development of anearly warning system’for harmful algal blooms using solid-phase adsorption toxin tracking (SPATT). In Proceedings of the Oceans 2007-Europe, Aberdeen, UK, 18–21 June 2007; pp. 1–6. [Google Scholar]
  172. Fux, E.; Marcaillou, C.; Mondeguer, F.; Bire, R.; Hess, P. Field and mesocosm trials on passive sampling for the study of adsorption and desorption behaviour of lipophilic toxins with a focus on OA and DTX1. Harmful Algae 2008, 7, 574–583. [Google Scholar] [CrossRef] [Green Version]
  173. Zhao, H.; Qiu, J.; Fan, H.; Li, A. Mechanism and application of solid phase adsorption toxin tracking for monitoring microcystins. J. Chromatogr. A 2013, 1300, 159–164. [Google Scholar] [CrossRef]
  174. Li, A.; Ma, F.; Song, X.; Yu, R. Dynamic adsorption of diarrhetic shellfish poisoning (DSP) toxins in passive sampling relates to pore size distribution of aromatic adsorbent. J. Chromatogr. A 2011, 1218, 1437–1442. [Google Scholar] [CrossRef]
  175. Tesser, R.; Di Serio, M.; Casale, L.; Carotenuto, G.; Santacesaria, E. Absorption of water/methanol binary system on ion-exchange resins. Can. J. Chem. Eng. 2010, 88, 1044–1053. [Google Scholar] [CrossRef]
  176. Peacock, M.B.; Gibble, C.M.; Senn, D.B.; Cloern, J.E.; Kudela, R.M. Blurred lines: Multiple freshwater and marine algal toxins at the land-sea interface of San Francisco Bay, California. Harmful Algae 2018, 73, 138–147. [Google Scholar] [CrossRef] [PubMed]
  177. Fux, E.; Bire, R.; Hess, P. Comparative accumulation and composition of lipophilic marine biotoxins in passive samplers and in mussels (M. edulis) on the West Coast of Ireland. Harmful Algae 2009, 8, 523–537. [Google Scholar] [CrossRef]
  178. Howard, M.D.; Nagoda, C.; Kudela, R.M.; Hayashi, K.; Tatters, A.; Caron, D.A.; Busse, L.; Brown, J.; Sutula, M.; Stein, E.D. Microcystin prevalence throughout lentic waterbodies in coastal Southern California. Toxins 2017, 9, 231. [Google Scholar] [CrossRef] [Green Version]
  179. Bueno, M.J.M.; Hernando, M.D.; Agüera, A.; Fernández-Alba, A.R. Application of passive sampling devices for screening of micro-pollutants in marine aquaculture using LC–MS/MS. Talanta 2009, 77, 1518–1527. [Google Scholar] [CrossRef] [PubMed]
  180. Pizarro, G.; Paz, B.; González-Gil, S.; Franco, J.M.; Reguera, B. Seasonal variability of lipophilic toxins during a Dinophysis acuta bloom in Western Iberia: Differences between picked cells and plankton concentrates. Harmful Algae 2009, 8, 926–937. [Google Scholar] [CrossRef]
  181. Rundberget, T.; Gustad, E.; Samdal, I.A.; Sandvik, M.; Miles, C.O. A convenient and cost-effective method for monitoring marine algal toxins with passive samplers. Toxicon 2009, 53, 543–550. [Google Scholar] [CrossRef]
  182. Rodríguez, P.; Alfonso, A.; Turrell, E.; Lacaze, J.-P.; Botana, L.M. Study of solid phase adsorption of paralytic shellfish poisoning toxins (PSP) onto different resins. Harmful Algae 2011, 10, 447–455. [Google Scholar] [CrossRef]
  183. Rundberget, T.; Aasen, J.A.; Selwood, A.I.; Miles, C.O. Pinnatoxins and spirolides in Norwegian blue mussels and seawater. Toxicon 2011, 58, 700–711. [Google Scholar] [CrossRef]
  184. Touzet, N.; Lacaze, J.; Maher, M.; Turrell, E.; Raine, R. Summer dynamics of Alexandrium ostenfeldii (Dinophyceae) and spirolide toxins in Cork Harbour, Ireland. Mar. Ecol. Prog. Ser. 2011, 425, 21–33. [Google Scholar] [CrossRef] [Green Version]
  185. MacKenzie, L.A.; Selwood, A.I.; McNabb, P.; Rhodes, L. Benthic dinoflagellate toxins in two warm-temperate estuaries: Rangaunu and Parengarenga Harbours, Northland, New Zealand. Harmful Algae 2011, 10, 559–566. [Google Scholar] [CrossRef]
  186. Wood, S.A.; Kuhajek, J.M.; de Winton, M.; Phillips, N.R. Species composition and cyanotoxin production in periphyton mats from three lakes of varying trophic status. FEMS Microbiol. Ecol. 2012, 79, 312–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Pizarro, G.; Moroño, Á.; Paz, B.; Franco, J.M.; Pazos, Y.; Reguera, B. Evaluation of passive samplers as a monitoring tool for early warning of Dinophysis toxins in shellfish. Mar. Drugs 2013, 11, 3823–3845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Garcia-Altares, M.; Casanova, A.; Bane, V.; Diogene, J.; Furey, A.; de la Iglesia, P. Confirmation of pinnatoxins and spirolides in shellfish and passive samplers from Catalonia (Spain) by liquid chromatography coupled with triple quadrupole and high-resolution hybrid tandem mass spectrometry. Mar. Drugs 2014, 12, 3706–3732. [Google Scholar] [CrossRef] [PubMed]
  189. Gibble, C.M.; Kudela, R.M. Detection of persistent microcystin toxins at the land–sea interface in Monterey Bay, California. Harmful Algae 2014, 39, 146–153. [Google Scholar] [CrossRef] [Green Version]
  190. Zendong, Z.; McCarron, P.; Herrenknecht, C.; Sibat, M.; Amzil, Z.; Cole, R.B.; Hess, P. High resolution mass spectrometry for quantitative analysis and untargeted screening of algal toxins in mussels and passive samplers. J. Chromatogr. A 2015, 1416, 10–21. [Google Scholar] [CrossRef]
  191. García-Altares, M.; Casanova, A.; Fernández-Tejedor, M.; Diogène, J.; De La Iglesia, P. Bloom of Dinophysis spp. dominated by D. sacculus and its related diarrhetic shellfish poisoning (DSP) outbreak in Alfacs Bay (Catalonia, NW Mediterranean Sea): Identification of DSP toxins in phytoplankton, shellfish and passive samplers. Reg. Stud. Mar. Sci. 2016, 6, 19–28. [Google Scholar] [CrossRef]
  192. Li, Z.; Sobek, A.; Radke, M. Fate of Pharmaceuticals and Their Transformation Products in Four Small European Rivers Receiving Treated Wastewater. Environ. Sci. Technol. 2016, 50, 5614–5621. [Google Scholar] [CrossRef] [Green Version]
  193. Kim, J.-H.; Tillmann, U.; Adams, N.G.; Krock, B.; Stutts, W.L.; Deeds, J.R.; Han, M.-S.; Trainer, V.L. Identification of Azadinium species and a new azaspiracid from Azadinium poporum in Puget Sound, Washington State, USA. Harmful Algae 2017, 68, 152–167. [Google Scholar] [CrossRef] [Green Version]
  194. Roué, M.; Darius, H.T.; Viallon, J.; Ung, A.; Gatti, C.; Harwood, D.T.; Chinain, M. Application of solid phase adsorption toxin tracking (SPATT) devices for the field detection of Gambierdiscus toxins. Harmful Algae 2018, 71, 40–49. [Google Scholar] [CrossRef]
  195. Wood, S.A.; Biessy, L.; Puddick, J. Anatoxins are consistently released into the water of streams with Microcoleus autumnalis-dominated (cyanobacteria) proliferations. Harmful Algae 2018, 80, 88–95. [Google Scholar] [CrossRef] [PubMed]
  196. Hattenrath-Lehmann, T.K.; Lusty, M.W.; Wallace, R.B.; Haynes, B.; Wang, Z.; Broadwater, M.; Deeds, J.R.; Morton, S.L.; Hastback, W.; Porter, L.; et al. Evaluation of Rapid, Early Warning Approaches to Track Shellfish Toxins Associated with Dinophysis and Alexandrium Blooms. Mar. Drugs 2018, 16, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Murray, J.S.; Boundy, M.J.; Selwood, A.I.; Harwood, D.T. Development of an LC–MS/MS method to simultaneously monitor maitotoxins and selected ciguatoxins in algal cultures and P-CTX-1B in fish. Harmful Algae 2018, 80, 80–87. [Google Scholar] [CrossRef] [PubMed]
  198. Smith, J.; Lie, A.A.Y.; Seubert, E.L.; Crowley, N.; Robertson, G.; Caron, D.A. Co-occurring dissolved algal toxins observed at multiple coastal sites in southern California via solid phase adsorption toxin tracking. Toxicon 2019, 171, 62–65. [Google Scholar] [CrossRef] [PubMed]
  199. Krock, B.; Schloss, I.R.; Trefault, N.; Tillmann, U.; Hernando, M.; Deregibus, D.; Antoni, J.; Almandoz, G.O.; Hoppenrath, M. Detection of the phycotoxin pectenotoxin-2 in waters around King George Island, Antarctica. Polar Biol. 2020, 43, 263–277. [Google Scholar] [CrossRef]
  200. Mathon, B.; Ferreol, M.; Togola, A.; Lardy-Fontan, S.; Dabrin, A.; Allan, I.J.; Staub, P.F.; Mazzella, N.; Miège, C. Polar organic chemical integrative samplers as an effective tool for chemical monitoring of surface waters—Results from one-year monitoring in France. Sci. Total Environ. 2022, 824, 153549. [Google Scholar] [CrossRef]
  201. Challis, J.K.; Almirall, X.O.; Helm, P.A.; Wong, C.S. Performance of the organic-diffusive gradients in thin-films passive sampler for measurement of target and suspect wastewater contaminants. Environ. Pollut. 2020, 261, 114092. [Google Scholar] [CrossRef]
  202. Sonavane, M.; Creusot, N.; Maillot-Marechal, E.; Pery, A.; Brion, F.; Aїt-Aїssa, S. Zebrafish-based reporter gene assays reveal different estrogenic activities in river waters compared to a conventional human-derived assay. Sci. Total Environ. 2016, 550, 934–939. [Google Scholar] [CrossRef]
  203. Galle, T.; Bayerle, M.; Pittois, D.; Huck, V. Allocating biocide sources and flow paths to surface waters using passive samplers and flood wave chemographs. Water Res. 2020, 173, 115533. [Google Scholar] [CrossRef]
  204. Vrana, B.; Smedes, F.; Prokeš, R.; Loos, R.; Mazzella, N.; Miege, C.; Budzinski, H.; Vermeirssen, E.; Ocelka, T.; Gravell, A.; et al. An interlaboratory study on passive sampling of emerging water pollutants. TrAC Trends Anal. Chem. 2016, 76, 153–165. [Google Scholar] [CrossRef] [Green Version]
  205. Harman, C.; Allan, I.J.; Vermeirssen, E.L. Calibration and use of the polar organic chemical integrative sampler--a critical review. Environ. Toxicol. Chem. 2012, 31, 2724–2738. [Google Scholar] [CrossRef] [PubMed]
  206. Morin, S.; Pesce, S.; Kim-Tiam, S.; Libert, X.; Coquery, M.; Mazzella, N. Use of polar organic chemical integrative samplers to assess the effects of chronic pesticide exposure on biofilms. Ecotoxicology 2012, 21, 1570–1580. [Google Scholar] [CrossRef] [PubMed]
  207. Godlewska, K.; Stepnowski, P.; Paszkiewicz, M. Pollutant analysis using passive samplers: Principles, sorbents, calibration and applications. A review. Environ. Chem. Lett. 2021, 19, 465–520. [Google Scholar] [CrossRef]
  208. Alvarez, D.A.; Huckins, J.N.; Petty, J.D.; Jones-Lepp, T.; Stuer-Lauridsen, F.; Getting, D.T.; Goddard, J.P.; Gravell, A. Tool for monitoring hydrophilic contaminants in water: Polar organic chemical integrative sampler (POCIS). Compr. Anal. Chem. 2007, 48, 171–197. [Google Scholar]
  209. Booij, K.; Sleiderink, H.M.; Smedes, F. Calibrating the uptake kinetics of semipermeable membrane devices using exposure standards. Environ. Toxicol. Chem. Int. J. 1998, 17, 1236–1245. [Google Scholar] [CrossRef]
  210. Godlewska, K.; Stepnowski, P.; Paszkiewicz, M. Application of the Polar Organic Chemical Integrative Sampler for Isolation of Environmental Micropollutants—A Review. Crit. Rev. Anal. Chem. 2020, 50, 1–28. [Google Scholar] [CrossRef]
  211. Branchet, P.; Arpin-Pont, L.; Piram, A.; Boissery, P.; Wong-Wah-Chung, P.; Doumenq, P. Pharmaceuticals in the marine environment: What are the present challenges in their monitoring? Sci. Total Environ. 2021, 766, 142644. [Google Scholar] [CrossRef]
  212. Martínez Bueno, M.; Herrera, S.; Munaron, D.; Boillot, C.; Fenet, H.; Chiron, S.; Gómez, E. POCIS passive samplers as a monitoring tool for pharmaceutical residues and their transformation products in marine environment. Environ. Sci. Pollut. Res. 2016, 23, 5019–5029. [Google Scholar] [CrossRef]
  213. Vrana, B.; Urik, J.; Fedorova, G.; Svecova, H.; Grabicova, K.; Golovko, O.; Randak, T.; Grabic, R. In situ calibration of polar organic chemical integrative sampler (POCIS) for monitoring of pharmaceuticals in surface waters. Environ. Pollut. 2021, 269, 116121. [Google Scholar] [CrossRef]
  214. Kaserzon, S.; Hawker, D.; Kennedy, K.; Bartkow, M.; Carter, S.; Booij, K.; Mueller, J. Characterisation and comparison of the uptake of ionizable and polar pesticides, pharmaceuticals and personal care products by POCIS and Chemcatchers. Environ. Sci. Process. Impacts 2014, 16, 2517–2526. [Google Scholar] [CrossRef] [Green Version]
  215. Björlenius, B.; Ripszám, M.; Haglund, P.; Lindberg, R.H.; Tysklind, M.; Fick, J. Pharmaceutical residues are widespread in Baltic Sea coastal and offshore waters–Screening for pharmaceuticals and modelling of environmental concentrations of carbamazepine. Sci. Total Environ. 2018, 633, 1496–1509. [Google Scholar] [CrossRef] [PubMed]
  216. Jones-Lepp, T.; Alvarez, D.; Petty, J.; Huckins, J. Polar organic chemical integrative sampling and liquid chromatography–electrospray/ion-trap mass spectrometry for assessing selected prescription and illicit drugs in treated sewage effluents. Arch. Environ. Contam. Toxicol. 2004, 47, 427–439. [Google Scholar] [CrossRef] [PubMed]
  217. Sultana, T.; Murray, C.; Ehsanul Hoque, M.; Metcalfe, C.D. Monitoring contaminants of emerging concern from tertiary wastewater treatment plants using passive sampling modelled with performance reference compounds. Environ. Monit. Assess. 2016, 189, 1. [Google Scholar] [CrossRef]
  218. Baz-Lomba, J.A.; Harman, C.; Reid, M.; Thomas, K.V. Passive sampling of wastewater as a tool for the long-term monitoring of community exposure: Illicit and prescription drug trends as a proof of concept. Water Res. 2017, 121, 221–230. [Google Scholar] [CrossRef] [Green Version]
  219. Iparraguirre, A.; Prieto, A.; Vallejo, A.; Moeder, M.; Zuloaga, O.; Etxebarria, N.; Paschke, A. Tetraphasic polar organic chemical integrative sampler for the determination of a wide polarity range organic pollutants in water. The use of performance reference compounds and in-situ calibration. Talanta 2017, 164, 314–322. [Google Scholar] [CrossRef] [PubMed]
  220. MacLeod, S.L.; McClure, E.L.; Wong, C.S. Laboratory calibration and field deployment of the polar organic chemical integrative sampler for pharmaceuticals and personal care products in wastewater and surface water. Environ. Toxicol. Chem. Int. J. 2007, 26, 2517–2529. [Google Scholar] [CrossRef] [PubMed]
  221. Li, M.; Sun, G.; Qiu, J.; Li, A. Occurrence and variation of lipophilic shellfish toxins in phytoplankton, shellfish and seawater samples from the aquaculture zone in the Yellow Sea, China. Toxicon 2017, 127, 1–10. [Google Scholar] [CrossRef]
  222. Bartelt-Hunt, S.L.; Snow, D.D.; Damon-Powell, T.; Brown, D.L.; Prasai, G.; Schwarz, M.; Kolok, A.S. Quantitative evaluation of laboratory uptake rates for pesticides, pharmaceuticals, and steroid hormones using POCIS. Environ. Toxicol. Chem. 2011, 30, 1412–1420. [Google Scholar] [CrossRef]
  223. Jacquet, R.; Miège, C.; Bados, P.; Schiavone, S.; Coquery, M. Evaluating the polar organic chemical integrative sampler for the monitoring of beta-blockers and hormones in wastewater treatment plant effluents and receiving surface waters. Environ. Toxicol. Chem. 2012, 31, 279–288. [Google Scholar] [CrossRef]
  224. Bailly, E.; Levi, Y.; Karolak, S. Calibration and field evaluation of Polar Organic Chemical Integrative Sampler (POCIS) for monitoring pharmaceuticals in hospital wastewater. Environ. Pollut. 2013, 174, 100–105. [Google Scholar] [CrossRef]
  225. Assoumani, A.; Lissalde, S.; Margoum, C.; Mazzella, N.; Coquery, M. In situ application of stir bar sorptive extraction as a passive sampling technique for the monitoring of agricultural pesticides in surface waters. Sci. Total Environ. 2013, 463, 829–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Amdany, R.; Chimuka, L.; Cukrowska, E. Determination of naproxen, ibuprofen and triclosan in wastewater using the polar organic chemical integrative sampler (POCIS): A laboratory calibration and field application. Water SA 2014, 40, 407–414. [Google Scholar] [CrossRef] [Green Version]
  227. Brown, D.; Snow, D.; Hunt, G.A.; Bartelt-Hunt, S.L. Persistence of pharmaceuticals in effluent-dominated surface waters. J. Environ. Qual. 2015, 44, 299–304. [Google Scholar] [CrossRef] [PubMed]
  228. Skodova, A.; Prokes, R.; Simek, Z.; Vrana, B. In situ calibration of three passive samplers for the monitoring of steroid hormones in wastewater. Talanta 2016, 161, 405–412. [Google Scholar] [CrossRef] [PubMed]
  229. Ory, J.; Bricheux, G.; Togola, A.; Bonnet, J.L.; Donnadieu-Bernard, F.; Nakusi, L.; Forestier, C.; Traore, O. Ciprofloxacin residue and antibiotic-resistant biofilm bacteria in hospital effluent. Environ. Pollut. 2016, 214, 635–645. [Google Scholar] [CrossRef]
  230. Vystavna, Y.; Frkova, Z.; Marchand, L.; Vergeles, Y.; Stolberg, F. Removal efficiency of pharmaceuticals in a full scale constructed wetland in East Ukraine. Ecol. Eng. 2017, 108, 50–58. [Google Scholar] [CrossRef]
  231. Lhotský, O.; Krákorová, E.; Linhartová, L.; Křesinová, Z.; Steinová, J.; Dvořák, L.; Rodsand, T.; Filipová, A.; Kroupová, K.; Wimmerová, L. Assessment of biodegradation potential at a site contaminated by a mixture of BTEX, chlorinated pollutants and pharmaceuticals using passive sampling methods–case study. Sci. Total Environ. 2017, 607, 1451–1465. [Google Scholar] [CrossRef]
  232. Camotti Bastos, M.; Rheinheimer dos Santos, D.; Monteiro de Castro Lima, J.A.; Le Guet, T.; Santanna dos Santos, M.A.; Zanella, R.; Aubertheau, E.; Mondamert, L.; Caner, L.; Labanowski, J. Presence of anthropogenic markers in water: A case study of the Guaporé River watershed, Brazil. Clean–Soil Air Water 2018, 46, 1700019. [Google Scholar] [CrossRef]
  233. Chaves-Barquero, L.G.; Luong, K.H.; Rudy, M.D.; Frank, R.A.; Hanson, M.L.; Wong, C.S. Attenuation of pharmaceuticals, nutrients and toxicity in a rural sewage lagoon system integrated with a subsurface filtration technology. Chemosphere 2018, 209, 767–775. [Google Scholar] [CrossRef]
  234. Li, Y.; Yang, C.; Zha, D.; Wang, L.; Lu, G.; Sun, Q.; Wu, D. In situ calibration of polar organic chemical integrative samplers to monitor organophosphate flame retardants in river water using polyethersulfone membranes with performance reference compounds. Sci. Total Environ. 2018, 610–611, 1356–1363. [Google Scholar] [CrossRef]
  235. Kim, H.; Homan, M. Evaluation of pharmaceuticals and personal care products (PPCPs) in drinking water originating from Lake Erie. J. Great Lakes Res. 2020, 46, 1321–1330. [Google Scholar] [CrossRef]
  236. Li, H.; Helm, P.A.; Metcalfe, C.D. Sampling in the Great Lakes for pharmaceuticals, personal care products, and endocrine-disrupting substances using the passive polar organic chemical integrative sampler. Environ. Toxicol. Chem. Int. J. 2010, 29, 751–762. [Google Scholar] [CrossRef] [PubMed]
  237. Khalid, F.; Hassan, S.M.F.; Mushtaque, M.; Noor, R.; Ghayas, S.; Muhamma, I.N.; Hassan, F. Comparative analysis of biopharmaceutic classification system (BCS) based biowaiver protocols to validate equivalence of a multisource product. Afr. J. Pharm. Pharmacol. 2020, 14, 212–220. [Google Scholar]
  238. Guibal, R.; Lissalde, S.; Brizard, Y.; Guibaud, G. Semi-continuous pharmaceutical and human tracer monitoring by POCIS sampling at the watershed-scale in an agricultural rural headwater river. J. Hazard. Mater. 2018, 360, 106–114. [Google Scholar] [CrossRef] [PubMed]
  239. Fauvelle, V.; Mazzella, N.; Belles, A.; Moreira, A.; Allan, I.J.; Budzinski, H. Optimization of the polar organic chemical integrative sampler for the sampling of acidic and polar herbicides. Anal. Bioanal. Chem. 2014, 406, 3191–3199. [Google Scholar] [CrossRef]
  240. Müller, A.-K.; Leser, K.; Kämpfer, D.; Riegraf, C.; Crawford, S.E.; Smith, K.; Vermeirssen, E.L.; Buchinger, S.; Hollert, H. Bioavailability of estrogenic compounds from sediment in the context of flood events evaluated by passive sampling. Water Res. 2019, 161, 540–548. [Google Scholar] [CrossRef]
  241. Alvarez, D.A.; Petty, J.D.; Huckins, J.N.; Jones-Lepp, T.L.; Getting, D.T.; Goddard, J.P.; Manahan, S.E. Development of a passive, in situ, integrative sampler for hydrophilic organic contaminants in aquatic environments. Environ. Toxicol. Chem. 2004, 23, 1640–1648. [Google Scholar] [CrossRef] [Green Version]
  242. Zhang, Z.; Hibberd, A.; Zhou, J.L. Analysis of emerging contaminants in sewage effluent and river water: Comparison between spot and passive sampling. Anal. Chim. Acta 2008, 607, 37–44. [Google Scholar] [CrossRef]
  243. Cernoch, I.; Franek, M.; Diblikova, I.; Hilscherova, K.; Randak, T.; Ocelka, T.; Blaha, L. POCIS sampling in combination with ELISA: Screening of sulfonamide residues in surface and waste waters. J. Environ. Monit. 2012, 14, 250–257. [Google Scholar] [CrossRef]
  244. Tapie, N.; Devier, M.H.; Soulier, C.; Creusot, N.; Le Menach, K.; Ait-Aissa, S.; Vrana, B.; Budzinski, H. Passive samplers for chemical substance monitoring and associated toxicity assessment in water. Water Sci. Technol. 2011, 63, 2418–2426. [Google Scholar] [CrossRef]
  245. Vystavna, Y.; Huneau, F.; Grynenko, V.; Vergeles, Y.; Celle-Jeanton, H.; Tapie, N.; Budzinski, H.; Le Coustumer, P. Pharmaceuticals in rivers of two regions with contrasted socio-economic conditions: Occurrence, accumulation, and comparison for Ukraine and France. Water Air Soil Pollut. 2012, 223, 2111–2124. [Google Scholar] [CrossRef]
  246. Jones, L.; Ronan, J.; McHugh, B.; Regan, F. Passive sampling of polar emerging contaminants in Irish catchments. Water Sci. Technol. 2019, 79, 218–230. [Google Scholar] [CrossRef] [PubMed]
  247. Berho, C.; Togola, A.; Coureau, C.; Ghestem, J.P.; Amalric, L. Applicability of polar organic compound integrative samplers for monitoring pesticides in groundwater. Environ. Sci. Pollut. Res. Int. 2013, 20, 5220–5228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Mijangos, L.; Ziarrusta, H.; Prieto, A.; Zugazua, O.; Zuloaga, O.; Olivares, M.; Usobiaga, A.; Paschke, A.; Etxebarria, N. Evaluation of polar organic chemical integrative and hollow fibre samplers for the determination of a wide variety of organic polar compounds in seawater. Talanta 2018, 185, 469–476. [Google Scholar] [CrossRef] [PubMed]
  249. Jeong, Y.; Schäffer, A.; Smith, K. A comparison of equilibrium and kinetic passive sampling for the monitoring of aquatic organic contaminants in German rivers. Water Res. 2018, 145, 248–258. [Google Scholar] [CrossRef]
  250. Balaam, J.L.; Grover, D.; Johnson, A.C.; Jürgens, M.; Readman, J.; Smith, A.J.; White, S.; Williams, R.; Zhou, J.L. The use of modelling to predict levels of estrogens in a river catchment: How does modelled data compare with chemical analysis and in vitro yeast assay results? Sci. Total Environ. 2010, 408, 4826–4832. [Google Scholar] [CrossRef]
  251. Rotter, S.; Sans-Piché, F.; Streck, G.; Altenburger, R.; Schmitt-Jansen, M. Active bio-monitoring of contamination in aquatic systems—An in situ translocation experiment applying the PICT concept. Aquat. Toxicol. 2011, 101, 228–236. [Google Scholar] [CrossRef]
  252. Criquet, J.; Dumoulin, D.; Howsam, M.; Mondamert, L.; Goossens, J.F.; Prygiel, J.; Billon, G. Comparison of POCIS passive samplers vs. composite water sampling: A case study. Sci. Total Environ. 2017, 609, 982–991. [Google Scholar] [CrossRef]
  253. Carpinteiro, I.; Schopfer, A.; Estoppey, N.; Fong, C.; Grandjean, D.; de Alencastro, L.F. Evaluation of performance reference compounds (PRCs) to monitor emerging polar contaminants by polar organic chemical integrative samplers (POCIS) in rivers. Anal. Bioanal. Chem. 2016, 408, 1067–1078. [Google Scholar] [CrossRef]
  254. Kaserzon, S.L.; Vermeirssen, E.L.; Hawker, D.W.; Kennedy, K.; Bentley, C.; Thompson, J.; Booij, K.; Mueller, J.F. Passive sampling of perfluorinated chemicals in water: Flow rate effects on chemical uptake. Environ. Pollut. 2013, 177, 58–63. [Google Scholar] [CrossRef]
  255. Metcalfe, C.D.; Helm, P.; Paterson, G.; Kaltenecker, G.; Murray, C.; Nowierski, M.; Sultana, T. Pesticides related to land use in watersheds of the Great Lakes basin. Sci. Total Environ. 2019, 648, 681–692. [Google Scholar] [CrossRef] [PubMed]
  256. Lehmann, E.; Fargues, M.; Nfon Dibié, J.-J.; Konaté, Y.; de Alencastro, L.F. Assessment of water resource contamination by pesticides in vegetable-producing areas in Burkina Faso. Environ. Sci. Pollut. Res. 2018, 25, 3681–3694. [Google Scholar] [CrossRef] [PubMed]
  257. Metcalfe, C.D.; Sultana, T.; Li, H.; Helm, P.A. Current-use pesticides in urban watersheds and receiving waters of western Lake Ontario measured using polar organic chemical integrative samplers (POCIS). J. Great Lakes Res. 2016, 42, 1432–1442. [Google Scholar] [CrossRef]
  258. Writer, J.H.; Barber, L.B.; Brown, G.K.; Taylor, H.E.; Kiesling, R.L.; Ferrey, M.L.; Jahns, N.D.; Bartell, S.E.; Schoenfuss, H.L. Anthropogenic tracers, endocrine disrupting chemicals, and endocrine disruption in Minnesota lakes. Sci. Total Environ. 2010, 409, 100–111. [Google Scholar] [CrossRef] [PubMed]
  259. Bayen, S.; Segovia, E.; Loh, L.L.; Burger, D.F.; Eikaas, H.S.; Kelly, B.C. Application of Polar Organic Chemical Integrative Sampler (POCIS) to monitor emerging contaminants in tropical waters. Sci. Total Environ. 2014, 482–483, 15–22. [Google Scholar] [CrossRef] [PubMed]
  260. Brophy, M.J.; Mackie, A.L.; Park, Y.; Gagnon, G.A. Exploring the detection of microcystin-LR using polar organic chemical integrative samplers (POCIS). Environ. Sci. Process. Impacts 2019, 21, 659–666. [Google Scholar] [CrossRef]
  261. Sultana, T.; Murray, C.; Kleywegt, S.; Metcalfe, C.D. Neonicotinoid pesticides in drinking water in agricultural regions of southern Ontario, Canada. Chemosphere 2018, 202, 506–513. [Google Scholar] [CrossRef] [PubMed]
  262. Miller, T.H.; Baz-Lomba, J.A.; Harman, C.; Reid, M.J.; Owen, S.F.; Bury, N.R.; Thomas, K.V.; Barron, L.P. The first attempt at non-linear in silico prediction of sampling rates for polar organic chemical integrative samplers (POCIS). Environ. Sci. Technol. 2016, 50, 7973–7981. [Google Scholar] [CrossRef]
  263. Morrison, S.A.; Belden, J.B. Calibration of nylon organic chemical integrative samplers and sentinel samplers for quantitative measurement of pulsed aquatic exposures. J. Chromatogr. A 2016, 1449, 109–117. [Google Scholar] [CrossRef]
  264. Morrison, S.A.; Belden, J.B. Characterization of performance reference compound kinetics and analyte sampling rate corrections under three flow regimes using nylon organic chemical integrative samplers. J. Chromatogr. A 2016, 1466, 1–11. [Google Scholar] [CrossRef]
  265. Fauvelle, V.; Mazzella, N.; Delmas, F.; Madarassou, K.; Eon, M.; Budzinski, H. Use of mixed-mode ion exchange sorbent for the passive sampling of organic acids by polar organic chemical integrative sampler (POCIS). Environ. Sci. Technol. 2012, 46, 13344–13353. [Google Scholar] [CrossRef] [PubMed]
  266. Morin, N.; Camilleri, J.; Cren-Olivé, C.; Coquery, M.; Miège, C. Determination of uptake kinetics and sampling rates for 56 organic micropollutants using “pharmaceutical” POCIS. Talanta 2013, 109, 61–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Magi, E.; Di Carro, M.; Mirasole, C.; Benedetti, B. Combining passive sampling and tandem mass spectrometry for the determination of pharmaceuticals and other emerging pollutants in drinking water. Microchem. J. 2018, 136, 56–60. [Google Scholar] [CrossRef]
  268. Jaša, L.; Sadílek, J.; Kohoutek, J.; Straková, L.; Maršálek, B.; Babica, P. Application of passive sampling for sensitive time-integrative monitoring of cyanobacterial toxins microcystins in drinking water treatment plants. Water Res. 2019, 153, 108–120. [Google Scholar] [CrossRef]
  269. Vercraene-Eairmal, M.; Lauga, B.; Saint Laurent, S.; Mazzella, N.; Boutry, S.; Simon, M.; Karama, S.; Delmas, F.; Duran, R. Diuron biotransformation and its effects on biofilm bacterial community structure. Chemosphere 2010, 81, 837–843. [Google Scholar] [CrossRef] [PubMed]
  270. Fox, J.T.; Adams, G.; Sharum, M.; Steelman, K.L. Passive sampling of bioavailable organic chemicals in Perry County, Missouri cave streams. Environ. Sci. Technol. 2010, 44, 8835–8841. [Google Scholar] [CrossRef]
  271. Miège, C.; Budzinski, H.; Jacquet, R.; Soulier, C.; Pelte, T.; Coquery, M. Polar organic chemical integrative sampler (POCIS): Application for monitoring organic micropollutants in wastewater effluent and surface water. J. Environ. Monit. 2012, 14, 626–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Kaserzon, S.L.; Kennedy, K.; Hawker, D.W.; Thompson, J.; Carter, S.; Roach, A.C.; Booij, K.; Mueller, J.F. Development and calibration of a passive sampler for perfluorinated alkyl carboxylates and sulfonates in water. Environ. Sci. Technol. 2012, 46, 4985–4993. [Google Scholar] [CrossRef]
  273. Charlestra, L.; Amirbahman, A.; Courtemanch, D.L.; Alvarez, D.A.; Patterson, H. Estimating pesticide sampling rates by the polar organic chemical integrative sampler (POCIS) in the presence of natural organic matter and varying hydrodynamic conditions. Environ. Pollut. 2012, 169, 98–104. [Google Scholar] [CrossRef]
  274. Munaron, D.; Tapie, N.; Budzinski, H.; Andral, B.; Gonzalez, J.-L. Pharmaceuticals, alkylphenols and pesticides in Mediterranean coastal waters: Results from a pilot survey using passive samplers. Estuar. Coast. Shelf Sci. 2012, 114, 82–92. [Google Scholar] [CrossRef] [Green Version]
  275. Schopfer, A.; Estoppey, N.; Omlin, J.; Udrisard, R.; Esseiva, P.; de Alencastro, L.F. The Use of Passive Samplers to Reveal Industrial and Agricultural Pollution Trends in Swiss Rivers. Chimia 2014, 68, 778–782. [Google Scholar] [CrossRef] [PubMed]
  276. Tiam, S.K.; Morin, S.; Pesce, S.; Feurtet-Mazel, A.; Moreira, A.; Gonzalez, P.; Mazzella, N. Environmental effects of realistic pesticide mixtures on natural biofilm communities with different exposure histories. Sci. Total Environ. 2014, 473, 496–506. [Google Scholar] [CrossRef] [PubMed]
  277. Lissalde, S.; Mazzella, N.; Mazellier, P. Polar organic chemical integrative samplers for pesticides monitoring: Impacts of field exposure conditions. Sci. Total Environ. 2014, 488, 188–196. [Google Scholar] [CrossRef] [PubMed]
  278. Dalton, R.L.; Pick, F.R.; Boutin, C.; Saleem, A. Atrazine contamination at the watershed scale and environmental factors affecting sampling rates of the polar organic chemical integrative sampler (POCIS). Environ. Pollut. 2014, 189, 134–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Di Carro, M.; Bono, L.; Magi, E. A simple recirculating flow system for the calibration of polar organic chemical integrative samplers (POCIS): Effect of flow rate on different water pollutants. Talanta 2014, 120, 30–33. [Google Scholar] [CrossRef]
  280. Barranger, A.; Akcha, F.; Rouxel, J.; Brizard, R.; Maurouard, E.; Pallud, M.; Menard, D.; Tapie, N.; Budzinski, H.; Burgeot, T.; et al. Study of genetic damage in the Japanese oyster induced by an environmentally-relevant exposure to diuron: Evidence of vertical transmission of DNA damage. Aquat. Toxicol. 2014, 146, 93–104. [Google Scholar] [CrossRef] [Green Version]
  281. Poulier, G.; Lissalde, S.; Charriau, A.; Buzier, R.; Cleries, K.; Delmas, F.; Mazzella, N.; Guibaud, G. Estimates of pesticide concentrations and fluxes in two rivers of an extensive French multi-agricultural watershed: Application of the passive sampling strategy. Environ. Sci. Pollut. Res. 2015, 22, 8044–8057. [Google Scholar] [CrossRef]
  282. Jaimes-Correa, J.C.; Snow, D.D.; Bartelt-Hunt, S.L. Seasonal occurrence of antibiotics and a beta agonist in an agriculturally-intensive watershed. Environ. Pollut. 2015, 205, 87–96. [Google Scholar] [CrossRef]
  283. Gonzalez-Rey, M.; Tapie, N.; Le Menach, K.; Devier, M.H.; Budzinski, H.; Bebianno, M.J. Occurrence of pharmaceutical compounds and pesticides in aquatic systems. Mar. Pollut. Bull. 2015, 96, 384–400. [Google Scholar] [CrossRef]
  284. Terzopoulou, E.; Voutsa, D. Active and passive sampling for the assessment of hydrophilic organic contaminants in a river basin-ecotoxicological risk assessment. Environ. Sci. Pollut. Res. Int. 2016, 23, 5577–5591. [Google Scholar] [CrossRef]
  285. Zhang, Z.; Troldborg, M.; Yates, K.; Osprey, M.; Kerr, C.; Hallett, P.D.; Baggaley, N.; Rhind, S.M.; Dawson, J.J.C.; Hough, R.L. Evaluation of spot and passive sampling for monitoring, flux estimation and risk assessment of pesticides within the constraints of a typical regulatory monitoring scheme. Sci. Total Environ. 2016, 569–570, 1369–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Bayen, S.; Estrada, E.S.; Juhel, G.; Kit, L.W.; Kelly, B.C. Pharmaceutically active compounds and endocrine disrupting chemicals in water, sediments and mollusks in mangrove ecosystems from Singapore. Mar. Pollut. Bull. 2016, 109, 716–722. [Google Scholar] [CrossRef] [PubMed]
  287. Soulier, C.; Coureau, C.; Togola, A. Environmental forensics in groundwater coupling passive sampling and high resolution mass spectrometry for screening. Sci. Total Environ. 2016, 563–564, 845–854. [Google Scholar] [CrossRef] [PubMed]
  288. Diamond, S.R.; Sultana, T.; Servos, M.R.; Metcalfe, C.D. Biological responses to contaminants in darters (Etheostoma spp.) collected from rural and urban regions of the Grand River, ON, Canada. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2016, 199, 126–135. [Google Scholar] [CrossRef] [PubMed]
  289. Zha, D.; Li, Y.; Wang, L.; Yang, C.; Lu, G. Occurrence and attenuation of pharmaceuticals and their transformation products in rivers impacted by sewage treatment plants. RSC Adv. 2017, 7, 40905–40913. [Google Scholar] [CrossRef] [Green Version]
  290. Aisha, A.A.; Hneine, W.; Mokh, S.; Devier, M.-H.; Budzinski, H.; Jaber, F. Monitoring of 45 pesticides in Lebanese surface water using polar organic chemical integrative sampler (POCIS). Ocean Sci. J. 2017, 52, 455–466. [Google Scholar] [CrossRef]
  291. Van Metre, P.C.; Alvarez, D.A.; Mahler, B.J.; Nowell, L.; Sandstrom, M.; Moran, P. Complex mixtures of Pesticides in Midwest U.S. streams indicated by POCIS time-integrating samplers. Environ. Pollut. 2017, 220, 431–440. [Google Scholar] [CrossRef] [Green Version]
  292. Guibal, R.; Lissalde, S.; Leblanc, J.; Cleries, K.; Charriau, A.; Poulier, G.; Mazzella, N.; Rebillard, J.-P.; Brizard, Y.; Guibaud, G. Two sampling strategies for an overview of pesticide contamination in an agriculture-extensive headwater stream. Environ. Sci. Pollut. Res. 2018, 25, 14280–14293. [Google Scholar] [CrossRef]
  293. Zha, D.; Li, Y.; Yang, C.; Yao, C. Assessment of organophosphate flame retardants in surface water and sediment from a freshwater environment (Yangtze River, China). Environ. Monit. Assess. 2018, 190, 222. [Google Scholar] [CrossRef]
  294. Challis, J.K.; Cuscito, L.D.; Joudan, S.; Luong, K.H.; Knapp, C.W.; Hanson, M.L.; Wong, C.S. Inputs, source apportionment, and transboundary transport of pesticides and other polar organic contaminants along the lower Red River, Manitoba, Canada. Sci. Total Environ. 2018, 635, 803–816. [Google Scholar] [CrossRef] [Green Version]
  295. Challis, J.K.; Stroski, K.M.; Luong, K.H.; Hanson, M.L.; Wong, C.S. Field evaluation and in situ stress testing of the organic-diffusive gradients in thin-films passive sampler. Environ. Sci. Technol. 2018, 52, 12573–12582. [Google Scholar] [CrossRef] [PubMed]
  296. Fauvelle, V.; Belles, A.; Budzinski, H.; Mazzella, N.; Plus, M. Simulated conservative tracer as a proxy for S-metolachlor concentration predictions compared to POCIS measurements in Arcachon Bay. Mar. Pollut. Bull. 2018, 133, 423–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Berton, A.; Brugnera, M.F.; Dores, E. Grab and passive sampling applied to pesticide analysis in the Sao Lourenco river headwater in Campo Verde—MT, Brazil. J. Environ. Sci. Health B 2018, 53, 237–245. [Google Scholar] [CrossRef] [PubMed]
  298. Ahrens, L.; Daneshvar, A.; Lau, A.E.; Kreuger, J. Concentrations, fluxes and field calibration of passive water samplers for pesticides and hazard-based risk assessment. Sci. Total Environ. 2018, 637–638, 835–843. [Google Scholar] [CrossRef] [PubMed]
  299. Branchet, P.; Cadot, E.; Fenet, H.; Sebag, D.; Ngatcha, B.N.; Borrell-Estupina, V.; Ngoupayou, J.R.N.; Kengne, I.; Braun, J.J.; Gonzalez, C. Polar pesticide contamination of an urban and peri-urban tropical watershed affected by agricultural activities (Yaounde, Center Region, Cameroon). Environ. Sci. Pollut. Res. Int. 2018, 25, 17690–17715. [Google Scholar] [CrossRef]
  300. Yabuki, Y.; Ono, J.; Nagai, T.; Inao, K.; Tanimori, S. Determining the suitability of a polar organic chemical integrated sampler (POCIS) for the detection of pesticide residue in the Ishikawa River and its tributary in Osaka, Japan. J. Pestic. Sci. 2018, 43, 18–23. [Google Scholar] [CrossRef] [Green Version]
  301. Rico, A.; Arenas-Sánchez, A.; Alonso-Alonso, C.; López-Heras, I.; Nozal, L.; Rivas-Tabares, D.; Vighi, M. Identification of contaminants of concern in the upper Tagus river basin (central Spain). Part 1: Screening, quantitative analysis and comparison of sampling methods. Sci. Total Environ. 2019, 666, 1058–1070. [Google Scholar] [CrossRef]
  302. Arenas-Sanchez, A.; Rico, A.; Rivas-Tabares, D.; Blanco, A.; Garcia-Doncel, P.; Romero-Salas, A.; Nozal, L.; Vighi, M. Identification of contaminants of concern in the upper Tagus river basin (central Spain). Part 2: Spatio-temporal analysis and ecological risk assessment. Sci. Total Environ. 2019, 667, 222–233. [Google Scholar] [CrossRef]
  303. Iwanowicz, L.R.; Pinkney, A.; Guy, C.; Major, A.; Munney, K.; Blazer, V.S.; Alvarez, D.; Walsh, H.L.; Sperry, A.; Braham, R.; et al. Temporal evaluation of estrogenic endocrine disruption markers in smallmouth bass (Micropterus dolomieu) reveals seasonal variability in intersex. Sci. Total Environ. 2019, 646, 245–256. [Google Scholar] [CrossRef]
  304. Tousova, Z.; Vrana, B.; Smutna, M.; Novak, J.; Klucarova, V.; Grabic, R.; Slobodnik, J.; Giesy, J.P.; Hilscherova, K. Analytical and bioanalytical assessments of organic micropollutants in the Bosna River using a combination of passive sampling, bioassays and multi-residue analysis. Sci. Total Environ. 2019, 650, 1599–1612. [Google Scholar] [CrossRef]
  305. Bernard, M.; Boutry, S.; Lissalde, S.; Guibaud, G.; Saut, M.; Rebillard, J.P.; Mazzella, N. Combination of passive and grab sampling strategies improves the assessment of pesticide occurrence and contamination levels in a large-scale watershed. Sci. Total Environ. 2019, 651, 684–695. [Google Scholar] [CrossRef] [PubMed]
  306. Hayden, K.R.; Preisendanz, H.E.; Elkin, K.R.; Saleh, L.B.; Weikel, J.; Veith, T.L.; Elliott, H.A.; Watson, J.E. Comparison of POCIS and grab sampling techniques for monitoring PPCPs in vernal pools in central Pennsylvania. Sci. Total Environ. 2022, 806, 150607. [Google Scholar] [CrossRef] [PubMed]
  307. Helm, P.A.; Howell, E.T.; Li, H.L.; Metcalfe, T.M.; Chomicki, K.D.; Metcalfe, C. Influence of nearshore dynamics on the distribution of organic wastewater-associated chemicals in Lake Ontario determined using passive samplers. J. Great Lakes Res. 2012, 38, 105–115. [Google Scholar] [CrossRef]
  308. Mhadhbi, T.; Pringault, O.; Nouri, H.; Spinelli, S.; Beyrem, H.; Gonzalez, C. Evaluating polar pesticide pollution with a combined approach: A survey of agricultural practices and POCIS passive samplers in a Tunisian lagoon watershed. Environ. Sci. Pollut. Res. 2019, 26, 342–361. [Google Scholar] [CrossRef]
  309. Metcalfe, C.; Hoque, M.E.; Sultana, T.; Murray, C.; Helm, P.; Kleywegt, S. Monitoring for contaminants of emerging concern in drinking water using POCIS passive samplers. Environ. Sci. Process. Impacts 2014, 16, 473–481. [Google Scholar] [CrossRef] [PubMed]
  310. Gobelius, L.; Persson, C.; Wiberg, K.; Ahrens, L. Calibration and application of passive sampling for per- and polyfluoroalkyl substances in a drinking water treatment plant. J. Hazard. Mater. 2019, 362, 230–237. [Google Scholar] [CrossRef] [PubMed]
  311. Ibrahim, I.; Togola, A.; Gonzalez, C. In-situ calibration of POCIS for the sampling of polar pesticides and metabolites in surface water. Talanta 2013, 116, 495–500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Buzier, R.; Guibal, R.; Lissalde, S.; Guibaud, G. Limitation of flow effect on passive sampling accuracy using POCIS with the PRC approach or o-DGT: A pilot-scale evaluation for pharmaceutical compounds. Chemosphere 2019, 222, 628–636. [Google Scholar] [CrossRef]
  313. Ibrahim, I.; Togola, A.; Gonzalez, C. Polar organic chemical integrative sampler (POCIS) uptake rates for 17 polar pesticides and degradation products: Laboratory calibration. Environ. Sci. Pollut. Res. 2013, 20, 3679–3687. [Google Scholar] [CrossRef] [Green Version]
  314. Yabuki, Y.; Nagai, T.; Inao, K.; Ono, J.; Aiko, N.; Ohtsuka, N.; Tanaka, H.; Tanimori, S. Temperature dependence on the pesticide sampling rate of polar organic chemical integrative samplers (POCIS). Biosci. Biotechnol. Biochem. 2016, 80, 2069–2075. [Google Scholar] [CrossRef] [Green Version]
  315. Togola, A.; Budzinski, H. Development of polar organic integrative samplers for analysis of pharmaceuticals in aquatic systems. Anal. Chem. 2007, 79, 6734–6741. [Google Scholar] [CrossRef] [PubMed]
  316. Ahrens, L.; Daneshvar, A.; Lau, A.E.; Kreuger, J. Characterization of five passive sampling devices for monitoring of pesticides in water. J. Chromatogr. A 2015, 1405, 1–11. [Google Scholar] [CrossRef] [PubMed]
  317. Belles, A.; Pardon, P.; Budzinski, H. Development of an adapted version of polar organic chemical integrative samplers (POCIS-Nylon). Anal. Bioanal. Chem. 2014, 406, 1099–1110. [Google Scholar] [CrossRef] [PubMed]
  318. Jakubus, A.; Tyma, M.; Stepnowski, P.; Paszkiewicz, M. Application of passive sampling devices based on multi-walled carbon nanotubes for the isolation of selected pharmaceuticals and phenolic compounds in water samples–possibilities and limitations. Talanta 2017, 164, 700–707. [Google Scholar] [CrossRef]
  319. Guibal, R.; Lissalde, S.; Guibaud, G. Experimental Estimation of 44 Pharmaceutical Polar Organic Chemical Integrative Sampler Sampling Rates in an Artificial River under Various Flow Conditions. Environ. Toxicol. Chem. 2020, 39, 1186–1195. [Google Scholar] [CrossRef]
  320. Noro, K.; Endo, S.; Shikano, Y.; Banno, A.; Yabuki, Y. Development and Calibration of the Polar Organic Chemical Integrative Sampler (POCIS) for Neonicotinoid Pesticides. Environ. Toxicol. Chem. 2020, 39, 1325–1333. [Google Scholar] [CrossRef]
  321. Environmental Sampling Technologies Inc. Available online: https://www.est-lab.com/ (accessed on 27 June 2022).
Figure 1. General form of analyte uptake in passive samplers [11].
Figure 1. General form of analyte uptake in passive samplers [11].
Molecules 27 07898 g001
Scheme 1. Illustration of marine biotoxin groups and applied analytical techniques. MBA (Mouse Bio Assay) and RBA (Rat Bio Assay).
Scheme 1. Illustration of marine biotoxin groups and applied analytical techniques. MBA (Mouse Bio Assay) and RBA (Rat Bio Assay).
Molecules 27 07898 sch001
Table 3. Extraction steps for most common resins.
Table 3. Extraction steps for most common resins.
SPATT ResinToxin GroupsEluteReferences (Refs.)
DIAION HP20, SEPABEADS SP825L, SP850 and SEPABEADS SP700Microcystins50% MeOH[6,165]
DIAION HP20, SP 207, HP2MGPTX, PTX 2 SA, PTX 11, PT11 SA, OA, OA-ester, YTXMeOH[5,156]
DIAION HP20DA, CTXsMeOH[5,171,177]
Ammonium acetate in 50% MeOH
Ammonium acetate in 50% MeOH
DIAION HP20Cyanotoxins, Okadaic acid, Saxitoxin and related PSTsMeOH[156]
Table 4. Extraction steps for modified resins [2,156].
Table 4. Extraction steps for modified resins [2,156].
SPATT PhasesElute 1Elute 2
Strata-X (200 mg)100% MeOH-
PAC G-60 (200 mg)5% Formic acid in 70% MeOH5% conc. HCL in MeOH (6 mL)
AG 50W-X4 (200 mg)5% NH3 in MeOH (5 mL)-
Amberlite IRP-64 (200 mg)5% Formic acid in 70% MeOH-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kamali, N.; Abbas, F.; Lehane, M.; Griew, M.; Furey, A. A Review of In Situ Methods—Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) for the Collection and Concentration of Marine Biotoxins and Pharmaceuticals in Environmental Waters. Molecules 2022, 27, 7898. https://doi.org/10.3390/molecules27227898

AMA Style

Kamali N, Abbas F, Lehane M, Griew M, Furey A. A Review of In Situ Methods—Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) for the Collection and Concentration of Marine Biotoxins and Pharmaceuticals in Environmental Waters. Molecules. 2022; 27(22):7898. https://doi.org/10.3390/molecules27227898

Chicago/Turabian Style

Kamali, Naghmeh, Feras Abbas, Mary Lehane, Michael Griew, and Ambrose Furey. 2022. "A Review of In Situ Methods—Solid Phase Adsorption Toxin Tracking (SPATT) and Polar Organic Chemical Integrative Sampler (POCIS) for the Collection and Concentration of Marine Biotoxins and Pharmaceuticals in Environmental Waters" Molecules 27, no. 22: 7898. https://doi.org/10.3390/molecules27227898

Article Metrics

Back to TopTop