Skip to main content

Occurrence of carcinogenic illudane glycosides in drinking water wells

Abstract

Background

Ptaquiloside (PTA), caudatoside (CAU) and ptesculentoside (PTE) are carcinogenic illudane glycosides found in bracken ferns (Pteridium spp.) world-wide. The environmentally mobile PTA entails both acute and chronic toxicity. A comparable risk might be associated with the structurally similar CAU and PTE. It is of great health concern if these compounds are present in drinking water, however, it is currently unknown if these compounds can detected in wells in bracken-dominated regions. This study investigates the presence of PTA, CAU, PTE, and their corresponding hydrolysis products pterosins B (PtB), A (PtA) and G (PtG) in water wells in Denmark, Sweden and Spain. Water samples from a total of 77 deep groundwater wells (40–100 m) and shallow water wells (8–40 m) were collected and preserved in the field, pre-concentrated in the laboratory and analysed by liquid chromatography–mass spectrometry (LC–MS).

Results

Deep groundwater wells contained neither illudane glycosides nor their pterosins. However, seven private shallow wells contained at least one of the illudane glycosides and/or pterosins at concentrations up to 0.27 µg L−1 (PTA), 0.75 µg L−1 (CAU), 0.05 µg L−1 (PtB), 0.03 µg L−1 (PtA) and 0.28 µg L−1 (PtG). This is the first finding of illudane glycosides and pterosins in drinking water wells.

Conclusions

Detected concentrations of illudane glycosides in some of investigated wells exceeded the suggested maximum tolerable concentrations of PTA, although they were used for drinking water purpose. Contaminated wells were shallow with neutral pH and lower electric conductivity compared to deep groundwater wells with no illudane glycosides nor pterosins.

Background

Anthropogenic chemicals, such as pesticides and pharmaceuticals, are of high concern as water contaminants because of their persistence and toxicity [1, 2]. However, living organisms can also produce hazardous substances, which are both mobile, persistent and toxic [3]. Plants synthesise a broad variety of toxic chemical compounds, e.g. as a mean of protection against pathogens and herbivores [4]. Such plant-produced compounds are typically not harmful to the plants themselves but can adversely affect human or animal health through exposure via air, soil, water or food. By their physicochemical properties, many of them can be compared to the environmentally persistent and mobile organic compounds (PMOCs), which are often not easily removed from water due to their high water solubility and having half-lives of weeks to months in the environment [5]. Plant-produced toxins have a potential to pose a higher risk to water quality than anthropogenic chemicals because of their continuous production and comparatively high concentration in plants [6]. Nonetheless, they are still an unexplored group of emerging contaminants and we currently lack the tools to monitor, regulate and remove them.

One of the best-studied phytotoxins with PMOC characteristics under environmental conditions is the carcinogen ptaquiloside (PTA) from bracken fern. Bracken (Pteridium spp.) is globally widespread and is highly abundant in many parts of the world. It has an extremely high rate of reproduction and re-growth due to its extensive rhizomes and can rapidly adapt to new ecological conditions [6,7,8,9]. Bracken is acutely toxic and classified as possibly carcinogenic towards humans by the International Agency of Research on Cancer/WHO [10, 11]. Intake of toxins takes place via bracken-based food in some countries such as Korea, China and Japan. In addition, contamination of drinking water is proposed as a vector for human exposure [12].

PTA is considered the principle compound responsible for both acute and chronic bracken toxicities [13]. PTA is an illudane glycoside with carcinogenic, mutagenic and clastogenic properties and the most studied compound found in bracken [14, 15]. Under alkaline conditions, PTA (1) is converted into an unstable compound, a dienone (2) which subsequently leads to the formation of non-carcinogenic pterosin B (3) (Fig. 1). The dienone (2) is assumed to be the ultimate agent responsible for carcinogenicity with a cyclopropyl group being responsible for reaction with amino acids and DNA, and causing mutations [9, 16, 17]. Other compounds with structural similarity to PTA have also been found in bracken (PTA analogues) with caudatoside (CAU) and ptesculentoside (PTE) being prevalent [18, 19] (Fig. 1). Although there is only limited knowledge about the toxic properties of these compounds, they share the reactive cyclopropyl group and are expected to show similar reactivity and toxic effects as PTA [20].

Fig. 1
figure 1

Hydrolysis of PTA (1) to PtB (3) and PTA analogues CAU and PTE (1a) to the corresponding pterosins (2a). The reactive cyclopropyl group is marked with a circle [20, 32]

PTA is a highly polar compound (logP = − 1.3; Fig. 1) and can be easily transferred to the soil by release from living or dead bracken [21, 22]. Due to the high water solubility of PTA and its low sorption to soil, it is prone to leaching from topsoil to surface and groundwater [21]-[23]. Hydrolysis rates of PTA are highly pH dependent, with the lowest degradation rates at neutral to slightly acidic conditions (pH 4.5–6.5) and at low temperatures [13, 21, 24]. PTA is prone to microbial degradation, which makes PTA hydrolysis slow in environments with low microbiological activity (e.g. in groundwater). Hence, PTA presence in groundwater should be governed by pH and temperature and PTA can sustain for months in cold water under neutral to slightly acid conditions [24,25,26]. PTA has already been detected in different water bodies. Concentrations of 0.6 µg L−1 have been reported in spring water wells in Ireland [27], up to 0.09 μg L−1 in shallow groundwater (0.1–2.5 m) in Denmark [28], and up to 2.2 µg L−1 in stream water in United Kingdom [29]. In contrast to PTA, occurrence of CAU and PTE in the aqueous environment has been much less investigated. A recent study reported the first occurrence of CAU and PTE in surface waters up to 0.07 and 5.3 µg L−1, respectively [30]. This finding suggests that assessments of water toxicity based on PTA only were underestimated. CAU and PTE are slightly more polar compounds and as stable as PTA in surface waters [30].

Recent attempt of modelling PTA fate in plant–soil matrix under Danish/South-Scandinavian conditions indicates that intense precipitation events during cold periods and with a fully developed canopy induce high probability of significant PTA leaching. Furthermore, predicted high concentration of PTA in groundwater is also attributed to macropore transport in soils, sediments and bedrock [33]. Macropores have proven to be important for fast leaching of organic contaminants such as pesticides, bypassing biologically active layers, and thus, resulting in low extent of biodegradation [34]. Leaching of pesticides is pronounced in fractured clayey till soils compared to sandy soils due to rapid well-connected macropores [35]. Hence, transport of organic contaminants from upper soil layers into deep groundwater is frequently observed for a variety of pesticides and herbicides [35,36,37]. It has never previously been investigated whether illudane glycosides are mobile and stable enough, to reach deep aquifers used for drinking water supply.

In Denmark, groundwater-based drinking water utilities mainly apply deep groundwater abstraction wells, and subsequently, water is treated at waterworks before distribution to the consumers. In contrast to this, in some areas, the consumers rely on private water wells, typically shallower, and the raw water might not be treated prior to consumption. While deep groundwater wells usually abstract the water from confined aquifers, shallow water wells typically abstract water from unconfined aquifers and may be affected by seepage water. Several studies suggest that shallow wells are more prone to contamination by organic contaminants [38, 39]. Pesticides are found in higher concentration in Danish shallow water wells as compared to deep groundwater wells [39]. Therefore, shallow water wells might also be more vulnerable to contamination by illudane glucosides.

It is currently unknown whether the bracken illudane glycosides PTA, CAU and PTE pose a threat to groundwater-based drinking water supplies. Thus, it is important to: (1) investigate if the illudane glycosides are found in wells used for drinking water production, and subsequently, (2) identify the characteristics of vulnerable wells. To do so, we studied the presence of PTA, CAU and PTE and their corresponding pterosins PtB, PtA and PtG in a variety of water wells in bracken-dominated areas in Denmark, Sweden and Spain, as function of well depth and water chemistry (pH and electric conductivity). This report is the first finding of illudane glycosides and their pterosins in drinking water wells.

Materials and methods

Solvents, chemicals and resins

Acids, bases and buffers (glacial acetic acid, formic and hydrochloric acids, sodium hydroxide and ammonium acetate) were all of analytical grade from Sigma-Aldrich (Germany). LC–MS grade acetonitrile was obtained from Merck Millipore (LiChrosolv hypergrade for LC–MS, Germany) and LC–MS grade methanol was purchased from Honeywell (LC–MS Chromasolv, Germany). MilliQ water (electrical resistivity 18.2 MΩ cm, TOC less than µg L−1) was produced by Sartorius Ultrapure water system (Sartorius Stedim Biotech GmbH, Germany). Polyamide for column chromatography was from Fluka Analytical, Sigma-Aldrich Co (Germany). Oasis MAX (20 cc, 60 mg Sorbent, 30 mm particle size) was purchased from Waters (Milford, USA).

Preparation of analytical standards

No certified reference materials of PTA, CAU and PTE exist, and therefore, analytical standards were produced in-house. They were prepared from bracken plant material by the method described by Kisielius et al. [40].

Analytical procedure

Prior to LC–MS analyses, all water samples were pre-concentrated by solid-phase extraction (SPE) by a factor 250. The SPE method used in this study was optimised for PTA and PtB and validated for various groundwater samples [41]. A SPE protocol with more details is provided in Additional file 1: Figure S2 of the supplementary material (SM). Together with each set of SPE samples, positive (for PTA and PtB spiked at concentration of 0.5 µg L−1) and negative controls (DI water) were assessed at the same time. To determine the efficiency of the SPE method for CAU, PTE, PtA and PtG, 50 mL of MilliQ water was spiked to a concentration of 0.5 µg L−1 for each compound. Following the SPE protocol, samples were processed in separate SPE cartridges and analysed by LC–MS (n = 2). The recovery results for each compound are presented in Table 1.

Table 1 Recovery of illudane glycosides and corresponding pterosins in MilliQ spiked samples. OASIS MAX column (60 mg)

The samples were analysed using Agilent 1260 Infinity HPLC System equipped with Agilent 6130 Single Quadrupole mass spectrometer by the method described by Kisielius et al. [40]. Chromatographic separation was performed using Agilent Poroshell 120 EC-C18 column and the eluent comprised water (eluent A) and acetonitrile (eluent B), both with 0.1% v/v formic acid at gradient elution mode. The linear range of quantification was 20–500 µg L−1 for PTA and PtB, and 10–250 for the remaining compounds (r > 0.999). The instrumental limits of detection (LOD) and quantification (LOQ) were: 0.22–0.68 µg L−1 (PTA), 0.26–0.78 µg L−1 (CAU) and 0.08–0.25 µg L−1 (PTE) respectively, and for pterosins, 0.03–0.09 µg L−1 (PtB), 0.02–0.05 µg L−1 (PtA) and 0.01–0.03 µg L−1 (PtG). The LOD of the entire method including pre-concentration and instrumental LOD was ≤ 0.001 µg L−1 for all illudane glycosides and pterosins.

For all samples with reported presence of illudane glycosides or pterosins, a qualitative confirmation protocol was used. Three levels of conformation were applied: (a) retention time of the compounds (always compared with retention time of the standards); (b) spiking test for suspected positive samples (spiked with corresponding compound that was detected, which resulted in appearance of only one larger peak in the chromatogram); (c) the presence of confirmation ions extracted from the total ion chromatogram, e.g. [M-glucose-H2O + H]+, [M-glucose + H]+, [M + Na]+, [M + K]+ reported by Kisielius et al. [40]. All samples with reported presence of illudane glycosides or pterosins were confirmed with spiking test and accurate retention times of the analytes. Furthermore, at least one of the confirmation ions was detected. More details on spiking test are provided in Additional file 1: Figure S1. The concentrations of the compounds are reported as average concentrations measured in duplicates. In case the compound could be quantified in only one duplicate (e.g. below LOQ), the concentration was reported as a trace.

Identity of illudane glycosides and pterosins in bracken

Bracken plant material was collected from 11 locations during the water sampling campaign to quantify the content of illudane glycosides (Table 2; Fig. 2). From each location, 40 cm tip of a fully matured frond was collected in duplicates. Plant materials were dried in an ordinary plant press and grinded to powder prior to extraction according to Kisielius et al. [40]. Aqueous extracts were passed through a filter vial with 0.2 μm pore size membrane (Syringeless filter device Mini-Uniprep, GE Healthcare Life Sciences UK) prior to LC–MS analyses by which the contents of PTA, CAU and PTE and the pterosins were quantified using in-house produced standards [40].

Table 2 Measured content of illudane glycosides and corresponding pterosins in bracken ferns
Fig. 2
figure 2

The investigated water wells in Denmark, Sweden and Spain. All sampling locations were associated with adjacently growing bracken ferns. Map is produced by Qgis 3.4 Madeira and background map is ESRI Shaded Relief

Origin of water samples

Water samples from 77 wells in Denmark, Sweden and Spain were collected for analyses of PTA, CAU and PTE and the corresponding pterosins between June and October 2019 (Fig. 2). Water chemistry (pH and EC), well depth and more details on each investigated location are provided in Additional file 1: Table S1. The majority of investigated wells belong to HOFOR utility (44 locations), the largest drinking water utility in Denmark, that is providing drinking water to approximately 1 million customers in the Greater Copenhagen area. The remaining investigated locations were selected based on the presence of dense bracken ferns in their vicinity, diverse depth of water wells, geographical area and water type. To identify water wells with these features, an advertising campaign was performed through different public media prior to monitoring. Raw water (with no treatment) was abstracted from single wells. In case of HOFOR wells, mixed groundwater was collected (raw water mixed from several single wells). Only one location (number 9) does not represent a water well but is seepage water collected from a creek (3–4 m deep) just after a rain event. Wells 68 and 69 were sampled on two occasions, summer and fall 2019.

Most of the monitored wells serve for drinking water purpose. The 44 groundwater wells operated by HOFOR utility included in the study were deep groundwater wells (40–100 m). The remaining 33 wells were less than 40 m deep. In the following text, HOFOR wells are classified as deep wells (44) and the others as shallow wells (33).

Water sample collection

Water samples were collected and preserved in the field according to the method of Skrbic et al. [41]. The preservation protocol was originally developed for PTA and PtB, but was further validated for CAU and PTE and their pterosins in this study. 50 mL groundwater was collected in amber glass bottles, buffered in the field by adding between 0.05 and 1 mL of 0.5 M ammonium acetate adjusted to pH 5 with glacial acid for stabilising the PTA against hydrolysis. Samples were transported to the lab without ice for transportation up to 2 h (otherwise, placed on ice and transported within 4 h) and stored at 4 °C. SPE was performed in the lab within 72 h after sampling. Final samples were stored in vials at − 18 °C until LC–MS analyses. Filtration with sterile syringe through 0.45 μm pore size filters was performed only on two occasions, when very shallow/murky water was collected (locations 9 and 62). The samples were collected in duplicates, except for the HOFOR samples that were collected as single samples. The field blanks (DI water) were collected and treated as field samples during transport, preparation and analysis. The samples from Spain were packed with dry ice and delivered to Denmark by express mail (within 2 days).

Results

Identity of illudane glycosides and pterosins in bracken

The most commonly found illudane glycoside in bracken plant material was PTA, which was detected at all eleven sampled locations, while at eight locations CAU could also be detected (Table 2). The highest PTA concentration was measured at location 12 (3.23 mg g−1 dry weight) and the highest CAU content was detected at location 66 (0.42 mg g−1 dry weight). Neither PTE nor its degradation product PtG was detected in any of the analysed plant samples.

Leaching of illudane glycosides after a rain event

The presence of illudane glycosides and the corresponding pterosins was determined in seepage water collected just after a rain event in a bracken-dominated region (location 9). At this location, both PTA (1.27 µg L−1) and CAU (0.28 µg L−1) and their hydrolysis products were detected (Table 3).

Table 3 Water wells (site number 62–77) and seepage water (site number 9) with the presence of illudane glycosides and/or corresponding pterosins

Water samples

We sampled 77 water wells in Denmark (64), Sweden (5) and Spain (8) representing different geological settings. Neither bracken illudane glycosides (PTA, CAU and PTE) nor their hydrolysis products (PtB, PtA and PtG) were detected in 70 of the investigated wells (91%). However, in seven wells (9%), at least one of the compounds was detected (Fig. 2).

The deep groundwater wells did not contain illudane glycoside or pterosins. All seven wells in which at least one of the compounds was detected were shallow wells (Fig. 3). In terms of geography, one of the shallow wells with the presence of illudane glycosides and/or pterosins was located in Denmark, four in Sweden and two in Spain (Fig. 2). Apart from well 62, all these wells serve for drinking water purpose for humans and/or livestock (Table 3).

Fig. 3
figure 3

Number of deep and shallow wells with and without illudane glycosides and/or pterosins

Three shallow wells contained CAU in concentrations up to 0.75 µg L−1 (Table 3). However, PtA, the degradation product of CAU, occurred in six out of the seven wells, indicating even more frequent presence of CAU. PTA was present in three wells up to 0.27 µg L−1, while two wells contained PtB. PTE was not detected in any of the wells, while its degradation product PtG was detected in three of them.

In wells 62, 68 and 70, we detected both CAU and PtA. Similarly, PTA and PtB were detected in well 77. Detection of glycosides and corresponding pterosin in the same well validates the presence of glycosides. Well 62 was very shallow and not used for drinking purpose, but contained CAU and PtA at concentration up to 0.03 µg L−1. The presence of PTA has previously been reported for this location [41]. Wells 72 and 77 showed the presence of PtA, PtB and trace levels of PTA. The depth of these wells is unknown.

To include seasonal variations, we sampled wells 68 and 69 both during summer and fall. In the summer, location 68 contained both CAU and PtA, while only PtA was found in the fall. At location 69, PtB and PtA were detected in the summer, but again only PtA appeared in the fall.

Characteristics of the water wells in relation to the presence of illudane glycosides

To investigate determining factors for the presence of illudane glycosides in water wells, we analysed three characteristics of the sampled wells: depth, electrical conductivity (EC) and pH. First, for the subset of wells with known depths (59 out of 77 wells), the shallow wells with the presence of illudane glycosides and/or pterosins were on average 19 m deep (three wells with known depths; Table 3). Shallow wells without detected compounds were on average 30 m deep, while deep wells were 40 + m.

Second, EC in the deep wells were 940 ± 80 µS cm−1, while the conductivity in the shallow wells were 516 ± 79 µS cm−1, showing different water chemistry in the deep versus the shallow wells (p value <  < 0.001)  (Fig. 4a). There was no statistically significant difference in-between the shallow wells containing illudane glycosides and/or pterosins (465 ± 181 µS cm−1) and the other shallow wells (530 ± 89 µS cm−1).

There was no statistically significant difference in pH between the different groups of wells all being around pH 7, where illudane glycosides are most stable (Fig. 4b).

Fig. 4
figure 4

Deep and shallow wells with and without illudane glycoside and/or pterosins in relation to water pH (a) and EC (b). The error bars represent 95% confidence interval

Discussion

The presence of illudane glycosides in drinking water wells and bracken

We have detected illudane glycosides and their corresponding hydrolysis products in drinking water wells for the first time. Interestingly, CAU occurred more frequently than PTA, even though PTA is considered the most abundant carcinogen in bracken [43, 44]. Based on estimated LogP values and order of chromatographic elution, CAU is observed to be more polar than PTA (Fig. 1) [31, 40], and is, thus, more prone to be mobile in the environment, and potentially end up in water recipients. In a recent study, concentrations of CAU and PTE in surface water were also found to be similar or higher than PTA [30]. In addition, the same study showed that content of CAU in bracken was higher in Sweden than in Denmark [30], which could explain higher CAU occurrence in wells in Sweden in this study.

We did not detect PTE and PtG in any of the collected bracken plant samples, and in accordance with this, none of the water samples contained PTE. In Northern Europe PTE has only been detected rarely in bracken and in very low concentrations compared to PTA and CAU [30]. Nevertheless, we detected PTE’s degradation product PtG in three wells, indicating prior presence of PTE or yet unknown soil transformations. The content of illudane glycosides in bracken is decreasing from June to November while transformations into pterosins occurs [27, 30]. Hence, pterosins might still be present in water sampled in autumn while the glycosides are not present anymore or with contents below LOD. The content of PTE in bracken populations is highly variable and with a random distribution [30]. Hence, plant materials collected for this study might not represent the real PTE occurrence.

Maximum measured illudane glycoside concentrations in drinking water wells in this study were 0.27 µg L−1 for PTA and 0.75 µg L−1 for CAU. The maximum estimated tolerable concentration of PTA in drinking water is 0.002 µg L−1 (based on a model with one cancer incidence per million) [45], and the PTA analogue CAU, is presumably equally toxic [20]. The tolerable concentration were, thus, exceeded with more than a factor of 100 in these specific drinking water samples, indicating a health threat to humans or animals who are supplied with this drinking water.

Leaching of illudane glycosides after a rain event—pulse effects

Both PTA (1.27 µg L−1) and CAU (0.28 µg L−1) as well as their hydrolysis products were detected in fresh seepage water extruding from clayey sand soil in a small creek (location 9). The sampling took place immediately after a rain event [27, 30]. The finding demonstrates the potential of bracken toxins to pass through the topsoil in line with previous findings [27, 29, 41]. Although the study was not designed to cover precipitation-related pulses, it should be noticed that it rained just prior to sampling well 66, 68 and 69 in Sweden, all being positive for illudane glycosides and/or pterosins (Table 3). The presence of fissured bedrock and/or fractured till could facilitate fast transport of CAU to deeper layers and explain CAU presence in a 40 m deep well (location 66 and 68) [33]. Such conditions are expected to be more likely for shallow wells compared to the deep wells.

Characteristics of the water wells in relation to the presence of illudane glycosides

Water wells with the presence of illudane glycosides and/or pterosins in this study shared several features. They were mostly privately owned, shallow and with neutral pH and EC lower than the deep wells. Electrical conductivity is typically higher in deep groundwater [46]. We observed significantly lower EC in the shallow wells indicating connection with rain and surface water (i.e. the presence of seepage water). In addition, water in shallow wells (0–40 m) that contain illudane glycosides and/or pterosins had slightly lower average pH in comparison to deep groundwater wells (40–100 m). Lower pH (pH 4.5–6.5) is known to have stabilising effect on PTA hydrolysis [24]. Furthermore, a very short distance to high bracken biomass (< 10 m) and rain event prior to sampling supported illudane glycoside occurrence.

Six out of seven wells with the presence of illudane glycosides serve for drinking water supply. In Denmark, there are 55,000 similar private single-abstraction drinking water wells [47] and it is estimated that that 1 in 10 citizens of the EU receives drinking water from small and local systems, including private wells [48]. These systems are more vulnerable to illudane glycoside contamination and might not have any sort of water treatment.

Illudane glycosides and their hydrolysis products were not detected in deep groundwater wells investigated in this study (40–100 m). This complies with the fact that PTA’s half-life in alkaline water is shorter than under neutral-acid conditions, e.g. the half-life of PTA at 8 °C is approx. 100 days at pH 6, 30 days at pH 7 but only 3 days at pH 8 [24], indicating PTA is not sufficiently stable to reach deep groundwater aquifers as travel times from soil surface to deep aquifers are 25–100 years [46]. In contrast, shallow wells investigated in this study showed the presence of illudane glycosides and/or pterosins. Shallow wells are easily affected by younger seepage water that is more vulnerable to illudane glycosides contamination. The presence of cracks or macropores in soil profiles of shallow wells could support fast transport of illudane glycosides in relation to pulses [35, 49]. Similar observations were noticed when leaching of pesticides was investigated in clayey tills with up to 5–6 m deep macropores [50]. In that study, leaching of pesticides into shallow groundwater was noticed and the transport occurred in macropores and was driven by precipitation events [50]. Hence, leaching of illudane glycosides to shallow drinking water wells is possible, it is site specific and likely occurs in pulses.

Conclusions

In this study, we detected at least one illudane glycoside or pterosin from bracken fern in seven of the 77 studied wells. Thereby we demonstrated that leaching of illudane glycosides to drinking water wells is possible and could pose a threat to human health. Deep groundwater wells did not contain bracken illudane glycosides, probably due to the long travelling time to these aquifers. However, we detected bracken illudane glycosides in shallow wells, which could be due to intrusion of seepage water. A dense bracken biomass aboveground is a risk factor, especially in relation to rain events.

Six out of seven wells with presence of illudane glycoside were used for drinking water supply. The concentrations in these drinking water wells were 0.27–0.75 μg L−1, which violate the maximum estimated tolerable PTA concentration in drinking water by approximately 100–300 fold. In this study, CAU was more frequently detected in drinking water wells than PTA. Hence risk assessments are likely to underestimate water toxicity if they are based on PTA only and should also include CAU and PTE. These results are of great importance for consumers supplied by shallow wells in bracken dominated regions.

Availability of data and materials

All data generated or analysed during this study are included in this published article and its additional information file.

Abbreviations

PTA:

Ptaquiloside

CAU:

Caudatoside

PTE:

Ptesculentoside

PtB:

Pterosin B

PtA:

Pterosin A

PtG:

Pterosin G

PMOC:

Persistent and mobile organic compounds

WHO:

World Health Organization

DNA:

Deoxyribonucleic acid

Kow:

Octanol–water partition coefficient

LC–MS:

Liquid chromatography–mass spectrometry

SPE:

Solid-phase extraction

LOD:

Limit of detection

LOQ:

Limit of quantification

dw:

Dry weight

EC:

Electrical conductivity

DK:

Denmark

SE:

Sweden

ES:

Spain

n.a:

Not available

CS:

Clayey sand soil

SC:

Sandy clay soil

BR:

Bedrock

SS:

Sandstone

Drink:

Drinking water

Tech:

Technical water

day b.s.:

Before sampling

EU:

European Union

References

  1. Rasheed T, Bilal M, Nabeel F, Adeel M, Iqbal HMN (2019) Environmentally-related contaminants of high concern: potential sources and analytical modalities for detection, quantification, and treatment. Environ Int 122:52–66

    Article  CAS  Google Scholar 

  2. Luo Y, Guo W, Ngo HH, Nghiem LD, Hai FI, Zhang J, Liang S, Wang XC (2014) A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment. Sci Total Environ 473–474:619–641

    Article  Google Scholar 

  3. Teuscher E, Lindequist U (1994) Biogene Gifte, 2nd edn. Gustav Fischer, Stuttgart-Jena-New York

    Google Scholar 

  4. Biere A, Marak HB, van Damme JMM (2004) Plant chemical defense against herbivores and pathogens: generalized defense or trade-offs? Oecologia 140:430–441

    Article  Google Scholar 

  5. Reemtsma T, Berger U, Arp HPH, Gallard H, Knepper TP, Neumann M, Quintana JB, de Voogt P (2016) Mind the Gap: Persistent and Mobile Organic Compounds—Water Contaminants That Slip Through. Environ Sci Technol 50:10308–10315

    Article  CAS  Google Scholar 

  6. Günthardt BF, Hollender J, Hungerbühler K, Scheringer M, Bucheli TD (2018) Comprehensive Toxic Plants-Phytotoxins Database and Its Application in Assessing Aquatic Micropollution Potential. J Agric Food Chem 66:7577–7588

    Article  Google Scholar 

  7. Wu JS, Clauson-Kaas F, Lindqvist DN, Rasmussen LH, Strobel BW, Hansen HCB (2021) Does the natural carcinogen ptaquiloside degrade readily in groundwater? Environ Sci Eur 33:24. https://doi.org/10.1186/s12302-021-00468-0

    Article  CAS  Google Scholar 

  8. Thomson JA (2004) Towards a taxonomic revision of Pteridium (Dennstaedtiaceae). Telopea 10:793–803

    Google Scholar 

  9. da Costa RMG, Bastos M, Oliveira PA, Lopes C (2012) Bracken-associated human and animal health hazards: Chemical, biological and pathological evidence. J Hazard Mater 203:1–12

    Article  Google Scholar 

  10. International Agency for Research on Cancer (IARC) - Summaries & Evaluations, Bracken fern (Pteridium aquilinum) and some of its constituents, 1998, http://www.inchem.org/documents/iarc/vol40/brackenfern.html Accessed 15 July 2020.

  11. Agents Classified by the IARC Monographs, 2018, https://monographs.iarc.fr/list-of-classifications-volumes/ Accessed 15 July 2020.

  12. Galpin OP, Whitaker CJ, Whitaker RH, Kassab JY (1990) Gastric cancer in Gwynedd. Possible links with bracken. Br J Cancer 61:737–740

    Article  CAS  Google Scholar 

  13. Yamada K, Ojika M, Kigoshi H (2007) Ptaquiloside, the major toxin of bracken, and related terpene glycosides: chemistry, biology and ecology. Nat Prod Rep 24:798–813

    Article  CAS  Google Scholar 

  14. Niwa H, Ojika M, Wakamatsu K, Yamada K, Ohba S, Saito Y, Hirono I, Matsushita K (1983) Stereochemistry of ptaquiloside, a novel norsesquiterpene glucoside from bracken Pteridium aquilinum var latiusculum. Tetrahedron Lett 24:5371–5372

    Article  CAS  Google Scholar 

  15. Van der Hoeven JCM, Lagerweij WJ, Posthumus MA, Van Veldhuizen A, Holterman HAJ (1983) Aquilide A, a new mutagenic compound isolated from bracken fern (Pteridium aquilinum (L.) Kuhn). Carcinogenesis 4:1587–1590

    Article  Google Scholar 

  16. Kushida T, Uesugi M, Sugiura Y, Kigoshi H, Tanaka H, Hirokawa J, Ojika M, Yamada K (1994) DNA damage by ptaquiloside, a potent bracken carcinogen: Detection of selective strand breaks and identification of DNA cleavage products. J Am Chem Soc 116:479–486

    Article  CAS  Google Scholar 

  17. Gomes J, Magalhães A, Michel V, Amado IF, Aranha P, Ovesen RG, Hansen HCB, Gärtner F, Reis CA, Touati E (2012) Pteridium aquilinum and its ptaquiloside toxin induce DNA damage response in gastric epithelial cells, a link with gastric carcinogenesis. Toxicol Sci 126:60–71

    Article  CAS  Google Scholar 

  18. Fletcher MT, Hayes PY, Somerville MJ, De Voss JJ (2010) Ptesculentoside, a novel norsesquiterpene glucoside from the Australian bracken fern Pteridium esculentum. Tetrahedron Lett 51:1997–1999

    Article  CAS  Google Scholar 

  19. Castillo UF, Wilkins AL, Lauren DR, Smith BL, Towers NR, Alonso-Amelot ME, Jaimes-Espinoza R (1997) Isoptaquiloside and caudatoside, illudane-type sesquiterpene glucosides from Pteridium aquilinum var. caudatum. Phytochemistry 44:901–906

    Article  CAS  Google Scholar 

  20. Fletcher MT, Brock IJ, Reichmann KG, McKenzie RA, Blaney BJ (2011) Norsesquiterpene Glycosides in Bracken Ferns (Pteridium esculentum and Pteridium aquilinum subsp. wightianum) from Eastern Australia: Reassessed Poisoning Risk to Animals. J Agric Food Chem 59:5133–5138

    Article  CAS  Google Scholar 

  21. Rasmussen LH, Hansen HCB, Lauren D (2005) Sorption, degradation and mobility of ptaquiloside, a carcinogenic Bracken (Pteridium sp.) constituent, in the soil environment. Chemosphere 58:823–835

    Article  CAS  Google Scholar 

  22. Rasmussen LH, Kroghsbo S, Frisvad JC, Hansen HCB (2003) Occurrence of the carcinogenic Bracken constituent ptaquiloside in fronds, topsoils and organic soil layers in Denmark. Chemosphere 51:117–127

    Article  CAS  Google Scholar 

  23. Rasmussen LH, Jensen LS, Hansen HCB (2003) Distribution of the carcinogenic terpene ptaquiloside in bracken fronds, rhizomes (Pteridiumaquilinum), and litter in Denmark. J Chem Ecol 29:771–778

    Article  CAS  Google Scholar 

  24. Ayala-Luis KB, Hansen PB, Rasmussen LH, Hansen HCB (2006) Kinetics of ptaquiloside hydrolysis in aqueous solution. Environ Toxicol Chem An Int J 25:2623–2629

    Article  CAS  Google Scholar 

  25. Clauson-Kaas F (2016) Environmental Fate and Analysis of Ptaquiloside from the Bracken Fern. PhD thesis. University of Copenhagen.

  26. Engel P, Brandt KK, Rasmussen LH, Ovesen RG, Sørensen J (2007) Microbial degradation and impact of Bracken toxin ptaquiloside on microbial communities in soil. Chemosphere 67:202–209

    Article  CAS  Google Scholar 

  27. O’Driscoll C, Ramwell C, Harhen B, Morrison L, Clauson-Kaas F, Hansen HCB, Campbell G, Sheahan J, Misstear B, Xiao L (2016) Ptaquiloside in Irish bracken ferns and receiving waters, with implications for land managers. Molecules 21:543

    Article  Google Scholar 

  28. Clauson-Kaas F, Jensen PH, Jacobsen OS, Juhler RK, Hansen HCB (2014) The naturally occurring carcinogen ptaquiloside is present in groundwater below bracken vegetation. Environ Toxicol Chem 33:1030–1034

    Article  CAS  Google Scholar 

  29. Clauson-Kaas F, Ramwell C, Hansen HCB, Strobel BW (2016) Ptaquiloside from bracken in stream water at base flow and during storm events. Water Res 106:155–162

    Article  CAS  Google Scholar 

  30. Kisielius V (2020) Toxic secondary metabolites from plants: A new aspect of water quality. PhD thesis. University of Copenhagen

  31. PubChema chemistry database. https://pubchem.ncbi.nlm.nih.gov/. Accessed 10 August 2020

  32. Ojika M, Wakamatsu K, Niwa H, Yamada K (1987) Ptaquiloside, a potent carcinogen isolated from bracken fern pteridiumaquilinum var. latiusculum: structure elucidation based on chemical and spectral evidence, and reactions with amino acids, nucleosides, and nucleotides. Tetrahedron 43:5261–5274

    Article  CAS  Google Scholar 

  33. García-Jorgensen DB, Hansen HCB, Abrahamsen P, Diamantopoulos E, Garc DB (2020) A novel model concept for modelling the leaching of natural toxins: results for the case of ptaquiloside. Environ Sci Process Impacts 22:1768–1779

    Article  Google Scholar 

  34. Rosenbom AE, Olsen P, Plauborg F, Grant R, Juhler RK, Brüsch W, Kjær J (2015) Pesticide leaching through sandy and loamy fields–Long-term lessons learnt from the Danish Pesticide Leaching Assessment Programme. Environ Pollut 201:75–90

    Article  CAS  Google Scholar 

  35. Brüsch W, Rosenbom AE, Badawi N, Olsen P (2016) Monitoring of pesticide leaching from cultivated fields in Denmark. Geol Surv Denmark Greenl Bull 17–22

  36. González-Pradas E, Ureña-Amate MD, Flores-Céspedes F, Fernández-Pérez M, Garratt J, Wilkins R (2002) Leaching of imidacloprid and procymidone in a greenhouse of southeast of Spain. Soil Sci Soc Am J 66:1821–1828

    Article  Google Scholar 

  37. Holt MS (2000) Sources of chemical contaminants and routes into the freshwater environment. Food Chem Toxicol 38:S21–S27

    Article  CAS  Google Scholar 

  38. Malaguerra F, Albrechtsen H-J, Binning PJ (2013) Assessment of the contamination of drinking water supply wells by pesticides from surface water resources using a finite element reactive transport model and global sensitivity analysis techniques. J Hydrol 476:321–331

    Article  CAS  Google Scholar 

  39. Water supply in Denmark. GEUS, Danish Ministry of the Environment. https://eng.ecoinnovation.dk/media/mst/8051461/Vandforsyning_artikel.pdf. Accessed 20 June 2020

  40. Kisielius V, Lindqvist DN, Thygesen MB, Rodamer M, Hansen HCB, Rasmussen LH (2020) Fast LC-MS quantification of ptesculentoside, caudatoside, ptaquiloside and corresponding pterosins in bracken ferns. J Chromatogr B 12:1966

    Google Scholar 

  41. Skrbic N, Pedersen A, Christensen SCB (2020) A Novel Method for Determination of the Natural Toxin Ptaquiloside in Ground and Drinking Water. Water 12:2852

    Article  CAS  Google Scholar 

  42. Adhikari K (2013) Soil Mapping in Denmark Using Digital Soil Mapping Techniques. Aarhus Universitet, Institut for Agroøkologi

    Google Scholar 

  43. Mori H, Sugie S, Hirono I, Yamada K, Niwa H, Ojika M (1985) Genotoxicity of ptaquiloside, a bracken carcinogen, in the hepatocyte primary culture/DNA-repair test. Mutat Res Lett 143:75–78

    Article  CAS  Google Scholar 

  44. Vetter J (2009) A biological hazard of our age: Bracken fern [Pteridium aquilinum (L.) Kuhn]—A review. Acta Vet Hung 57:183–196

    Article  CAS  Google Scholar 

  45. Rasmussen LH (2003) Ptaquiloside: an environmental hazard? The Royal and Veterinary and Agricultural University

  46. National groundwater database Jupiter, GEUS database (Denmark), https://www.geus.dk/produkter-ydelser-og-faciliteter/data-og-kort/national-boringsdatabase-jupiter/. Accessed 18 September 2020

  47. Who supplies the drinking water? Danish Environmental Protection Agency, https://mst.dk/natur-vand/vand-i-hverdagen/drikkevand/hvem-leverer-drikkevandet/. Accessed 10 June 2020

  48. Hulsmann A (2005) Small systems large problems. A European Inventory of small water systems and associated problems, Nieuwegein, Web-based European Knowledge Network on Water

    Google Scholar 

  49. Kördel W, Egli H, Klein M (2008) Transport of pesticides via macropores (IUPAC technical report). Pure Appl Chem 80:105–160

    Article  Google Scholar 

  50. Jørgensen PR, Krogh PH, Hansen S, Petersen CT, Nielsen MH, Rasmussen SB, Heinrichson K, Spliid NH (2017) Dybe bioporers forekomst og betydning for pesticidudvaskning i moræneler

Download references

Acknowledgements

This project has received funding from the European Union’s Horizon 2020 research and Innovation Programme under the Marie Sklodowska-Curie Grant agreement No. 722493 (NaToxAq). The authors are grateful to many people and institutions that helped in localization and sampling from the water wells: Associate professor Elena Maria Fernandez Gonzales (Univeristy of Ovieedo, Spain), Anne Esbjørn (VCS Denmark), Niels Peter Dalsgård Jensen (SLS A/S), Tommy Larsen (HOFOR A/S), employees of Kagerup, Gadevang and Nødebo vandværk. The authors would also like to thank Jesper Jørgen Nielsen (HOFOR A/S), technicians Birgitte Munch Nielsen and Steen Dam Larsen (HOFOR A/S), as well as laboratory technicians Helle Sparwath (HOFOR A/S) and Jimena Bertorello Martinez (University College Copenhagen).

Funding

This project was funded by the European Union’s Horizon 2020 research and Innovation Programme under the Marie Sklodowska-Curie Grant agreement No. 722493 (NaToxAq).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: NS, A-KP, SCBC, MJH, HCBH and LHR. Investigations: NS, VK and LHR. Data curation: NS and LHR. Visualisation: NS. Funding acquisition: A-KP, HCBH and LHR. Supervision: A-KP, SCBC, MJH, HCBH and LHR. Writing—original draft: NS. Writing—review and editing: NS, A-KP, SCBC, MJH, VK, HCBH and LHR. All the authors read and approved the final manuscript.

Corresponding author

Correspondence to Natasa Skrbic.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1.

Supplementary material (additional tables and figures).

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Skrbic, N., Kisielius, V., Pedersen, AK. et al. Occurrence of carcinogenic illudane glycosides in drinking water wells. Environ Sci Eur 33, 44 (2021). https://doi.org/10.1186/s12302-021-00486-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12302-021-00486-y

Keywords