Figure 1 - available via license: Creative Commons Zero 1.0
Content may be subject to copyright.
Zoanthid colony responsible for a severe respiratory reaction, collected from a home aquarium in 2008. [Referred to as VAZOA in Figure 2A and Virginia zoanthid in text]. Sample was found 

Zoanthid colony responsible for a severe respiratory reaction, collected from a home aquarium in 2008. [Referred to as VAZOA in Figure 2A and Virginia zoanthid in text]. Sample was found 

Source publication
Article
Full-text available
Zoanthids (Anthozoa, Hexacorallia) are colonial anemones that contain one of the deadliest toxins ever discovered, palytoxin (LD(50) in mice 300 ng/kg), but it is generally believed that highly toxic species are not sold in the home aquarium trade. We previously showed that an unintentionally introduced zoanthid in a home aquarium contained high co...

Contexts in source publication

Context 1
... 48 agarose gels (Invitrogen Catalog No. G8008-01) according to manufacturer’s protocols with the E-Base H Integrated power supply (Invitrogen Catalog No. EB-M03). Gels were run for 15 min and then visualized using a Bio-Rad gel documentation system (Gel Doc 2000, Hercules, CA). The gel was also photographed with this instrument and the picture retained for records. Successfully amplified products were purified with Exosap-IT (USB, Cleveland, OH) by adding 2 m l of Exosap-IT to 5 m l of PCR product, incubating at 37 u C for 15 minutes followed by 15 minutes at 80 C. At least two sequencing reactions (in the forward and reverse direction) were run on each sample at the Smithsonian Institution’s Laboratories of Analytical Biology (Suitland, MD, USA). Sequencing reactions consisted of 0.5 m l of BigDye H Terminator v3.1 (Applied Biosystems), 1.75 m l of 5X sequencing buffer (Applied Biosystems), 0.5 m l of primer (either M13F-29 or M13R), 6.25 m l of molecular grade water for a total of 9 m l to which 1 m l of cleaned up PCR product was added. Sequencing reactions were conducted on an Eppendorf Mastercycler H ep gradient S Thermocycler with the following conditions with the following conditions: 95 u C for 2 min, 30 cycles of 95 u C for 30 sec, 50 u C for 15 sec, and 60 u C for 4 min, followed by a 10 u C hold. Reactions were cleaned using a sephedex column, and then dried down. Finally 10 m l of Hi-Di formamide (Applied Biosystems # 4311320) was added to each sample and heated at 95 u C for 2 min. Samples were then sequenced on an ABI 3730xL sequencer (Applied Biosystems, Foster City, CA), which was maintained according to manufacturer’s specifications. Resulting sequences were edited in Sequencher 4.9 (Gene Codes Corp., Ann Arbor, MI). Phylogenetic Analysis. Nucleotide sequences of 16S and COI from samples were manually aligned with previously published 16S and COI sequences from various zoanthid species representing the genera Palythoa , Zoanthus and Isaurus . These published sequences were the top matches to the unknown samples using the BLAST algorithm [16] to search the National Center for Biotechnology Information (NCBI) database located at Out-group sequences for both 16S and COI trees were from the genera Hydrozoanthus (within suborder Macrocnemina). All alignments were inspected by eye and manually edited. All ambiguous sites of the alignments were removed from the dataset for phylogenetic analyses. Consequently, two alignment datasets were generated: 1. 817 sites of 30 sequences (16S) and 2. 461 sites of 26 sequences (COI). The alignment data are available on request from the corresponding author, and also at the webpage For the phylogenetic analyses of the two alignments, the same methods were applied independently. Alignments were subjected to analyses with maximum-likelihood method (ML) with PhyML [17] and neighbor-joining (NJ) method with CLC Free Work- bench 3 [18]. PhyML was performed using an input tree generated by BIONJ with the general time-reversible (GTR) model [19] of nucleotide substitution incorporating invariable sites and a discrete gamma distribution (eight categories) (GTR + I + C ). The proportion of invariable sites, a discrete gamma distribution, and base frequencies of the model were estimated from the dataset. PhyML bootstrap trees (1000 replicates) were constructed using the same parameters as the individual ML tree. The distances were calculated using a Kimura’s 2-parameter model [20]. Support for NJ branches was tested by bootstrap analysis [21] of 1000 replicates. Bayesian trees were reconstructed by using the program MrBayes 3.1.2 under the GTR model [22]. One cold and three heated Markov chain Monte Carlo (MCMC) chains with default- chain temperatures were run for 20 million generations, sampling log-likelihoods (InLs), and trees at 100-generation intervals (20,000 InLs and trees were saved during MCMC). The first 15% of all runs were discarded as ‘‘burn-in’’ for all datasets. The likelihood plots for both datasets also showed that MCMC reached the stationary phase by this time (PSRF = 1.000 for both data sets). Thus, the remaining 170,000 trees (17 million generations) of 16S and COI were used to obtain posterior probabilities and branch- length estimates, respectively. After our initial finding of a highly toxic Palythoa sp. in a home aquarium in Virginia [sample ‘‘Virginia zoanthid’’, described in Deeds and Schwartz [8], (Fig. 1)], we visited a local aquarium store known to carry zoanthids (aquarium store # 1) and purchased every visually distinct variety of zoanthid they had for sale ( n = 7). These included specimens visually consistent with both Palythoa/Protopalythoa spp. and Zoanthus spp. (according to [23]). Of these, only one proved to be as toxic as the sample from the Virginia home aquarium (sample 305.11.2) (Table 3). Among the specimens purchased from aquarium store # 1, only sample 305.11.2 was morphologically similar to the Virginia zoanthid sample (Fig. 1, Fig. 2C). With this information, two additional aquarium stores were visited (aquarium store # 2 and # 3), targeting specimens that were visually consistent with our two toxic Palythoa sp. samples. These stores were identified as the most likely in the area to carry zoanthids based on discussions with local marine aquarium hobbyists. Between the two stores, eight additional specimens were acquired. Of these, three were found to be highly toxic (Table 3). The first sample consisted of a few individual polyps growing on a small fragment of coral at the bottom of a large zoanthid display tank (sample 306.37.3) (Fig. 2D) in aquarium store # 2. The small specimen was located at the bottom of the tank, under a display rack, but was retrieved by the store owner and sold for a minimal price. The last two samples, acquired from aquarium store # 3, were either growing as disperse polyps among a colony of another zoanthid species (sample 306.39.2), or as a few random polyps growing on the side of a small piece of plastic pipe sitting on the bottom of another display tank (sample 306.39.3) (Fig. 2E,F respectively). Using high performance liquid chromatography (HPLC) with UV detection, with quantification against a PLTX standard, we found that all four of the purchased zoanthids that were visually consistent with the Virginia home aquarium specimen were themselves highly toxic (approx. 500-3500 m g crude toxin/g wet zoanthid) (Fig. 3, Table 3). In our previous study, we found the Virginia home aquarium sample to contain approx. 600 m g crude toxin/g wet zoanthid [8]. For comparison, the original Hawaiian P. toxica collections in the 1960’s yielded approximately 300 m g pure toxin/g wet zoanthid [2]. Using electrospray ionization liquid chromatography mass spectrometry (ESI/LC/MS) with a high resolution mass spectrometer, for confirmation of PLTX type, four of the samples were shown to contain only palytoxin (Virginia zoanthid, 305.11.2, 306.39.2, 306.39.3) while the last (306.37.3) contained primarily deoxy-palytoxin with a lesser amount of palytoxin (Table 2). Minor PLTXs could not be quantified due to a lack of chromatographic resolution. None were found to contain 42-hydroxy-palytoxin, the primary toxin recently reported from samples collected in the 1990s from the original Hana tidepools from, presumably, P. toxica [24]. As expected, our PLTX standard, isolated from P. tuberculosa , contained primarily palytoxin with a minor amount of 42-hydroxy-palytoxin [22] (Table 2). Several deoxy-palytoxins have been previously described: one as a minor toxin from P. tuberculosa [25], and more recently, several from cultures of the dinoflagellate Ostreopsis cf. ovata and Ostreopsis ovata (Rossi et al. [26], therein called ovatoxin-b, and Ciminiello et al. [27], therein called ovatoxins d and e, respectively). This is the first report of a deoxy-palytoxin being the primary toxin from any zoanthid. 42-hydroxy-palytoxin has been shown to be similar in potency with palytoxin [24]. None of the described deoxy- palytoxins have been assessed for potency. In this study, the position of deoxygenation was not determined, therefore the exact relationship of the deoxy-palytoxin described here with those previously described could not be established. All of the other samples collected ( n = 11), both additional Palythoa spp. and other specimens visually and genetically (see below) consistent with Zoanthus spp., were either non-toxic or weakly-toxic (i.e. shown to have an extractable hemolytic component with properties consistent with PLTX but too low in concentration to confirm as PLTX by other chemical means – detailed methods can be found in [8]). Mitochondrial 16S and COI markers were successfully generated for 10 of the 16 specimens from this study (Table 3). For the remaining six specimens either 16S or COI sequences were generated. All failed sequences were attempted at least twice. Phylogenetic trees generated using both markers showed the same result that all of our toxic samples ( n = 5) formed a closely related clade that was distinct from the additional non- weakly-toxic Palythoa spp. ( n = 3) and the non- weakly-toxic Zoanthus spp. ( n = 8) (Fig. 2A,B). The most closely related species to our toxic specimens based on available molecular data is Palythoa heliodiscus . It should be noted that relatively few of the described species of zoanthids have been analyzed genetically. To our knowledge, no molecular data ...
Context 2
... 0.062 m l of 10 mM dNTPs (New England Biolabs Catalog No. N0447L), and 0.060 m l of Platinum Taq (5 U/ m l, Invitrogen Catalog No. 10966- 034) and 1 m l of undiluted DNA template per reaction, according to Ivanova et al. [15]. An Eppendorf Mastercycler H ep gradient S Thermocycler was used for the PCRs with the following conditions: 95 u C for 1 min, 35 cycles of 95 u C for 1 min, 50 u C for 40 sec (for 16S, 40 u C for COI), and 72 u C for 1.5 min, with a final extension at 72 u C for 7 min. All products were verified using pre-cast 1% E-gel 48 agarose gels (Invitrogen Catalog No. G8008-01) according to manufacturer’s protocols with the E-Base H Integrated power supply (Invitrogen Catalog No. EB-M03). Gels were run for 15 min and then visualized using a Bio-Rad gel documentation system (Gel Doc 2000, Hercules, CA). The gel was also photographed with this instrument and the picture retained for records. Successfully amplified products were purified with Exosap-IT (USB, Cleveland, OH) by adding 2 m l of Exosap-IT to 5 m l of PCR product, incubating at 37 u C for 15 minutes followed by 15 minutes at 80 C. At least two sequencing reactions (in the forward and reverse direction) were run on each sample at the Smithsonian Institution’s Laboratories of Analytical Biology (Suitland, MD, USA). Sequencing reactions consisted of 0.5 m l of BigDye H Terminator v3.1 (Applied Biosystems), 1.75 m l of 5X sequencing buffer (Applied Biosystems), 0.5 m l of primer (either M13F-29 or M13R), 6.25 m l of molecular grade water for a total of 9 m l to which 1 m l of cleaned up PCR product was added. Sequencing reactions were conducted on an Eppendorf Mastercycler H ep gradient S Thermocycler with the following conditions with the following conditions: 95 u C for 2 min, 30 cycles of 95 u C for 30 sec, 50 u C for 15 sec, and 60 u C for 4 min, followed by a 10 u C hold. Reactions were cleaned using a sephedex column, and then dried down. Finally 10 m l of Hi-Di formamide (Applied Biosystems # 4311320) was added to each sample and heated at 95 u C for 2 min. Samples were then sequenced on an ABI 3730xL sequencer (Applied Biosystems, Foster City, CA), which was maintained according to manufacturer’s specifications. Resulting sequences were edited in Sequencher 4.9 (Gene Codes Corp., Ann Arbor, MI). Phylogenetic Analysis. Nucleotide sequences of 16S and COI from samples were manually aligned with previously published 16S and COI sequences from various zoanthid species representing the genera Palythoa , Zoanthus and Isaurus . These published sequences were the top matches to the unknown samples using the BLAST algorithm [16] to search the National Center for Biotechnology Information (NCBI) database located at Out-group sequences for both 16S and COI trees were from the genera Hydrozoanthus (within suborder Macrocnemina). All alignments were inspected by eye and manually edited. All ambiguous sites of the alignments were removed from the dataset for phylogenetic analyses. Consequently, two alignment datasets were generated: 1. 817 sites of 30 sequences (16S) and 2. 461 sites of 26 sequences (COI). The alignment data are available on request from the corresponding author, and also at the webpage For the phylogenetic analyses of the two alignments, the same methods were applied independently. Alignments were subjected to analyses with maximum-likelihood method (ML) with PhyML [17] and neighbor-joining (NJ) method with CLC Free Work- bench 3 [18]. PhyML was performed using an input tree generated by BIONJ with the general time-reversible (GTR) model [19] of nucleotide substitution incorporating invariable sites and a discrete gamma distribution (eight categories) (GTR + I + C ). The proportion of invariable sites, a discrete gamma distribution, and base frequencies of the model were estimated from the dataset. PhyML bootstrap trees (1000 replicates) were constructed using the same parameters as the individual ML tree. The distances were calculated using a Kimura’s 2-parameter model [20]. Support for NJ branches was tested by bootstrap analysis [21] of 1000 replicates. Bayesian trees were reconstructed by using the program MrBayes 3.1.2 under the GTR model [22]. One cold and three heated Markov chain Monte Carlo (MCMC) chains with default- chain temperatures were run for 20 million generations, sampling log-likelihoods (InLs), and trees at 100-generation intervals (20,000 InLs and trees were saved during MCMC). The first 15% of all runs were discarded as ‘‘burn-in’’ for all datasets. The likelihood plots for both datasets also showed that MCMC reached the stationary phase by this time (PSRF = 1.000 for both data sets). Thus, the remaining 170,000 trees (17 million generations) of 16S and COI were used to obtain posterior probabilities and branch- length estimates, respectively. After our initial finding of a highly toxic Palythoa sp. in a home aquarium in Virginia [sample ‘‘Virginia zoanthid’’, described in Deeds and Schwartz [8], (Fig. 1)], we visited a local aquarium store known to carry zoanthids (aquarium store # 1) and purchased every visually distinct variety of zoanthid they had for sale ( n = 7). These included specimens visually consistent with both Palythoa/Protopalythoa spp. and Zoanthus spp. (according to [23]). Of these, only one proved to be as toxic as the sample from the Virginia home aquarium (sample 305.11.2) (Table 3). Among the specimens purchased from aquarium store # 1, only sample 305.11.2 was morphologically similar to the Virginia zoanthid sample (Fig. 1, Fig. 2C). With this information, two additional aquarium stores were visited (aquarium store # 2 and # 3), targeting specimens that were visually consistent with our two toxic Palythoa sp. samples. These stores were identified as the most likely in the area to carry zoanthids based on discussions with local marine aquarium hobbyists. Between the two stores, eight additional specimens were acquired. Of these, three were found to be highly toxic (Table 3). The first sample consisted of a few individual polyps growing on a small fragment of coral at the bottom of a large zoanthid display tank (sample 306.37.3) (Fig. 2D) in aquarium store # 2. The small specimen was located at the bottom of the tank, under a display rack, but was retrieved by the store owner and sold for a minimal price. The last two samples, acquired from aquarium store # 3, were either growing as disperse polyps among a colony of another zoanthid species (sample 306.39.2), or as a few random polyps growing on the side of a small piece of plastic pipe sitting on the bottom of another display tank (sample 306.39.3) (Fig. 2E,F respectively). Using high performance liquid chromatography (HPLC) with UV detection, with quantification against a PLTX standard, we found that all four of the purchased zoanthids that were visually consistent with the Virginia home aquarium specimen were themselves highly toxic (approx. 500-3500 m g crude toxin/g wet zoanthid) (Fig. 3, Table 3). In our previous study, we found the Virginia home aquarium sample to contain approx. 600 m g crude toxin/g wet zoanthid [8]. For comparison, the original Hawaiian P. toxica collections in the 1960’s yielded approximately 300 m g pure toxin/g wet zoanthid [2]. Using electrospray ionization liquid chromatography mass spectrometry (ESI/LC/MS) with a high resolution mass spectrometer, for confirmation of PLTX type, four of the samples were shown to contain only palytoxin (Virginia zoanthid, 305.11.2, 306.39.2, 306.39.3) while the last (306.37.3) contained primarily deoxy-palytoxin with a lesser amount of palytoxin (Table 2). Minor PLTXs could not be quantified due to a lack of chromatographic resolution. None were found to contain 42-hydroxy-palytoxin, the primary toxin recently reported from samples collected in the 1990s from the original Hana tidepools from, presumably, P. toxica [24]. As expected, our PLTX standard, isolated from P. tuberculosa , contained primarily palytoxin with a minor amount of 42-hydroxy-palytoxin [22] (Table 2). Several deoxy-palytoxins have been previously described: one as a minor toxin from P. tuberculosa [25], and more recently, several from cultures of the dinoflagellate Ostreopsis cf. ovata and Ostreopsis ovata (Rossi et al. [26], therein called ovatoxin-b, and Ciminiello et al. [27], therein called ovatoxins d and e, respectively). This is the first report of a deoxy-palytoxin being the primary toxin from any zoanthid. 42-hydroxy-palytoxin has been shown to be similar in potency with palytoxin [24]. None of the described deoxy- palytoxins have been assessed for potency. In this study, the position of deoxygenation was not determined, therefore the exact relationship of the deoxy-palytoxin described here with those previously described could not be established. All of the other samples collected ( n = 11), both additional Palythoa spp. and other specimens visually and genetically (see below) consistent with Zoanthus spp., were either non-toxic or weakly-toxic (i.e. shown to have an extractable hemolytic component with properties consistent with PLTX but too low in concentration to confirm as PLTX by other chemical means – detailed methods can be found in [8]). Mitochondrial 16S and COI markers were successfully generated for 10 of the 16 specimens from this study (Table 3). For the remaining six specimens either 16S or COI sequences were generated. All failed sequences were attempted at least twice. Phylogenetic trees generated using both markers showed the same result that all of our toxic samples ( n = 5) formed a closely related clade that was distinct from the additional non- weakly-toxic Palythoa spp. ( n = 3) and the non- weakly-toxic Zoanthus spp. ( n = 8) (Fig. 2A,B). The most closely related species to our toxic specimens based on available molecular data is Palythoa heliodiscus . It should be noted that relatively few of the described species of zoanthids have been analyzed genetically. ...
Context 3
... the late 1950’s, Dr. Albert H. (Hank) Banner began a program at the University of Hawaii to search for the elusive cause of ciguatera fish poisoning [1]. Dr. Philip Helfrich, a young researcher hired by Banner, began this search by investigating an entry in the Hawaiian dictionary for the ‘‘ limu-make-o-Hana ’’ (rough translation deadly seaweed of Hana) [2]. This legend dates back to Hawaiian antiquity with tales of Shark Gods, sacred pools, and a seaweed when applied to a warrior’s spear would ‘‘bring sure death’’ to their enemies. The pool became ‘‘kapu’’ or taboo to the local Hawaiians and it was said that an ill fate would befall anyone who disturbed the sacred site. Every legend holds some basis in fact, and in 1961, Helfrich, accompanied by graduate student John Shupe, tracked down the fabled pool near the village of Mu 9 olea on the island of Maui and introduced to the world a new species of cnidarian zoanthid (colonial anemone) known as Palythoa toxica [3,4]. This research led to the discovery of palytoxin (PLTX), one of the most potent natural products ever discovered [5]. Although much of the structural elucidation of palytoxin would be determined from less toxic, but far more abundant, species such as Palythoa tuberculosa [6,7] none were ever found to be as potent as the samples collected from the tidepools at Mu 9 olea. Now the legendary limu appears to be exacting its ancient curse once again, but this time upon unsuspecting marine home aquarists. In 2007, we assisted the Georgia poison center in an investigation into a potential dermal intoxication with palytoxin from zoanthids in a home aquarium [8]. During the course of the investigation, we learned of another marine aquarium hobbyist in Virginia who had recently experienced a severe respiratory reaction while trying to eradicate brown non-descript colonies of zoanthids that had arrived as hitchhikers with live rock and were overgrowing more desirable organisms in the tank (Fig. 1) [account can be found at: http:// www.reefcentral.com/forums/showthread.php?t = 1083843 – ac- cessed 12/09/10]. A sample (here called Virginia zoanthid or VAZOA) was found to contain high levels of palytoxin (600 m g crude toxin/g wet zoanthid). Zoanthids are a notoriously problematic taxonomic group [9–11] and it proved difficult at the time to identify to the species level the sample from the home aquarium in Virginia. Histologic evaluation of preserved polyps could only confirm that it was a Palythoa sp. (D. Fautin, University of Kansas, personal communication). In an attempt to determine the prevalence of palytoxin in aquarium store zoanthids, we purchased several colonies from local aquarium stores and analyzed them for palytoxin. We further performed a molecular analysis in an attempt to identify the colonies to the species level. Fifteen zoanthid colonies visually consistent with Palythoa/ Protopalythoa spp. or Zoanthus spp. were purchased live from 3 aquarium stores in the Washington DC metro area. Samples, with associated seawater, were placed in individual glass beakers under fluorescent light for 24 hours in an attempt to induce the opening of individual polyps for photographic documentation. After photographic documentation, several polyps from each colony were removed from their supporting substrate using forceps and fine dissecting scissors and placed in 10 ml of 80% ethanol for 24 hours at 4 u C to extract PLTXs. Some of the larger polyps were split in half to facilitate extraction but no additional homogeni- zation of tissues was performed. After extraction, polyps were removed, blotted dry, weighed, and placed in fresh 95% ethanol for DNA analysis. The total amount of tissue extracted varied depending on the size of the colony purchased and the size of the individual polyps, but the range for all samples extracted for toxin analysis in this study was 0.2–0.8 g/specimen. Palytoxin Quantification. High performance liquid chromatography (HPLC) analysis was performed using an Agilent 1200 series HPLC system (Agilent, Wilmington, DE) with UV detection following Ciminiello et al. [12]. Samples were diluted 1:10 using a solution of 80% acetonitrile in HPLC grade H 2 O containing 30 mM acetic acid and 30 m l was injected onto a Gemini C18 column (5 m m, 110 A ̊ , 150 mm 6 2 mm) (Phenomenex, Torrance, CA). The sample was eluted using a gradient of 20% solvent A/80% solvent B to 100% solvent A over 10 min at 250 m l/min. at 30 u C [Solvent A: 95% acetonitrile in HPLC grade water, Solvent B: HPLC grade water, 30 mM acetic acid added to both]. Beginning and ending gradient conditions were maintained for 5 min. before and after the gradient run, respectively. PLTX-like compounds were detected at 263 nm and quantified through linear regression of a PLTX standard (2680 Da. from P. tuberculosa purchased from Wako Pure Chemical Industries, Ltd., Japan, dissolved in 50% ethanol according to manufacturer’s instructions) at concentrations of 10, 5, 2.5, 1.25, 0.625, 0.3125 m g/ml. Linear regression analysis was performed using GraphPad Prism software (ver. 4.03, GraphPad Software, Inc., San Diego, CA). Determination of palytoxin type. High resolution liquid chromatography mass spectrometry was used to distinguish palytoxin from 42-hydroxy-palytoxin and deoxy-palytoxin. A Thermo Exactive mass spectrometer coupled with an Accela liquid chromatography system (ThermoScientific, Waltham, MA) was used for this analysis. LC conditions were as follows: 10 ul injections were made onto a Gemini C18 column (5 m m, 110 A ̊, 150 mm 6 2 mm) (Phenomenex, Torrance, CA) and eluted with a linear gradient, as described previously, at a flow rate of 200 m l/ min. Eluent was electrosprayed via the heated probe (HESI-II) into the mass spectrometer (MS). MS conditions were as follows: MS was operated in the full scan positive ion mode from 150–3000 daltons at the ultrahigh resolution setting of 100,000 @ 1 Hz. Other MS conditions are listed in Table 1. PLTX type was confirmed if the total ion chromatogram contained each of the calculated masses in Table 2 as the dominant ion in an isotopic cluster, out to two decimal places. The only predicted ion never detected for any toxin was (M + 3H) 3 + . Minor PLTXs were considered present if select ions for that type were also present as the dominant ion in an isotopic cluster, typically (M + 2H- H 2 O) 2 + and (M + 2H + K) 3 + , the most abundant ions present for each PLTX type in all samples. Molecular Methods. A small piece of tissue ( 10 mg) from one representative polyp from each of the 16 zoanthid samples (15 collected here plus the Virginia zoanthid sample from Deeds and Schwartz [8]) was removed using flame sterilized (with ethanol) tweezers and/or scissors and added to a sterile 1.5 ml microcentrifuge tube. DNA was extracted from tissue through the use of a DNeasy Blood and Tissue Kit (Qiagen # 69506). Reagent volumes were reduced to a quarter of the volume listed in ...

Citations

... Thirdly, as Palythoa and Zoanthus are not listed under CITES, sampling, permitting, and international specimen shipment which are logistically easier than for scleractinians. Colonies of both genera can be easily fragmented and shipped live, as is regularly done in the pet trade (Deeds et al., 2011), and thus, aquaria experiments should be comparatively straightforward. Thus, we propose that some Palythoa and Zoanthus should be examined in more detail, not replacing scleractinian species, but as additional model species to gather more needed information as stated above, and to provide wider comparisons to scleractinian-focused datasets or experiments. ...
Article
Full-text available
Whilst natural analogues for future ocean conditions such as CO2 seeps and enclosed lagoons in coral reef regions have received much recent research attention, most efforts in such locations have focused on the effects of prolonged high CO2 levels on scleractinian corals and fishes. Here, we demonstrate that the three species of zooxanthellate zoantharians, hexacorallian non-calcifying “cousins” of scleractinians, are common across five coral reef natural analogue sites with high CO2 levels in the western Pacific Ocean, in Japan (n = 2), Palau, Papua New Guinea, and New Caledonia (n = 1 each). These current observations support previously reported cases of high Palythoa and Zoanthus abundance and dominance on various impacted coral reefs worldwide. The results demonstrate the need for more research on the ecological roles of zooxanthellate zoantharians in coral reef systems, as well as examining other “understudied” taxa that may become increasingly important in the near future under climate change scenarios. Given their abundance in these sites combined with ease in sampling and non-CITES status, some zoantharian species should make excellent hexacoral models for examining potential resilience or resistance mechanisms of anthozoans to future high pCO2 conditions.
... Like brevetoxins, CTXs act on site 5 of the alpha subunit of sodium channel receptors to increase excitability and prolong refractory periods of sensory and motor nerves. The result of the high-affinity interaction of CTXs with such sodium channels is the opposite of the effects of tetrodotoxin or saxitoxin, which block sodium channels (Cameron et al., 1991a,b;Dechraoui et al., 1999). In vitro, CTX induces contraction of the ileum and exerts a positive inotropic effect on guinea pig cardiac muscle ). ...
... Exposures of humans to PLTXs from aerosols of seawater that had blooms of Ostreopsis, and handling or pouring hot water onto zoanthid corals while tending saltwater reef aquaria, have been associated with respiratory distress, rhinorrhea, cough, fever, ocular irritation, and dermatitis (Deeds and Schwartz, 2010;Deeds et al., 2011;Hall et al., 2015;Pelin et al., 2016). Cleaning PLTX-containing aquaria has also been linked to tachypnea, wheezing, hemoptysis, radiographically evident pulmonary opacities, hypoxemia, tachycardia, electrocardiographic abnormalities, leukocytosis, myalgia, weakness, muscle spasms, nausea, paresthesias, ataxia, and tremors. ...
... The genus Palythoa is known for comprising highly toxic species due to the presence of palytoxin (PLTX) in their tissues (Béress et al., 1983;Deeds et al., 2011;Moore & Scheuer, 1971). PLTX, one of the most toxic natural compounds ever discovered, is a non-protein marine toxin which consists of a long, partially unsaturated (with eight double bonds) aliphatic backbone with spaced cyclic ethers and 64 chiral centres (Uemura et al., 1985). ...
... Moreover, several analogues of PLTX were discovered in various microorganisms (see Table 1). To date, Palythoa heliodiscus (Ryland & Lancaster, 2003) is the largest PLTX producer known among zoantharians with the highest recorded yield (1 mg of PLTX /g wet Palythoa) and deoxy-PLTX (3.51 mg/g wet Palythoa (Deeds et al., 2011)). However, in most cases the toxin yield in other Palythoa species is low. ...
... One of the goals of this study was to use the evolutionary relationships revealed by molecular phylogenetic analyses, as well as the comparative analysis of PLTX contents, to predict zoantharian toxicity in a phylogenetic context and investigate the relation between toxicity and Symbiodiniaceae strains. To date, only one study of the relation between toxicity and phylogeny in Zoantharia has been conducted, with a reduced-scale taxonomic sampling (Deeds et al., 2011). To assess the potential exposure of PLTX to marine aquarium hobbyists, specimens were identified through genetic analysis of 16S and COI markers. ...
Article
Anemone-like animals belonging to the order Zoantharia are common anthozoans widely distributed from shallow to deep tropical and subtropical waters. Some species are well-known because of their high toxicity due to the presence of palytoxin (PLTX) in their tissues. PLTX is a large polyhydroxylated compound and one of the most potent toxins known. Currently, the PLTX biosynthetic pathway in zoantharians and the role of the host or the putative symbiotic organism(s) involved in this pathway are entirely unknown. To better understand the presence of PLTX in some Zoantharia, twenty-nine zoantharian colonies were analysed in this study. All zoantharian samples and their endosymbiotic dinoflagellates (Symbiodiniaceae = Zooxanthellae) were identified using DNA barcoding and phylogenetic reconstructions. Quantification of PLTX and its analogues showed that the yields contained in Palythoa heliodiscus, Palythoa aff. clavata and one potentially undescribed species of Palythoa are among the highest ever found (up to > 2 mg/g of wet zoantharian). Mass spectrometry imaging was used for the first time on Palythoa samples and revealed that in situ distribution of PLTX is mainly located in ectodermal tissues such as the epidermis of the body wall and the pharynx. Moreover, high levels of PLTX have been detected in histological regions where few or no Symbiodiniaceae cells could be observed. Finally, issues such as host‐specificity and environmental variables driving biogeographical patterns of hosted Symbiodiniaceae in zoantharian lineages were discussed in light of our phylogenetic results as well as the patterns of PLTX distribution. It was concluded that (1) the variability of Symbiodiniaceae diversity may be related to ecological divergence in Zoantharia, (2) all Palythoa species hosted Cladocopium Symbiodiniaceae (formerly clade C), (3) the sole presence of Cladocopium is not sufficient to explain the presence of high concentrations of PLTX and/or its analogues, and (4) the ability to produce high levels of PLTX and/or its analogues highlighted in some Palythoa species could be a plesiomorphic character inherited from their last common ancestor and subsequently lost in several lineages.
... Other pathways through which PlTX accumulates also exist ( Table 2). Poisoning cases related to PlTX-containing seafood have been reported in tropical and subtropical regions, and there are reports on the correlations of Ostreopsis and PlTX exposure pathways based on soft coral trade for aquarium decoration purposes [70]. Another route is fish ingesting Ostreopsis spp. or accumulation of Ostreopsis in fish while feeding on algae to which Ostreopsis spp. ...
Article
Full-text available
Among marine biotoxins, palytoxins (PlTXs) and cyclic imines (CIs), including spirolides, pinnatoxins, pteriatoxins, and gymnodimines, are not managed in many countries, such as the USA, European nations, and South Korea, because there are not enough poisoning cases or data for the limits on these biotoxins. In this article, we review unregulated marine biotoxins (e.g., PlTXs and CIs), their toxicity, causative phytoplankton species, and toxin extraction and detection protocols. Due to global warming, the habitat of the causative phytoplankton has expanded to the Asia-Pacific region. When ingested by humans, shellfish that accumulated toxins can cause various symptoms (muscle pain or diarrhea) and even death. There are no systematic reports on the occurrence of these toxins; however, it is important to continuously monitor causative phytoplankton and poisoning of accumulating shellfish by PlTXs and CI toxins because of the high risk of toxicity in human consumers.
... Von Relevanz für Ophthalmolog*innen in diesem Fall ist das Gift Palytoxin, welches in Korallen der Spezies Palythoa zu finden ist und zu den tödlichsten Toxinen zählt, welche der Menschheit momentan bekannt sind. Die tödliche Dosis ("lethal dose" [LD]) ist in Mäusen mit LD = 30 ng/kg beschrieben [2]. Das Toxin selbst kann nach Kontakt eine massive Inflammation der Augenoberfläche bis hin zur kornealen Perforation verursachen [3][4][5] ...
... Zoantharians of the genus Palythoa are popular in the saltwater aquarium trade due to their bright coloration, ease of propagation, comparatively low cost and wide availability (Deeds et al. 2011), but at the same time also produce one of the most potent marine toxins, palytoxin (PTX) (Moore and Scheuer 1971). Retailers frequently purchase unknown species of Palythoa based on coloration from suppliers who provide vague documentation of origin and rarely have robust species identifications (Deeds et al. 2011). ...
... Zoantharians of the genus Palythoa are popular in the saltwater aquarium trade due to their bright coloration, ease of propagation, comparatively low cost and wide availability (Deeds et al. 2011), but at the same time also produce one of the most potent marine toxins, palytoxin (PTX) (Moore and Scheuer 1971). Retailers frequently purchase unknown species of Palythoa based on coloration from suppliers who provide vague documentation of origin and rarely have robust species identifications (Deeds et al. 2011). This lack of information poses serious threats to aquarists because it is unknown what species are currently being distributed through the aquarium trade, their potential toxicity, or their human health risks. ...
... This lack of information poses serious threats to aquarists because it is unknown what species are currently being distributed through the aquarium trade, their potential toxicity, or their human health risks. There have been multiple reports of accidental poisoning through the marine aquarium trade, however inconsistent identification and taxonomic uncertainty of Palythoa species impede our understanding of the risks to aquarium hobbyists (Hoffmann et al. 2008;Deeds and Schwartz 2010;Deeds et al. 2011;Tartaglione et al. 2016). Due to the current uncertain state of Palythoa taxonomy, it is unclear which species are being exported and imported around the world, along with the distribution of PTX among these species. ...
Article
Full-text available
Zoantharians (Cnidaria: Hexacorallia: Zoan-tharia) of the genus Palythoa are ubiquitous species that occupy reef habitats in every tropical ocean. Disagreements among classifications based on morphology, reproductive traits, and molecular techniques have generated taxonomic challenges within this group. Molecular studies provide limited phylogenetic resolution between species, and dis-cordance is frequently attributed to slow mitochondrial rates and lack of resolution among molecular markers. Here we conducted the first phylogenomic survey of Pa-lythoa, using a reduced representation genomic approach (ezRAD) to resolve relationships among eight described and four putative Palythoa species (N = 22 plus two out-groups) across the Pacific and Atlantic Oceans. We constructed nearly complete mitochondrial genomes and assembled transcriptome loci datasets by reference mapping. A de novo assembly was performed for the holobiont dataset, and we compared a range of filtering strategies from unfiltered data down to 136 unlinked high-quality biallelic SNPs shared by all samples to resolve evolutionary lineages within Palythoa. Across all these datasets, the resulting Bayesian and ML trees revealed six highly con-cordant and well-supported clades, however, the phyloge-nomic data were inconclusive in resolving species relationships within the clades. We detected putative species complexes within two well sampled Palythoa clades (clades I and II), but species delimitation results were inconsistent in whether these clades contain multiple nominal species or represent a single variable species. Polyphyly in the broadly distributed species Palythoa tuberculosa and P. mutuki highlight the need for additional study. Consistency among nuclear and mitogenomic data-sets points to a lack of biological understanding of species boundaries among these zoantharians rather than limitations of the molecular markers. More complete taxonomic sampling of nominal species across the geographic ranges of distribution is necessary to resolve species boundaries and evolutionary histories among members of this genus.
... Zoantharians of the genus Palythoa are popular in the saltwater aquarium trade due to their bright coloration, ease of propagation, comparatively low cost and wide availability (Deeds et al. 2011), but at the same time also produce one of the most potent marine toxins, palytoxin (PTX) (Moore and Scheuer 1971). Retailers frequently purchase unknown species of Palythoa based on coloration from suppliers who provide vague documentation of origin and rarely have robust species identifications (Deeds et al. 2011). ...
... Zoantharians of the genus Palythoa are popular in the saltwater aquarium trade due to their bright coloration, ease of propagation, comparatively low cost and wide availability (Deeds et al. 2011), but at the same time also produce one of the most potent marine toxins, palytoxin (PTX) (Moore and Scheuer 1971). Retailers frequently purchase unknown species of Palythoa based on coloration from suppliers who provide vague documentation of origin and rarely have robust species identifications (Deeds et al. 2011). This lack of information poses serious threats to aquarists because it is unknown what species are currently being distributed through the aquarium trade, their potential toxicity, or their human health risks. ...
... This lack of information poses serious threats to aquarists because it is unknown what species are currently being distributed through the aquarium trade, their potential toxicity, or their human health risks. There have been multiple reports of accidental poisoning through the marine aquarium trade, however inconsistent identification and taxonomic uncertainty of Palythoa species impede our understanding of the risks to aquarium hobbyists (Hoffmann et al. 2008;Deeds and Schwartz 2010;Deeds et al. 2011;Tartaglione et al. 2016). Due to the current uncertain state of Palythoa taxonomy, it is unclear which species are being exported and imported around the world, along with the distribution of PTX among these species. ...
Article
Full-text available
Zoantharians (Cnidaria: Hexacorallia: Zoantharia) of the genus Palythoa are ubiquitous species that occupy reef habitats in every tropical ocean. Disagreements among classifications based on morphology, reproductive traits, and molecular techniques have generated taxonomic challenges within this group. Molecular studies provide limited phylogenetic resolution between species, and discordance is frequently attributed to slow mitochondrial rates and lack of resolution among molecular markers. Here we conducted the first phylogenomic survey of Palythoa, using a reduced representation genomic approach (ezRAD) to resolve relationships among eight described and four putative Palythoa species (N = 22 plus two outgroups) across the Pacific and Atlantic Oceans. We constructed nearly complete mitochondrial genomes and assembled transcriptome loci datasets by reference mapping. A de novo assembly was performed for the holobiont dataset, and we compared a range of filtering strategies from unfiltered data down to 136 unlinked high-quality biallelic SNPs shared by all samples to resolve evolutionary lineages within Palythoa. Across all these datasets, the resulting Bayesian and ML trees revealed six highly concordant and well-supported clades, however, the phylogenomic data were inconclusive in resolving species relationships within the clades. We detected putative species complexes within two well sampled Palythoa clades (clades I and II), but species delimitation results were inconsistent in whether these clades contain multiple nominal species or represent a single variable species. Polyphyly in the broadly distributed species Palythoa tuberculosa and P. mutuki highlight the need for additional study. Consistency among nuclear and mitogenomic datasets points to a lack of biological understanding of species boundaries among these zoantharians rather than limitations of the molecular markers. More complete taxonomic sampling of nominal species across the geographic ranges of distribution is necessary to resolve species boundaries and evolutionary histories among members of this genus.
... Zoantharians also engage in interspecific competition. The zoantharian Palythoa caribaeorum is a notoriously aggressive competitor, due to its fast and continuous growth ratewhich is far superior to that of co-ocurring scleractinian corals - (Silva et al., 2015), and its ability to produce an allelochemical known as palytoxin (Gleibs et al., 1995;Deeds et al., 2011). ...
Article
Full-text available
A R T I C L E I N F O Keywords: 18th century Artificial reef Crustose coralline algae Turf algae Palythoa caribaeorum Millepora spp. A B S T R A C T With increasing maritime activities in the proximity of coral reefs, a growing number of manmade structures are becoming available for coral colonisation. Yet, little is known about the sessile community composition of such artificial reefs in comparison with that of natural coral reefs. Here, we compared the diversity of corals and their competitors for substrate space between a centuries-old manmade structure and the nearest natural reef at St. Eustatius, eastern Caribbean. The artificial reef had a significantly lower species richness and fewer competitive interactions than the natural reef. The artificial reef was dominated by a cover of crustose coralline algae and zoantharians, instead of turf algae and fire corals on the natural reef. Significant differences in species composition were also found between exposed and sheltered sites on both reefs. Our study indicates that even a centuries-old manmade reef cannot serve as a surrogate for natural reefs.
... Since these corals are traded and sold without distinction as to their origin, identifying the presence of toxin based on the specific coral is not possible without genetic testing, and all zoanthid corals should be presumed to have toxic potential. 5,6 Human exposure to the toxin is rare but has occurred with increased frequency within the last decade. Exposure can occur through direct contact with the skin or eyes, inhalation, or ingestion. ...
Article
Full-text available
Palytoxin is one of the most lethal natural toxins ever discovered. This molecule has been isolated from various marine animals, including zoanthid corals. This popular organism is commonly found in many home saltwater aquariums due to its beauty and survivability. As a result of an increase in popularity, an increased number of individuals are at risk for exposure to this potentially deadly toxin. Affected patients may experience various symptoms based on the route of exposure (ie, cutaneous contact, inhalation of aerosolized toxin, ocular exposure, or ingestion). Ocular exposure can occur in various ways (eg, contact with contaminated water, rubbing the eye with a dirtied hand, or direct spraying into the eye), and incidence rates have dramatically risen in recent years. In this review, we discuss a case of systemic toxicity from inhalation and ocular exposure to presumed palytoxin on a zoanthid coral which resulted in an intensive care unit (ICU) stay, and corneal perforation which required a corneal transplant. Additionally, we review what is known about the mechanism of action of this toxin, propose a comprehensive hypothesis of its effects on corneal cells, and discuss the prognosis and clinical management of patients with systemic symptoms secondary to other routes of exposure.
... ovata have been frequently reported in temperate areas, such as the Mediterranean Sea and the Atlantic coasts of Portugal, and were often associated with adverse effects in the respiratory tract, eyes, and skin [12][13][14][15][16]. In addition, increasing reports of adverse effects after inhalational and/or cutaneous exposure to water and/or vapors from aquaria containing Palythoa and Zoanthus corals-widely used as decorative elements-are documented worldwide [17,18]. On the other hand, the main problem in tropical areas is represented by PLTX accumulation in edible marine organisms, the consumption of which has been associated with a series of severe human poisonings, sometimes with fatal outcomes. ...
Article
Full-text available
The marine polyether palytoxin (PLTX) is one of the most toxic natural compounds, and is involved in human poisonings after oral, inhalation, skin and/or ocular exposure. Epidemiological and molecular evidence suggest different inter-individual sensitivities to its toxic effects, possibly related to genetic-dependent differences in the expression of Na+/K+-ATPase, its molecular target. To identify Na+/K+-ATPase subunits, isoforms correlated with in vitro PLTX cytotoxic potency, sensitivity parameters (EC50: PLTX concentration reducing cell viability by 50%; Emax: maximum effect induced by the highest toxin concentration; 10−7 M) were assessed in 60 healthy donors’ monocytes by the MTT (methylthiazolyl tetrazolium) assay. Sensitivity parameters, not correlated with donors’ demographic variables (gender, age and blood group), demonstrated a high inter-individual variability (median EC50 = 2.7 × 10−10 M, interquartile range: 0.4–13.2 × 10−10 M; median Emax = 92.0%, interquartile range: 87.5–94.4%). Spearman’s analysis showed significant positive correlations between the β2-encoding ATP1B2 gene expression and Emax values (rho = 0.30; p = 0.025) and between Emax and the ATP1B2/ATP1B3 expression ratio (rho = 0.38; p = 0.004), as well as a significant negative correlation between Emax and the ATP1B1/ATP1B2 expression ratio (rho = −0.30; p = 0.026). This toxicogenetic study represents the first approach to define genetic risk factors that may influence the onset of adverse effects in human PLTX poisonings, suggesting that individuals with high gene expression pattern of the Na+/K+-ATPase β2 subunit (alone or as β2/β1 and/or β2/β3 ratio) could be highly sensitive to PLTX toxic effects.