First Bloom Benthic

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Marine Pollution Bulletin 158 (2020) 111313

Contents lists available at ScienceDirect

Marine Pollution Bulletin


journal homepage: www.elsevier.com/locate/marpolbul

The first benthic harmful dinoflagellate bloom in China: Morphology and T


toxicology of Prorocentrum concavum

Jian Zoua,b,1, Qun Lia,b,1, Songhui Lua,b,c, , Yuelei Donga,b, Heng Chena,b, Chengzhi Zhenga,b,

Lei Cuia,b,
a
Research Center of Harmful Algae and Marine Biology, Jinan University, Guangzhou 510632, China
b
Key Laboratory of Eutrophication and Red Tide Prevention of Guangdong Higher Education Institutes, Jinan University, Guangzhou 510632, China
c
Southern Marine Science and Engineering Guangdong Laboratory, Zhuhai, China

A R T I C LE I N FO A B S T R A C T

Keywords: More frequent events and geographic expansion of benthic harmful algal blooms have been reported in recent
Benthic harmful algae bloom years. An unexpected bloom of benthic P. concavum occurred in Xincun Bay, Hainan Island, the South China Sea
Prorocentrum concavum was monitored in August 2018. Species identification, toxin analysis and toxicity test were conducted in the
Morphology study. Quantitative study revealed that P. concavum had a high cell density on the surface of substrates and in
Phylogenetic analysis
water column. The bloom forming species was identified based on the morphology and phylogeny. Toxin
Toxicity assessments
analysis indicated that there were no detectable DSP toxins either in algae or in shellfish samples. The result of
toxicity test revealed that the extracts of P. concavum caused the mortality of brine shrimp larvae (Artemia
salina). The results from this study may provide more insight into the rising threats of harmful dinoflagellate
blooms to marine benthic ecosystems.

1. Introduction nine are benthic (An et al., 2010; Holmes et al., 2001; Hoppenrath
et al., 2013; Luo et al., 2017; Murakami et al., 1982; Nascimento et al.,
Benthic harmful algal blooms (BHABs) are emergent marine disaster 2016). There were also reports from some species of Prorocentrum for
to affect benthic ecosystem (Berdalet et al., 2017; GEOHAB, 2012). production of various bioactive compounds (Faust and Gulledge, 2002).
Differed with the planktonic blooms, BHABs are induced mainly by Though Prorocentrum is one group of species for production of toxins
epibenthic dinoflagellates (Accoroni and Totti, 2016; Litaker et al., responsible for diarrhetic shellfish poisoning (DSP) in humans through
2010) and negatively altering marine benthic ecosystem (Accoroni food train, reports of toxic blooms and DSP incidents related to benthic
et al., 2016; Mangialajo et al., 2017). The most reported harmful Prorocentrum species were scarce (Foden et al., 2005; Gayoso et al.,
benthic dinoflagellates are Gambierdiscus, which are producers of toxin 2002). It is partly because of the difficulty to detect the outbreaks of
ciguatera (Chan, 2015; Litaker et al., 2010; Russell and Egen, 1991), toxic Prorocentrum species in the benthic nature.
and Ostreopsis, which are origin of ovatoxins and its derivatives (Amzil P. concavum is one of the toxic Prorocentrum species (Dickey et al.,
et al., 2012; Tibiriçá et al., 2019). 1990; Hu et al., 1992). It was originally described by Fukuyo (1981) in
The genus Prorocentrum, one of the largest groups in dinoflagellates, French Polynesia, New Caledonia and Ryukyu Islands and reexamined
was established in 1834 by Ehrenberg with Prorocentrum micans as type by Mohammad-Noor et al. (2007). This species was synonymized as P.
species (Ehrenberg, 1834). So far 75 species has been identified, of arabianum (Mohammad-Noor et al., 2007) and P. faustiae (Chomérat
which around 30 species are associated with benthic habitats et al., 2019). The production of OA by P. concavum was first reported in
(Hoppenrath et al., 2013; Lim et al., 2019; Rodríguez et al., 2018). a culture strain from Caribbean (Dickey et al., 1990). The production of
Some presumable epibenthic species (e.g., P. arabianum Morton & Faust OA and diol ester derivatives of OA has also been confirmed from ex-
(=P. concavum Fukuyo)) are tychoplanktonic, which can also be col- tracts of culture strains of P. concavum (Hu et al., 1992). Some studies
lected from the plankton (Morton et al., 2002). So far ten species from also reported that P. concavum had cytotoxic, ichthyotoxic, and hae-
Prorocentrum has been reported to produce diarrhetic toxins, okadaic molytic activity, which were lethal to mice (Morton et al., 2002;
acid (OA) and the methyl derivatives dinophysistoxins (DTXs), of which Yasumoto et al., 1987). However, it seems that the toxin productions


Corresponding authors at: Research Center for Harmful Algae and Marine Biology, Jinan University, Guangzhou 510632, China.
E-mail addresses: lusonghui1963@163.com (S. Lu), leicui@jnu.edu.cn (L. Cui).
1
These authors contributed equally.

https://doi.org/10.1016/j.marpolbul.2020.111313
Received 7 December 2019; Received in revised form 19 May 2020; Accepted 21 May 2020
Available online 17 June 2020
0025-326X/ © 2020 Elsevier Ltd. All rights reserved.
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

and the toxicity effects of P. concavum are probably strains specific. A Ostreopsis, Gambierdiscus, Coolia, and Amphidinium) based on their
recent study on P. concavum revealed that there were no OA, DTX-1 and morphologic characteristics. The final bloom density data were ex-
DTX-2 detected in a tropical strain from Australia (Verma et al., 2019). pressed as cells g−1 wet weight for the seaweed.
The toxin analysis of four strains of P. concavum collected from tropical Live P. concavum cells were surveyed and photographed under a
waters of Hainan Island, China indicated that all strains did not produce fluorescence microscope (Olympus BX 61, Olympus Corporation,
detectable level of OA and DTX-1 (Luo et al., 2017). Tokyo, Japan) coupled with a QImaging Retiga 400R digital camera
There were numerous studies which revealed high diversity of (QImaging, Surrey, BC, Canada). Over forty P. concavum cells were
Prorocentrum in tropical benthic reef ecosystems (Irola-Sansores et al., measured at 400× magnification using Image-Pro Plus v. 6.0 (Media
2018; Lim et al., 2019; Verma et al., 2019). Previous study revealed that Cybernetics, Rockville, MD, USA). The cells stained with SYBR Green
there were seven Prorocentrum species, P. lima, P. rhathymum, P. con- were used to observe the shape and location of the nucleus and chlor-
cavum, P. cf. emarginatum, P. fukuyoi, P. cf. maculosum (synonymized oplasts. To collect the cells for scanning electron microscopy (SEM), a
with P. caipirignum by Nascimento et al. (2017)) and P. panamense have 1 ml exponential culture was centrifuged in a Sorvall Biofuge Primo R
been described from the northern South China Sea, of which OA was (Thermo Fisher Scientific, Waltham, MA, USA) at 5000 × g for 2 min at
detected in all strains of P. lima and P. caipirignum (Luo et al., 2017). 25 °C. The supernatant was removed, and the pellet was cleaned with
During our routine benthic dinoflagellate survey conducted on the filtered seawater. The pellet was fixed, dehydrated, and then coated
tropical Hainan Island, China in 2018, a high density of P. concavum as described by Zhang et al. (2015). Then, the samples were photo-
bloom on surface of the substrates was encountered in Xincun Bay, graphed with a scanning electron microscope (Zeiss ULTRA™ 55, Carl
southeast of Hainan Island. The bloom was the first identified benthic Zeiss Inc., Oberkochen, Germany).
bloom in China. In this paper, the benthic bloom was described. The
detailed morphological observations of the bloom forming species, to- 2.4. PCR amplifications and phylogenetic analyses
gether with phylogenetics have been studied. The environmental
parameters where the bloom occurred were monitored. The culture A unialgal strain of P. concavum (HN437) was isolated from the live
strain of the species was succeeded established and the toxicity of the samples using the methods of Luo et al. (2017). Isolated P. concavum
species was tested. cells were grown in filtered seawater with the L1 medium (Guillard and
Hargraves, 1993). Unialgal cultures were incubated at 25 °C under a
2. Materials and methods 12/12 h light/dark regime with an irradiance of 100 μmol pho-
tons m−2 s−1. The exponential cells of P. concavum were centrifuged at
2.1. Site description and sample collection 10,000 ×g for 2 min at 4 °C. The DNA of P. concavum was extracted and
amplified as described Luo et al. (2017). The D1-D3 regions of large
Xincun Bay is a tropical lagoon located in the southeast of Hainan subunit (LSU) was amplified using general primers (D1R: 5′-ACCCGC
Island, South China Sea (Fig. 1). The bay (area of 13.1pkm2) is almost TGAATTTAAGCATA-3′, D3B: 5′-TCGGAGGGAACCAGCTACTA-3′)
closed, with only one narrow canal connected to the open sea (Yang and (Scholin et al., 1994). The PCR procedure was as follows: an initial
Yang, 2009). The substrata of Xincun Bay mainly contains seagrasses, denaturation step at 94 °C for 4 min; 36 cycles of denaturation at an-
macroalgae, silt, and sand, which are suitable for the growth of benthic nealing at 94 °C for 20 s, annealing at 56 °C for 30 s, and extension at
dinoflagellates (Fig. 2). In this study, samples were collected during the 72 °C for 45 s; and a final extension of 7 min at 72 °C. In addition,
benthic dinoflagellate bloom at site 1, where there are lots of seagrasses internal transcribed spacer (ITS) was amplified using the general pri-
(Thalassia hemprichii and Enhalus acoroides) and seaweed (Ulva lactuca). mers (ITS1F: 5′-TCGTAACAAGGTTTCCGTAG-3′, ITS1R: ATATGCTTA
Site 2 located in the mangroves without the presence of the bloom, was AGCTCAGCGGG) (Pin et al., 2001). PCR conditions composed 3 min at
designated as the reference site. Triplicate water samples were collected 94 °C followed by 40 cycles of 30 s each at 94 °C, 30 s at 56 °C, 1.5 min
using 100 ml sample bottles near the benthic substrates. The macro- at 72 °C, and a final extension of 10 min at 72 °C. Sequencing was
phyte samples were collected in three replicates at a depth of 0.4 m. The conducted in both directions, using an ABI PRISM 3730XL (Applied
epibenthic microalgae were washed several times using fresh seawater Biosystems, Foster City, CA, USA). The partial LSU and ITS sequences of
to ensure they were entirely separated from the macrophyte. Some live P. concavum from Xincun Bay were deposited into the GenBank data-
cells were transferred into a polycarbonate bottle and others reserved in base under the accession numbers MN698951 and MN699564, re-
the 15 ml centrifuge tubes. After the tubed samples were preserved in a spectively. The LSU and ITS sequences were aligned using the multiple
1.5% Lugol's solution, the fresh seaweed was weighed. In addition, we sequence alignment program (https://www.ebi.ac.uk/Tools/msa/
collected the field benthic microalgae assemblages and two shellfish muscle/). The aligned results were inputted into maximum likelihood
(Callista erycina and Gafrarium pectinatum) in the study area. (ML) analyses using RAxML-HPC2 and XSEDE v. 8.2.10 on the CIPRES
website (https://www.phylo.org/). In addition, the best models of the
2.2. Hydrographic and nutrient parameters LSU and ITS sequences selected by PAUP (version 4.0) were SYM + G
and Bayesian Inference (BI), respectively, for which MrBayes (version
The water temperature, pH, and salinity were measured using a YSI 3.0) was used and run for 1,000,000 generations with sampling every
meter (YSI-6600; YSI Incorporated, USA). The concentrations (three 100 generations.
replicates) of total nitrogen (TN), total phosphate (TP), nitrate (NO3-N),
nitrite (NO2-N), ammonium (NH3−N), phosphate (PO4-P), and silicate 2.5. Extraction and detection of DSP toxins
(Si) were measured using spectrophotometry according to the methods
of Sagi (1966), Murphy and Riley (1962), Bendschneider (1952), Ebina Field samples of benthic microalgal assemblages attached to mac-
et al. (1983), and Mullin and Riley (1955), respectively. rophytes were concentrated and collected from Xincun Bay. The con-
centrated samples were centrifuged at 10,000 ×g for 10 min. Then,
2.3. Cell counting and morphological observations 10 ml of methanol was added to the pellets, before being ultrasonicated
thrice, for 20 min each time. The cells slurry was observed in an in-
Epibenthic dinoflagellates were identified and counted using an verted microscope and centrifuged at 10,000 ×g for 15 min. Samples of
inverted microscope (ECLIPSE TE 2000-U, Nikon, Japan) at 100× shellfish were collected from sampling site 1. Wet tissue samples of
magnification. The identification of P. concavum was based on the Callista erycina and Gafrarium pectinatum (100 g each) were individually
morphology of the periflagellar area and dimensions. Other benthic homogenized using a T25 Ultra Turrax mixer at 24,000 rpm (IKA
dinoflagellates were differentiated into five genera (Prorocentrum, Works, Wilmington, NC, USA). The homogenates were extracted and

2
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 1. Sampling sites in Xincun Bay, South China Sea. Site 1 is in the seagrass bed; and Site 2 is in the mangroves, designated as the reference site.

purified according to the methods of Liu et al. (2019). To remove the chromatography-tandem mass spectrometry (LC-MS/MS), as described
disturbance of DSP toxin analogues, 1 ml of each pure extracted solu- by Liu et al. (2019). The detection limits of OA, DTX-1, and DTX-2 were
tion was hydrolysed by 125 μl of 2.5 mol l−1 NaOH at 76 °C for 40 min. 1.53, 3.04, and 1.34 ng ml−1, respectively.
Then, the solutions were neutralized by adding 125 μl of 2.5 mol ml−1
HCl. 2.6. Toxicity assessments
For the laboratory experiment, P. micans (strain: HSJ19) and P. lima
(strain: XS326), obtained from Algal Culture Collection, Research Dried brine shrimp (Artemia salina) cysts were incubated in filtered
Center of Harmful Algae and Marine Biology, Jinan University, artificial seawater (pore size: 0.22 μm; salinity: 30 ± 1 ppt) for 24 h.
Guangzhou, China, were selected as the negative and positive control, Then, the cysts were aerated and maintained at 25 ± 1 °C,
respectively. Three litre mid-exponential cultures of P. concavum, P. 100 μmol photons m−2 s−1, under a 12 h light: 12 h dark cycle. The
micans, and P. lima were individually cultivated and extracted ac- brine shrimp larvae were transferred into 12-well tissue culture plates
cording to the methods of Luo et al. (2017). The field and cultured for the toxic bioassay. The toxicity of the field and cultured P. concavum
pellets of P. concavum, P. lima, and P. micans were discarded and the samples were examined using bioassays with brine shrimp larvae. The
supernatants were collected and dried with N2. Dried algae sediments concentration series of the treatment was 2.5 × 102, 2.5 × 103,
were suspended separately in 5 ml of methanol and filtered with a spin- 2.5 × 104, 5 × 104, and 2.5 × 105 cells ml−1. The 4 ml algal ex-
filter (0.22 μm pore-size; Millipore Ultrafree, Eschborn, Germany). tractions were dried with N2 and suspended using 1% Tween-80 dis-
Three lipophilic DSP toxins (OA, DTX-1, and DTX-2) of P. concavum, P. solved with phosphate-buffered solution for the toxicity test. The arti-
micans, P. lima, and two shellfish were detected using liquid ficial seawater, solvent, and negative and positive controls consisted of

3
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 2. Habitat during the P. concavum bloom in Xincun Bay, South China Sea. A, B, C, and D exhibit the seagrass bed, Enhalus acoroides, Thalassia hemprichii, and Ulva
lactuca, respectively.

filtered seawater, 1% Tween-80, and cultured P. micans and P. lima, while U. lactuca was the lowest (3.2 × 104 cells g−1 wet weight)
respectively. The concentrations of the solvent controls were diluted (Fig. 3). Similarly, the water column of the seaweed beds contained a
according to the concentrations of the algal lysate solution. The con- high amount of P. concavum cells (1.7 × 104 cells l−1). The total ratios
centrations of the negative and positive controls were similar to those of of P. concavum were 93.5%, 99.7%, 97.0%, and 99.7% in the T. hem-
the treatment. Twenty brine shrimp larvae were assigned to individual prichii, E. acoroides, Ulva lactuca and water column samples, respec-
wells of 12-well tissue culture plates and 2 ml of filtered seawater tively. In addition to P. concavum, four potentially toxic benthic dino-
(salinity: 30 ppt) was added for the toxicity test (48 h test). They were flagellates (P. rhathymum, Coolia spp., Ostreopsis spp., and Amphidinium
maintained at 25 ± 1 °C, 100 μmol photons m−2 s−1, under a 12/12 h spp.) were found on these substrata and in the water column. As shown
light/dark cycle. The mortality and abnormality rates were recorded at in Table 1, the water temperature, pH value, and salinity were 28.0 °C,
48 h. Deceased individuals were identified by a lack of movement in the 8.40, and 31.92 psu in the bloom area, respectively. Moreover, all of the
larvae appendages. Abnormal individuals were classified by abnormal nutrient concentrations were lower in the bloom area than those in the
swimming behaviours or immobility (Leung et al., 2017). non-bloom area.

2.7. Statistical analysis


3.2. Morphology of P. concavum
The concentrations that led to a mortality of 50% of the field and
Exhibiting the typical morphological characteristics of P. concavum,
cultured P. concavum were counted using a sigmoidal-log mode curve
the organism was symmetrically oval and had two pusules (Pu) situated
formed by Origin (version 9.60) (OriginLab software, USA). The sta-
towards the anterior of the cells (Fig. 4A and B). The length (L) and
tistical results were analysed using IBM SPSS Statistics 25 (IBM cor-
width (W) of the cells were 43.5–55.6 (mean = 48.84 ± 2.7 μm,
poration, Armonk, NY, USA).
n = 41) and 34.3–45.8 μm (mean = 40.95 ± 2.5, n = 41), respec-
tively. In addition, the L/W ratio ranged from 1.11 to 1.30. The
3. Results chloroplasts (Chl) were well-distributed with central pyrenoids (P), as
shown in Fig. 4C. The oblong nucleus (N) was located in the posterior of
3.1. Bloom detection the cell (Fig. 4D). SEM micrographs of P. concavum displayed the
foveate thecal plates of the cells (Fig. 4E, F, and G). The plates were
The benthic Prorocentrum bloom was observed by chance on 27th ornamented with numerous depressions that decorated many pores
August 2018 in the south of Xincun Bay, Hainan Island. During the (Fig. 4H). The anterior V–shaped periflagellar area consisted of a fla-
bloom period, the seagrasses and macroalgae were covered with nu- gella pore and accessory pore, as well as nine platelets (1a, 1b, 2, 3, 4,
merous Prorocentrum cells (Fig. S1). Among them, the density of P. 5, 6, 7, and 8), according to the periflagellar descriptions of Hoppenrath
concavum on T. hemprichii was the highest (3.0 × 105 cells g−1 wet et al. (2013) (Fig. 4I).
weight), followed by E. acoroides (2.2 × 105 cells g−1 wet weight),

4
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 3. Abundance of benthic dinoflagellates on the seagrasses, on the macroalgae, and in the water column during the bloom period. The unit of abundance for P.
concavum, Thalassia hemprichii, and Enhalus acoroides is cells g−1 wet weight and the unit of abundance in the water column is cells l−1.

Table 1 3.4. Detection of toxins causing DSP


Features of the bloom in Xincun Bay.
Parameters Site 1 (bloom area) Site 2 (non-bloom area)
LC-MS/MS experiments were conducted on the extracts of the field
and cultured samples using external calibration. There were no obvious
Temperature (°C) 28.0 – peaks observed in the extracts of the field and cultured P. concavum
Salinity (psu) 31.92 – samples (Fig. S3). Similarly, no DSP toxin signals were detected in the
pH 8.40 –
NO3-N (μmol l−1) 1.04 ± 0.04 1.90 ± 0.14
shellfish samples (C. erycina and G. pectinatum) or cultured P. micans.
NO2-N (μmol l−1) 0.02 ± 0.00 0.13 ± 0.04 However, OA and DTX-1 were detected in P. lima. The concentrations of
NH3-N (μmol l−1) 12.69 ± 1.83 15.77 ± 0.97 OA and DTX-1 for P. lima were 1871.31 and 80.88 pg cell−1, respec-
PO4-P (μmol l−1) 0.90 ± 0.11 1.19 ± 0.02 tively.
SiO3-Si (μmol l−1) 9.03 ± 1.22 12.04 ± 0.46
TN (μmol l−1) 17.74 ± 3.71 22.44 ± 0.40
TP (μmol l−1) 0.718 ± 0.22 0.83 ± 0.17 3.5. Toxicity of P. concavum lysate to brine shrimp

A lethal effect was detected in the field and cultured samples (P.
3.3. Molecular phylogeny concavum, P. lima, and P. micans), following the concentrations of
2.5 × 102, 2.5 × 103, 2.5 × 104, 5 × 104, and 2.5 × 105 cells ml−1
To determine the phylogenetic position of P. concavum HN437, 69 (Fig. 6). Brine shrimp larvae mortality rates of over 50% were observed
and 52 sequences of LSU and ITS of the genus Prorocentrum were ob- for the field samples and cultured P. concavum in the concentrations of
tained from the GenBank, respectively. The LSU sequence of Adenoides 2.5 × 104, 5 × 104, and 2.5 × 105 cells ml−1. For cultured P. lima and
eludens ADE15 France (LC002847) and ITS sequence of Karena brevis P. micans, high mortality rates (> 80%) were observed for the treat-
CCMP719 USA (AF352827) were used for outgroups. Phylogenetic trees ment with P. lima, only at the of concentration of 2.5 × 105 cells ml−1.
inferred from the ML and BI methods were constructed for the nu- Based on the 48 h treatment of the field samples and cultured P. con-
cleotide sequences of LSU and ITS, respectively (Figs. 5 and S2). The cavum and P. lima, mortality rates of 50% (LC50 value) were estimated
phylogenetic tree based on LSU sequences revealed two main clades (A using sigmoidal log-response curves (Fig. 7). The 48 h-LC50 values were
and B) (Fig. 5). The clade A consisted of 13 species, which consisted of 6 1.50 × 104 cells ml−1 for the field samples, 1.54 × 104 cells ml−1 for
kinds of benthic species (P. elegans, P. fukuyoi, P. scupltile, P. emargi- the cultured P. concavum, and 5.34 × 104 cells ml−1 for the cultured P.
natum, P. tsawwassenense and P. rhathymum) and 7 kinds of planktonic lima.
species (P. dentatum, P. minimum, P. triesinum, P. koreanum, P. micans, P.
texanum and P. gracile). Eleven species of Prorocentrum were consisted
4. Discussion
of the clade B, which were benthic species except for P. playfairi. P.
concavum, which branches into three independent sub-branches, shares
4.1. The first P. concavum bloom in Chinese coastal waters
a close ancestry with P. leve. The first sub-branch consisted of the strains
IFR12–251 (MG701855) and NMN013 (EF566744), isolated from
In this study, the greatest density of P. concavum was
Martinique and Malaysia, respectively. The HN437 P. concavum strain
3.0 × 105 cells g−1 wet weight on T. hemprichii. Simultaneously, P.
and the four P. concavum strains isolated from Hainan Island
concavum on E. acoroides and U. lactuca also had a high density
(KY010227, KY010228, KY010226, and KY010229) comprised the
(2.2 × 105 cells g−1 wet weight and 3.2 × 104 cells g−1 wet weight,
second sub-branch. The last sub-branch included other strains isolated
respectively). The density of P. concavum in Xincun Bay was much
from China (AJ567464), Malaysia (EF566751), and the Arabian Sea
higher than other Prorocentrum blooms of previous studies (e.g. the P.
(EF566752). Moreover, the phylogenetic analysis results based on ITS
lima blooms (approximately 4.5 × 104 cells g−1 wet weight) on mac-
regions were consistent with the phylogenetic tree based on LSU se-
roalgae in the French coast and the P. lima (1.0 × 104 and
quences. The P. concavum strain HN437 was the sister clade of the other
3.0 × 104 cells g−1 wet weight, respectively) on Phaeophyta in
four P. concavum strains (KY010227, KY010228, KY010226, and
Caribbean Sea and on Thalassia in Gulf of Mexico) (Blanfuné et al.,
KY010229) isolated from Hainan Island (Fig. S2).
2015; Delgado et al., 2006). The average ratio of P. concavum

5
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 4. Light and scanning electron micrographs of P. concavum. (A and B) Light microscopy (LM); (A) a live field cell showing the shape of the cell and pusule, and
(B) a live cultured cell showing the shape of the cell. (C and D) Fluorescence LM; (C) a cell showing the location of the pyrenoids (P) and arrangement of the
chloroplasts (Chl), and (D) a cell stained with SYBR Green exhibiting the shape and location of the nucleus (N). (E–I) SEM; (E and F) the right and left valve showing
the V-shaped periflagellar area, countless depressions, and thecal pores, (G) the intercalary band showing the transverse striations, (H) the details of the thecal pores,
and (I) the details of the periflagellar area showing nine platelets, the flagella pore (fp), and the accessory pore (ap). Scale bars: 10 μm (A–G), 1 μm (H and I). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

abundance was 96.7% in the benthos and even reached 99.7% in the Totti, 2016). Thus, the same phenomenon could exist in P. concavum
water column. Additionally, we found that all of the nutrient con- and Prorocentrum other species. Due to occasional blooms and diffi-
centrations were lower in the bloom area than in the non-bloom area, culties in sampling, there were little reports paying attention to benthic
indicating that the numerous P. concavum exhausted lots of nutrients Prorocentrum so far. As one of the main results in present study, the
during the bloom. Thus, it is rare that the P. concavum bloom in Chinese benthic P. concavum bloom was firstly reported in the substrate and
coastal waters had a high density and amount. Although the P. ara- water column in Xincun Bay. The P. concavum bloom has not been
bianum (sym. P. concavum) bloom was reported previously in the water noted as present in China before and the abundance of the causative
column in the Gulf of Oman, 1995, this study did not concern the species had not previously been reported anywhere in the world during
concentration and hazard of the bloom (Morton et al., 2002). Besides, P. blooms as well. Therefore, it need further attention that if this benthic
arabianum was found in water column and was originally reported as a P. concavum bloom occurred again in this region.
planktonic species. As other benthic dinoflagellates (e.g. Ostreopsis), the
cell may detach from the substrate to water column by the hydro-
graphic conditions such as tide, wave and turbulence (Accoroni and

6
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 5. Phylogenetic tree for the genus Prorocentrum based on the LSU nucleotide sequences, using Bayesian inference (BI) and maximum likelihood (ML) methods.
The outgroup is Adenoides eludens ADE15 France (LC002847). The numbers above the branches refer to the ML bootstrap values (left) and Bayesian posterior
probability values (right). The black dots represent support values equal to 100/1.00. The values are > 50 and 0.8 for the ML bootstrap analysis and Bayesian
posterior probabilities, respectively. The scale bar represents the substitutions per site.

7
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Fig. 6. Toxicology of the field and cultured samples (P. concavum, P. lima, and P. micans) to brine shrimp (48 h treatment). Solvent (SC) and filtered seawater (SW)
were provided for eliminating the effects of 1% Tween80 and artificial seawater.

(Calado and Hansen, 1999; Hansen et al., 2000). According to the


phylogenetic trees, the species P. concavum branched with the species P.
faustiae, P. foraminosum and P. leve (Figs. 5 and S2), which was con-
sistent with the previous phylogenetic results (Luo et al., 2017; Verma
et al., 2019). P. concavum strain HN437 closely formed a sister clade
with another P. concavum strains (DS4C10 and DS4E11) from Lingshui,
Hainan Island which was closed to this bloom area. The P. concavum
strains of Hainan Island were separated with other strains from Marti-
nique, Malaysia, and the Arabian Sea. This result supports the view that
Prorocentrum species from different geographical positions may have
morphological differences (Herrera-Sepúlveda et al., 2015; Hoppenrath
et al., 2013; Zhang et al., 2015). In the original descriptions by Morton
(1998), strain NMN013 was a new species P. faustiae which was first
found in Heron Island, Australia. However, some studies suggested that
P. faustiae was synonymized as P. concavum due to that the two species
were extremely similar in cell size, shape, pores and periflagellar area
(1a, 1b, 2, 3, 4, 5, 6, 7 and 8) (Chomérat et al., 2019; Hoppenrath et al.,
2013). Our phylogenetic analysis also revealed that the strain NMN013
branched into an independent sub-branch with the eight P. concavum
strains, shared a close ancestry with the strain of P. concavum
(IFR12–251) (Fig. 5). Hence the two species P. concavum and P. faustiae
Fig. 7. Sigmoidal-log mode curve stating the toxicity of the field and cultured
samples (P. concavum and P. lima) to brine shrimp (48 h treatment). might be conspecific based on a combination of morphology and phy-
logeny.

4.2. Identification of P. concavum


4.3. Toxicity of P. concavum
P. concavum was identified based on the morphology and phylogeny
in this study. Both the field specimens and cultured P. concavum were The three DSP toxins (OA, DTX-1, and DTX-2) were not detected in
generally consistent with the original morphological descriptions that both the cultured and field P. concavum in this study, which indicated
this species possessed a wide V-shaped periflagellar area (1a, 1b, 2, 3, 4, that the strain HN437 of P. concavum isolated from Xincun Bay might
5, 6, 7 and 8) and many depressions distributed in the thecal plates not contain the three toxins or could produce rare toxins. But it is in-
except the center (Fukuyo, 1981). In the descriptions by Mohammad- conclusive whether P. concavum could produce OA, DTX-1 and DTX-2
Noor et al. (2007), P. concavum had been reported as possessing two and there are controversial data in some researches so far. Several
types of pores, however, only one type of pore was distributed into the previous study found the extract of P. concavum could produce OA and
depressions and these pores had a different diameter in the present its diol ester derivatives (Hu et al., 1992). Similarly, OA and DTX-1 was
study. There were no pores in the margins of P. concavum and the found in P. faustiae (sym. P. concavum) by using the HPLC-fluorescent
periflagellar area was divided into nine platelets, as platelet 1 was se- method (Morton, 1998). However, P. concavum in the tropical Australia
parated into platelets 1a and 1b. These results were found in the pre- was not detected OA, DTX-1 and DTX-2 were in recent research (Verma
vious researches (Luo et al., 2017; Mohammad-Noor et al., 2007; Verma et al., 2019). Luo et al. (2017) found that P. concavum strains in waters
et al., 2019). In addition, the morphology of field specimens had similar near Xincun Bay did not produce the three DSP toxins, except for two of
characteristics with other Prorocentrum species (P. lima and P. hoff- which were detected extremely low concentration of OA
mannianum) that two obvious pusules were situated below the acces- (< 0.031 fg cells−1) below detectable level. The strain-specific varia-
sory pore (Fig. 4A) (Zhou and Fritz, 1993). Visible pusule had not been tion of toxin production may exist in P. concavum. Some report showed
observed in the cultured cells and this phenomenon could be associated that the toxin concentration of Prorocentrum could be influenced by the
with a change in the growth environment during the laboratory culture cultured conditions and this phenomenon should be considered in the

8
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

toxin investigation of P. concavum (Aquino-Cruz et al., 2018). Besides species (Prorocentrales, Dinophyceae) from Anse Dufour (Martinique Island, eastern
the reason for strains and growth conditions, possible misidentification Caribbean Sea). Mar. Biodivers. 49, 1299–1319.
Delgado, G., Lechuga-Devéze, C.H., Popowski, G., Troccoli, L., Salinas, C.A., 2006.
of strains may lead to the difference in toxin detection. For instance, Epiphytic dinoflagellates associated with ciguatera in the northwestern coast of Cuba.
Dickey et al. (1990) and Lillian et al. (1997) reported that P. concavum Rev. Biol. Trop. 54, 299–310.
could produce OA, but the species was similar to toxic species P. hoff- Dickey, R.W., Bobzin, S.C., Faulkner, D.J., Bencsath, F.A., Andrzejewski, D., 1990.
Identification of okadaic acid from a Caribbean dinoflagellate, Prorocentrum con-
mannianum based on the later morphology and phylogeny. Though no cavum. Toxicon 28, 371–377.
DSP toxins were detected in present study, the results of the toxicity to Ebina, J., Tsutsui, T., Shirai, T., 1983. Simultaneous determination of total nitrogen and
brine shrimp larvae showed that algae lysates were toxic and lethal to total phosphorus in water using peroxodisulfate oxidation. Water Res. 17,
1721–1726.
marine invertebrate larvae. The LC50 value of P. concavum was Ehrenberg, C., 1834. Organisation in der Richtung des kleinsten Raumes. Dritter Beitrag
1.5 × 104 cell ml−1, which was lower than some toxic dinoflagellate zur Erkenntnis grosser Organisationen in der Richtung des kleinsten Raumes. Abh.
species, such as Coolia malayensis and P. lima for brine shrimp larvae Akad. Wiss. 307, 307–308.
was approximately 3 × 105 cell ml−1 (Leung et al., 2017) and Faust, M.A., Gulledge, R.A., 2002. Identifying harmful marine dinoflagellates.
Smithsonian Institution.Contrib. U.S. Natl. Herbarium 42, 1–144.
5.34 × 104 cells ml−1 (present study), respectively. Thus, the bloom Foden, J., Purdie, D.A., Morris, S., Nascimento, S., 2005. Epiphytic abundance and
strain of P. concavum isolated from Xincun Bay may be velogenic. Si- toxicity of Prorocentrum lima populations in the Fleet Lagoon, UK. Harmful Algae 4,
milarly, the previous research suggested that P. concavum from Oki- 1063–1074.
Fukuyo, Y., 1981. Taxonomical study on benthic dinoflagellates collected in coral reefs.
nawa, Japan didn't produce detectable OA, but the species had pow- Bull. Jpn. Soc. Sci. Fish. 47, 967–978.
erful ichthyotoxicity (Yasumoto et al., 1987). Morton et al. (2002) also Gayoso, A.M., Dover, S., Morton, S., Busman, M., Moeller, P., Fulco, V.K., Maranda, L.,
showed that P. concavum from the Gulf Oman, Arabian Sea could induce 2002. Diarrhetic shellfish poisoning associated with Prorocentrum lima (Dinophyceae)
in Patagonian gulfs (Argentina). J. Shellfish Res. 21, 461–463.
death in mice because of cytotoxic and haemolytic activity, although GEOHAB, 2012. Global ecology and oceanography of harmful algal blooms, GEOHAB
the type of toxin was not clear. Based on above mentioned, P. concavum Core Research Project: HABs in benthic systems. In: Berdalet, E., Tester, P., Zin- gone,
from Xincun Bay may contain uncertain toxin rather than DSP toxins, A. (Eds.), IOC of UNESCO and SCOR, Paris and Newark, (64 pp).
Guillard, R.R.L., Hargraves, P.E., 1993. Stichochrysis immobilis is a diatom, not a chryso-
which were lethal to brine shrimp larvae. Hence, it is worth noting that phyte. Phycologia 32, 234–236.
this benthic P. concavum had potential harm for the benthic marine Hansen, G., Daugbjerg, N., Henriksen, P., 2000. Comparative study of Gymnodinium mi-
ecosystem and public health. In addition, the type of toxin needed for kimotoi and Gymnodinium aureolum, comb. nov (=Gyrodinium aureolum) based on
morphology, pigment composition, and molecular data. J. Phycol. 36, 394–410.
further analyses to demonstrate toxic mechanism of P. concavum in the Herrera-Sepúlveda, A., Medlin, L.K., Murugan, G., Sierra-Beltrán, A.P., Cruz-Villacorta,
future. A.A., Hernández-Saavedra, N.Y., 2015. Are Prorocentrum hoffmannianum and
Supplementary data to this article can be found online at https:// Prorocentrum belizeanum (Dinophyceae, Prorocentrales), the same species? An in-
tegration of morphological and molecular data. J. Phycol. 51, 173–188.
doi.org/10.1016/j.marpolbul.2020.111313.
Holmes, M.J., Lee, F.C., Khoo, H.W., Teo, S.L.M., 2001. Production of 7-deoxy-okadaic
acid by a new caledonian strain of Prorocentrum lima (Dinophyceae). J. Phycol. 37,
Acknowledgements 280–288.
Hoppenrath, M., Chomérat, N., Horiguchi, T., Schweikert, M., Nagahama, Y., Murray, S.,
2013. Taxonomy and phylogeny of the benthic Prorocentrum species
This work was supported by Special Foundation for National (Dinophyceae)—A proposal and review. Harmful Algae 27, 1–28.
Science and Technology Basic Research Program of China Hu, T., Marr, J., de Freitas, A.S., Quilliam, M.A., Walter, J.A., Wright, J.L., Pleasance, S.,
(2018FY100200 and 2018FY100100) and National Natural Science 1992. New diol esters isolated from cultures of the dinoflagellates Prorocentrum lima
and Prorocentrum concavum. J. Nat. Prod. 55, 1631–1637.
Foundation of China (41876173 and 41576162). Irola-Sansores, E.D., Delgado-Pech, B., Garcia-Mendoza, E., Núñez-Vázquez, E.J., Olivos-
Ortiz, A., Almazán-Becerril, A., 2018. Population dynamics of benthic-epiphytic di-
Declaration of competing interest noflagellates on two macroalgae from coral reef systems of the northern Mexican
Caribbean. Front. Mar. Sci. 5, 487.
Leung, P.T.Y., Yan, M., Yiu, S.K.F., Lam, V.T.T., Ip, J.C.H., Au, M.W.Y., Chen, C.-Y., Wai,
The authors declare no competing financial interests. T.-C., Lam, P.K.S., 2017. Molecular phylogeny and toxicity of harmful benthic di-
noflagellates Coolia (Ostreopsidaceae, Dinophyceae) in a sub-tropical marine eco-
system: the first record from Hong Kong. Mar. Pollut. Bull. 124, 878–889.
References
Lillian, Douglas, Phd, J.-M., Bs, C., Phd, S., Phd, C., 1997. Isolation and separation of
toxins produced by Gambierdiscus toxicus and Prorocentrum concavum. Journal of
Accoroni, S., Totti, C., 2016. The toxic benthic dinoflagellates of the genus Ostreopsis in Aquatic Food Product Technology 6, 5–25.
temperate areas: a review. Adv. Oceanogr. Limnol. 7, 1–15. Lim, Z.F., Luo, Z., Lee, L.K., Hii, K.S., Teng, S.T., Chan, L.L., Chomérat, N., Krock, B., Gu,
Accoroni, S., Romagnoli, T., Pichierri, S., Totti, C., 2016. Effects of the bloom of harmful H., Lim, P.T., 2019. Taxonomy and toxicity of Prorocentrum from Perhentian Islands
benthic dinoflagellate Ostreopsis cf. ovata on the microphytobenthos community in (Malaysia), with a description of a non-toxigenic species Prorocentrum malayense sp.
the northern Adriatic Sea. Harmful Algae 55, 179–190. nov.(Dinophyceae). Harmful Algae 83, 95–108.
Amzil, Z., Sibat, M., Chomerat, N., Grossel, H., Marco-Miralles, F., Lemee, R., Nezan, E., Litaker, R.W., Vandersea, M.W., Faust, M.A., Kibler, S.R., Nau, A.W., Holland, W.C.,
Sechet, V., 2012. Ovatoxin-a and palytoxin accumulation in seafood in relation to Chinain, M., Holmes, M.J., Tester, P.A., 2010. Global distribution of ciguatera
Ostreopsis cf. ovata blooms on the French Mediterranean coast. Marine Drugs 10, causing dinoflagellates in the genus Gambierdiscus. Toxicon 56, 711–730.
477–496. Liu, Y., Yu, R.-C., Kong, F.-Z., Li, C., Dai, L., Chen, Z.-F., Geng, H.-X., Zhou, M.-J., 2019.
An, Tianying, Winshell, Jamie, Scorzetti, Gloria, W. Fell, Jack, S. Rein, Kathleen, 2010. Contamination status of lipophilic marine toxins in shellfish samples from the Bohai
Identification of okadaic acid production in the marine dinoflagellate Prorocentrum Sea, China. Environ. Pollut. 249, 171–180.
rhathymum from Florida Bay. Toxicon 55, 653–657. https://doi.org/10.1016/j. Luo, Z., Zhang, H., Krock, B., Lu, S., Yang, W., Gu, H., 2017. Morphology, molecular
toxicon.2009.08.018. In press. phylogeny and okadaic acid production of epibenthic Prorocentrum (Dinophyceae)
Aquino-Cruz, A., Purdie, D.A., Morris, S., 2018. Effect of increasing sea water temperature species from the northern South China Sea. Algal Res. 22, 14–30.
on the growth and toxin production of the benthic dinoflagellate Prorocentrum lima. Mangialajo, L., Fricke, A., Perez-Gutierrez, G., Catania, D., Jauzein, C., Lemee, R., 2017.
Hydrobiologia 813, 103–122. Benthic Dinoflagellate Integrator (BEDI): a new method for the quantification of
Bendschneider, K., 1952. A new spectrophotometric method for the determination of benthic harmful algal blooms. Harmful Algae 64, 1–10.
nitrite in sea water. J. Mar. Res. 11, 87–96. Mohammad-Noor, N., Moestrup, Ø., Daugbjerg, N., 2007. Light, electron microscopy and
Berdalet, E., Tester, P.A., Chinain, M., Fraga, S., Lemée, R., Litaker, W., Penna, A., Usup, DNA sequences of the dinoflagellate Prorocentrum concavum (syn. P. arabianum) with
G., Vila, M., Zingone, A., 2017. Harmful algal blooms in benthic systems recent special emphasis on the periflagellar area. Phycologia 46, 549–564.
progressand future research. Oceanography 30, 36–45. Morton, S.L., 1998. Morphology and Toxicology of Prorocentrum faustiae sp. nov., a Toxic
Blanfuné, A., Boudouresque, C.F., Grossel, H., Thibaut, T., 2015. Distribution and abun- Species of Non-Planktonic Dinoflagellate from Heron Island, Australia. Bot. Mar. 41,
dance of Ostreopsis spp. and associated species (Dinophyceae) in the northwestern 565–569.
Mediterranean: the region and the macroalgal substrate matter. Environ. Sci. Pollut. Morton, S.L., Faust, M.A., Fairey, E.A., Moeller, P.D., 2002. Morphology and toxicology of
Res. 22, 12332–12346. Prorocentrum arabianum sp. nov.,(Dinophyceae) a toxic planktonic dinoflagellate
Calado, A., Hansen, G., 1999. Architecture of the flagellar apparatus and related struc- from the Gulf of Oman, Arabian Sea. Harmful Algae 1, 393–400.
tures in the type species of Peridinium, P. cinctum (Dinophyceae). Eur. J. Phycol. 34, Mullin, J.B., Riley, J.P., 1955. The Spectrophotometric determination of nitrate in natural
179–191. waters, with particular references to see water. Anal. Chim. Acta 12, 464–480.
Chan, T.Y., 2015. Ciguatera fish poisoning in east Asia and southeast Asia. Marine Drugs Murakami, Y., Oshima, Y., Yasumoto, T., 1982. Identification of okadaic acid as a toxic
13, 3466–3478. component of a marine dinoflagellate Prorocentrum lima. Nippon Suisan Gakkaishi 48,
Chomérat, N., Bilien, G., Zentz, F., 2019. A taxonomical study of benthic Prorocentrum 69–72.

9
J. Zou, et al. Marine Pollution Bulletin 158 (2020) 111313

Murphy, J., Riley, J.P., 1962. A modified single solution method for the determination of Scholin, C.A., Herzog, M., Sogin, M., Anderson, D.M., 1994. Identification of group-and
phosphate in natural waters. Anal. Chim. Acta 27, 31–36. strain-specific genetic markers for globally distributed Alexandrium (Dinophyceae). ii.
Nascimento, S.M., Salgueiro, F., Menezes, M., de Andréa Oliveira, F., Magalhães, V.C.P., sequence analysis of a fragment of the LSU rRNA gene 1. J. Phycol. 30, 999–1011.
De Paula, J.C., Morris, S., 2016. Prorocentrum lima from the South Atlantic: mor- Tibiriçá, E.J.A.C., Leite, P.I., Batista, V.V.T., Fernandes, F.L., Chomérat, N., Herve, F.,
phological, molecular and toxicological characterization. Harmful Algae 57, 39–48. Hess, P., Mafra, L.L., 2019. Ostreopsis cf. ovata bloom in Currais, Brazil: phylogeny,
Nascimento, S.M., Mendes, M.C.Q., Menezes, M., Rodríguez, F., Alves-de-Souza, C., toxin profile and contamination of mussels and marine plastic litter. Toxins 11, 446.
Branco, S., Riobó, P., Franco, J., Nunes, J.M.C., Huk, M., 2017. Morphology and Verma, A., Kazandjian, A., Sarowar, C., Harwood, D.T., Murray, J.S., Pargmann, I.,
phylogeny of Prorocentrum caipirignum sp. nov.(Dinophyceae), a new tropical toxic Hoppenrath, M., Murray, S.A., 2019. Morphology and phylogenetics of benthic
benthic dinoflagellate. Harmful Algae 70, 73–89. Prorocentrum species (Dinophyceae) from Tropical Northwestern Australia. Toxins
Pin, L.C., Teen, L.P., Ahmad, A., Usup, G., 2001. Genetic diversity of Ostreopsis ovata 11, 571.
(Dinophyceae) from Malaysia. Mar. Biotechnol. 3, 246–255. Yang, D., Yang, C., 2009. Detection of seagrass distribution changes from 1991 to 2006 in
Rodríguez, F., Riobó, P., Crespín, G.D., Daranas, A.H., de Vera, C.R., Norte, M., Xincun Bay, Hainan, with satellite remote sensing. Sensors 9, 830–844.
Fernández, J.J., Fraga, S., 2018. The toxic benthic dinoflagellate Prorocentrum ma- Yasumoto, T., Seino, N., Murakami, Y., Murata, M., 1987. Toxins produced by benthic
culosum Faust is a synonym of Prorocentrum hoffmannianum Faust. Harmful Algae dinoflagellates. Biol. Bull. 172, 128–131.
78, 1–8. Zhang, H., Li, Y., Cen, J., Wang, H., Cui, L., Dong, Y., Lu, S., 2015. Morphotypes of
Russell, F.E., Egen, N.B., 1991. Ciguateric fishes, ciguatoxin (CTX) and ciguatera poi- Prorocentrum lima (Dinophyceae) from Hainan Island, South China Sea: morpholo-
soning. Journal of Toxicology: Toxin Reviews 10, 37–62. gical and molecular characterization. Phycologia 54, 503–516.
Sagi, T., 1966. Determination of Ammonia in Sea Water by the Indophenol Method and Its Zhou, J., Fritz, L., 1993. Ultrastructure of two toxic marine dinoflagellates, Prorocentrum
Application to the Coastal and Off-shore Waters. lima and Prorocentrum maculosum. Phycologia 32, 444–450.

10

You might also like