Home Morphological and molecular characterization of Kareniaceae (Dinophyceae, Gymnodiniales) in Kuwait’s waters
Article Publicly Available

Morphological and molecular characterization of Kareniaceae (Dinophyceae, Gymnodiniales) in Kuwait’s waters

  • Manal Al-Kandari

    Manal Al-Kandari (PhD) is a marine biologist with over 23 years of experience in the taxonomy and molecular biology of planktonic microalgae from a wide geographic range. Since 2012, she has been engaged in research on marine microalgae bloom of Kuwait at the Kuwait Institute for Scientific Research. Her research interests are focused on microalgal biodiversity, taxonomy, molecular biology and ecology, with a particular emphasis on harmful bloom species and their impact on coastal ecosystems.

    ORCID logo EMAIL logo
    , Maria Saburova

    Maria Saburova (PhD) is a marine biologist with over 30 years of experience in the taxonomy and ecology of planktonic and benthic microalgae from a wide geographic range. Since 2005, she has been engaged in research on marine microalgae in the Middle East region as a long-term scientific consultant at the Kuwait Institute for Scientific Research. Her research interests are focused on microalgal biodiversity, taxonomy, and ecology, with a particular emphasis on harmful bloom species and their impact on coastal ecosystems.

    , Igor Polikarpov

    Igor Polikarpov is a biological oceanographer with over 35 years of research experience, working over 19 years at the Kuwait Institute for Scientific Research. His current research interests include biological oceanography of the marginal seas of the Indian Ocean, marine phytoplankton and microplankton ecology, and long-term natural and human-induced environmental changes in the coastal waters. He has co-authored several books on marine ecology and biological oceanography of the north-western Indian Ocean.

    , Jacob Larsen

    Jacob Larsen (PhD) is a marine biologist with over 40 years’ experience in the taxonomy planktonic and benthic dinoflagellates from a wide geographic range. Since 1995, he has mainly been involved in the development, organization, and teaching of IOC courses on taxonomy and identification of harmful microalgae.

    , Nina Lundholm and Sumaiah Hussain
Published/Copyright: March 14, 2025

Abstract

Kuwait’s waters were first faced with a bloom of ichthyotoxic dinoflagellate Karenia and fish kill in 1999. Since then, there have been sporadic blooms and records of various kareniacean taxa, but accurate species identification has been challenging due to the high morphological similarity among morphotypes and the lack of molecular studies. Ten clonal kareniacean strains isolated from Kuwait’s coastal waters were characterized based on light and fluorescence microscopy and the LSU rDNA gene sequencing. Seven strains represented the distinctive morphology of Karenia papilionacea and showed LSU rDNA sequences with > 99 % similarity to the type material of this species. One Karenia strain represented the typical morphology of Karenia selliformis and the LSU rDNA sequence fell within a well-supported K. selliformis clade. Two other strains showed the typical morphology of Karlodinium ballantinum and phylogenetic analysis supported the morphological results. Consequently, a combination of morphological and molecular analyses confirmed the presence of K. papilionacea and K. selliformis in Kuwait’s waters, thereby resolving the previous regional taxonomic uncertainty regarding these species. The identification of K. ballantinum represents the first regional record. Investigating the morphology and phylogeny of commonly occurring Kareniaceae enhances the monitoring and risk assessment of harmful algal blooms in Kuwait’s marine environment.

1 Introduction

The family Kareniaceae was initially established to encompass a distinct evolutionary lineage (de Salas et al. 2003) with three genera of unarmored gymnodinioid dinoflagellates, namely Karenia and Takayama, sharing a single synapomorphic character: fucoxanthin and its derivatives as the major accessory pigments instead of peridinin (Bergholtz et al. 2006). Later, the genera Asterodinium (Benico et al. 2019), Gertia (Takahashi et al. 2019), and Shimiella (Ok et al. 2021) were assigned to the family based on morphological and genetic similarity. Morphological diagnostic features such as cell shape, path and length of the apical groove, presence of a ventral pore, sulcal intrusion into the epicone, and position of the nucleus serve to distinguish Kareniaceae species (e.g., Botes et al. 2003; Daugbjerg et al. 2000; de Salas et al. 2003; Haywood et al. 2004; Steidinger et al. 2008).

Many species belonging to the family Kareniaceae are of ecological and socio-economic importance due to their ability to produce toxins and form harmful blooms, resulting in a significant impact on coastal marine ecosystems worldwide. A wide range of Karenia, Karlodinium, and Takayama species produce lipids, sterols, and polyunsaturated fatty acids that are potentially toxic to various marine animals, leading to the deterioration of gill membranes. In addition to ichthyotoxicity, these dinoflagellates are capable of producing high biomasses during blooming phases, resulting in respiratory interferences in bloom-exposed animals by algal mucilage and hypoxia in the water column (e.g., Adolf et al. 2009; Brand et al. 2012; de Salas et al. 2003; Furuya et al. 2018; Hallegraeff 2023; Larsen 2023; Lassus et al. 2016; Lundholm et al. 2009 onwards; Mooney et al. 2007, 2010; Oh et al. 2023; Steidinger et al. 1998). The adverse impacts of the Kareniaceae species on marine fauna have been documented worldwide. Recurrent blooms of Karenia selliformis have been associated with severe mortalities of marine organisms in the Mediterranean Sea, Tunisia (Feki et al. 2013; Zingone et al. 2021), and along the coast of Chilean Patagonia (Baldrich et al. 2024; Clément et al. 2001; Mardones et al. 2020; Uribe and Ruíz 2001). Intensive K. selliformis-dominated harmful blooms have recently occurred in the Bering Sea, Kamchatka, Russia (Orlova et al. 2022) and in the region of Hokkaido, Japan (Iwataki et al. 2022; Siswanto et al. 2024), causing marine mortalities and damage to fishery industries. Additionally, harmful blooms and associated fish mortality have been widely reported for Karlodinium species, including in the waters of Malaysia, Singapore, China, and the Sea of Oman, Iran (Asefi et al. 2023; Cen et al. 2019; Leong et al. 2015; Lim et al. 2014; Sakamoto et al. 2020; Yñiguez et al. 2021).

A number of Karenia species producing potent natural polyether neurotoxins, termed ‘brevetoxins’, are responsible for mammalian respiratory distress when exposed to aerosols and neurotoxic shellfish poisoning through the consumption of contaminated seafood. Of these, Karenia brevis is the best-known dinoflagellate due to its recurrent dense toxic blooms along the west coast of Florida in the Gulf of Mexico, USA (e.g., Landsberg and Steidinger 1998; Lin et al. 1981; Steidinger 2009; Tester et al. 2000, 2008). Another neurotoxin producer, Karenia papilionacea, is a widely distributed species, although its bloom events have been scarcely documented worldwide. This species is characterized by lower toxicity than found in K. brevis, with both toxic and non-toxic strains described (Fowler et al. 2015; Haywood et al. 2004; Wolny et al. 2024). The species has been documented to bloom in Japan (Yamaguchi et al. 2016) and in coastal waters of the Delmarva Peninsula and Chesapeake Bay, USA (Wolny et al. 2024).

The marine environment of the Arabian/Persian Gulf (hereafter referred to as the Gulf following Sheppard et al. 2010; Polikarpov et al. 2019) first experienced the adverse effect of Karenia blooms in 1999, when a major fish kill occurred in northwestern Iranian coastal waters, which was immediately followed by fish kills in Kuwait Bay (Heil et al. 2001; Husain and Faraj 2000). Although the cause of marine mortality in Iranian waters has not been confirmed, the fish kill in Kuwait Bay has been attributed to a bloom of gymnodinioid dinoflagellate, initially thought to be Gymnodinium cf. mikimotoi due to its close morphological resemblance to this species (Husain and Faraj 2000). However, further taxonomic examination of bloom specimens by A. Haywood and K. Steidinger recognized the morphotype from Kuwait as resembling the previously undescribed Gymnodinium sp. sensu Mackenzie et al. (1995, 1996) from New Zealand waters, which was tentatively named Gymnodinium selliformis (Heil et al. 2001) before Haywood et al. (2004) described it as K. selliformis. The taxonomic and nomenclatural confusion in relation to the dinoflagellate responsible for the 1999 bloom in Kuwait has resulted in a number of names under which this taxon has been reported in the relevant literature, including Gymnodinium cf. mikimotoi (Husain and Faraj 2000), Gymnodinium sp. (Heil et al. 2001), Gymnodinium selliforme (figure 2B, C and table 1 in Heil et al. 2001), Karenia sp. (Glibert et al. 2002), and finally K. selliformis (e.g., Al-Kandari et al. 2009; Al-Yamani et al. 2004, 2012; Al-Yamani and Saburova 2019; Devlin et al. 2019; Polikarpov et al. 2020).

In the latter half of the 2000s, a K. brevis-like dinoflagellate appeared in Kuwait’s waters, which was clearly morphologically distinct from K. selliformis in cell shape and nucleus location. As with K. selliformis, there was some taxonomic confusion regarding the identity of this species. It was first identified as Gymnodinium breve (Al-Yamani et al. 2004), and afterwards as Karenia cf. brevis (Al-Kandari et al. 2009), K. brevis (Devlin et al. 2019), or K. papilionacea (Al-Yamani and Saburova 2019; Al-Yamani et al. 2012; Polikarpov et al. 2020). Due to the challenges of identifying these poorly fixed and fragile species, and the lack of molecular evidence to confirm species identity, taxonomic uncertainty has long persisted for kareniacean taxa in Kuwait’s waters.

Since the 2010s, sporadic blooms of Karenia, Karlodinium, and Takayama species (>105–107 cells L−1) have been observed in the inshore waters along Kuwait’s coast (Al-Yamani and Saburova 2019; Polikarpov et al. 2020). These kareniacean taxa have until now been distinguished only on morphological observations and their identifications have never been supported by molecular data. However, an unambiguous species recognition based on morphological analysis alone is often difficult and usually requires molecular tools for confirmation (e.g., Haywood et al. 2004; Iwataki 2023; Iwataki et al. 2022). The present paper focuses on the Karenia and Karlodinium morphotypes, which have formed blooms and caused fish mortality in Kuwait’s waters, in order to resolve previous taxonomic uncertainties regarding these species and to clarify their identity. These taxa are characterized here based on field samples and clonal cultures using high-resolution light and fluorescence microscopy and analyses of the large subunit (LSU) nuclear rDNA gene sequences.

2 Materials and methods

2.1 Study area

Kuwait’s marine area occupies the northwestern corner of the Gulf, stretching for about 180 km from Iraq in the north to Saudi Arabia in the south (Figure 1A). Kuwait’s waters are hypersaline, with an average salinity of 41.6, shallow, with a maximum depth of about 30 m, and are generally well mixed year round (Al-Yamani 2021; Al-Yamani et al. 2004, 2017, 2021). Kuwait’s subtropical location and arid climate results in marked amplitudes of summer/winter temperature differences (Sheppard et al. 2010). The average annual seawater temperature in Kuwait is 23.8 °C, with seasonal variations ranging from an average of 14 °C in winter to 30.5 °C in the summer months (Al-Yamani et al. 2004). One of the most prominent features of Kuwait’s waters is the semi-enclosed Kuwait Bay, which extends approximately 46 km westward, accounting for nearly half of the country’s coastline and covering an area of approximately 720 km2 (Figure 1B). The bay is a shallow, meso-tidal hypersaline inverse estuary, with an average depth of 5.2 m and a maximum depth at its mouth of 23 m and experiences a high tidal range of 4 m (Al-Yamani et al. 2004).

Figure 1: 
Area of investigation: (A) general location of the sampling area; (B) map of Kuwait’s marine area; (C) satellite image of Kuwait Bay and adjacent waters showing location of sampling sites.
Figure 1:

Area of investigation: (A) general location of the sampling area; (B) map of Kuwait’s marine area; (C) satellite image of Kuwait Bay and adjacent waters showing location of sampling sites.

2.2 Sampling

During 2014, 2016, and 2021, morphological and molecular data were obtained from field seawater samples and from cultures of Karenia and Karlodinium species collected or isolated at several sampling sites during routine phytoplankton monitoring or in response to phytoplankton bloom events (Figure 1C; Table 1).

Table 1:

Data on Karenia and Karlodinium strains used in the molecular analysis for characterization of Kareniaceae species in Kuwaiti waters.

Strain Species Isolation date Site LSU rDNA
D1/D2 D1/D3
KW-E1–014 K. papilionacea Oct 2014 Kuwait Bay +
KW-F3-015 K. papilionacea Oct 2014 Kuwait Bay + +
KW-F4-061 K. papilionacea Mar 2021 Qibla Marina +
KW-G4-062 K. papilionacea Mar 2021 Qibla Marina +
KW-G5-063 K. papilionacea Mar 2021 Qibla Marina + +
Kw-G8-064 K. papilionacea Mar 2021 Qibla Marina +
KW-G9-065 K. papilionacea Mar 2021 Qibla Marina + +
KW-D10-060 K. selliformis Mar 2021 Qibla Marina +
KW-E9-046 Karl. ballantinum Oct 2014 Kuwait Bay +
KW-JL-07 Karl. ballantinum Jun 2016 KISR Salmiya Marina +

One-liter seawater samples were collected from the surface layer (1 m depth) using a 5-liter Niskin bottle sampler and were carefully transferred to sampling containers through a flexible hose. Samples were preserved with 4 % acidified Lugol’s solution. A replicate sampling was applied to obtain both preserved and living material simultaneously. Live samples were screened and utilized for taxonomic purposes and the establishment of cultures. Associated water temperature and salinity were measured at the sampling sites using a CTD profiler JFE AAQ1183 (JFE Advantech Co., Japan).

2.3 Species identification

Species identification was primarily based on live material obtained from unpreserved environmental samples or clonal cultures. For a detailed examination, as soon as possible after collection, when cells were in a live state, they were isolated by micropipetting and transferred to a glass slide in preparation for high-magnification photomicroscopy. Cells were examined and photographed using a Leica DM LM compound microscope (Leica Microsystems GmbH, Wetzlar, Germany) equipped with DFC425 color digital camera at 630–1,000 × magnification. A range of microscopic tools were used to examine the cells, including brightfield (BF) illumination to show the natural color, and Nomarski differential interference contrast (DIC) to visualize a three-dimensional appearance of cellular content and cell surface structure in unstained, transparent specimens. Autofluorescence of chloroplasts and nuclei stained by 4ʹ,6-diamidino-2-phenylindole (DAPI, Sigma Chemical Co.) were observed under fluorescence illumination (100 W short-arc mercury lamp, ultraviolet excitation light). Cell dimensions were measured from the images using Leica Application Suite v. 3.7 software (Leica Microsystems (Switzerland) Ltd).

2.4 Cell isolation and maintenance of cultures

Ten monoclonal cultures were established from samples collected in the central part of Kuwait Bay and artificial small semi-enclosed marinas in October 2014, June 2016, and March 2021 (Figure 1C; Table 1). Single cells were isolated using sterilized drawn out capillary pipettes under an inverted microscope, washed at least three times in a drop of sterilized seawater, and were then transferred to a 96-well plate filled with 350 μl of sterile K medium without silicate (Keller et al. 1987) prepared with filtered (0.22 μm, Millipore) and autoclaved natural seawater with a salinity of approximately 40. The isolated cells were incubated at 23.5 °C under 80 μmol photons·m−2·s−1 of light and a 12:12 h light:dark photoperiod and checked regularly. When the cell concentration became sufficiently high, the contaminant-free isolates with dividing and motile cells were transferred to 25-cm2 tissue culture flasks containing 40 ml sterile K medium as stock cultures. The established cultures are stored and maintained at the Laboratories of Phytoplankton Taxonomy of the Kuwait Institute for Scientific Research (KISR, Kuwait) and the IOC Science and Communication Centre on Harmful Algae, University of Copenhagen, Denmark. They are checked biweekly to maintain healthy growing cultures, and 1 ml of each strain is transferred to fresh medium monthly.

2.5 DNA amplification and sequencing

DNA was extracted from eight monoclonal strains of Karenia isolated from Kuwait’s waters in 2014 and 2021 and from two monoclonal strains of Karlodinium isolated in 2014 and 2016. Cells were harvested as pellets from 15-ml aliquots of late exponential phase cultures via gentle centrifugation at 4,000g for 5 min at room temperature. Total genomic DNA was extracted from the cell pellets using the DNeasy Blood & Tissue Kit (Qiagen, Hilden, Germany) according to the manufacturer’s instructions. DNA quality was assessed by agarose gel electrophoresis and quantified using a Nanodrop™ 1,000 spectrophotometer (Thermo Scientific, Labtech International Ltd, UK). Aliquots of DNA were kept at 4 °C for immediate PCR analysis, and replicate stocks were held at −80 °C for long-term storage.

Genomic DNA resulting from the 10 microalgal isolates was examined by PCR amplification using the D1/D2 region of the large nuclear subunit (LSU) of the ribosomal operon with the forward primer D1R-F and the reverse primer D2C-R (Scholin et al. 1994) for the eight isolates and the D1/D3 region of the LSU rDNA with the forward primer D1C-F (Scholin et al. 1994) and the reverse primer, either 28–1483R (Daugbjerg et al. 2000) or D3B-R (Nunn et al. 1996), for the five isolates (Tables 1 and 2).

Table 2:

Primers and sequences used in this study to amplify the LSU rDNA of Kareniaceae species.

Amplifies Primer Primer sequence (5′–3′) Reference
LSU rDNA D1R-F ACCCGCTGAATTTAAGCATA Scholin et al. (1994)
D2C-R CCTTGGTCCGTGTTTCAAGA Scholin et al. (1994)
D1C-F ACCCGCTGAATTTAAGCATA Scholin et al. (1994)
28–1483R GCTACTACCACCAAGATCTGC Daugbjerg et al. (2000)
D3B-R TCGGAGGGAACCAGCTACTA Nunn et al. (1996)

PCRs were carried out using a Bio-Rad C1000 thermal cycler (Bio-Rad Laboratories Inc., Hercules, CA, USA). PCR amplification conditions were as follows: initial denaturation at 95 °C for 5 min, followed by 35 cycles of denaturation at 95 °C for 30 s, annealing at 60 °C for 1 min and extension at 72 °C for 1 min, and a final extension step at 72 °C for 5 min. Reactions were performed in 50 µl of the AccuPower PCR Premix (Bioneer, Daejeon, Korea) and 5X HOT FIREPol Blend Master Mix, containing 1 µl template DNA (3.3–47.5 ng), 20 pmol forward primer, 20 pmol reverse primer. All the PCR products were separated through 1.6 % (w/v) agarose gels in 1 × TAE (Tris-acetate-EDTA) buffer, stained with GelRed (Biotium Inc., Hayward, CA, USA), and visualized using a UV fluorescent gel documentation system (Bio-Rad, UK).

Amplicons were sequenced after two-step qualitative PCR with the KAPA2G Robust HotStart Kit using Buffer B and Enhancer A (Roche, Kit Code KK5525, Cat. No 07961120001), then amplicons showing a clear single band were sequenced using the ABI 3730XL DNA Analyzer (PE Biosystems, Vernon Hills, IL, USA). The identity of PCR products was confirmed by using BLASTN analysis (www.ncbi.nlm.nih.gov/BLAST). The consensus sequences were assembled and edited by pair-wise alignment using BioEdit 7.7.1 (Hall 1999). New sequence data were deposited in the GenBank (National Center for Biotechnology Information, NCBI) under the following accession numbers: PP951870-PP951879.

2.6 Phylogenetic analyses

The obtained sequences of partial LSU rDNA were aligned with other related Kareniaceae species retrieved from the NCBI nucleotide database (http://www.ncbi.nlm.nih.gov/) using the ClustalW algorithm (Thompson et al. 1994) using MEGA 10.2.6 (Kumar et al. 2018) and were subsequently verified manually. The LSU rDNA dataset of this study contained 1,002 characters and comprised 10 nucleotide sequences of Kareniaceae from Kuwait’s waters and 47 sequences of other closely related taxa, which were used for phylogenetic analyses. Phylogenetic relationships were analyzed by the maximum likelihood (ML) method using MEGA 10.2.6 and by Bayesian inference (BI) using MrBayes v. 3.1.2 (Ronquist et al. 2012). The Tamura-Nei model of nucleotide substitution (Tamura and Nei 1993) with gamma-distributed rates among sites (TN93 + G) was considered best to describe the substitution pattern for ML based on the lowest Bayesian Information Criterion (BIC) scores and corrected Akaike Information Criterion (AICc) values. The initial tree for the heuristic search was obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances, and then selecting the topology with the superior log likelihood value. A discrete gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.4071)). Bootstrap analysis was carried out with 1,000 replicates to assess statistical reliability. For Bayesian inference, the GTR substitution model with gamma-distributed rate variation across sites and a proportion of invariable sites (GTR + I + Г) was used. Four independent Markov chain Monte Carlo simulations were run simultaneously for 1,000,000 generations, and trees were sampled at every 1,000 generations. The first 25 % of the trees were discarded as burn-in. Convergence was judged by the average standard deviation of split frequencies (less than 0.01). The remaining trees were used to generate a consensus tree and calculate the posterior probabilities of all branches using a majority-rule consensus approach. FigTree 1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/) was used to view and edit the trees. The ML topology was used to visualize the phylogenetic tree. The tree with the highest log likelihood (−3,690.53) is shown. GenBank accession numbers and collection geographic region for Kareniaceae sequences used in this study are provided in the phylogenetic tree. Genetic distances (p-distances) between sequences were calculated using an uncorrected genetic distance (UGD) model with a pairwise deletion option (MEGA 10.2.6).

3 Results

3.1 Species morphology

Morphological descriptions were based on the results of observations using light and epifluorescence microscopy of cultured material gathered during this study and field samples collected in the coastal waters of Kuwait. Conservative diagnostic characters used to distinguish between different kareniacean morphotypes included cell shape (length/width ratio), apical groove length, degree of extension of the sulcal groove onto the epicone, presence of a ventral pore, and shape and position of the nucleus (Table 3). Based on morphological characteristics, three distinct morphotypes from Kuwait’s waters (Figures 24; Table 3) were identified, matching well the descriptions of K. papilionacea, K. selliformis, and Karlodinium ballantinum (de Salas et al. 2008; Haywood et al. 2004; Steidinger et al. 2008).

Table 3:

Comparison of morphological characteristics among the species of Karenia and Karlodinium in Kuwaiti waters described in this study and previous reports.

K. papilionacea K. selliformis Karl. ballantinum
Cell length (μm) 34–40a

18–32 (24.7)b

19–38d

20.7–38.5e

19–29.5f

38–41g

20–25h

24.89–53.01 (37.35)i

20.23–45.85 (30.97)i
30–33a

20–27 (23.6)b

21–30.8 (26.1)f

30–34g

17–37k

25–30l

35.3–43.6 (39.4)m

38.6–48.8 (43.0)m
14–18a

11–18n

10-22 (12.50)o
Cell width (μm) 38–44a

18–48 (32.4)b up to 60c

21–50d

20.0–44.9e

34.1–52.5f

42–46g

20–30h

22.44–44.33 (31.54)i

20.68–44.17 (31.48)i

18–56 (33.7)j
24–27a

16–24 (18.5)b

19.3–29.3 (24.2)f

25–28g

13–35k

20–25l

31.8–44.7 (37.3)m

30.9–40.1 (36.3)m
10–13a

8–14n

7–12 (9.5)o
Cell depth (μm) 10–15b

10–18d

13.6–25.7f
5–10b

13.7–17.7 (15.6)f

18–22l
8–10 (9.00)o
Length to width ratio 0.9a,g

0.8b

1.0h

1.09–1.52 (1.185)i

0.73–1.09 (0.98)i
1.2a

1.3b

0.9–1.1f
1.4a

1.32p
Dorsoventral compression Moderate Strong None
Apical protrusion Pointed

Slightly/typically pointedd
None None
Antapical excavation Pronounced Moderate

Negligiblel
None
Vetral pore Absent Absent Absent (LM)a

Absent or inconspicuousn

Present (pore/pore-like depression)p
Lateral pore Singleb Single dorso-lateralb,c

Absentf,l
Absent
Cingulum position and displacement Median

1 cingular widtha,f

1–2 cingular widthb,c
Pre-median

1.5–2 cingular widtha

2 cingular widthb,f

1–1.5 cingular widthc

1.5 cingular widthg

1 cingular widthl

Median, 2–3 cingular widthm
Median, 2 cingular widtha

1/3 cingular widthn

1 cingular widtho
Sulcus Open sulcus extends to left of apex Open narrow sulcus extends to left of apex Sigmoid, short sulcal extension onto the epiconea

“S” bend, short but evident sulcal extensionn
Apical groove, ventral 0.25a

0.25–0.33b

0.23–0.25 – 0.33–0.36d
Longa

0.33–0.66b
Very shorta

<1/2n
Apical groove, dorsal 1/3a

0.25–0.33b

0.23–0.26 – 0.33–0.36d
1/3a

0.25–0.33b
Very short
Nucleus shape Spherical Oblong, transversely elongateda

Reniform to ellipticalb
Large, round
Nucleus location Left lobe of hypocone Central hypoconea

Upper middle hypoconeb
Centrally, toward dorsal side
Chloroplasts 16–26, peripheral, reniforma

2–20, peripheral, round to reniformb

Several, peripheral, almost roundd

Few, large, round to reniformf

Several, reniformg
48–56, plate-like, elongated, with pyrenoidsa

Several, peripheral, reniformb,c,l

Numerous, peripheral, reniformf

Numerous, reniform to elongatedg

Numerous (46–105), peripheral, granular or strap-shaped, irregularly curved, with internal pyrenoidm
Numerous (8–19), irregularly shaped, with internal pyrenoidsa

Numerous, long and narrow, with internal pyrenoids, more abundant in hypoconen

7–12, peripheral, irregularly shapedo
  1. Cell dimensions – range (average values where available); variable morphological characters are cited as follows: athis study; b Haywood et al. (2004) (original description); c Steidinger et al. (2008); d Yamaguchi et al. (2016); e Kim et al. (2023); f Orlova et al. (2022); g Al-Yamani and Saburova (2019); h Yeung et al. (2005); i Stuart (2011); j Gómez (2006); k Heil et al. (2001); l Mardones et al. (2020); m Iwataki et al. (2022); n de Salas et al. (2008) (original description); o Siano et al. (2009); p Benico et al. (2020).

Figure 2: 
Light micrographs of Karenia papilionacea cells (Figure 2A–C: strain KW-F3-015; Figure 2D–G: strain KW-E1-014; Figure 2N: strain KW-G4-062; Figure 2H–M, O-R: isolates from field samples; arrows point to the apical groove; arrowheads point to the open sulcal extension onto the epicone; n marks the nucleus): (A, B) ventral view showing apical groove, open-ended sulcus onto the epicone, and nucleus in the left lobe of hypocone; (C) epifluorescence microscopy under blue-light illumination showing numerous kidney-shaped chloroplasts; (D, E) ventral and dorsal views of the same cell showing apical groove, sulcus, and nucleus; (F) apical view in deep focus showing dorso-ventral compression; (G) epifluorescence microscopy showing several elongated chloroplasts; (H) two cells in ventral and dorsal view; (I–K) cells in left lateral, apical, and antapical views showing dorso-ventral compression; (L) ventral view showing apical groove and open-ended sulcus onto the epicone; (M) dorsal view showing apical groove over epicone; (N) Lugol’s-fixed cell in ventral view; (O–R) contracted cells during butterfly-like movement. Scale bars: 10 µm.
Figure 2:

Light micrographs of Karenia papilionacea cells (Figure 2A–C: strain KW-F3-015; Figure 2D–G: strain KW-E1-014; Figure 2N: strain KW-G4-062; Figure 2H–M, O-R: isolates from field samples; arrows point to the apical groove; arrowheads point to the open sulcal extension onto the epicone; n marks the nucleus): (A, B) ventral view showing apical groove, open-ended sulcus onto the epicone, and nucleus in the left lobe of hypocone; (C) epifluorescence microscopy under blue-light illumination showing numerous kidney-shaped chloroplasts; (D, E) ventral and dorsal views of the same cell showing apical groove, sulcus, and nucleus; (F) apical view in deep focus showing dorso-ventral compression; (G) epifluorescence microscopy showing several elongated chloroplasts; (H) two cells in ventral and dorsal view; (I–K) cells in left lateral, apical, and antapical views showing dorso-ventral compression; (L) ventral view showing apical groove and open-ended sulcus onto the epicone; (M) dorsal view showing apical groove over epicone; (N) Lugol’s-fixed cell in ventral view; (O–R) contracted cells during butterfly-like movement. Scale bars: 10 µm.

Figure 3: 
Light micrographs of Karenia selliformis cells (Figure 3A–D, L: strain KW-D10-060; Figure 3C–K: isolates from field samples; arrows point to the apical groove; arrowheads point to the sulcal intrusion onto the epicone; n marks the nucleus): (A, B) ventral view showing apical groove and sulcal intrusion onto the epicone; (C) squashed cell showing numerous elongated chloroplasts and transversely elongated nucleus in the hypocone; (D) epifluorescence microscopy under blue-light illumination showing numerous peripherally located, elongated chloroplasts; (E, F) ventral view showing apical groove and sulcal intrusion onto the epicone; (G) lateral view showing strong dorso-ventral compression; (H, I) dorsal view of different cells showing apical groove over epicone and nucleus position; (J) epifluorescence microscopy of DAPI-stained cell showing transversely elongated nucleus in hypocone; (K) epifluorescence microscopy showing numerous peripherally located, elongated chloroplasts; (L) Lugol’s-fixed cell in ventral view. Scale bars: 10 µm.
Figure 3:

Light micrographs of Karenia selliformis cells (Figure 3A–D, L: strain KW-D10-060; Figure 3C–K: isolates from field samples; arrows point to the apical groove; arrowheads point to the sulcal intrusion onto the epicone; n marks the nucleus): (A, B) ventral view showing apical groove and sulcal intrusion onto the epicone; (C) squashed cell showing numerous elongated chloroplasts and transversely elongated nucleus in the hypocone; (D) epifluorescence microscopy under blue-light illumination showing numerous peripherally located, elongated chloroplasts; (E, F) ventral view showing apical groove and sulcal intrusion onto the epicone; (G) lateral view showing strong dorso-ventral compression; (H, I) dorsal view of different cells showing apical groove over epicone and nucleus position; (J) epifluorescence microscopy of DAPI-stained cell showing transversely elongated nucleus in hypocone; (K) epifluorescence microscopy showing numerous peripherally located, elongated chloroplasts; (L) Lugol’s-fixed cell in ventral view. Scale bars: 10 µm.

Figure 4: 
Light micrographs of Karlodinium ballantinum cells (Figure 4A–C, F, G: strain KW-JL-07; Figure 4D, E, H–L: strain KW-E9-046; n marks the nucleus): (A–C) ventral view showing apical groove (arrow) and short finger-like sulcal intrusion onto the epicone (arrowhead); (D) dorsal view showing apical groove (arrow) over epicone; (E) apical view showing apical groove (arrow) over epicone and no dorso-ventral compression; (F) dividing motile cell; (G) squashed cell in dorsal view showing the knob-like structures lining the lower margin of the cingulum (arrows); (H) squashed cell showing large central nucleus surrounded by several chloroplasts; (I) epifluorescence microscopy of DAPI-stained cell in dorsal view showing centrally located nucleus; (J) epifluorescence microscopy of DAPI-stained cell in lateral view showing nucleus close to the dorsal side; (K) epifluorescence microscopy showing numerous peripherally located, elongated chloroplasts with internal pyrenoids (white arrows); (L) Lugol’s-fixed cell in ventral view. Scale bars: 10 µm.
Figure 4:

Light micrographs of Karlodinium ballantinum cells (Figure 4A–C, F, G: strain KW-JL-07; Figure 4D, E, H–L: strain KW-E9-046; n marks the nucleus): (A–C) ventral view showing apical groove (arrow) and short finger-like sulcal intrusion onto the epicone (arrowhead); (D) dorsal view showing apical groove (arrow) over epicone; (E) apical view showing apical groove (arrow) over epicone and no dorso-ventral compression; (F) dividing motile cell; (G) squashed cell in dorsal view showing the knob-like structures lining the lower margin of the cingulum (arrows); (H) squashed cell showing large central nucleus surrounded by several chloroplasts; (I) epifluorescence microscopy of DAPI-stained cell in dorsal view showing centrally located nucleus; (J) epifluorescence microscopy of DAPI-stained cell in lateral view showing nucleus close to the dorsal side; (K) epifluorescence microscopy showing numerous peripherally located, elongated chloroplasts with internal pyrenoids (white arrows); (L) Lugol’s-fixed cell in ventral view. Scale bars: 10 µm.

3.1.1 Karenia papilionacea A.J.Haywood et K.A.Steidinger 2004

The cells were solitary, transversely elongated, wider than long, dorsoventrally compressed, dorsally convex and ventrally concave, 34–41 μm long, and 38–44 μm wide, with a length to width ratio of about 0.9 (Figure 2). The epicone possessed a pointed apical protrusion (carina). The hypocone was bilobed with a deeply excavated antapex (Figure 2A, B, D–F, H–N). The cingulum was slightly pre-median to median, descending, and displaced by about one cingular width (Figure 2B, I, L). The sulcus continued as a short open extension onto the epicone (Figure 2B, D, L). A short linear apical groove bisected the apex and extended to about one-third of the dorsal epicone (Figure 2A, D, I, L, M). The nucleus was spherical and located in the left lobe of the hypocone, as seen in the ventral view (Figure 2A, B, E, H, K, N). The cytoplasm contained numerous (16–26) kidney-shaped to elongated, golden-brown chloroplasts located peripherally (Figure 2C, G). Swimming cells were able to bend along the longitudinal axis, resulting in slow movements resembling a flying butterfly (Figure 2O–R).

3.1.2 Karenia selliformis A.J.Haywood, K.A.Steidinger et L.MacKenzie 2004

The cells were solitary, oval, longer than wide, strongly dorsoventrally compressed, 30–33 μm long, and 24–27 μm wide, with a length to width ratio of about 1.2 (Figure 3). The epicone was broadly dome-shaped with slightly convex sides. The hypocone was hemispherical with a bilobed, centrally excavated antapex and a slightly protruding right lobe (Figure 3A, B, E–I). The cingulum was slightly premedian, descending, and displaced by 1.5–2 cingular widths. The sulcus was narrow; its left margin opened onto the epicone as a narrow extension (Figure 3A, B, E, F). The narrow and linear apical groove originated just above and slightly to the right of the proximal end of the cingulum and extended across the apex to about one-third down on the dorsal side of the epicone. The apex was slightly indented by the apical groove (Figure 3A, B, E, F, H, I). The nucleus was large, oblong, transversely elongated, and located in the center of the hypocone (Figure 3C, I, J, L). The cytoplasm contained numerous (48–56) plate-like, elongated, golden-brown chloroplasts with internal pyrenoids (Figure 3C, D, K).

3.1.3 Karlodinium ballantinum Salas 2008 (Figure 4)

The cells were solitary, more or less oval in ventral view, 14–18 μm long, and 10–13 μm wide, with a length to width ratio of about 1.4. The cingulum was equatorial, and the epicone and hypocone were both hemispherical and nearly equal in size (Figure 4A–D, L). The cingulum originated at the longitudinal axis of the cell, descended, and displaced about two cingular widths or 25–30 % of the cell length. The sulcus was sigmoid; it formed a short finger-like projection onto the epicone and continued a narrow, oblique connection between the two ends of the cingulum before fading and becoming wider posteriorly (Figure 4A–D). A short apical groove was linear and extended across the apex and briefly down the dorsal epicone (Figure 4B–E). In squashed cells, a pattern of knob-like structures lining the lower margin of the cingulum was discerned (Figure 4G). No ventral pore was observed. The nucleus was located in the central part of the cell, close to the dorsal side (Figure 4H–J). Numerous (8–19) irregularly shaped chloroplasts with internal pyrenoids were present (Figure 4K). Cell division occurred at the motile stage (Figure 4F).

3.2 Phylogenetic relationships

A total of 10 kareniacean strains isolated from Kuwait’s waters were sequenced during this study and included in the phylogenetic analyses. The construction of phylogenetic trees using Maximum Likelihood (ML) and Bayesian Inference (BI) methods based on the D1/D2 and D1/D3 domains of the LSU rDNA resulted in a consistent topology, with high bootstrap support and posterior probability for most clades (Figure 5).

Figure 5: 
Maximum-likelihood (ML) phylogenetic tree of the D1-D2/D1-D3 LSU rDNA sequences of the Kareniaceae from Kuwait’s waters and closely related species from other locations around the world. The numbers on the branches are statistical support values (ML bootstrap support (%), n = 1,000/Bayesian posterior probability, n = 1,000,000). GenBank accession numbers and collection geographic regions follow the taxon names. New sequences are shown in bold. Scale bar = number of nucleotide substitutions per site.
Figure 5:

Maximum-likelihood (ML) phylogenetic tree of the D1-D2/D1-D3 LSU rDNA sequences of the Kareniaceae from Kuwait’s waters and closely related species from other locations around the world. The numbers on the branches are statistical support values (ML bootstrap support (%), n = 1,000/Bayesian posterior probability, n = 1,000,000). GenBank accession numbers and collection geographic regions follow the taxon names. New sequences are shown in bold. Scale bar = number of nucleotide substitutions per site.

The newly sequenced Karenia strains were placed in well-supported monophyletic clades of K. selliformis or K. papilionacea (Figure 5) and distinct from related Karenia species. Within the highly supported (ML bootstrap value of 99 and BI posterior probability of 0.97) clade of K. selliformis strains with a worldwide distribution, Kuwait’s strain KW-D10-060 was closest to those originating from New Zealand, including the holotype sequence U92250, forming a subclade with moderate bootstrap (62) and high posterior probability (0.99) values. Other K. selliformis strains formed three distinct, geographically consistent subclades, including either 1) Asian strains originating from Japan and China (66/0.94), 2) Chilean strains (87/0.99), or 3) Russian strains isolated from Kamchatka, Bering Sea (86/0.65).

The highly supported clade of K. papilionacea (100/1.00) comprised three subclades. The seven strains from Kuwait were found in a large subclade among strains originating from New Zealand (the type locality), Australia, Japan, China, Hong Kong, and Korea (89/0.99) (Figure 5). The remaining K. papilionacea sequences split into two well-supported subclades consisting of either European (86/0.99) or Japanese (85/0.99) strains. The strains isolated in Kuwait Bay in 2014 (KW-E1-14 and KW-F3-15) and those obtained during the Karenia bloom in 2021 (KW-G4-062, KW-G5-063, Kw-G8-064, and KW-G9-065) were identical (the pairwise genetic distance = 0).

Sequences obtained for Karlodinium from Kuwait were found in a clade of other K. ballantinum strains with strong support (99/1.00). The two Kuwait strains appeared in a subclade with Asian strains from Japan and the Philippines and the Australian strain, representing the holotype. A minor genetic divergence (p-distance = 0.0044, 3 bp difference) was detected between K. ballantinum strains isolated in Kuwait Bay in 2014 (KW-E9-46) and in the artificial KISR Salmiya marina to the south in 2016 (KW-JL-07), but in the phylogenetic analyses they clustered together (Figure 5).

4 Discussion

The family Kareniaceae comprises poorly preservable athecate species, most of which can only be distinguished morphologically by minor and sometimes subtle morphological differences. Identifying these species using light microscopy poses a challenge due to their morphological plasticity and the frequent co-occurrence of various kareniacean taxa in environmental samples. Consequently, an unambiguous species identification based on morphological examination alone is often impossible and usually requires support by genetic analyses (e.g., Haywood et al. 2004; Iwataki et al. 2022).

The composition of Kareniaceae taxa in Kuwait’s waters still needs to be understood. Until recently, K. selliformis and K. papilionacea were the only kareniacean species reported from Kuwait (Al-Kandari et al. 2009; Al-Yamani et al. 2004, 2012; Devlin et al. 2019; Glibert et al. 2002; Heil et al. 2001), although several tiny unidentified gymnodinioid dinoflagellates similar to Karlodinium or Takayama have been reported previously based on their morphology (e.g., Al-Kandari et al. 2009; Al-Yamani and Saburova 2019). These dinoflagellates have occasionally been observed in bloom proportions (Al-Yamani and Saburova 2019; Polikarpov et al. 2020, as Gymnodinium-like group). To date, the known diversity of the Kareniaceae in Kuwait has grown to at least 16 taxa belonging to Karenia, Karlodinium, and Takayama based on morphological observation of live cells (Al-Yamani and Saburova 2019; also see Table 4) However, the identities of these species are pending confirmation by molecular analysis.

Table 4:

Chronological summary of taxonomic and nomenclatural records of the Kareniaceae taxa in Kuwait’s waters.

Species Locality Period Material Testing method Evidence (abundance, cells l−1) Reference (illustrations)
Gymnodinium cf. mikimotoi Kuwait’s waters Sept-Oct 1999 Lugol’s-fixed field samples LM Bloom (2.7–7.6 × 106)

Fish kill
Husain and Faraj (2000)
Gymnodinium sp./Gymnodinium selliforme a sensu MacKenzie et al. 1995, 1996, Haywood et al. unpubl. Kuwait Bay Sept-Oct 1999 Lugol’s-fixed field samples, live cells LM Bloom (>106)

Fish kill
Heil et al. 2001 (fig. 2B, C)
Karenia sp.a sensu Haywood 2002; Heil et al. 2001 Kuwait Bay Aug–Sept 2001 Lugol’s-fixed field samples LM Presence in low abundance (400) Glibert et al. 2002 (fig. 5A, inset)
Karenia selliforme

Gymnodinium breve (Karenia breve)
Kuwait’s waters 1999–2002 Lugol’s-fixed field samples LM Listed Al-Yamani et al. (2004)
Karenia selliformis*

Karenia cf. brevis*
Kuwait’s waters 2003–2007 Lugol’s-fixed field samples LM, SEM Occurrence Al-Kandari et al. 2009 (pl. 7, 8A)
Karenia selliformis

Karenia papilionacea

Karenia mikimotoi

Takayama cf. pulchella
Kuwait’s waters 2004–2009 Lugol’s-fixed field samples LM Listed Al-Yamani et al. (2012)
Karenia selliformis

Karenia brevis

Karenia mikimotoi
Kuwait’s waters 2007–2016 Lugol’s-fixed field samples LM 3.3–65.1 × 103

1.2–2.3 × 105

4.3–31.0 × 103
Devlin et al. (2019)
Karenia digitata a

Karenia mikimotoi a

Karenia papilionacea a

Karenia selliformis a

Karenia umbella a

Karlodinium australe a

Karlodinium decipiens a

Karlodinium gentienii a

Karlodinium veneficum a

Takayama acrotrocha a

Takayama cladochroma a

Takayama helix a

Takayama tasmanica a

Takayama tuberculata a

Takayama spp. a
Kuwait’s waters 2004–2017 Lugol’s-fixed field samples, live cells LM, epifluorescence Rarely occurred

Occasionally occurred

Bloom-forming (2.4 × 107)

Frequently occurred

Rarely occurred

Rarely occurred

Rarely occurred

Occasionally occurred

Occasionally occurred

Rarely occurred

Rarely occurred

Rarely occurred

Occasionally occurred

Occasionally occurred

Rarely occurred
Al-Yamani and Saburova 2019 (pls. 58–72)
Karenia papilionacea a

Karenia selliformis a

Karlodinium ballantinum a
Kuwait Bay, artificial marinas 2014, 2016, 2021 Field samples, cultured strains, live cells LM, epifluorescence, LSU rDNA sequencing Morphological and molecular characterization This study (Figures 25)
  1. aIllustrated report.

In this study, the commonly occurring bloom-forming Karenia and Karlodinium species isolated from the coastal waters of Kuwait were characterized by LM and, for the first time, by molecular data inferred from partial LSU rDNA sequences. Based on LSU rDNA phylogenetic analyses, 10 strains were identified as three different species within the family Kareniaceae, namely K. selliformis, K. papilionacea, and Karl. ballantinum. The molecular data obtained in the present study confirmed the presence of these species in Kuwait’s waters and their identity.

4.1 Karenia selliformis

K. selliformis was the first species of the family Kareniaceae identified in Kuwait’s waters. Its appearance was linked to red tide and large fish kills in Kuwait Bay in September-October 1999 (Heil et al. 2001; Husain and Faraj 2000). It was subsequently recorded again in 2001 but did not form a bloom and was reported at a concentration of 400 cells l−1 (Glibert et al. 2002). This species was initially reported as Gymnodinium cf. mikimotoi (Husain and Faraj 2000) and then as Gymnodinium sp. or G. selliforme (Heil et al. 2001), and Karenia sp. (Glibert et al. 2002) (Table 4). In the course of a taxonomic examination of bloom material by A. Haywood and K. Steidinger, the Kuwaiti taxon was believed to be conspecific with yet undescribed Gymnodinium sp. from New Zealand coastal waters, which possessed the laterally elongated nucleus in the hypocone (Chang 1995; MacKenzie et al. 1995, 1996). Upon the description of the genus Karenia (Daugbjerg et al. 2000), this taxon was recognized as fitting the newly erected genus and redescribed as K. selliformis (Haywood et al. 2004). The authors noted the conspecificity of material from Kuwait due to similar morphology (Haywood et al. 2004; Steidinger et al. 2008). K. selliformis was recurrently found in Kuwait’s waters during the phytoplankton taxonomic survey in 2003–2007, when its morphology was illustrated by LM and SEM (Al-Kandari et al. 2009, pl. 7) (Table 4).

The observed morphological characters of K. selliformis cells from field samples and monoclonal culture (strain KW-D10-60) isolated from Kuwait’s waters are consistent with the original description of this species (Haywood et al. 2004) and previous records from Kuwait (Al-Yamani and Saburova 2019; Al-Kandari et al. 2009; Heil et al. 2001) in terms of their shape, apical groove length and path, and position of the nucleus (Figure 3; Table 3). The cells observed in this study were slightly larger than those described initially. The cells were 30–33 µm long, as compared to 26–32 µm (MacKenzie et al. 1996) and 20–27 μm (Haywood et al. 2004), but our measurements fall within the range previously reported from Kuwait (17–37 μm, Heil et al. 2001; 30–34 μm, Al-Yamani and Saburova 2019). The most notable characteristics distinguishing this species from the other Karenia are the large, transversely elongated nucleus located posteriorly and the higher number of chloroplasts (Table 3).

Molecular phylogenetic analysis using partial LSU rDNA showed that the strain from Kuwait clearly belonged to the K. selliformis clade and was genetically distinct from other Karenia species (Figure 5). Kuwait’s strain was most closely related to isolates from New Zealand and was 99.86 % identical (1 bp difference) to the holotype sequence of K. selliformis (U92250). This study provides the first molecular data on K. selliformis isolated from Kuwait’s waters and confirms the taxonomic identity of this dinoflagellate that has caused harmful blooms.

Recently, the genetic variability among the LSU rDNA and ITS sequences of K. selliformis strains originating from distant geographical areas worldwide was found to be high enough to form distinct nested subclades, suggesting the existence of two or three distinct phylotypes, depending on the molecular marker used (Iwataki et al. 2022; Mardones et al. 2020; Orlova et al. 2022). Moreover, the delineated phylotypes were distinguished phenotypically and by their temperature tolerance (Iwataki et al. 2022; Mardones et al. 2020; Orlova et al. 2022), pointing to the potential K. selliformis “species complex” (Mardones et al. 2020).

Our morphological and molecular data support the hypothesis of phylotype separation within K. selliformis, differentiating phenotypically and by temperature preference. In the phylogeny inferred from partial LSU rDNA, the topology of the K. selliformis clade is consistent with that previously described (Mardones et al. 2020). The K. selliformis strains were grouped into distinct nested subclades representing the different phylotypes (Figure 5). The basal subclade (phylotype I) comprised strains isolated from K. selliforms bloom along the coast of the Kamchatka Peninsula in 2020. The strain from subtropical Kuwait clustered with other warm-water strains originating from New Zealand, which were closely related to the Asian strains, belonging to phylotype II, and separated from another subclade that included the Chilean strains (phylotype III). The warm-water strain from Kuwait was represented by strongly dorso-ventrally flattened cells with a distinct antapical notch consistent with the described initially K. selliformis from New Zealand waters. In contrast, strains blooming in Chile (1999), Kamchatka, Russia (2020), and Hokkaido, Japan (2021), which belong to phylotype I, exhibited cells with abundant granular to strap-shaped chloroplasts and bloomed at low temperatures <20 °C (Iwataki et al. 2022; Orlova et al. 2022; Uribe and Ruíz 2001). The strains isolated during the K. selliformis bloom in Chile in 2018 (MN203220 and MN203221) were genetically distinct enough to form phylotype III. Morphologically, cells of this phylotype differed from the holotype strain by weak dorsal-ventral compression, shallower antapical excavation, and a smaller number of lateral pores in the hyposome (Mardones et al. 2020).

4.2 Karenia papilionacea

The taxonomic survey by Al-Kandari et al. (2009) revealed the second morphotype of Karenia in Kuwait’s waters. It was distinguished from K. selliformis by its cell shape, being wider rather than oblong. The small globular nucleus of this morphotype was located in the left lobe of the hypocone versus the larger, transversely elongated, and posteriorly located nucleus in K. selliformis (Table 3). While the morphology of this taxon resembled K. brevis, its unambiguous identification remained uncertain for a long time. In taxonomic and ecological phytoplankton studies in Kuwait’s waters, this taxon has been variously referred to as either Gymnodinium breve (Al-Yamani et al. 2004), Karenia cf. brevis (Al-Kandari et al. 2009, plate 8A, LM), K. brevis (Devlin et al. 2019), or K. papilionacea (Al-Yamani and Saburova 2019; Al-Yamani et al. 2012, plate 60, LM; Polikarpov et al. 2020) (Table 4).

Morphologically, K. papilionacea is similar to K. brevis, and these two species can be distinguished only by minor details. The cells of K. papilionacea are wider (24–36 µm vs 18–48 µm), possess pointed rather than bulbous apical protrusions (carina), and have slightly deeper antapical excavations (Haywood et al. 2004). Given the high morphological variability observed in both species, even within the same strain (Persson et al. 2013; Stuart 2011), it is doubtful whether species discrimination based on morphological criteria alone can be considered reliable. Nevertheless, these two species are genetically distinct (Haywood et al. 2004).

K. papilionacea may have previously been misidentified as K. brevis worldwide. The geographic range of K. brevis appears to be limited to the Gulf of Mexico (Brand et al. 2012; Magaña et al. 2003; Steidinger 2009). Previous reports of fish mortality caused by K. brevis in Asia and the Mediterranean Sea require molecular analysis to verify if they are actually misidentifications of K. papilionacea or other Karenia species (e.g., Tsikoti and Genitsaris 2021; Yamaguchi et al. 2016; Yeung et al. 2005).

The seven strains of K. papilionacea examined in this study displayed similarity to each other, and the morphological characters observed under LM were consistent with previous reports of this species (e.g., Al-Yamani and Saburova 2019; Kim et al. 2023; Yamaguchi et al. 2016), including its original description (Haywood et al. 2004). Despite some morphological variability within and among strains in cell size and shape, the typical cells were wide, dorsoventrally flattened, with short apical grooves, pointed apices, and round nuclei in the left lobe of the hypocone (Figure 2; Table 3).

In the phylogenetic tree based on partial LSU rDNA sequences, the strains from Kuwait were grouped together within the monophyletic clade of K. papilionacae among strains isolated from New Zealand (the type locality) and Japan with strong statistical support (Figure 5). The sequences obtained from Kuwait’s strains were identical, sharing 99.7 % similarity (2 bp differences) to that of the type strain of K. papilionacea (U92252), and were clearly genetically distinct from the K. brevis clade. Therefore, our results provide the first unambiguous molecular identification of K. papilionacea from Kuwait’s coastal waters and resolve the previous regional taxonomic uncertainty for this species. This implies that earlier records of K. brevis-like species in Kuwait’s waters need to be reconsidered and attributed to K. papilionacea. Furthermore, the genetic similarity among the strains isolated seven years apart (first in 2014 and then in 2021) and from different sampling sites (Kuwait Bay and semi-enclosed marina, Figure 1C; Table 1) suggests the persistent presence of the same population of this species throughout the study area.

Similar to K. selliforms, recent phylogenetic analyses based on both LSU and ITS have revealed genetic divergence among the K. papilionacea strains, resulting in the distinction of at least two phylotypes within this species (Kim et al. 2023; Yamaguchi et al. 2016). No morphological differences were observed between phylotypes, while different physiological growth traits were assumed to distinguish them (Yamaguchi et al. 2016). In our LSU-based phylogeny, the topology of the K. papilionacea clade was almost the same as previously reported by Yamaguchi et al. (2016) and Kim et al. (2023) and consisted of three distinct, well-supported subclades (Figure 5). All of Kuwait’s strains were placed in a large, diverse, and strongly supported clade (89/0.99) among strains from Australia, New Zealand, Japan, China, and Korea, corresponding to the original phylotype (as per Yamaguchi et al. 2016). The second distinct subclade, with high nodal support (85/0.99), solely comprised K. papilionacea strains restricted to the western Japanese coast (Yamaguchi et al. 2016), representing phylotype I. Two strains originating from the French Atlantic and north-western Mediterranean Sea, Spain, were genetically divergent from those of the original phylotype and phylotype I, clustering separately into a well-supported subclade (86/0.99).

4.3 Karlodinium ballantinum

Like Karenia, Karlodinium species possess the characteristic straight apical groove but differ in the presence of a ventral pore on the epicone in most species and a unique type of amphiesma with plugs. Typical Karlodinium cells are small, globular rather than flattened like most Karenia, and their straight versus sigmoid apical groove distinguishes them from the morphologically similar Takayama species (Daugbjerg et al. 2000; de Salas et al. 2003). Most Karlodinium species share similar morphology and often overlap in size, shape, and diagnostic morphological characters. Species differentiation within this genus is based on cell shape, the presence/absence of a ventral pore, the position of the nucleus, and the arrangement of chloroplasts. These characters are often difficult to distinguish in preserved samples but may be discerned by observing live cells (e.g., Bergholtz et al. 2006; de Salas et al. 2008).

Until recently, dinoflagellates belonging to the genus Karlodinium have not been recognized from Kuwait’s waters due to research being restricted to Lugol’s preserved samples and possible misidentification with other small gymnodinioid taxa. Five Karlodinium morphotypes have recently been reported from Kuwait based on the observations of live cells in freshly collected water samples (Table 4), including Karl. australe, Karl. decipiens, Karl. digitatum (as Karenia digitata), Karl. gentienii, and Karl. veneficum, as reported by Al-Yamani and Saburova (2019); however, no molecular data were provided for these species. In this study, two Kuwait strains, KW-JL-07 and KW-E9-046, exhibited the typical Karlodinium morphology, with small, nearly spherical to ellipsoidal cells, possessing a short straight apical groove, a large centrally located nucleus, and several elongated peripheral chloroplasts. The lack of a ventral pore in this species (as discerned by LM) sets it apart from most Karlodinium taxa previously reported in Kuwait, suggesting the finding of an unrecorded taxon. Among the currently described Karlodinium species, an inconspicuous or lacking ventral pore has been reported in Karl. antarcticum, Karl. ballantinum, Karl. digitatum, Karl. jejuense, and Karl. zhouanum (de Salas et al. 2008; Li and Shin 2018; Luo et al. 2018; Siano et al. 2009; Yang et al. 2000). The cell morphology of Kuwait’s strains differed from that described in four out of five pore-lacking Karlodinium species in terms of shape (ellipsoidal instead of elongated in Karl. antarcticum), nucleus position (centrally located nucleus versus those in the epicone in Karl. zhouanum or in the hypocone in Karl. antarcticum and Karl. digitatum), the apical groove length that extends for a short distance onto the dorsal epicone compared to its longer path (1/4–1/2 down the dorsal epicone) in Karl. antarcticum, Karl. digitatum, Karl. jejuense, and Karl. zhouanum, and the knob-like structures (microprocesses) lining the lower margin of the cingulum (Figure 4G) have not been observed in other Karlodinium species. The morphology observed in Kuwait’s strains corresponds well with that described for Karl. ballantinum (Table 3). Although de Salas et al. (2008) originally reported this species as lacking a ventral pore, further SEM analysis by Benico et al. (2020) of strains from the Philippines and Japan revealed a pore or pore-like shallow depression in this species. As our morphological observations of this species were limited to light microscopy, further SEM examination is required to determine whether this character varies in Kuwait’s material.

The phylogenetic tree inferred from the LSU rDNA sequences in the present study also showed a high genetic similarity between the sequences obtained from Kuwait and the type material of Karl. ballantinum from Tasmania (Figure 5). Kuwait’s strains shared 98.19–99.27 % similarity (5 bp differences) with the holotype strain EF469232. Through combined evidence from molecular phylogeny and morphology, this study reports the identification of K. ballantinum in Kuwait’s waters, marking the first record of this species in the entire Gulf.

4.4 Risk assessment of kareniacean harmful algal blooms

The family Kareniaceae is of particular global ecological concern because it contains many toxigenic bloom-forming species with a wide range of negative impacts. However, the types of harmful algal blooms (HABs), the toxins they produce, and hence the impacts of the blooms are species-specific and vary significantly from region to region. This makes a robust taxonomic identification of species within the Kareniaceae ecologically and economically important for regional HAB risk assessments.

A long-term phytoplankton monitoring program revealed that several kareniacean species are seasonally part of the phytoplankton community along Kuwait’s coast, often present at background to low concentrations (100–10,000 cells l−1) (Al-Kandari et al. 2009; Al-Yamani et al. 2012; Al-Yamani and Saburova 2019; Devlin et al. 2019; Polikarpov et al. 2020). Three species belonging to the genus Karenia, namely K. selliformis, K. papilionacea, and K. mikimotoi, as well as Karlodinium spp. and Takayama spp., can be considered bloom-forming due to their ability to form sporadic outbreaks (>105 cell l−1) in Kuwait’s waters resulting in surface water discolouration (red tides) (Al-Yamani and Saburova 2019; Al-Yamani et al. 2012; Heil et al. 2001; Polikarpov et al. 2020).

The dinoflagellate species of particular concern for Kuwait is the ichthyotoxic K. selliformis. Its intense bloom in Kuwait Bay in 1999 resulted in massive fish kills and considerable economic losses (Al-Yamani et al. 2004, 2012; Heil et al. 2001). Although no harmful blooms of this species have been recorded since then, Kuwait’s marine environment remains at risk of further K. selliformis blooms. This species routinely co-occurs with K. papilionacea and K. mikimotoi in Kuwait’s coastal waters, forming sporadic multi-species blooms (Polikarpov et al. 2020). The presence of these toxigenic Karenia species in the phytoplankton composition could potentially contribute to new HAB events in response to nutrient surplus and increased seawater temperatures, particularly in areas with low water exchange rates, such as the shallow, semi-enclosed Kuwait Bay and numerous artificial marinas along Kuwait’s coast.

The bloom-forming toxigenic kareniacean species can be considered a latent threat to Kuwait’s marine environment and requires more detailed and comprehensive studies of their taxonomy, dynamics, and toxicity. Accurate identification and quantification of cell abundance in the water column is essential for routine monitoring of coastal waters. Potential misidentification of tiny, fragile, and poorly preservable kareniacean cells can result in significant underestimation of abundance and consequently affect the quality of the risk assessment. The key diagnostic characters that distinguish different kareniacean morphotypes, in particular the apical and sulcal grooves and the ventral pore, can be discerned by high-resolution microscopy of live cells. However, these taxonomically important morphological characters are often indistinct in preserved samples, resulting in the loss of data when Lugol’s-fixed samples are used for routine monitoring programs. To facilitate further phytoplankton monitoring in Kuwait’s waters, representative images of Lugol’s-fixed cells were documented from the cultured strains of K. papilionacea, K. selliformis, and Karl. ballantinum, which were identified through close morphological and molecular analyses (Figures 2N, 3L, and 4L).

Routine monitoring based on microscopy alone is insufficient for unambiguous species identification among the species belonging to the family Kareniaceae; therefore, supporting molecular data is necessary to evaluate potential risks, particularly for species verification of high biomass blooms. Establishing a HAB culture collection at KISR with the morphological and genetic characterization of the cultured species is the first valuable resource for HAB research in Kuwait and the entire Gulf region. Testing the ichthyotoxicity of these local strains to fish and invertebrates could be a priority topic for further research.

The long-term uncertainty surrounding the identity of K. brevis/K. papilionacea has affected the recent regional checklist of phytoplankton (Attaran-Fariman and Asefi 2022), in which both species were listed. Further molecular studies are required to ascertain whether previous records of the K. brevis-like morphotype should be attributed to K. papilionacea or if both species co-occur across the Gulf, as has been observed in the Gulf of Mexico (Steidinger et al. 2008) and along Florida’s east coast (Wolny et al. 2015).

Further investigations should focus on the development of an efficient molecular marker to discriminate among the isolates of Kareniaceae species from Kuwait, the region, and worldwide in order to track their origin and possible spreading pathways. In this regard, the plastid gene ribulose-bisphosphate carboxylase (rbcL) sequences have been found to be more discriminating than the LSU and ITS nuclear markers (e.g., Daugbjerg et al. 1995) and have been successfully used to distinguish Karenia mikimotoi isolates originating from different geographical regions (Al-Kandari et al. 2011).


Corresponding author: Manal Al-Kandari, Environment & Life Sciences Research Center, Coastal and Marine Resources Program, Kuwait Institute for Scientific Research, P.O. Box 1638, Salmiya 22017, Kuwait, E-mail:

Funding source: Kuwait Foundation for the Advancement of Sciences (KFAS)

Award Identifier / Grant number: FM113C

Award Identifier / Grant number: Kuwait Institute for Scientific Research

About the authors

Manal Al-Kandari

Manal Al-Kandari (PhD) is a marine biologist with over 23 years of experience in the taxonomy and molecular biology of planktonic microalgae from a wide geographic range. Since 2012, she has been engaged in research on marine microalgae bloom of Kuwait at the Kuwait Institute for Scientific Research. Her research interests are focused on microalgal biodiversity, taxonomy, molecular biology and ecology, with a particular emphasis on harmful bloom species and their impact on coastal ecosystems.

Maria Saburova

Maria Saburova (PhD) is a marine biologist with over 30 years of experience in the taxonomy and ecology of planktonic and benthic microalgae from a wide geographic range. Since 2005, she has been engaged in research on marine microalgae in the Middle East region as a long-term scientific consultant at the Kuwait Institute for Scientific Research. Her research interests are focused on microalgal biodiversity, taxonomy, and ecology, with a particular emphasis on harmful bloom species and their impact on coastal ecosystems.

Igor Polikarpov

Igor Polikarpov is a biological oceanographer with over 35 years of research experience, working over 19 years at the Kuwait Institute for Scientific Research. His current research interests include biological oceanography of the marginal seas of the Indian Ocean, marine phytoplankton and microplankton ecology, and long-term natural and human-induced environmental changes in the coastal waters. He has co-authored several books on marine ecology and biological oceanography of the north-western Indian Ocean.

Jacob Larsen

Jacob Larsen (PhD) is a marine biologist with over 40 years’ experience in the taxonomy planktonic and benthic dinoflagellates from a wide geographic range. Since 1995, he has mainly been involved in the development, organization, and teaching of IOC courses on taxonomy and identification of harmful microalgae.

Acknowledgments

The authors thank all seagoing staff at Kuwait Institute for Scientific Research, Alan Lennox and his team, for collecting samples. We would like to express our gratitude to the Editor in Chief, Prof. Matthew J. Dring, and the two anonymous reviewers for their valuable and constructive comments and suggestions, which have helped us to improve the quality of the manuscript. Gratitude to the Kuwait Foundation for the Advancement of Sciences and the Kuwait Institute for Scientific Research for providing financial support.

  1. Research ethics: All research procedures were conducted in accordance with local laws. All data were obtained and collected in accordance with local research protocols.

  2. Informed consent: Not applicable.

  3. Author contributions: M. Al-Kandari: original concept, resources, methodology, sampling, analysis of molecular data, drafting and editing manuscript, supervision; M. Saburova: original concept, methodology, sampling, isolation and culturing, light microscopy, analysis of molecular data, visualization, drafting and editing manuscript; I. Polikarpov: sampling, culturing, methodology, software, drafting and editing manuscript; J. Larsen: sampling, isolation and culturing, light microscopy, editing manuscript; N. Lundholm, S. Hussain: analysis of molecular data, editing manuscript. All authors commented on and contributed to the final manuscript. The authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The authors state no conflict of interest.

  6. Research funding: Kuwait Foundation for the Advancement of Sciences (KFAS) and the Kuwait Institute for Scientific Research, project FM113C.

  7. Data availability: Partial nuclear encoded 28S (LSU) rDNA sequences were deposited in GenBank with accession numbers PP951870-PP951879.

References

Adolf, J.E., Bachvaroff, T.R., and Place, A.R. (2009). Environmental modulation of karlotoxin levels in strains of the cosmopolitan dinoflagellate, Karlodinium veneficum (Dinophyceae). J. Phycol. 45: 176–192, https://doi.org/10.1111/j.1529-8817.2008.00641.x.Search in Google Scholar PubMed

Al-Kandari, M., Al-Yamani, F., and Al-Rifaie, K. (2009). Marine phytoplankton atlas of Kuwait’s waters. Kuwait Institute for Scientific Research, Kuwait.Search in Google Scholar

Al-Kandari, M., Highfield, A.C., Hall, M.J., Hayes, P., and Schroeder, D.C. (2011). Molecular tools separate harmful algal bloom species, Karenia mikimotoi, from different geographical regions into distinct sub-groups. Harmful Algae 10: 636–643, https://doi.org/10.1016/j.hal.2011.04.017.Search in Google Scholar

Al-Yamani, F.Y. (2021). Fathoming the Northwestern Arabian Gulf: oceanography and marine biology. Kuwait Institute for Scientific Research, Kuwait.Search in Google Scholar

Al-Yamani, F. and Saburova, M. (2019). Marine phytoplankton of Kuwait’s waters. Cyanobacteria, dinoflagellates, flagellates, Vol. 1. Kuwait Institute for Scientific Research, Kuwait.Search in Google Scholar

Al-Yamani, F.Y., Bishop, J., Ramadhan, E., Al-Husaini, M., and Al-Ghadban, A.N. (2004). Oceanographic atlas of Kuwait waters. Kuwait Institute for Scientific Research, Kuwait.Search in Google Scholar

Al-Yamani, F., Saburova, M., and Polikarpov, I. (2012). A preliminary assessment of harmful algal blooms in Kuwait’s marine environment. Aquat. Ecosyst. Health Manag. 15: 64–72, https://doi.org/10.1080/14634988.2012.679450.Search in Google Scholar

Al-Yamani, F., Yamamoto, T., Al-Said, T., and Alghunaim, A. (2017). Dynamic hydrographic variations in northwestern Arabian Gulf over the past three decades: temporal shifts and trends derived from long-term monitoring data. Mar. Pollut. Bull. 122: 488–499, https://doi.org/10.1016/j.marpolbul.2017.06.056.Search in Google Scholar PubMed

Al-Yamani, F., Polikarpov, I., and Saburova, M. (2021) Northern Gulf marine biodiversity in relevance to the river discharge. In: Jawad, L.A. (Ed.). Southern Iraq’s marshes. Coastal research library, Vol. 37. Springer Nature Switzerland AG, pp. 379–438.10.1007/978-3-030-66238-7_20Search in Google Scholar

Asefi, M.A., Attaran-Fariman, G., and Mohammadpour, G. (2023). Harmful algal bloom of Karlodinium cf. veneficum (Dinophyceae) and marine organism mortality from the Northern coastal waters of the Oman Sea in Iran (2019). Iran. J. Fish. Sci. 22: 962–985, https://doi.org/10.22092/ijfs.2023.130263.Search in Google Scholar

Attaran-Fariman, G. and Asefi, M.A. (2022). Checklist of phytoplankton of the tropical Persian Gulf and Sea of Oman. Nova Hedwigia 114: 251–301, https://doi.org/10.1127/nova_hedwigia/2022/0687.Search in Google Scholar

Baldrich, Á.M., Díaz, P.A., Rosales, S.A., Rodríguez-Villegas, C., Álvarez, G., Pérez-Santos, I., Díaz, M., Schwerter, C., Araya, M., and Reguera, B. (2024). An unprecedented bloom of oceanic dinoflagellates (Karenia spp.) inside a fjord within a highly dynamic multifrontal ecosystem in Chilean Patagonia. Toxins 16: 77, https://doi.org/10.3390/toxins16020077.Search in Google Scholar PubMed PubMed Central

Benico, G., Takahashi, K., Lum, W.M., and Iwataki, M. (2019). Morphological variation, ultrastructure, pigment composition and phylogeny and the star-shaped dinoflagellate Asterodinium gracile (Kareniaceae, Dinophyceae). Phycologia 58: 405–418, https://doi.org/10.1080/00318884.2019.1601948.Search in Google Scholar

Benico, G., Takahashi, K., Lum, W.M., Yñiguez, A.T., and Iwataki, M. (2020). The harmful unarmored dinoflagellate Karlodinium in Japan and Philippines, with reference to ultrastructure and micropredation of Karlodinium azanzae sp. nov. (Kareniaceae, Dinophyceae). J. Phycol. 56: 1264–1282, https://doi.org/10.1111/jpy.13030.Search in Google Scholar PubMed

Bergholtz, T., Daugbjerg, N., Moestrup, Ø., and Fernandez-Tejedor, M. (2006). On the identity of Karlodinium veneficum and description of Karlodinium armiger sp. nov. (Dinophyceae), based on light and electron microscopy, nuclear-encoded LSU rDNA, and pigment composition. J. Phycol. 42: 170–193, https://doi.org/10.1111/j.1529-8817.2006.00172.x.Search in Google Scholar

Botes, L., Sym, S.D., and Pitcher, G.C. (2003). Karenia cristata sp. nov. and Karenia bicuneiformis sp. nov. (Gymnodiniales, Dinophyceae): two new Karenia species from the South Africa coast. Phycologia 42: 563–571, https://doi.org/10.2216/i0031-8884-42-6-563.1.Search in Google Scholar

Brand, L.E., Campbell, L., and Bresnan, E. (2012). Karenia: the biology and ecology of a toxic genus. Harmful Algae 14: 156–178, https://doi.org/10.1016/j.hal.2011.10.020.Search in Google Scholar PubMed PubMed Central

Cen, J., Wang, J., Huang, L., Ding, G., Qi, Y., Cao, R., Cui, L., and Lü, S. (2019). Who is the “murderer” of the bloom in coastal waters of Fujian, China, in 2019? J. Oceanol. Limnol. 38: 722–732, https://doi.org/10.1007/s00343-019-9178-6.Search in Google Scholar

Chang, F.H. (1995) The first records of Gymnodinium sp. nov. (cf. breve) (Dinophyceae) and other harmful phytoplankton species in the early 1993 blooms in New Zealand. In: Lassus, P., Arzul, G., Erard, G., Gentien, P., and Marcaillou, C. (Eds.). Harmful marine algalblooms. Lavoisier Publishing, Paris, pp. 145–150.Search in Google Scholar

Clément, A., Seguel, M., Arzul, G., Guzmán, L., and Alarcón, C. (2001) A widespread outbreak of a haemolytic, ichthyotoxic Gymnodinium sp. in southern Chile. In: Hallegraeff, G.M., Blackburn, S.I., Bolch, C.J., and Lewis, R.J. (Eds.). Harmful algal blooms 2000. Intergovernmental oceanographic commission. UNESCO, Paris, pp. 66–69.Search in Google Scholar

Daugbjerg, N., Moestrup, Ø., and Arctander, P. (1995). Phylogeny of genera of Prasinophyceae and Pedinophyceae (Chlorophyta) deduced from molecular analysis of the rbcL gene. Phycol. Res. 43: 203–213, https://doi.org/10.1111/j.1440-1835.1995.tb00026.x.Search in Google Scholar

Daugbjerg, N., Hansen, G., Larsen, J., and Moestrup, Ø. (2000). Phylogeny of some of the major genera of dinoflagellates based on ultrastructure and partial LSU rDNA sequence data, including the erection of three new genera of unarmoured dinoflagellates. Phycologia 39: 302–317, https://doi.org/10.2216/i0031-8884-39-4-302.1.Search in Google Scholar

de Salas, M.F., Bolch, C.J.S., Botes, L., Nash, G., Wright, S.W., and Hallegraeff, G.M. (2003). Takayama gen. nov. (Gymnodiniales, Dinophyceae), a new genus of unarmoured dinoflagellates with sigmoid apical grooves, including the description of two new species. J. Phycol. 39: 1233–1246, https://doi.org/10.1111/j.0022-3646.2003.03-019.x.Search in Google Scholar

de Salas, M.F., Laza-Martínez, A., and Hallegraeff, G.M. (2008). Novel unarmored dinoflagellates from the toxigenic family Kareniaceae (Gymnodiniales): five new species of Karlodinium and one new Takayama from the Australian sector of the Southern Ocean. J. Phycol. 44: 241–257, https://doi.org/10.1111/j.1529-8817.2007.00458.x.Search in Google Scholar PubMed

Devlin, M.J., Breckels, M., Graves, C.A., Barry, J., Capuzzo, E., Huerta, F.P., Al Ajmi, F., Al-Hussain, M.M., LeQuesne, W.J.F., and Lyons, B.P. (2019). Seasonal and temporal drivers influencing phytoplankton community in Kuwait marine waters: Documenting a changing landscape in the Gulf. Front. Mar. Sci. 6: 141, https://doi.org/10.3389/fmars.2019.00141.Search in Google Scholar

Feki, W., Hamza, A., Frossard, V., Abdennadher, M., Hannachi, I., Jacquot, M., Belhassen, M., and Aleya, L. (2013). What are the potential drivers of blooms of the toxic dinoflagellate Karenia selliformis? A 10-year study in the Gulf of Gabes, Tunisia, southwestern Mediterranean Sea. Harmful Algae 23: 8–18, https://doi.org/10.1016/j.hal.2012.12.001.Search in Google Scholar

Fowler, N., Tomas, C., Baden, D., Campbell, L., and Bourdelais, A. (2015). Chemical analysis of Karenia papilionacea. Toxicon 101: 85–91, https://doi.org/10.1016/j.toxicon.2015.05.007.Search in Google Scholar PubMed

Furuya, K., Iwataki, M., Lim, P.T., Lu, S., Leaw, C.-P., Azanza, R.V., Kim, H.-G., and Fukuyo, Y. (2018) Overview of harmful algal blooms in Asia. In: Glibert, P.M., Berdalet, E., Burford, M.A., Pitcher, G.C., and Zhou, M. (Eds.). Global ecology and oceanography of harmful algal blooms. Ecological Studies. Springer, Cham, pp. 289–308.10.1007/978-3-319-70069-4_14Search in Google Scholar

Glibert, P.M., Landsberg, J.H., Evans, J.J., Al-Sarawi, M.A., Faraj, M., Al-Jarallah, A., Haywood, A., Ibrahem, S., Klesius, P., Powell, C., et al.. (2002). A fish kill of massive proportion in Kuwait Bay, Arabian Gulf, 2001: the roles of bacterial disease, harmful algae and eutrophication. Harmful Algae 1: 215–231, https://doi.org/10.1016/s1568-9883(02)00013-6.Search in Google Scholar

Gómez, F. (2006). The dinoflagellate genera Brachidinium, Asterodinium, Microceratium and Karenia in the open SE Pacific ocean. Algae 21: 445–452, https://doi.org/10.4490/algae.2006.21.4.445.Search in Google Scholar

Hall, T.A. (1999). BioEdit: a user-friendly biological sequence alignment editor and analysis program for windows 95/98/NT. Nucleic Acids Symp. Ser. 41: 95–98.Search in Google Scholar

Hallegraeff, G. (2023) Introduction. In: Hallegraeff, G.M., Anderson, D.M., Davidson, K., Gianella, F., Hansen, P.J., Hegaret, H., Iwataki, M., Larsen, T.O., Mardones, J., MacKenzie, L., et al.. (Eds.). GlobalHAB. 2023. Fish-killing marine algal blooms: causative organisms, ichthyotoxic mechanisms, impacts and mitigation, IOC Manuals and Guides, 93, UNESCO-IOC/SCOR, Paris, pp. 11–16.Search in Google Scholar

Haywood, A.J. (2002). Morphological and molecular systematics of unarmoured dinoflagellates (Gymnodiniales, Dinophyta) from New Zealand, Ph.D. thesis. University of Auckland, Auckland, New Zealand.Search in Google Scholar

Haywood, A.J., Steidinger, K.A., Truby, E.W., Bergquist, P.R., Bergquist, P.L., Adamson, J., and MacKenzie, L. (2004). Comparative morphology and molecular phylogenetic analysis of three new species of the genus Karenia (Dinophyceae) from New Zealand. J. Phycol. 40: 165–179, https://doi.org/10.1111/j.0022-3646.2004.02-149.x.Search in Google Scholar

Heil, C.A., Glibert, P.M., Al-Sarawi, M.A., Faraj, M., Behbehani, M., and Husain, M. (2001). First record of a fish-killing Gymnodinium sp. bloom in Kuwait Bay, Arabian Sea: chronology and potential causes. Mar. Ecol. Prog. Ser. 214: 15–23, https://doi.org/10.3354/meps214015.Search in Google Scholar

Husain, M. and Faraj, M. (2000) Blooms of dinoflagellate Gymnodinium c.f. mikimotoi in Kuwaiti waters. In: Proceedings of the ninth international conference on harmful algal blooms, February 7-11, 2000: conference abstracts. Hobart, Tasmania, Australia, pp. 166.Search in Google Scholar

Iwataki, M. (2023) Taxonomy and identification of fish-killing harmful algae. In: Hallegraeff, G.M., Anderson, D.M., Davidson, K., Gianella, F., Hansen, P.J., Hegaret, H., Iwataki, M., Larsen, T.O., Mardones, J., MacKenzie, L., et al.. (Eds.). GlobalHAB. 2023. Fish-killing marine algal blooms: causative organisms, ichthyotoxic mechanisms, impacts and mitigation. UNESCO-IOC/SCOR, Paris, pp. 22–27, IOC Manuals and Guides, 93.Search in Google Scholar

Iwataki, M., Lum, W.M., Kuwata, K., Takahashi, K., Arima, D., Kuribayashi, T., Kosaka, Y., Hasegawa, N., Watanabe, T., Shikata, T., et al.. (2022). Morphological variation and phylogeny of Karenia selliformis (Gymnodiniales, Dinophyceae) in an intensive cold-water algal bloom in Eastern Hokkaido, Japan. Harmful Algae 114: 102204, https://doi.org/10.1016/j.hal.2022.102204.Search in Google Scholar PubMed

Keller, M.D., Selvin, R.C., Claus, W., and Guillard, R.R.L. (1987). Media for the culture of oceanic ultraphytoplankton. J. Phycol. 23: 633–638, https://doi.org/10.1111/j.1529-8817.1987.tb04217.x.Search in Google Scholar

Kim, S., Cho, M., Yoo, J., and Park, B. (2023). Application of a quantitative PCR to investigate the distribution and dynamics of two morphologically similar species, Karenia mikimotoi and K. papilionacea (Dinophyceae) in Korean coastal waters. Toxins 15: 469, https://doi.org/10.3390/toxins15070469.Search in Google Scholar PubMed PubMed Central

Kumar, S., Stecher, G., Li, M., Knyaz, C., and Tamura, K. (2018). MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35: 1547–1549, https://doi.org/10.1093/molbev/msy096.Search in Google Scholar PubMed PubMed Central

Landsberg, J.H. and Steidinger, K.A. (1998) A historical review of Gymnodinium breve red tides implicated in mass mortalities of the manatee (Trichechus manatus latirostris) in Florida USA. In: Reguera, B., Blanco, J., Fernandez, M., and Wyatt, T. (Eds.). Harmful algae. Xunta de Galicia. IOC UNESCO, Paris, pp. 97–100.Search in Google Scholar

Larsen, T.O. (2023) Chemistry and analytic methods - how to deal with the many congeners? In: Hallegraeff, G.M., Anderson, D.M., Davidson, K., Gianella, F., Hansen, P.J., Hegaret, H., Iwataki, M., Larsen, T.O., Mardones, J., MacKenzie, L., et al.. (Eds.). GlobalHAB. 2023. Fish-killing marine algal blooms: causative organisms, ichthyotoxic mechanisms, impacts and mitigation. IOC Manuals and Guides, 93, UNESCO-IOC/SCOR, Paris, pp. 28–35.Search in Google Scholar

Lassus, P., Chomérat, N., Hess, P., and Nézan, E. (2016). Toxic and harmful microalgae of the World Ocean/Micro-algues toxiques et nuisibles de L’océan Mondial. IOC manuals and Guides 68. International Society for the Study of Harmful Algae/Intergovernmental Oceanographic Commission of UNESCO, Denmark.Search in Google Scholar

Leong, S.C.Y., Lim, L.P., Chew, S.M., Kok, J.W.K., and Teo, S.L.M. (2015). Three new records of dinoflagellates in Singapore’s coastal waters, with observations on environmental conditions associated with microalgal growth in the Johor Straits. Raffles Bull. Zool. S31: 24–36.Search in Google Scholar

Li, Z. and Shin, H.H. (2018). Morphology and phylogeny of an unarmored dinoflagellate, Karlodinium jejuense sp. nov. (Gymnodiniales), isolated from the Northern East China Sea. Phycol. Res. 66: 318–328, https://doi.org/10.1111/pre.12334.Search in Google Scholar

Lim, H.C., Leaw, C.P., Tan, T.H., Kon, N.F., Yek, L.H., Hii, K.S., Teng, S.T., Razali, R.M., Usup, G., Iwataki, M., et al.. (2014). A bloom of Karlodinium australe (Gymnodiniales, Dinophyceae) associated with mass mortality of cage-cultured fishes in West Johor Strait, Malaysia. Harmful Algae 40: 51–62, https://doi.org/10.1016/j.hal.2014.10.005.Search in Google Scholar

Lin, Y., Risk, M., Ray, S.M., Van Engen, D., Clardy, J., Golik, J., James, J.C., and Nakanishi, K. (1981). Isolation and structure of brevetoxin B from the “red tide” dinoflagellate Ptychodiscus brevis (Gymnodinium breve). J. Am. Chem. Soc. 103: 6773, https://doi.org/10.1021/ja00412a053.Search in Google Scholar

Lundholm, N., Churro, C., Fraga, S., Hoppenrath, M., Iwataki, M., Larsen, J., Mertens, K., Moestrup, Ø., and Zingone, A. (Eds.). (2009 onwards). IOC-UNESCO taxonomic reference list of harmful micro algae, https://www.marinespecies.org/hab (Accessed 14 February 2024).Search in Google Scholar

Luo, Z., Wang, L., Chan, L., Lu, S., and Gu, H. (2018). Karlodinium zhouanum, a new dinoflagellate species from China, and molecular phylogeny of Karenia digitate and Karenia longicanalis (Gymnodiniales, Dinophyceae). Phycologia 57: 401–412, https://doi.org/10.2216/17-106.1.Search in Google Scholar

MacKenzie, L., Rhodes, L., Till, D., Chang, F.H., Kaspar, H., Haywood, A., Kapa, J., and Walker, B. (1995) A Gymnodinium sp. bloom and the contamination of shellfish with lipid soluble toxins in New Zealand, Jan-April 1993. In: Lassus, P., Arzul, G., Erard, G., Gentien, P., and Marcaillou, C. (Eds.). Harmful marine algal blooms. Lavoisier Publishing, Paris, pp. 795–800.Search in Google Scholar

MacKenzie, L., Haywood, A., Adamson, J., Truman, P., Till, D., Seki, T., Satake, M., and Yasusmoto, T. (1996) Gymnodimine contamination of shellfish in New Zealand. In: Yasumoto, T., Oshima, Y., and Fukuyo, Y. (Eds.). Harmful and toxic algal blooms. Inter-Governmental Oceanographic Commission of UNESCO, Tokyo, pp. 97–100.Search in Google Scholar

Magaña, H.A., Contreras, C., and Villareal, T.A. (2003). A historical assessment of Karenia brevis in the western Gulf of Mexico. Harmful Algae 2: 163–171, https://doi.org/10.1016/s1568-9883(03)00026-x.Search in Google Scholar

Mardones, J.I., Norambuena, L., Paredes, J., Fuenzalida, G., Dorantes-Aranda, J.J., Chang, K.J.L., Guzmán, L., Krock, B., and Hallegraeff, G. (2020). Unraveling the Karenia selliformis complex with the description of a non-gymnodimine producing Patagonian phylotype. Harmful Algae 98: 101892, https://doi.org/10.1016/j.hal.2020.101892.Search in Google Scholar PubMed

Mooney, B.D., Nichols, P.D., de Salas, M.F., and Hallegraeff, G.M. (2007). Lipid, fatty acid and sterol composition of eight species of Kareniaceae (Dinophyta): chemotaxonomy and putative phycotoxins. J. Phycol. 43: 101–111, https://doi.org/10.1111/j.1529-8817.2006.00312.x.Search in Google Scholar

Mooney, B.D., Hallegraeff, G.M., and Place, A.R. (2010). Ichthyotoxicity of four species of gymnodinioid dinoflagellates (Kareniaceae, Dinophyta) and purified karlotoxins to larval sheepshead minnow. Harmful Algae 9: 557–562, https://doi.org/10.1016/j.hal.2010.04.005.Search in Google Scholar

Nunn, G.B., Theisen, B.F., Christensen, B., and Arctander, P. (1996). Simplicity-correlated size growth of the nuclear 28S ribosomal RNA D3 expansion segment in the crustacean order Isopoda. J. Mol. Evol. 42: 211–223, https://doi.org/10.1007/bf02198847.Search in Google Scholar

Oh, J.W., Pushparaj, S.S.C., Muthu, M., and Gopal, J. (2023). Review of harmful algal blooms (HABs) causing marine fish kills: toxicity and mitigation. Plants (Basel) 12: 3936, https://doi.org/10.3390/plants12233936.10.3390/plants12233936Search in Google Scholar PubMed PubMed Central

Ok, J.H., Jeong, H.J., Lee, S.Y., Park, S.A., and Noh, J.-H. (2021). Shimiella gen. nov. and Shimiella gracilenta sp. nov. (Dinophyceae, Kareniaceae), a kleptoplastidic dinoflagellate from Korean waters and its survival under starvation. J. Phycol. 57: 70–91, https://doi.org/10.1111/jpy.13067.Search in Google Scholar PubMed

Orlova, T.Y., Aleksanin, A.I., Lepskaya, E.V., Efimova, K.V., Selina, M.S., Morozova, T.V., Stonik, I.V., Kachur, V.A., Karpenko, A.A., Vinnikov, K.A., et al.. (2022). A massive bloom of Karenia species (Dinophyceae) off the Kamchatka coast, Russia, in the fall of 2020. Harmful Algae 120: 102337, https://doi.org/10.1016/j.hal.2022.102337.Search in Google Scholar PubMed

Persson, A., Smith, B.C., Morton, S., Shuler, A., and Wikfors, G.H. (2013). Sexual life stages and temperature dependent morphological changes allow cryptic occurrence of the Florida red tide dinoflagellate Karenia brevis. Harmful Algae 30: 1–9, https://doi.org/10.1016/j.hal.2013.08.004.Search in Google Scholar

Polikarpov, I., Al-Yamani, F., and Saburova, M. (2019) Remote sensing of phytoplankton variability in the Arabian/Persian Gulf. In: Barale, V., and Gade, M. (Eds.). Remote sensing of the Asian seas. Springer, Cham, pp. 485–504.10.1007/978-3-319-94067-0_27Search in Google Scholar

Polikarpov, I., Saburova, M., and Al-Yamani, F. (2020). Decadal changes in diversity and occurrence of microalgal blooms in the NW Arabian/Persian Gulf. Deep-Sea Res. II: Top. Stud. Oceanogr. 179: 104810, https://doi.org/10.1016/j.dsr2.2020.104810.Search in Google Scholar

Ronquist, F., Teslenko, M., Mark, P.V.D., Ayres, D.L., Darling, A., Höhna, S., Liu, L., Suchard, M.A., and Huelsenbeck, J.P. (2012). MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61: 539–542, https://doi.org/10.1093/sysbio/sys029.Search in Google Scholar PubMed PubMed Central

Sakamoto, S., Lim, W.A., Lu, D., Dai, X., Orlova, T., and Iwataki, M. (2020). Harmful algal blooms and associated fisheries damage in East Asia: current status and trends in China, Japan, Korea and Russia. Harmful Algae 102: 101787, https://doi.org/10.1016/j.hal.2020.101787.Search in Google Scholar PubMed

Scholin, C.A., Herzog, M., Sogin, M., and Anderson, D.M. (1994). Identification of group- and strain-specific genetic markers for globally distributed Alexandrium (Dinophyceae). II. Sequence analysis of a fragment of the LSU rDNA. J. Phycol. 30: 999–1011, https://doi.org/10.1111/j.0022-3646.1994.00999.x.Search in Google Scholar

Sheppard, C., Al-Husiani, M., Al-Jamali, F., Al-Yamani, F., Baldwin, R., Bishop, J., Benzoni, F., Dutrieux, E., Dulvy, N.K., Subba Rao, D.V., et al.. (2010). The Gulf: a young sea in decline. Mar. Pollut. Bull. 60: 13–38, https://doi.org/10.1016/j.marpolbul.2009.10.017.Search in Google Scholar PubMed

Siano, R., Kooistra, W.H.C.F., Montresor, M., and Zingone, A. (2009). Unarmoured and thin-walled dinoflagellates from the Gulf of Naples, with the description of Woloszynskia cincta sp. nov. (Dinophyceae, Suessiales). Phycologia 48: 44–65, https://doi.org/10.2216/08-61.1.Search in Google Scholar

Siswanto, E., Luang-on, J., Ogata, K., Higa, H., and Toratani, M. (2024). Observations of water optical properties during red tide outbreaks off southeast Hokkaido by GCOM-C/SGLI: implications for the development of red tide algorithms. Remote Sens. Lett. 15: 121–132, https://doi.org/10.1080/2150704x.2024.2302948.Search in Google Scholar

Steidinger, K.A. (2009). Historical perspective on Karenia brevis red tide research in the Gulf of Mexico. Harmful Algae 8: 549–561, https://doi.org/10.1016/j.hal.2008.11.009.Search in Google Scholar

Steidinger, K.A., Landsberg, J.H., Truby, E.W., and Roberts, B.S. (1998). First report of Gymnodinium pulchellum (Dinophyceae) in North America and associated fish kills in the Indian River, Florida. J. Phycol. 34: 431–437, https://doi.org/10.1046/j.1529-8817.1998.340431.x.Search in Google Scholar

Steidinger, K.A., Wolny, J.L., and Haywood, A.J. (2008). Identification of Kareniaceae (Dinophyceae) in the Gulf of Mexico. Nova Hedwegia 133: 269–284.Search in Google Scholar

Stuart, M.R. (2011). Morphological studies of the dinoflagellate Karenia papilionacea in culture, Ph.D. thesis. University of North Carolina Wilmington, Wilmington, NC, USA.Search in Google Scholar

Takahashi, K., Benico, G., Lum, W.M., and Iwataki, M. (2019). Gertia stigmatica gen. et sp. nov. (Kareniaceae, Dinophyceae), a new marine unarmored dinoflagellate possessing the peridinin-type chloroplast with an eyespot. Protist 170: 1–19, https://doi.org/10.1016/j.protis.2019.125680.Search in Google Scholar PubMed

Tamura, K. and Nei, M. (1993). Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Mol. Biol. Evol. 10: 512–526, https://doi.org/10.1093/oxfordjournals.molbev.a040023.Search in Google Scholar PubMed

Tester, P.A., Turner, J.T., and Shea, D. (2000). Vectorial transport of toxins from the dinoflagellate Gymnodinium breve through copepods to fish. J. Plankton Res. 22: 47–61, https://doi.org/10.1093/plankt/22.1.47.Search in Google Scholar

Tester, P.A., Shea, D., Kibler, S.R., Varnam, S.M., Black, M.D., and Litaker, R.W. (2008). Relationships among water column toxins, cell abundance and chlorophyll concentrations during Karenia brevis blooms. Cont. Shelf Res. 28: 59–72, https://doi.org/10.1016/j.csr.2007.04.007.Search in Google Scholar

Thompson, J.D., Higgins, D.G., and Gibson, T.J. (1994). ClustalLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22: 4673–4680, https://doi.org/10.1093/nar/22.22.4673.Search in Google Scholar PubMed PubMed Central

Tsikoti, C. and Genitsaris, S. (2021). Review of harmful algal blooms in the coastal Mediterranean Sea, with a focus on Greek waters. Diversity 13: 396, https://doi.org/10.3390/d13080396.Search in Google Scholar

Uribe, J.C. and Ruíz, M. (2001). Gymnodinium brown tide in the Magellanic fjords, southern Chile. Rev. Biol. Mar. Oceanogr. 36: 155–164, https://doi.org/10.4067/s0718-19572001000200004.Search in Google Scholar

Wolny, J., Scott, P., Tustison, J., and Brooks, C. (2015). Monitoring the 2007 Florida East coast Karenia brevis (Dinophyceae) red tide and neurotoxic shellfish poisoning (NSP) event. Algae 30: 49–58, https://doi.org/10.4490/algae.2015.30.1.049.Search in Google Scholar

Wolny, J.L., Whereat, E.B., Egerton, T.A., Gibala-Smith, L.A., McKay, J.R., O’Neil, J.M., Wazniak, C.E., and Mulholland, M.R. (2024). The occurrence of Karenia species in mid-Atlantic coastal waters: data from the Delmarva Peninsula, USA. Harmful Algae 132: 102579, https://doi.org/10.1016/j.hal.2024.102579.Search in Google Scholar PubMed

Yamaguchi, H., Hirano, T., Yoshimatsu, T., Tanimoto, Y., Matsumoto, T., Suzuki, S., Hayashi, Y., Urabe, A., Miyamura, K., Sakamoto, S., et al.. (2016). Occurrence of Karenia papilionacea (Dinophyceae) and its novel sister phylotype in Japanese coastal waters. Harmful Algae 57: 59–68, https://doi.org/10.1016/j.hal.2016.04.007.Search in Google Scholar PubMed

Yang, Z.B., Takayama, H., Matsuoka, K., and Hodgkiss, I.J. (2000). Karenia digitata sp. nov. (Gymnodiniales, Dinophyceae), a new harmful algal bloom species from the coastal waters of West Japan and Hong Kong. Phycologia 39: 463–470, https://doi.org/10.2216/i0031-8884-39-6-463.1.Search in Google Scholar

Yeung, P.K.K., Hung, V.K.L., Chan, F.K.C., and Wong, J.T.Y. (2005). Characterization of a Karenia papilionacea-like dinoflagellate from the South China sea. J. Mar. Biol. Assoc. U.K. 85: 779–781, https://doi.org/10.1017/s0025315405011690.Search in Google Scholar

Yñiguez, T.Y., Lim, P.T., Leaw, C.P., Jipanin, S.J., Iwataki, M., Benico, G., and Azanza, R.V. (2021). Over 30 years of HABs in the Philippines and Malaysia: what have we learned? Harmful Algae 102: 101776, https://doi.org/10.1016/j.hal.2020.101776.Search in Google Scholar PubMed

Zingone, A., Escalera, L., Aligizaki, K., Fernández-Tejedor, M., Ismael, A., Montresor, M., Mozetič, P., Taş, S., and Totti, C. (2021). Toxic marine microalgae and noxious blooms in the Mediterranean Sea: a contribution to the global HAB status report. Harmful Algae 102: 101843, https://doi.org/10.1016/j.hal.2020.101843.Search in Google Scholar PubMed

Received: 2024-10-04
Accepted: 2025-01-24
Published Online: 2025-03-14
Published in Print: 2025-04-28

© 2025 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 11.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/bot-2024-0083/html
Scroll to top button