You are on page 1of 27

Harmful Algae 98 (2020) 101850

Contents lists available at ScienceDirect

Harmful Algae
journal homepage: www.elsevier.com/locate/hal

Review

Bioluminescence and toxicity as driving factors in harmful algal blooms: T


Ecological functions and genetic variability
Kathleen D. Cusicka, , Edith A. Widderb

a
University of Maryland Baltimore County, Department of Biological Sciences, 1000 Hilltop Circle, Baltimore, MD 21250, United States
b
Ocean Research and Conservation Association, 1420 Seaway Dr, Fort Pierce, FL 34949, United States

ARTICLE INFO ABSTRACT

Keywords: Dinoflagellates are an ecologically important group of marine microbial eukaryotes with a remarkable array of
Bioluminescence adaptive strategies. It is ironic that two of the traits for which dinoflagellates are best known, toxin production
Dinoflagellates and bioluminescence, are rarely linked when considering the ecological significance of either. Although dino-
Harmful algal blooms flagellate species that form some of the most widespread and frequent harmful algal blooms (HABs) are bio-
Luciferase
luminescent, the molecular and eco-evolutionary associations between these two traits has received little at-
Saxitoxin
tention. Here, the major themes of biochemistry and genetics, ecological functions, signaling mechanisms, and
Toxins
evolution are addressed, with parallels and connections drawn between the two. Of the 17 major classes of
dinoflagellate toxins, only two are produced by bioluminescent species: saxitoxin (STX) and yessotoxin. Of these,
STX has been extensively studied, including the identification of the STX biosynthetic genes. While numerous
theories have been put forward as to the eco-evolutionary roles of both bioluminescence and toxicity, a general
consensus is that both function as grazing deterrents. Thus, both bioluminescence and toxicity may aid in HAB
initiation as they alleviate grazing pressure on the HAB species. A large gap in our understanding is the genetic
variability among natural bloom populations, as both toxic and non-toxic strains have been isolated from the
same geographic location. The same applies to bioluminescence, as there exist both bioluminescent and non-
bioluminescent strains of the same species. Recent evidence demonstrating that blooms are not monoclonal
events necessitates a greater level of understanding as to the genetic variability of these traits among sub-
populations as well as the mechanisms by which cells acquire or lose the trait, as sequence analysis of STX+ and
STX- species indicate the key gene required for toxicity is lost rather than gained. While the extent of genetic
variability for both bioluminescence and toxicity among natural HAB sub-populations remains unknown, it is an
area that needs to be explored in order to gain greater insights into the molecular mechanisms and environ-
mental parameters driving HAB evolution.

1. Introduction discussing the ecological significance of either. Although it is well


known that many of the dinoflagellates that form harmful algal blooms
Dinoflagellates are a common component of the marine plankton, (HABs) are bioluminescent, the significance of light emission on the
where they are best known for the roles they play in biogeochemical generation and persistence of HABs has received little attention com-
cycling, toxin production, and bioluminescence (Hackett et al., 2004). pared to an ever expanding array of approaches focused on under-
Autotrophic and heterotrophic species contribute to cycling of key standing these complex phenomena (Anderson et al., 2012b). This
elements such as C, N and P. Toxin-producing species can significantly paper examines the relationship between bioluminescence and toxicity
alter food webs and impact both human and ecosystem health via the in dinoflagellates under the major themes of: Occurrence; Biochemistry
bioaccumulation and transfer of toxins through the food chain. Biolu- and Genetics; Energetic Costs; Ecological Functions, including the ge-
minescent species are responsible for brilliant, eye-catching displays netic variability of both traits within natural populations; Signal
that capture intense public interest and are frequently a focus of eco- Transduction; and Evolution.
tourism.
It is ironic that the two features for which dinoflagellates are best
known, light emission and toxin production, are rarely linked when


Corresponding author.
E-mail addresses: kcusick@umbc.edu (K.D. Cusick), ewidder@teamorca.org (E.A. Widder).

https://doi.org/10.1016/j.hal.2020.101850
Received 15 September 2019; Received in revised form 29 May 2020; Accepted 2 June 2020
Available online 29 July 2020
1568-9883/ © 2020 Elsevier B.V. All rights reserved.
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 1
List of known bioluminescent dinoflagellate species.a and associated toxin production. Presence of luciferin binding protein (LBP), when known, and corresponding
GenBank accession number are indicated.
Species Toxic Toxin produced LBP GenBank LBP

Alexandrium ostenfeldii Y saxitoxin, spirolides, gymnodimines n/a


Alexandrium acatenella Y Saxitoxin n/a
Alexandrium affine Y Saxitoxin Y AFN26992.1
Alexandrium cf. catenella Y Saxitoxin Y ABY78836.1
Alexandrium fundyense Y Saxitoxin n/a
Alexandrium monilatum Y Saxitoxin Y AFN26995.1
Alexandrium tamarense Y Saxitoxin Y AFN27013.1
Ceratium breve N/A
Ceratium furca N
Ceratium fusus N
Ceratium gibberum N
Ceratium horridum N
Ceratium longipes N
Ceratocorys horrida N
Dissodinium lunula N
Fragilidium heterolobum N
Gonyaulax catenata N
Gonyaulax digitale N
Gonyaulax excavata Y Saxitoxin n/a
Gonyaulax hyalina N/A
Gonyaulax monacantha N/A
Gonyaulax parva N
Gonyaulax ploygramma N
Gonyaulax scrippsae N
Gonyaulax sphaeroides N/A
Gonyaulax spinifera Y Yessotoxin Y JQ946866.1
Gymnodinium flavum N
Gymnodinium sanguineum N
Lingulodinium polyedrum Y Yessotoxin Y AAA29166.1
Noctiluca scintillans N
Peridinium elegans N
Polykrikos kofoidii N
Polykrikos schwartzii N
Protoceratium reticulatum Y Yessotoxin Y AFN27016.1
Protoperidinium antarcticum N/A
Protoperidinium brevipes/breve N
Protoperidinium brochii N/A
Protoperidinium ceraseus N/A
Protoperidinium claudicans N
Protoperidinium concoides N
Protoperidinium conicum N
Protoperidinium crassipes N
Protoperidinium curtipes N/A
Protoperidinium depressum N
Protoperidinium divergens N
Protoperidinium excentricum N
Protoperidinium exiquipes N/A
Protoperidinium globulus N/A
Protoperidinium granii N/A
Protoperidinium huberi N/A
Protoperidinium leonis N
Protoperidinium minutum N
Protoperidinium oceanicum N
Protoperidinium ovatum N/A
Protoperidinium pallidum N
Protoperidinium pellucidum N
Protoperidinium pentagonum N
Protoperidinium punctulatum N
Protoperidinium pyriforme N/A
Protoperidinium saltans N/A
Protoperidinium sinaicum N/A
Protoperidinium sournia N/A
Protoperidinium steinii N/A
Protoperidinium subinerme N
Protoperidinium tubum N/A
Pyrocystis acuta N
Pyrocystis fusiformis N
Pyrocystis lunula N
Pyrocystis noctiluca N
Pyrodinium bahamense Y Saxitoxin Y GAIO01054099.1
Pyrophacus steinii N

a
Species summarized from (Cusick and Widder, 2014) and (Marcinko et al., 2013). N/A in Toxic column indicates no information available on whether species is
toxic; n/a in LBP column indicates no information available on whether the species possesses an LBP.

2
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

2. Occurrence (Dominguez et al., 2010; Paz et al., 2008). The majority of YTXs syn-
thesized from dinoflagellates were first identified in Protoceratium re-
2.1. Bioluminescence ticulatum (Alfonso et al., 2016). YTXs were originally included as
Diarrhetic Shellfish Poisoning (DSP) Toxins, as they were commonly
The phenomenon of bioluminescence is widespread in the marine found with okadaic acid during DSP-screening assays (Tubaro et al.,
environment. Bioluminescence is the chemical generation of light by 2010). However, they have recently been re-classified as they do not
organisms and occurs across diverse taxa from bacteria to fish in more share the same mechanism of action as okadaic acid. No human toxicity
than 700 genera (Haddock et al., 2010; Widder, 2010). Although the has been reported for YTXs (Tubaro et al., 2010). The precise me-
number of bioluminescent species is relatively rare, most sampling chanism of action of YTX remains unknown (Paz et al., 2008), though in
schemes have revealed that the total number of light emitters in the vitro studies indicate it modulates calcium homeostasis and is able to
ocean significantly outnumber non-bioluminescent individuals inhibit entry of Ca+ ions (Paz et al., 2008). The majority of YTX-pro-
(Martini and Haddock, 2017; Widder et al., 1999). The relative success ducing dinoflagellates are bioluminescent and include P. reticulatum
of overall bioluminescent species is also evident in the high degree of (syn Gonyaulax grindleyi), Gonyaulax spinifera, and Lingulodinium poly-
convergent evolution of the trait with multiple origins (>40) of dif- edra (Draisci et al., 1999; Howard et al., 2009; Paz et al., 2008;
ferent light producing systems distributed across phyla (Haddock et al., Paz et al., 2004) (Table 1). In keeping with other toxin-producing di-
2010; Widder, 2010). noflagellate species, there exist both toxic and non-toxic strains of all
Dinoflagellates are the primary source of bioluminescence in surface three YTX-producers as well (Armstrong and Kudela, 2006;
waters. The overall number of described dinoflagellate species is esti- Howard et al., 2008, 2009; Ramstad et al., 2001; Rhodes et al., 2005).
mated to be between 2000 and 2500, assigned to >30 genera, with
approximately 1555 recognized as marine species (Gomez, 2005). 2.3. Toxic, bioluminescent dinoflagellates
However, only a small percentage are bioluminescent: 70 species
(Table 1), or less than 3%. Yet these include some of the best known Neither bioluminescence nor toxicity are confined to a specific order
and extensively researched of the dinoflagellata. These species have or genus; when considered individually, both are scattered among the
attracted attention due to their ability to form HABs (often referred to dinoflagellate tree of life (Fig. 1). However, the two traits occur to-
as “red tides”), their toxic impacts on shellfish and humans, and their gether in a more defined cluster of genera, all within the order Go-
production of brilliant displays of bioluminescence that illuminate nyaulacales. Additionally, a general trend is that toxic, bioluminescent
every disturbance in the water: outlining the forms of swimming fish dinoflagellates all occur in the similar environmental niche of coastal
and dolphins, illuminating breaking waves and leaving shimmering areas. Of the ca. 70 bioluminescent dinoflagellate species, 12 are es-
wakes behind ocean vessels. tablished toxin producers (Table 1). Of the bioluminescent toxin-pro-
ducers, Alexandrium sp. and P. bahamense are notorious STX-producing
2.2. Toxicity HAB species, while L. polyedra, G. spinifera, and P. reticulatum produce
YTX. Most of these toxic bioluminescent species have caused significant
There are approximately seventeen major types of dinoflagellate HAB outbreaks worldwide.
toxins (Table 2). As with bioluminescence, only a small portion of
marine dinoflagellates are toxic - ca. 100 species (Table 3). These 3. Biochemistry and genetics
“toxins” are comprised of an array of structurally and mechanistically
diverse secondary metabolites. A general trend is that within the same 3.1. Biochemistry: bioluminescence
species there are toxic and non-toxic strains. Additionally, toxin profile
and composition frequently differ among toxic strains of the same Dinoflagellate light emission occurs as a flash, triggered by a
species, and can be influenced by many parameters (Hoppenrath et al., membrane deformation that activates an action potential in the va-
2013). There is also the unique case of Alexandrium: among HAB cuolar membrane (Eckert and Sibaoka, 1968; Widder and Case, 1981a).
genera, Alexandrium is one of the most well-known in terms of bloom In general, the reaction occurs through the oxidation of the luciferin
severity, diversity, and frequency. Of its ca. 30 species, at least half are substrate by molecular oxygen via the luciferase enzyme. In dino-
confirmed toxin-producers. It is the only HAB genus with three different flagellates, the reaction occurs in many (~400 per cell) small (~0.4
families of toxins produced by its species: saxitoxin, spirolides, and um) organelles known as scintillons (DeSa and Hastings, 1968), which
goniodomins (Anderson et al., 2012a). contain the luciferin and luciferase (LCF) enzyme, and in some genera,
Of the 17 major classes of dinoflagellate toxins, only two are pro- a luciferin-binding protein (LBP) (Johnson et al., 1985; Nicolas et al.,
duced by bioluminescent species: saxitoxin and yessotoxin (Table 1), 1991; Schmitter et al., 1976). Mechanical stimulation of the cells cre-
and so are briefly expanded upon here. Saxitoxin (STX) (Table 2) is the ates an action potential that opens membrane channels, resulting in a
parent molecule in a class of compounds known as paralytic shellfish rapid and transient drop in pH within the scintillons due to a rapid
toxins (PSTs) (Table 4). STX and its derivatives target voltage-gated ion influx of protons from the acidic vacuole. The pH decrease activates the
channels in humans: sodium, potassium, and calcium, though the me- enzyme, and the oxidation of the luciferin results in a flash of light with
chanism of action differs among the three (Cusick and Sayler, 2013). It peak emission in the 472–475 nm wavelength range (Hastings, 1983;
is the causative agent of paralytic shellfish poisoning (PSP) and sax- Widder et al., 1983; Wilson and Hastings, 1998). The entire mechan-
itoxin pufferfish poisoning (SPFP) (Landsberg et al., 2006). Filter- otransductive process from stimulus to the initiation of the flash is
feeding mollusks and crustaceans ingest the toxic cells, concentrating rapid, requiring only 12–20 ms (Eckert, 1965; Latz et al., 2008;
the toxins within their organs and tissues. Currently, three genera of Widder and Case, 1981a). Flash kinetics and photon flux cover a re-
marine dinoflagellates are established STX producers: Alexandrium spp., markably wide range depending on species. The smaller (~20–40 µm)
Pyrodinium bahamense, and Gymnodinium catenatum (Cusick and armored dinoflagellates like L. polyedra emit brief (~100 ms), often
Sayler, 2013). Of these, Alexandrium spp. and P. bahamense are biolu- solitary dim flashes of between 3.1 × 107 and 1.2 × 108 photons cell−1
minescent (Table 1). (Biggley et al., 1969; Seliger et al., 1969; Swift et al., 1973;
Yessotoxin (YTX) and its ca. 100 analogues are polycyclic ether Widder et al., 1993) while the brightest dinoflagellates, such as Pyr-
polyketides (Table 2); the basic structure of YTX is that of a disulfated ocystis noctiluca (~350 µm) and Pyrocystis fusiformis (up to 1000 µm),
polyether, with the characteristic ladder-shape comprised of 11 ad- can emit multiple longer (200–500 ms) and brighter flashes of between
jacent ether rings of different sizes and a terminal acyclic unsaturated 3.7 × 1010 to 6.5 × 1011 photons cell−1 (Swift et al., 1973; Widder and
side chain comprised of 9 carbons and 2 sulfate ethers Case, 1981b; Widder et al., 1993).

3
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 2
List of major dinoflagellate toxins.

(continued on next page)

4
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 2 (continued)

Spirolide (MW = 694)


macrocyclic Alexandrium neuronal competitive binding to unknown
imine ostenfeldii, A. nicotinic nAChR
peruvianum acetylcholine
receptors
(nAChR)

Pinnatoxin (MW = 711.9)


macrocyclic Vulcanodinium neuronal competitive binding to unknown
imine rugosum nicotinic nAChR
acetylcholine
receptors
(nAChR)

Palytoxin (MW = 2680.1)


linear Ostreopsis cf. Na:K ATPase channel modifier, Palytoxin Poisoning
superchain ovata allowing non-selective - toxic in humans
polyether ion passage; binds to following ingestion,
extracellular portion of inhalation, or
protein cutaneous contact

Amphidinol (MW = 1327.7)


linear Amphidinium cell pore antifungal and
superchain klebsii membranes formation/disruption of hemolytic; no human
polyether biomembranes - apolar illnesses
side chain penetrates
lipid bilayer

(continued on next page)

5
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 2 (continued)

(continued on next page)

6
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 2 (continued)

Yessotoxin (MW = 1143.4)


ladder-frame Protoceratium unknown exact mechanism Neurotoxic Shellfish
polyether reticulatum unknown; potentially Poisoning via
targets cardiac muscle consumption of
cells, cytoskeleton contaminated
shellfish

Gymnocins A (top) and B


ladder-frame Karenia mikimotoi unknown unknown unknown
polyether

Maitotoxin (MW = 3425.9)


super-sized Gambierdiscus cation/calcium channel modifier, allowing unknown
ladder-like toxicus channels non-selective ion passage
polyether

Brevisculan F (2054.39)
super-sized Karenia unknown unknown lethal to mice; fish
ladder-like brevisulcata and humans
polyether unknown
linear Coolia monotis unknown unknown symptoms similar to
polyether NSP
(structure
unknown)
linear Ostreopsis unknown inactivation gating symptoms similar to
polyether lenticularis modifier NSP
(structure
unknown)

7
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 3
List of toxic dinoflagellate species.
Toxin Species Ref

Amphidinol −17, −18 Amphidinium carterae (Meng et al., 2010; Nuzzo et al., 2014)
Amphidinol −2, −3 Amphidinium klebsii (Iwamoto et al., 2017; Paul et al., 1995)
Azaspiracid Azadinium spinosum (Tillmann et al., 2009)
Azaspiracids Azadinium poporum (Luo et al., 2016)
Brevetoxins Karenia bicuneiformis (Brand et al., 2012)
Brevetoxins Karenia brevis (Brand et al., 2012)
Brevetoxins Karenia papilonacea (Brand et al., 2012)
Brevetoxins Karenia selliformis (Harju et al., 2016; Brand et al., 2012)
Ciguatoxins Gamberdiscus australes (Parsons et al., 2012)
Ciguatoxins Gamberdiscus pacificus (Parsons et al., 2012)
Ciguatoxins Gamberdiscus polynesiensis (Parsons et al., 2012)
Ciguatoxins Gamberdiscus toxicus (Parsons et al., 2012)
Ciguatoxins Gambierdiscus excentricus (Fraga et al., 2011)
Cooloatoxin Coolia monotis (Holmes et al., 1995)
Dinophysistoxin derivatives Prorocentrum maculosm (Hoppenrath et al., 2013; Lee et al., 2016)
Goniodomins Alexandrium hiranoi (Anderson et al., 2012a)
Goniodomins Alexandrium monilatum (Anderson et al., 2012a)
Gymnocin-A, -B Karenia mikimotoi (Brand et al., 2012)
Gymnodimine Alexandrium ostenfeldii (Molgo et al., 2017)
Gymnodimine Alexandrium peruvianum (Molgo et al., 2017)
Gymnodimine Karenia selliformis (Harju et al., 2016; Brand et al., 2012)
Karlotoxins Karlodinium veneficum (Place et al., 2012)
Maitotoxins Gambierdiscus excentricus (Fraga et al., 2011)
Neurotoxins Prorocentrum borbinucum (Hoppenrath et al., 2013; Lee et al., 2016; Glibert et al., 2012)
Okadaic Acid Prorcentrum hoffmannianum (Hoppenrath et al., 2013; Lee et al., 2016; Glibert et al., 2012)
Okadaic Acid Prorocentrum belizeanum (Hoppenrath et al., 2013; Lee et al., 2016; Glibert et al., 2012)
Okadaic Acid Prorocentrum caipirignum (Luo et al., 2017)
Okadaic Acid Prorocentrum cf. maculosum (Luo et al., 2017; Glibert et al., 2012)
Okadaic Acid Prorocentrum concavum (Hoppenrath et al., 2013; Glibert et al., 2012; Lee et al., 2016)
Okadaic Acid Prorocentrum faustiae (Hoppenrath et al., 2013; Lee et al., 2016; Glibert et al., 2012)
Okadaic Acid Prorocentrum lima (Hoppenrath et al., 2013; Lee et al., 2016; Luo et al., 2017; Glibert et al.,
2012)
Okadaic Acid Prorocentrum rhathymum (Hoppenrath et al., 2013; Lee et al., 2016; Glibert et al., 2012)
Okadaic Acid/Dinophysistoxins Phalacroma mitra (Reguera et al., 2014, 2012)
Okadaic Acid/Dinophysistoxins Phalacroma rotundatum (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis acuminata (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis acuta (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis caudata (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis fortii (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis miles (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis norvegica (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis ovum (Reguera et al., 2012; Rugeura et al., 2014)
Okadaic Acid/Dinophysistoxins Dinophysis sacculus (Reguera et al., 2012; Rugeura et al., 2014)
Ostreotoxin 3 Ostreopsis lenticularis (Parsons et al., 2012)
Palytoxin/ostreopsis toxins Osteropsis siamensis (Parsons et al., 2012)
Palytoxin/ostreopsis toxins Ostreopsis mascarenensis (Parsons et al., 2012)
Palytoxin/ostreopsis toxins Ostreopsis cf. ovata (Parsons et al., 2012)
Pinnatoxins Vulcanodinium rugosum (Molgo et al., 2017)
Pinnatoxins Vulcanodinium rugosum (Molgo et al., 2017)
Prorocentrolide Prorocentrum caipirignum (Nascimento et al., 2017)
Prorocentrolides A,B; Spiro-prorocentrimine Prorocentrum lima (Molgo et al., 2017)
Prorocentrolides A,B; Spiro-prorocentrimine Prorocentrum maculosum (Molgo et al., 2017)
Saxitoxin, spirolides, gymnodimines Alexandrium ostenfeldii (Kremp et al., 2014; Tillmann et al., 2014)
Saxitoxins Alexandrium acatenella (Anderson et al., 2012a)
Saxitoxins Alexandrium affine (Anderson et al., 2012a)
Saxitoxins Alexandrium andersonii Balech (Anderson et al., 2012a)
Saxitoxins Alexandrium angustitabulatum (Anderson et al., 2012a)
Saxitoxins Alexandrium catenella (Anderson et al., 2012a)
Saxitoxins Alexandrium cf. catenella (Tillmann and John, 2002)
Saxitoxins Alexandrium cohorticula (Anderson et al., 2012a)
Saxitoxins Alexandrium fundyense (Anderson et al., 2012a)
Saxitoxins Alexandrium minutum (Anderson et al., 2012a)
Saxitoxins Alexandrium tamarense (Andeson et al., 2012a)
Saxitoxins Alexandrium tamiyavanichii (Anderson et al., 2012a)
Saxitoxins Alexandrium taylori (Anderson et al., 2012a)
Saxitoxins Gonyaulax digitale (Swift et al., 1995)
Saxitoxins Gonyaulax excavata (White, 1986)
Saxitoxins Gymnodinium catenatum (Hallegraeff et al., 2012)
Saxitoxins Pyrodinium bahamense (Lewitus et al., 2012)
Spirolides Alexandrium ostenfeldii (Molgo et al., 2017)
Spirolides, saxitoxins Alexandrium peruvianum (Molgo et al., 2017; Anderson et al., 2012a; Tomas et al., 2012)
Yessotoxin Gonyaulax spinifera (Paz et al., 2008)
Yessotoxin Gonyaulax taylorii (Alvarez et al., 2016)
Yessotoxin Lingulodinium polyedra (Paz et al., 2008)
Yessotoxin Protoceratium reticulatum (syn.:)Gonyaulax grindleyi (Paz et al., 2008; Satake et al., 1997)

8
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Table 4 was also sequenced in the heterotrophic dinoflagellate Noctiluca scin-


Saxitoxin and major derivatives, which collectively are referred to as Paralytic tillans, a bloom-forming species that is phylogenetically very distant
Shellfish Toxins (PSTs). from the seven photosynthetic species. This is reflected in its LCF gene,
which codes for a polypeptide containing two distinct domains: that of
the LCF found in photosynthetic species, and the LBP, which occurs as a
separate gene in the photosynthetic species (Liu and Hastings, 2007).

3.3. Biochemistry: saxitoxin

Saxitoxin is unique in that it is produced by organisms encom-


passing two kingdoms of life inhabiting different aquatic systems: eu-
karyotic dinoflagellates in marine systems and cyanobacteria in fresh-
water systems (Cusick and Sayler, 2013). The route of biosynthesis is
similar between the two (Shimizu, 1993), and was initially proposed
based on extensive studies using labeled precursors with the dino-
flagellate Alexandrium tamarense and the cyanobacterium Aphanizo-
menan flos-aquae (Shimizu, 1993, 1996). However, more recent genetic
and bioinformatic analyses (Kellmann et al., 2008b; Mihali et al., 2011,
Division Namea R1 R2 R3 R4
STX H H H OCONH2 2009), coupled with screening of the biosynthetic intermediates and in-
NeoSTX OH H H OCONH2 vitro biosynthesis of saxitoxin (Kellmann and Neilan, 2007), has re-
GTX1 OH OSO3− H OCONH2 sulted in modifications of the original pathway. The gene cluster coding
Carbamate GTX2 H OSO3− H OCONH2
for STX biosynthesis, spanning approximately 35 kb and encompassing
GTX3 H H OSO3− OCONH2
GTX4 OH H OSO3− OCONH2
26 proteins, was first identified in the toxic cyanobacterium Cylin-
GTX5 (B1) H H H OCONHSO3− drospermopsis raciborskii T3 (Kellmann et al., 2008b) followed by find-
GTX6 (B2) OH H H OCONHSO3− ings of homologous gene clusters in several other cyanobacteria (Mihali
C1 H OSO3− H OCONHSO3− et al., 2011, 2009). The findings with C. raciborskii T3 revealed that
N-sulfocarbamoyl C2 H H OSO3− OCONHSO3−
saxitoxin biosynthesis is initiated by SxtA, a novel polyketide synthase
C3 OH OSO3− H OCONHSO3−
C4 OH H OSO3− OCONHSO3− (described in detail in the Genetics section below) (Kellmann et al.,
dcSTX H H H OH 2008b). The original model proposed condensation of arginine to
dcNeoSTX OH H H OH acetate as the initial step, with the methyl side chain introduced later.
dcGTX1 OH OSO3− H OH
However, in the revised reaction scheme (Fig. 3), the acyl carrier pro-
Decarbamoyl dcGTX2 H OSO3− H OH
dcGTX3 H H OSO3− OH
tein (ACP) is loaded with acetate from acetyl Co-A, followed by me-
dcGTX4 OH H OSO3− OH thylation of acetyl-ACP by SxtA3. SxtA4 then performs the Claisen
doSTX H H H H condensation (steps 1–2 in Fig. 3). The presence of the putative inter-
Deoxydecarbamoyl doGTX2 H H OSO3− H mediate, designated A’, was confirmed via LC-MS-MA, while the ori-
doGTX3 H OSO3− H H
ginally-proposed intermediate was not detected. The third step in the
a
Abbreviations: STX, saxitoxin, GTX, gonyautoxin. pathway, the amidino transfer from arginine to Ɐ-amino A’ group (B’)
occurs via SxtG. This is followed by the first cyclicization via retroaldol
Differences exist among dinoflagellate species with regards to cleavage of ammonia from cytidine via SxtB (step 4). Hydroxylation
physiological characteristics such as the circadian rhythm and bio- and bicyclicization occur via SxtD, SxtS, and SxtU (steps 5–8). The final
chemistry of bioluminescence. In L. polyedra, one of the best-studied reactions (steps 9–10) involve carbamoyl transfer and dihydroxylation,
dinoflagellate species in terms of both toxicity and bioluminescence, though the exact order remains unknown (Kellmann et al., 2008b).
both LCF and LBP are broken down at the end of the night phase and Intracellular STX quotas are influenced by environmental and nu-
then synthesized again in the next cycle (Wilson and Hastings, 1998). tritional factors. Phosphorus limitation results in an increase in cellular
The scintillons themselves are also broken down and reformed each day levels, while nitrogen limitation results in a decrease (Van de Waal
(Fritz et al., 1990). Additionally, LBP has been shown to be a necessary et al., 2014). Cellular toxin quotas and composition also fluctuate based
component in the biochemical reaction. However, in other species, such on growth phase; in general, higher levels and more diverse profiles
as the non-toxic Pyrocystis species, the daily levels of LCF remain con- have been recorded in mid-exponential and stationary phase (Lim and
stant, there is no LBP (Knaust et al., 1998) and the scintillons do not Ogata, 2005; Parkhill and Cembella, 1999). Salinity has also been
break down, but they do exhibit rhythmic changes in excitability and shown to influence both toxin profile and overall toxin content, though
cellular distribution (Widder and Case, 1982). results vary among species; for example, an inverse correlation was
shown to exist between salinity and toxin content in A. minutum, while
in A. tamiyavanichii, toxin content was greatest at salinities optimal for
3.2. Genetics: bioluminescence growth (Lim and Ogata, 2005). Toxin analyses with batch cultures of a
Malaysian isolate of P. bahamense showed both toxin content and pro-
The full-length lcf gene has been sequenced in seven photosynthetic file were influenced by factors (temperature, salinity, light intensity)
species of dinoflagellates: L. polyedra, P. lunula, P. fusiformis, P. nocti- that alter growth rate, and that toxin content was likely to be de-
luca, A. affine, A. tamarense, and P. reticulatum (Liu et al., 2004). The termined by the manner in which the cells allocate their nitrogen and
predicted proteins all possess the same unique organization: a single carbon to various cellular components and processes under various
polypeptide consisting of an N-terminal region of unknown function growth conditions (Usup et al., 1994). Overall, toxin content and profile
followed by three homologous domains (“D1,” “D2,” “D3”), each with is influenced by growth conditions (Anderson et al., 1990a, 1990b;
catalytic activity (Liu et al., 2004; Okamoto et al., 2001). The individual Lim et al., 2006; Lim and Ogata, 2005), nutrient limitation (Boyer et al.,
domains among species (i.e. D1 A. tamarense, D1 A. affine) are more 1987; John and Flynn, 2000), and intracellular arginine concentration
similar than among the three different domains of the same species (i.e., (Anderson et al., 1990b; John and Flynn, 2000).
D1, D2, D3 of A. tamarense) (Liu et al., 2004) (Fig. 2). The full-length lcf

9
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 1. Representative distribution of toxic (box) and bioluminescent (blue star) species across the dinoflagellate tree of life. Transparent blue flashes indicate
bioluminescence in a closely related species not included in tree. Orders are designated as vertical lines. Modified from Orr et al. (2012). (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

10
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

3.5. Biochemistry: yessotoxin

Yessotoxins are polyketides. Great structural diversity exists among


the dinoflagellate polyketides. This diversity can be achieved through
different mechanisms, including alternative extender units (acetate,
malonate, propionate, etc.); modification or omission of different re-
actions within the overall biosynthetic pathway; and the integration of
enzymes from other metabolic pathways (Kellmann et al., 2010).
Polyketides are synthesized by polyketide synthases (PKS). PKS en-
zymes are multifunctional complexes that act on acyl monomers; the
minimal set of catalytic domains required for function includes ketoa-
cylsynthase, acyl transferase, and acyl carrier protein. The presence of
three additional domains accounts for the variation in polyketide
structure: ketoacylreductases, dehydrases, and enolreductases
(Gokhale and Dipika, 2001). Polyether ladders are specific to dino-
flagellates, and their close structural similarities indicate shared bio-
synthetic mechanisms (Kellmann et al., 2010). Most biosynthesis stu-
dies have been conducted with brevetoxin (BTX); however, the single
Fig. 2. Maximum likelihood tree of the catalytic domains of the dinoflagellate
study done with YTX demonstrated that the incorporation of acetate
luciferase. The corresponding domains of different LCFs group together more
closely than the individual domains of each species. At = A. tamarense, Aa = A. units, which form the backbone of the structure, was similar to that of
affine, Pr = P. reticulatum, Pf = P. fusiformis, Pn = P. noctiluca, Pl = P. lunula, BTX. The cyclic trans-polyethers were suggested to be formed via a
Lp = L. polyedra. D= Domain. Modified from Liu et al. (2004). polyepoxide cascade mechanism (Fig. 4). The little information avail-
able on YTX biosynthesis has shown that a portion of the ring fragment
can be obtained with a suitable polyepoxide system (Jamison and
3.4. Genetics: saxitoxin
Vilotijevic, 2007), supporting previous findings that suggested YTX and
other dinoflagellate polyether toxins may all be synthesized in a similar
Saxitoxin biosynthesis genes have been identified in all three di-
fashion from a polyepoxide (Nakanishi, 1985).
noflagellate genera (Alexandrium, Pyrodinium, Gymnodinium) capable of
toxin production (Hackett et al., 2013; Stuken et al., 2011), albeit with
3.6. Genetics: yessotoxin
varying degrees of coverage. The core genes of cyanobacterial STX
biosynthesis have been confirmed in dinoflagellates, along with genes
Recent sequencing efforts have identified genes coding for PKS-re-
related to toxin transport and modification (Hackett et al., 2013).
lated enzymes in multiple species of dinoflagellates that produce
Overall, 14 “core” genes have been identified (with the designation as
polyether-based toxins ((Beedessee et al., 2015; Kimura et al., 2015;
“core” based on characterization of STX biosynthesis in cyanobacteria);
Kohli et al., 2017, 2015; Meyer et al., 2015; Pawlowiez et al., 2014).
along with approximately 10 genes involved in transport and/or ana-
Even with these advances, no sequence data currently exist as to the
logue modification and two potential regulators involved in signal
genes and/or proteins involved in YTX biosynthesis.
transduction (as summarized in (Akbar et al., 2020)). Of the 14 core
biosynthesis genes, nine have been putatively identified in STX-pro-
3.7. Correlating bioluminescence and toxicity
ducing dinoflagellate species, though with the exceptions of sxtA and
sxtG, their roles in dinoflagellate STX biosynthesis remain to be de-
The potential molecular and/or biochemical links between toxin
termined (Akbar et al., 2020). SxtA is the first gene in the pathway
production and bioluminescence remain unexplored and so are con-
(Stuken et al., 2011), coding for a novel polyketide synthase comprised
sidered here. As the genes and pathways involved in YTX biosynthesis
of four catalytic domains (SxtA1-SxtA4) (Kellmann et al., 2008a). In
are not known, the link is explored with STX and bioluminescence. One
dinoflagellates, SxtA occurs as two isoforms: SxtA1-4, and SxtA1-3
of the first steps of STX biosynthesis is the claisen condensation of ar-
(Stuken et al., 2011). The gene coding for the SxtA1 domain is wide-
ginine and acetate, performed by a previously unrecognized type of
spread, occurring in both toxic and non-toxic strains (Hackett et al.,
polyketide synthase (Kellmann et al., 2008b). Claisen condensations are
2013; Murray et al., 2015). SxtA4 appears to be specific for toxin
rare on amino acids, but also do occur in the synthesis of tetrapyrroles
synthesis, as screening of over 40 isolates (35 Alexandrium, four Gym-
(Kellmann et al., 2008b). The substrate in the bioluminescence reaction
nodinium, and a single Pyrodinium) demonstrated its presence and high
of dinoflagellates is luciferin, which is an open-chain tetrapyrrole si-
sequence conservation in toxic strains; conversely, it was not found in
milar in structure to chlorophyll. While the synthesis pathway for di-
67 strains of 54 species of non-STX producing species (Murray et al.,
noflagellate luciferin is not known, the structural similarities between
2014, 2011b; Orr et al., 2013a; Stuken et al., 2011). SxtG codes for an
dinoflagellate luciferin and chlorophyll indicate luciferin formation via
amidinotransferase that putatively encodes the second step in STX
chlorophyll catabolism: P. lunula was shown to incorporate radio-
biosynthesis. While sxtG has been detected in Alexandrium species not
actively labeled chlorophyll precursors into chlorophyll and luciferin,
known to synthesize STX, it has not been found in numerous non-STX-
suggesting biosynthesis of the two are linked (Topalov and Kishi, 2001;
producing dinoflagellate species other than those of the genus Alexan-
Wu et al., 2003). Nonphotosynthetic dinoflagellates contain the com-
drium or G. catenatum (Orr et al., 2013a). Overall, transcriptome ana-
ponents of the plastid tetrapyrrole pathway, which accounts for the
lyses from multiple toxin-producing dinoflagellate spp. suggest that the
mechanism of luciferin biosynthesis in nonphotosynthetic biolumines-
genes involved in the first three steps of saxitoxin biosynthesis (sxtA,
cent dinoflagellates (Janouškovec et al., 2017). In any case, both STX
sxtG, sxtB) in dinoflagellates and cyanobacteria are derived from
synthesis and what may be considered the start of the bioluminescence
common ancestral proteins not involved in STX synthesis, as they are
reaction – that of luciferin formation – require an enzyme with a do-
also found in organisms that do not produce the toxin (Hackett et al.,
main that has a somewhat unique catalytic activity. It is also worth
2013).
noting that the sxtA gene encoding polyketide synthase contains an
additional domain specific to the toxin-producing strains that occurred
likely as a gene duplication event (Murray et al., 2015). Could there be
a shared enzyme or associated molecular/genetic element between STX

11
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 3. Saxitoxin biosynthetic pathway with putative gene functions. The original pathway, derived from radioisotope tracing experiments, has been revised based on
collective genetic, biochemical, and bioinformatic analyses. Core and accessory genes are as initially identified in cyanobacteria; thus far, sxtA and sxtG have been
characterized in dinoflagellates. Modified from Cusick and Sayler (2013) and Kellmann et al. (2008b).

12
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 4. Polyepoxide reaction proposed to occur during YTX synthesis. Image redrawn from Kellmann et al. (2010).

biosynthesis and bioluminescence? Intriguingly, a proteomics analysis bioluminescence potential, while release of eggs dropped to zero
between a toxic A. catenella strain and a toxicity-lost lab mutant showed (Buskey and Stearns, 1991). In dinoflagellates, it was found that when
that multiple proteins similar to the LBP were down-regulated in the the heterotrophic dinoflagellate Protoperidinium cf. divergens was de-
non-toxic mutant (Wang et al., 2012). More recently, proteomic ana- prived of food it still exhibited high bioluminescence potential even as
lysis of a toxic Alexandrium strain revealed bioluminescence-associated growth rate dropped to zero (Latz and Jeong, 1996). This points to a
proteins were upregulated during toxin production (Zhang et al., 2018). strong selective advantage for light emission.
The large and complex nature of dinoflagellate genomes presents a
formidable challenge for the identification of potential shared mole-
cular elements in toxic, bioluminescent dinoflagellates, but is a link that 4.2. Toxicity
requires further analysis in light of the contributions by both biolumi-
nescence and toxin production on HAB dynamics. The costs of toxin production are influenced by multiple factors,
including the concentration of grazers and their cues (Selander et al.,
2006) and resource (nitrogen, light) availability. Multiple studies trying
4. Energetic costs to quantify STX costs produced similar results, in that cells with in-
creased toxin production grew as fast as uninduced cells, and toxic and
4.1. Bioluminescence non-toxic strains of the same species grow at the same rate (Pančić and
Kiørboe, 2018). As for the role of nutrients in toxin production, when
Bioluminescent reactions are energetically costly, requiring the nitrogen or light is limiting, STX production is reduced, while nutrient-
equivalent of as many as 60 ATP molecules per photon replete or phosphorus-limited cells produce toxins at a high rate
(Hastings, 1978). The extreme energy demands of light production are (Anderson et al., 1990b; John and Flynn, 2000). Toxin levels have also
evident in mixed cultures of luminescent and dark mutants of the same been shown to increase with direct or indirect exposure to grazers.
species of bacteria (Vibrio harveyi) where the dark mutants rapidly Overall, there exists much variability as to the responses of the grazers
overrun the culture. By contrast, the bioluminescent strain dominates if to the toxic dinoflagellates, with variability stemming from both the
the mixture is irradiated with ultraviolet light of sufficient energy to particular predator-prey combination as well as the strain involved
damage the microbial DNA. The selective advantage of the biolumi- (Pančić and Kiørboe, 2018). The influence of predator and nutrient are
nescent strain in this case is apparently that it initiates DNA repair via not mutually exclusive: waterborne cues from Acartia sp. were shown to
the enzyme photolyase, which is activated by blue light emitted by the result in increased STX production in A. minutum in N-replete but not
bioluminescent strain (Czyż et al., 2003; Czyz et al., 2000). It appears low-N environments. Thus, the cost of defense may be neglible under
that a strong selective advantage is necessary to overcome the high- nutrient-rich conditions, but could be a factor in environments where
energy demands associated with light production. nitrogen is limited.
Even when food deprived, organisms that employ bioluminescence Extrapolating lab findings to environmental settings, the potential
as the first line of defense will continue to invest energy in light pro- exists for toxic species to gain an advantage under high nitrogen and
duction at the expense of other functions. For example, starvation of high grazer biomass - a scenario typical of coastal environs – as growth
adult female copepods (Metridia longa) had no significant effect on rates can double due to the benefit of toxin production at a minimal

13
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

production cost. The environmental conditions that promote cell toxi- concentrations used in these studies were one to two orders of magni-
city, i.e., high grazer biomass, coincide with the time of the year when tude less than those used in experiments like those of Esaias and
nitrogen levels are low; thus ecological modeling predicts that the cost Curl (1972) that used the much smaller (~20–50 µm) and dimmer cells
of toxin production (quantified as a reduction in cell division rate of (~108 photons s−1) associated with blooms. Cusick and Widder (2014)
defended relative to undefended cells) is substantial, leading to >20% demonstrated that the foraging efficiency of a nocturnal fish predator
decrease in cell division rate. However, the investment in defense pays was enhanced in the presence of the bright emitter P. noctiluca at low
off since defended cells experience lower grazing mortality, and the net cell concentrations (5–40 cells ml−1) but not at similar low cell con-
growth rates can be up to twice or more of that of undefended cells centrations of the dim emitter L. polyedra. Only at cell concentrations
(Chakraborty et al., 2019). greater than 500 cells ml−1 did the bioluminescence of L. polyedra
function as a burglar alarm. However, the grass shrimp Palaemonetes
5. Ecological functions pugio rather than copepods served as the primary predator in these
experiments, and chemical cues (such as the copepodamides produced
5.1. Bioluminescence by copepods) were not measured, nor was the flash intensity of single
cells at different cell concentrations. Overall, the results from that study
The function of the flash has been long debated. While early in- suggest the function of the flash is cell-concentration dependent, and
vestigators could imagine no selective advantage for bioluminescence that in “dim-emitting” bloom-forming bioluminescent dinoflagellates,
in dinoflagellates (Biggley et al., 1969), the energy requirements asso- high cell concentrations are required for a predator to be able to trace
ciated with light production and the existence of complex cellular the path of a copepod through the water (Cusick and Widder, 2014).
machinery, which insures that light production occurs at night and is Thus, it is plausible that flash function is dependent on both cell
inhibited during the day, suggests otherwise. Early experiments showed number and the main dinoflagellate predators within that specific
that bioluminescence reduces grazing by zooplankton predators ecosystem. While earlier experimental studies provide clear support for
(Buskey and Swift, 1983; Esaias and Curl, 1972) and that nocturnally the selective advantage of flash avoidance by primary predators, recent
grazing copepods consumed fewer dinoflagellates when these were work with the copepodamide-induced bioluminescence of L. polyedra
offered in their bioluminescent night phase compared to their non- demonstrated the advantage to the dinoflagellate as well, in that the
bioluminescent day phase (Esaias and Curl, 1972; White, 1979). More cells are seemingly rejected unharmed by the copepod predator.
recently, bioluminescence has been shown to increase upon exposure to Since many bloom-forming species are also toxin producers an al-
the chemical cues of predators (Lindstrom et al., 2017a). The nature of ternative hypothesis is that the bioluminescent flash in these species
the flash is such that the light is emitted both as brief flashes (~100 ms) functions as an aposematic warning, signaling unpalatability to grazers.
and as a low intensity glow (Wilson and Hastings, 1998) depending Bioluminescence has been shown to function as an aposematic warning
upon both species and stimulus. Additionally, there exists within single in firefly larvae (de Cock and Matthysen, 1999, 2002; Underwood et al.,
dinoflagellate species both bioluminescent and non-bioluminescent 1997), millipedes (Marek et al., 2011), and brittle-stars (Grober, 1988a,
strains (Marcinko et al., 2013). Collectively, these findings are evidence 1988b). In a dark environment such as the ocean at night, the light
for the adaptive value of light emission. substitutes for aposematic coloration, with the flash warning potential
Three major hypotheses have been proposed regarding the role of predators of toxicity or some other unpalatable trait. It is worth noting
bioluminescence in dinoflagellate ecology. All three are rooted in the that most copepod feeding occurs during the night, coinciding with the
premise that bioluminescence acts as an anti-grazing strategy bioluminescent phase. The basis for deterrence is that predators learn to
(Lindstrom et al., 2017a). The difference among the three hypotheses associate the flash with the undesirable trait, resulting in decreased
lies in the function of the flash. The flash may act as (A) a startle re- predation on the bioluminescent prey (Schantz, 1971).
sponse display, (B) a burglar alarm, or (C) an aposematic warning, in Support for the aposematic function of the flash emerged from a
which the flash serves as a warning of toxicity or unpalatability. As a series of experiments that looked at copepod grazing on toxic and non-
startle response display, the flash is hypothesized to stimulate an innate toxic bioluminescent dinoflagellates in the presence of non-luminescent
photophobic response in the grazer (Esaias and Curl, 1972). However, diatoms (Hanley and Widder, 2017). Although the existence of a non-
while flash-induced burst swimming occurs in marine copepods toxic bioluminescent dinoflagellate might seem to undermine the
(Buskey et al., 1983; Buskey and Swift, 1983, 1985) it is absent in co- aposematic hypothesis, it is important to recognize the likely existence
pepods from freshwater environments, where planktonic biolumines- of Batesian mimics, palatable prey species that mimic the flash of un-
cence is not found (Buskey et al., 1987). Also, without any direct benefit palatable species, thereby benefitting from the learned avoidance be-
to the marine copepods of avoiding the flash, there is no selective ad- havior of the predator (Lindström et al., 1997). A Batesian mimic does
vantage to retaining the photophobic behavior. For the dinoflagellate not long remain aversive to predators when in the absence of the or-
flash and the copepod burst swimming response to be retained, a se- ganisms whose threat it mimics (Waldbauer and Sternbury, 1987).
lective advantage must accrue to both. Evidence that this was in fact the case for the non-toxic bioluminescent
According to the burglar alarm hypothesis the adaptive value of dinoflagellate (L. polyedra), came from the fact that copepods consumed
copepod burst swimming is to evade higher order visual predators this dinoflagellate in its non-luminescent phase, but not during its lu-
alerted to the copepod's presence by dinoflagellate flashing minescent phase. By contrast the toxic dinoflagellate (P. bahamense)
(Burkenroad, 1943). Laboratory studies have demonstrated an in- was not consumed in either its bioluminescent or non-bioluminescent
creased susceptibility of zooplankton to visual predators such as fish phase. It was also shown that the feeding efficiency on the alternative
and cephalopods in the presence of bioluminescent dinoflagellates, prey (diatoms) was lower in the presence of the non-bioluminescent
thereby supporting the adaptive value of flash avoidance by the pri- phase of toxic dinoflagellates than in the presence of the non-biolumi-
mary predator (Abrahams and Townsend, 1993; Fleisher and nescent phase of non-toxic dinoflagellates, presumably because of the
Case, 1995; Mensinger and Case, 1992). increased time needed to handle and reject the toxic dinoflagellates.
For bioluminescence to function as a burglar alarm, sufficient light In all likelihood, toxic dinoflagellates employ multiple deterrents
must be emitted by the dinoflagellates to alert visual predators. It is thereby improving effectiveness in warding off attacks; however, when
therefore noteworthy that the dinoflagellate used in the experiments by toxic dinoflagellates can be rejected based on the flash alone the
Mensinger and Case (1992) and Fleisher and Case (1995) was P. fusi- grazing rate on alternative prey is increased. Hanley and Widder (2017)
formis, which is one of the largest (~1 mm long) and brightest (peak also corroborated the previously demonstrated concentration depen-
photon flux >1011 photons s − 1) of all known bioluminescent dino- dence of the burglar alarm effect for dim emitters (Cusick and
flagellates. P. fusiformis is not a bloom-forming species and the cell Widder, 2014), because while the co-occurring diatom was consumed

14
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

at low cell concentrations of the bioluminescent dinoflagellates, at high pheromones, with roles in mating and cyst formation (Wyatt and
(bloom-level) concentrations virtually all grazing ceased, apparently Jenkinson, 1997). They possess similar characteristics as established
because of the escape responses of the grazers. pheromones, including (1) profiles akin to those in terrestrial animal
species, in which a mix of compounds rather than a single congener is
5.2. How bioluminescence can drive HAB development produced and (2) low levels of secretion (10−9–10−10 M). Most of the
toxin is retained within the cells during exponential growth, and re-
A recent study documenting avoidance or rejection of biolumines- leased during senescence, which, in natural populations, would coin-
cent cells by predators provides strong support for the role of biolu- cide with the induction of sexuality during bloom decline
minescence in HAB initiation and persistence. In this study, high-speed, (Cembella, 2003). This hypothesis explains the diversity in toxin pro-
low-light video recordings of individual copepod behavior showed L. files among strains and geographical locations; what is not accounted
polyedra cells flashed on contact with the copepod and was then re- for, however, is the mechanism by which mating is mediated in non-
jected unharmed (Prevett et al., 2019). In this study L. polyedra was toxic strains of the same species (Cembella, 2003). Immunofluorescent
induced to brighter bioluminescence by the addition of copepodamides. labeling studies with toxic A. fundyense demonstrated chromosomal
When the dinoflagellates were not induced to the brighter biolumi- localization of the toxin (Anderson and Cheng, 1988; Doucette and
nescence, then in mixed populations of phytoplankton where L. poly- Anderson, 1993); this close proximity to the chromosomes led to the
edra was only dimly luminescent and contributed 25% of the available hypothesis that it may play a role in chromosome structural organiza-
prey biovolume, the grazer consumed it in proportion to its fraction in tion, with the two positively-charged guanidinium groups on the mo-
the community (24%). However, with copepodamide induction of lecule binding in a manner similar to that of polyamines. The large
brighter bioluminescence the proportion of L. polyedra in the grazer's number of nitrogen atoms contained within the structure has led some
diet dropped to 2%. An important point in this regard is the observation to suggest it functions as a form of nitrogen storage, though a coun-
that after the copepod rejected the flashing cell it exhibited “forceful terargument is that nitrogen storage could be achieved in a more
beating with its swimming appendages” which is a behavior associated bioenergetically efficient manner via lower molecular weight com-
with the burglar alarm response. While this study did not determine the pounds such as urea or amino acids (Cembella, 1998).
mechanism (aposematism, startle response, or burglar alarm) by which Two additional theories as to the ecological function of STX in di-
the flash may function in the initial stages of a bloom, it demonstrated noflagellates are derived from findings demonstrating its role in copper
the means by which bioluminescence can facilitate bloom development. homeostasis. A combination of genomic, transcriptomic, physiological
Given that the doubling rate of dinoflagellates is one third that of and bioinformatic analyses showed that STX inhibited copper uptake
diatoms (Banse, 1982), selective feeding by grazers on non-luminescent across multiple microbial species, including yeast and photosynthetic
prey would favor bioluminescent dinoflagellates, thereby contributing algae, while not affecting their growth (Cusick et al., 2009, 2012,
to bloom initiation. These “windows of opportunity” afforded by 2013). These studies identified the copper transporter (CTR), a hybrid
openings or relief from grazing pressure are postulated to play an im- ion channel/transporter protein, as an additional target of STX. Copper
portant role in dinoflagellate bloom formation (Stoecker et al., 2008). is an essential trace metal required for key metabolic and physiological
However, the function of the flash among bioluminescent species may processes in most organisms, including marine phytoplankton. Thus,
differ based on the dinoflagellate predators within that ecosystem. Once one theory is that it targets membrane proteins (via their “selectivity
flash intensities and/or flash number reaches a threshold value, the filter”) involved in trace metal assimilation in co-occurring phyto-
burglar alarm response would remove grazers from the vicinity of the plankton species. Copper is also toxic to living cells through a variety of
bloom thereby contributing to its persistence. In the situation described mechanisms (Rademacher and Masepohl, 2012). It has become pre-
above, copepodamide-induced bioluminescence protects L. polyedra valent in marine systems due to its use as an algaecide/biocide and as
from copepod grazers, who seek alternate prey in the presence of bio- an anti-fouling (AF) agent in marine vessel coatings (Borkow and
luminescent cells. This allows L. polyedra to successfully compete with Gabbay, 2005; Yebra et al., 2004). Studies with non-toxic Alexandrium
faster-growing phytoplankton competitors. Monitoring of natural po- spp. demonstrated that exposure to copper compromised membrane
pulations of L. polyedra over a ten-year period showed a positive cor- integrity, resulting in leakage and cell death (Li et al., 2008). Thus, the
relation with copepod biomass (Prevett et al., 2019), suggesting that second theory is that it alleviates copper stress in toxin-producers by
bioluminescence serves as a defense that provides an advantage to the inhibiting copper entry. However, both of these theories remain to be
slower-growing dinoflagellate species, allowing them to successfully experimentally verified.
compete with faster-growing phytoplankton competitors and poten- The predominant hypothesis is that the molecule functions as a
tially achieve bloom numbers. grazing deterrent. Similar to bioluminescence, toxin production may
aid in HAB initiation as it alleviates grazing pressure on the HAB spe-
5.3. Toxicity: saxitoxin cies. Numerous studies have been conducted with organisms from
multiple trophic levels – including micro- and mesozooplankton, fish,
The ecological role of both STX and YTX in the species that produce and macroinvertebrates - and results vary as to the effects on health and
them remains unknown. Numerous studies have been done to try and whether the organisms actively reject the toxic cells (Barreiro et al.,
ascertain the function of STX. As yet no studies exist as to the potential 2006; da Costa et al., 2005; Frangoulos et al., 2000; Oikawa et al., 2005;
function of YTX, and so STX is discussed here. The well-established Robineau et al., 1991; Teegarden, 1999; Zimmer and Ferrer, 2007). One
target of STX is the voltage-gated sodium channel in humans and other of the most well-studied interactions is that of copepod grazing on di-
mammals, where it binds with high affinity and can result in death due noflagellates. However, as with other organisms, grazing studies with
to respiratory paralysis (Catterall, 1985). Hence, due to its detrimental copepods have produced contradictory results: experiments with toxic
impacts on human health, STX and its congeners are typically referred Alexandrium spp, range from no adverse effects upon consumption of
to as potent neurotoxins. However, it is highly unlikely that toxin the toxic cells to increased mortality of the grazer (Bagøien et al., 1996;
biosynthesis evolved with the purpose of blocking mammalian sodium Teegarden and Cembella, 1996). These results indicate that both the
channels. The sxt genes are estimated to have emerged at least 2.1 strain and grazer must be considered when examining the role of toxin
billion years ago (Murray et al., 2011a), pre-dating the evolution of function. Additionally, natural populations of Alexandrium are com-
voltage-gated sodium channels. prised of a mix of strains, differing in both toxin content and profiles
Multiple hypotheses have been put forward regarding the eco-evo- (Alpermann et al., 2010). The copepod species and its co-occurrence
lutionary role of STX and its congeners. Studies on Alexandrium popu- with toxic dinoflagellates also influences the effects of the toxin, as
lations dynamics led to the suggestion that they function as some copepods have been demonstrated to detoxify STX

15
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

(Teegarden et al., 2003). Additionally, copepod populations previously typically utilize a strain possessing that phenotype. In general, a strain
exposed to toxic blooms are less affected than naïve populations is obtained by establishing clonal cultures, each of which are derived
(Colin and Dam, 2004). These results indicate that toxin induction and from a single-cell isolate from an environmental water sample
function are a combined result of both the strain and grazer. Collec- (Menden-Deuer and Montalbano, 2015). Thus, each isolate is an in-
tively, these studies point towards a predator-prey co-evolution, and in dependent, clonal strain of the species under investigation, and in
interpreting the potential role of the toxin, as with bioluminescence, theory the genetic composition of that cell dictates the genetic make-up
one must remain cognizant of the ecosystem view: strain of dino- of the cells used in subsequent lab-based studies.
flagellate, type of grazer, and environmental parameters. The advancement of molecular tools in dinoflagellate ecology has
While a large number of lab studies have examined the ecophysio- resulted in population studies that show blooms are not monoclonal
logical requirements of toxin production, these results are not sufficient events. The use of genetic markers to link genetic diversity with trait
to predict the succession of HABs in the environment, as the net growth diversity is emerging as a valuable mechanism in phytoplankton
performance of a species is affected by complex interactions with other ecology. Microsatellite markers, long used to assess population genetics
organisms (Zingone and Oksfeldt Enevoldsen, 2000). Indeed, the pre- of macroorganisms (Jarne and Lagoda, 1996), are now being used as a
sence of copepods and their waterborne cues have also been shown to tool to examine the ecology and evolution of microorganisms
stimulate toxin production (Selander et al., 2006; Wohlrab et al., 2010; (Pettay and LaJeunesse, 2013). Microsatellite markers have shown ge-
Yang et al., 2011) (as discussed in detail below). Thus, it is also plau- netic diversity in field populations of phytoplankton, and this clonal
sible that the toxin may serve multiple ecological functions. When diversity has been shown to be associated with differences in genome
taking into account the metabolic demands associated with toxin size (Whittaker et al., 2012), morphology (Saravanan and
synthesis and modification, as well as the fluctuations in both in- Godhe, 2010), and growth rate (Rynearson and Virginia
tracellular content and profile based on environmental and nutritional Armbrust, 2004). Genetic diversity and bloom dynamics of A. fundyense
influences, it is logical to propose that natural selection would favor blooms assessed using rapidly evolving microsatellite markers revealed
multi-use compounds (Wink, 2003). a high level of genetic diversity in a geographically restricted area
(Richlen et al., 2012). The combined phenotypic and genetic diversity
5.4. Toxicity: yessotoxin occurring within the same HAB population is further exemplified in
studies with the HAB dinoflagellate species Akashiwo sanguinea. Mul-
As with STX, the eco-evolutionary role of YTX remains unknown. As tiple strains of A. sanguinea were obtained from a single water sample in
yet no studies exist as to the potential function of YTX. However, in which a bloom was occurring, and broad intra-specific variability in
vitro studies with various cell models indicate YTX modulates calcium multiple traits was shown to exist among strains all isolated from the
homeostasis, and is able to inhibit entry of Ca+ ions (Paz et al., 2008). same water sample (Menden-Deuer and Montalbano, 2015).
Based on the emerging evidence indicating a role for calcium in dino- In the experiments by Abrahams and Townsend (1993) and
flagellate signal transduction mechanisms (discussed below), one in- Mensinger and Case (1992), the authors proposed that even if the di-
triguing hypothesis as to the eco-evolutionary role of YTX is a role in noflagellate was consumed, its genetic clones would still benefit from
signal transduction, be it within the cells that produce it or a means of the reduction in grazing pressure. However, evidence demonstrating
inhibiting the ability to sense and respond to environmental stimuli in that dinoflagellates are not monoclonal, even in a monospecific bloom
co-occurring species. (Menden-Deuer and Montalbano, 2015; Richlen et al., 2012) under-
mines this hypothesis, as does the fact that bioluminescence occurs in
5.5. Saxitoxin resistance in predators non-bloom-forming species. Without any direct selective advantage to
the individual dinoflagellate among the dim emitters, there is no clear
While one of the most studied hypotheses as to the function of STX evolutionary path to flash production. Based on the recent work de-
is as an anti-predator mechanism, it is also noteworthy that some pre- monstrating the flash emitted by L. polyedra upon predator contact al-
dator species are able to build resistance to the toxin. Natural popula- lows the cell to escape unharmed, it seems logical for individual cells to
tions of organisms consistently exposed to STX-producing cells have harbor and control the LCF genes, as it results in a direct benefit to the
been shown to build resistance, resulting in inter-population variation individual cell.
within a species due to genetic adaptations (Bricelj et al., 2005). Soft- Minimal field data exist on bioluminescence capacity on strains
shell clams (Mya arenaria, a vector for toxin transfer) routinely exposed from the same geographic area and so studies on lab isolates can only be
to “red tides” are more resistant and accumulate the toxin at a greater considered here in relation to genetic diversity and phenotypic differ-
rate than clams from unexposed areas. The genetic basis was found to ences. Within a single dinoflagellate species, there exist bioluminescent
be a single amino acid mutation in the sodium channel, resulting in a and non-bioluminescent strains, such as has been demonstrated for
1000-fold decrease in toxin binding affinity (Bricelj et al., 2005). Gonyaulax excavata (Schmidt et al., 1978) and N. scintillons
Adaption to STX has also been shown to occur in the copepod A. hud- (Valiadi et al., 2019). The genetic and biochemical basis for loss of
sonica, which frequently co-occurs with toxic Alexandrium spp. Inges- bioluminescence in N. scintillons strains has recently been identified.
tion and egg production rates, as well as overall population fitness, These included the presence of the lcf gene but undectable transcripts,
were greater in exposed populations than in naïve populations and potentially deleterious mutations within the gene sequence
(Colin and Dam, 2005, 2002, 2007). However, resistance was not due to (Valiadi et al., 2019). While it remains to be determined whether these
sodium channel mutations as with clams (Chen et al., 2015; same mechanisms are the reason for differences in light and dark strains
Finiguerra et al., 2014, 2015), and the genetic basis for resistance in across species, it is intriguing to consider the influences that initiated
this species remains to be identified. and maintained these mutations. Did bioluminescence no longer confer
an advantage to that population? How frequently does this occur
5.6. The role of genetic diversity among sub-populations?
The question of genetic diversity among sub-populations also ap-
When considering the ecological functions of both bioluminescence plies to toxicity. Twenty years ago, it was feasible to hypothesize that
and toxin production, one feature that is often overlooked is the genetic toxicity functioned via a kin- or group-selection mechanism to decrease
variability for either trait within sub-populations of a natural bloom of grazing predation; in this scenario, toxin production by a cell, while
the species – i.e. do all the cells possess the genes necessary for that decreasing that cell's chance of survival, would increase the survival
trait, or only some of the cells? The theories as to the functions of both rate of other cells in the population. However, in order for this type of
bioluminescence and toxicity have been derived from lab studies, which strategy to continue requires the other cells in the population to be

16
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

capable of toxin production. P. bahamense, L. polyedra) have been isolated from the same geographic
While the core genes in the dinoflagellate STX biosynthetic pathway area. Linking this back to bioluminescence, could these non-toxic iso-
are known (Hackett et al., 2013; Stuken et al., 2011), the genetic basis lates be a modified form of Batesian mimics?
for variability in STX production within natural populations is un- The presence and frequency of the sxtA4 gene in natural populations
defined. SxtA4 has been found to occur in multiple, slightly different of dinoflagellate species capable of STX production remains unknown.
genomic copies in multiple Alexandrium strains, suggesting adaptive It is well-established that high levels of genetic diversity exist among
plasticity and the potential basis for phenotypic diversity in terms of toxic (Alexandrium) dinoflagellate species from the same geographic
STX profile and toxicity (Alpermann et al., 2010; Kremp et al., 2014.). area (Anderson et al., 2012a; Richlen et al., 2012). Additionally, toxin
Differences in toxicity in natural populations highlight the need to production differs in genetically similar populations: population genetic
examine the genetic variability among sub-populations and the me- studies with Alexandrium showed substantial variation in a range of
chanisms by which cells acquire or lose the trait. The factors driving traits including toxin production both within and across populations,
genetic variability in toxicity, and, to a lesser extent, bioluminescence with differences in trait variation between genetically similar popula-
in natural HAB populations remain unexplored. Sxt gene duplication tions. This trait variation may promote HAB formation and persistence
and loss appear to be very common in dinoflagellates (Murray et al., under fluctuating environmental conditions (Brandenburg et al., 2018).
2015) and, in keeping with other dinoflagellate genes, are present in All the studies examining the genetic basis of toxin potential with STX-
numerous copies due to processes such as the recycling of processed producing species thus far have been based on lab cultures. The fre-
cDNAs back into the genome through reverse transcription, retro- quency and potential for toxin production among natural sub-popula-
position, (Bachvaroff and Place, 2008; Slamovits and Keeling, 2008; tions of HABs remains unknown yet is an area that needs to be explored
Zhang et al., 2007) and potentially environmental stress (Verma et al., in order to gain greater insights into the genetic mechanisms driving
2019). In recent years horizontal gene transfer has been shown to be a HAB evolution.
significant driver in dinoflagellate genome evolution (Nosenko and
Bhattacharya, 2007; Wisecaver et al., 2013).
One striking example as to the influences of genetic diversity in 5.7. Correlating bioluminescence and toxicity
relation to ecological functions of toxicity is an event that occurred with
P. bahamense. This species occurs in both the Atlantic-Caribbean and No data currently exist as to the correlation between biolumines-
the Indo-Pacific (Phlips et al., 2011; Usup et al., 2012), and toxic out- cence intensity and genetic/biochemical features and toxin production.
breaks are well-documented from P. bahamense in the Indo-Pacific Complicating factors in attempting an analysis of intensity among
(Azanza and Miranda, 2001; Azanza and Taylor, 2001; Gacutan et al., species include not only the existence of different strains and Batesian
1985; Harada et al., 1982; Llewellyn et al., 2006; Montojo et al., 2006). mimics, but also different measurement protocols for measuring light
In contrast, P. bahamense bloomed along the east coast of Florida for output. A common methodology involves stirring a population of cells
years with no known record of STX production. However, in the mid- to exhaustion, dividing the integrated light output by the number of
2000s, STX-contaminated seafood outbreaks across the United States cells and then reporting the results as photons per cell. This is not a
were traced back to Florida, and P. bahamense was identified as the meaningful measure when considering the eyes that the flash impacts.
source (Bodager, 2002; Landsberg et al., 2006), marking the first oc- In terms of visibility, photon flux at the peak of the flash is the most
currence of toxin production in the Western Atlantic. The mystery as to pertinent value, followed by decay time and number of flashes. In a
the variability for toxin production among natural HAB populations study where all of these variables were measured using the same pro-
extends beyond the events in Florida. Analysis of numerous strains of A. tocols (Latz et al., 2004), it is noteworthy how much these different
minutum in Irish coastal waters showed strains were indistinguishable parameters varied between species (Table 5). L. polyedra, which is be-
both morphologically and via sequencing of common phylogenetic lieved to be a Batesian mimic of toxic species, is the dimmest of the
markers; however, strains from southern areas proved to be toxic via four, producing only 2 or 3 short duration flashes. By contrast, P. fu-
HPLC analysis, while no toxins were detected in west coast strains siformis, which is non-toxic and evidence suggests employs its biolu-
(Touzet et al., 2007). Studies covering a large range of Gymnodinium minescence as a burglar alarm (Fleisher and Case, 1995; Mensinger and
strains indicated a high level of intra-population and regional variation Case, 1992), produces a flash intensity almost 3 orders of magnitude
in both the presence and amounts of STX congeners (Hallegraeff et al., greater, a longer flash duration and many more flashes per cell. In
2012). Additionally, in screening numerous strains of P. bahamense keeping with the burglar alarm hypothesis it is also noteworthy how
from the same geographical area in the Gulf of Mexico, toxicity was much brighter the first flash is to subsequent flashes in this species
confirmed in only one isolate (Morquecho, 2019). A very recent study (Widder and Case, 1981b). All aspects of light production in this species
with L. polyedra found high intra-specific variability in YTX composi- seem optimized for detection over long distance. The other two species,
tion among strains from the same geographical region (Peter et al., Ceratium fusus and Ceratocorys horrida produce less light and fewer
2018). This is further illustrated by findings of broad intra-specific flashes than P. fusiformis but more than L. polyedra. Evidence suggests
variability in multiple traits among strains of a HAB species all isolated that they are non-toxic however both species exhibit atypical mor-
from the same water sample in which a bloom was occurring (Menden- phology with spines that may be aversive to predators.
Deuer and Montalbano, 2015). Could this discrepancy in toxin pro- The genetic diversity of lcf within HAB sub-populations has recently
duction be linked to bioluminescence? StxA4 appears to be a necessary been examined at the single-cell level with P. bahamense (Cusick et al.,
component for cells to have the genetic capacity for toxin production. 2016). This study utilized single-cell PCR targeting the LCF conserved
Toxic and non-toxic bioluminescent strains from the same species (e.g. catalytic region of D3 on P. bahamense cells from various locations.
Sequences from the Western Atlantic formed two distinct clusters, while

Table 5
Flash characteristics among toxic and non-toxic bioluminescent dinoflagellates. Flash values obtained from (Latz et al., 2004).
Species Max flash intensity p/s Flash durastion (ms) # flashes/cell Toxic Cell surface

Ceratium fusus 1.10E+09 239 2 N thecate


Ceratocorys horrida 9.20E+09 184 7 N thecate
Lingulodinium polyedra 1.90E+08 100–150 2–3 Y thecate
Pyrocystis fusiformis 6.90E+11 210 23–62 N non thecate

17
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 5. (A.) Neighbor-joining tree of lcf sequences obtained from D3 of P. bahamense cells collected from the Western Atlantic. Available lcf sequences from the Indo-
Pacific were included in the analysis. (B.) Amino acid differences between cluster Lcf-WA1 and Lcf-WA2. Reprinted from Cusick et al. (2016).

sequences from an Indo-Pacific strain formed a third distinct cluster containing β-strands 5–14 (Fig. 7A). Comparisons between the two
(Fig. 5A). The clusters were defined by a set of core non-synonymous types of LCF retrieved from the D3 of P. bahamense with this model
substitutions (Fig. 5B). Sequence results obtained from environmental demonstrated that ca. 67% of these entailed substitutions that yielded
samples encompassing multiple dinoflagellate species revealed a similar different physical properties (i.e., polar to non-polar, polar to basic,
trend as that obtained for the P. bahamense LCF D3: significant discrete etc.), and more importantly, that nearly all of the amino acid changes
lcf clusters were observed for both L. polyedra and Protoceratium re- occurred within a β-strand (5, 7, 8, and 11) or α-helix 5 – i.e. active
ticulatum (albeit from the D2 domain) (Valiadi et al., 2014), while those sites of the domain. Additionally, these non-synonymous substitutions
from non-toxic species clustered together. A. ostenfeldii is a toxic bio- were specific to the Lcf-WA1 sequence cluster, and did not occur among
luminescent dinoflagellate with a wide geographic distribution and Lcf-WA2, P. lunula LcfA and LcfB, L. polyedra, A. tamarense, A. affine,
extensive bloom formation in coastal waters of Europe and the Indo-Pacific strain of P. bahamense (Fig. 7B) (Cusick et al.,
(Almandoz et al., 2014; Gribble et al., 2005; MacKenzie et al., 1996). 2016).
Lcf gene sequences and bioluminescence were examined from toxic While the biochemical ramifications of the active site domain dif-
bloom populations in the Baltic and North Seas of Europe, and the re- ferences found in P. bahamense remain to be determined, this may allow
sulting phylogenetic analysis demonstrated that lcf sequences from for the cell to control the level of light produced. This concept of amino
Baltic Sea strains clustered separately from North Sea strains (Le acid variation altering the functional properties is in keeping with
Tortorec et al., 2016) (Fig. 6). findings of the light-harvesting complex proteins of the Symbiodinium,
Assimilating the collective data on flash characteristics (intensity, in which molecular diversity may translate into functionally distinct
duration), lcf sequence data, and toxicity leads to the question as to the proteins that enable the cells to utilize the different light conditions of
influence of toxicity on bioluminescence. The clusters obtained for P. their environment (Boldt et al., 2012). While the overall ecological role
bahamense lcf sequences (Fig. 5A) are defined by non-synonymous of bioluminescence in a single HAB species may be to alleviate preda-
substitutions resulting in a set of core amino acid differences (Fig. 5B) tion pressure, this may be achieved in different ways based on the toxin
that occur within the active site of the D3 domain in P. bahamense. capabilities of the cells and reflected in differences in LCF corre-
These differences were compared with the crystal structure of the L. sponding to light output.
polyedra luciferase D3, which has been resolved at 1.8-Å (Schultz et al.,
2005). Briefly, the region is comprised of seven α-helices and 16 β-
strands, organized into two subdomains: a regulatory domain defined
by a three-helix bundle at the top followed by a β-barrel below

18
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 6. Lcf sequences cluster by geographic location in A. ostenfeldii. Lcf sequences amplified from A. ostenfeldii isolates from the Baltic and North Sea. Sequences were
amplified using primers targeting the N-terminal and initial portion of the D1 domain. Figure reprinted from (Le Tortorec et al., 2016).

6. Signaling receptors, and enzyme-linked receptors, while calcium and cyclic AMP
are key molecules used in mediating these processes (Wolfe, 2000).
A signal is the transfer of information between two organisms While some work has been done on cell-surface signal receptors in the
through a biogenic stimulus that can be perceived by a sensory system protozoans (mainly ciliates and flagellates), little is known about such
and elicit an adaptive response (Wolfe, 2000). Within this context sig- signaling mechanisms in marine dinoflagellates (Wolfe, 2000).
naling refers to communication between individuals and includes both
visual and chemical signals. Visual signaling was discussed in the last 6.1. Signal transduction in bioluminescence
section with regard to ecological functions. In this section the focus is
on chemical signals. As reviewed in the Biochemistry section, bioluminescence is trans-
The main mechanisms involved in cell surface sensing and trans- duced via mechanical stimulation of the cells. The flash is the culmi-
duction include ion-channel-linked receptors, G protein-linked nation of intricate signal transduction events that begins with an

19
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Fig. 7. (A) L. polyedrum LCF D3 sequence with alpha-helix (α) and B-strand (β) noted. The two underlined six-residue sequences correspond to the beginning and end
of the highly conserved catalytic domain region. Red box designates approximate region amplified by degenerate lcf primers. Redrawn from Schultz et al. (2005). (B)
Amino acid sequence logo of highlighted D3 region. Included are LCF sequences from P. bahamense clusters Lcf-WA1, Lcf-WA2, an Indo-Pacific strain, P. lunula LcfA
and LcfB, A. tamarense, and A. affine. Blue arrows designate amino acid changes in Lcf-WA1.

increase in membrane fluidity due to a high mechanical stress. channels are common elements of mammalian mechanosensing path-
Deformation of the membrane creates an action potential that opens ways. Plasma membrane mechanical stress was previously shown to
membrane channels and the rapid influx of protons from the acidic activate a TRP family protein in mammalian cells (Shen et al., 2015).
vacuole results in an intracellular pH drop that activates LCF and, in TRP-like channels have recently been identified as components of the
dinoflagellates with an LBP, dissociates luciferin from the LBP. In 1972, mechanotransduction (bioluminescent) signaling pathway in L. poly-
J. Woodland Hastings and colleagues proposed the existence of a proton edra (Lindstrom et al., 2017b).
selective channel that opens in response to depolarizing voltage across
the vacuole membrane of bioluminescent dinoflagellates, conducting
protons into the scintillons and resulting in a pH drop that triggers light 6.2. Copepodamides enhance bioluminescence
emission (Fogel and Hastings, 1972). The discovery and analysis of a
voltage-gated proton channel in the toxic species Karlodinium veneficum Recent studies revealed that copepodamides influence dino-
revealed that while it contained enough signature elements to be flagellate bioluminescence, as both L. polyedra and A. tamarense were
identified as a proton channel, its expression exhibited novel properties, found to increase their total bioluminescence capacity in response to
suggesting that these channels in dinoflagellates serve very different copepodamides. Additional experiments with L. polyedra showed bio-
functions than those in mammalian and other systems (Smith et al., luminescence increased in a dose-dependent manner over time in re-
2011). However, Hastings’ early hypothesis was recently confirmed sponse to the copepod cues resulting in a brighter flash upon stimula-
with the discovery of a voltage-gated proton channel in L. polyedra (LpH tion and markedly affecting grazing preference by the copepod predator
(v)1) (Rodriguez et al., 2017). Hv1 sequences have also been identified A. tonsa (Lindstrom et al., 2017a; Prevett et al., 2019). L. polyedra is the
in A. tamarense and N. scintillans, providing strong evidence that Hv1, in preferred prey when non-bioluminescent; however, when biolumines-
addition to LCF and LBP, is part of the signal transduction pathway cent, the cells flash upon contact with the A. tonsa and are rejected,
(Rodriguez et al., 2017). Upon stimulation via shear stress, this signal without apparent harm to the cell (Prevett et al., 2019). How are these
transduction system is likely initiated by stretch activated channels chemical cues being sensed and transduced by the cells to enhance
(Jin et al., 2013) at the surface of L. polyedra, and relays a signal bioluminescence? The study by Lindstrom (Lindstrom et al., 2017a)
through intracellular calcium signaling via G-proteins (Chen et al., exposed the dinoflagellate cells to cell-free copepodamides (i.e. no co-
2007; von Dassow and Latz, 2002). However, the other members of this pepods in treatment vessels), thus removing the mechanical stimula-
signal transduction system remain unknown. tion. However, bioluminescence was measured following acidification
The ability of cells to sense and respond to their physical environ- of the cultures; therefore, while these data demonstrate the ability of
ment involves mechanically activated ion channels as a part of me- copepodamides to enhance bioluminescence, based on the methods, it
chanotransduction signaling pathways. In prokaryotes, these are appears that bioluminescence is augmented after entry of the copepo-
stretch-activated ion channels that regulate cell turgor for osmor- damides into the cell, such that bioluminescence is not enhanced via a
egulation, while eukaryotes can sense varied environmental mechanical cell-surface signaling network.
stresses in addition to stretch. Transient receptor potential (TRP) ion

20
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

6.3. Copepodamides enhance toxin production GPCRs are more divergent from known GPCRs than are diatom GPCRs
(Mojib and Kubanek, 2020) and while this divergence may present a
The presence of copepods and their waterborne cues have also been challenge it also provides a platform for future studies to examine the
shown to boost toxin production (Selander et al., 2006; Wohlrab et al., role of these proteins in dinoflagellate signal transduction.
2010; Yang et al., 2011). Comparisons among multiple factors, in-
cluding nitrogen sources, conspecific and interspecific species alarm 7. Evolution
cues, and grazer (copepod) presence in a toxic strain of A. catenella,
showed grazer presence enhanced toxin production by an order of 7.1. Bioluminescence
magnitude greater than any other variable (Griffins et al., 2019). A.
minutum has been shown to produce STX in response to waterborne Characterization of the LCF from seven species of photosynthetic
cues from the calanoid copepod A. tonsa, with toxin production corre- dinoflagellates demonstrated that they share a common origin
lated to an increased resistance to further grazing (Selander et al., (Liu et al., 2004). While LCF domain organization is similar among all,
2006). Intriguingly, it has been shown that grazer-induced toxin pro- differences exist in the distribution of the synonymous substitution
duction is a grazer-specific response, as exposure of toxic A. minutum rates across the lcf gene, as well as the intergenic regions (Liu and
strains to different species of calanoid copepods elicited significantly Hastings, 2006; Liu et al., 2004; Okamoto et al., 2001). Synonymous
increased cell-specific toxicity to two of the three copepod species (A, substitutions rates were found to be greatly reduced in L. polyedra and
clausi, Centropages typicus, but not Pseudocalanus sp.) tested P. reticulatum but not the other five (A. affine, A. tamarense, P. fusiformis,
(Bergkvist et al., 2008). These species-specific responses have also been P. noctiluca, P. lunula) (Liu et al., 2004).The significance of these dif-
shown to occur in a toxic A. tamarense strain exposed to three different ferences remains unknown, and how they relate to the evolutionary
copepod species (Calanus helgolandicus, A. clausii, and Oithona similis) constraints of this system in dinoflagellates. The luciferase of L. polyedra
(Wohlrab et al., 2010), indicating that co-evolutionary processes may likely diverged early from the other six based on lcf and ribosomal DNA
be involved in these responses. The chemical basis for this grazer-spe- analysis (Liu et al., 2004). Comparison of the LCF among photo-
cific response can be explained by copepodamides (taurine connected synthetic species indicate ancient triplication of one of the domains
via an amide to an isoprenoid fatty acid conjugate of varying compo- occurred in an ancestor common to all that was then carried forward
sition), a suite of lipids produced by copepod species. Toxic A. minutum during the evolution of the different photosynthetic species (Liu et al.,
was shown to increase its toxin production 20-fold in response to pico- 2004), with the entire lcf gene retained in descendants in which bio-
to nanomolar concentrations of copepodamides. Furthermore, the spe- luminescence provided a selective advantage under strong selective
cies-specific defense response of toxic cells can be accounted for by the pressure for maintaining catalytic activity (Liu et al., 2004). However,
species-specific blends emitted by the different copepod species the differences in synonymous substitution rates indicate a second se-
(Selander et al., 2015). Since their initial discovery in three species of lective pressure must have occurred, likely after the divergence of L.
calanoid copepods (Selander et al., 2015), more than 20 additional polyedra and P. reticulatum, to account for the very low rate of synon-
copepodamides have been identified, (Grebner et al., 2019), providing ymous substitutions within these species (Liu et al., 2004). The func-
further support as to the findings that the species-specific defense re- tional implications of this constraint, as evidenced by enhanced stabi-
sponse of toxic cells is accounted for by the species-specific blends lity of RNA secondary structure in the LCF of L. polyedra and P.
emitted by the different copepod species (Selander et al., 2015). reticulatum, may be associated with a regulatory mechanism (Liu et al.,
The signal transduction systems by which dinoflagellate cells sense 2004), such as the translational control of the circadian rhythm in L.
and respond to these chemical cues are only beginning to be elucidated. polyedra. As discussed in the Biochemistry section, L. polyedra has an
Using functional genomics, serine/threonine kinase signaling pathways LBP and an LCF, both of which are under circadian control. Biolumi-
were identified as significant elements in the response to copepod cues, nescent measurements with P. bahamense, which also possesses an LBP,
and in most cases, serine/threonine kinase pathway regulation was also suggest its bioluminescence is under circadian control
specific to the copepod species (Wohlrab et al., 2010). Additionally, (Biggley et al., 1969). This is in contrast to P. lunula, which contains the
transcripts for proteins dependent on calcium ions as second messen- greatest rate of synonymous substitutions (Liu et al., 2004): there is no
gers were differentially expressed (Wohlrab et al., 2010), in keeping LBP, and the daily levels of LCF remain constant (Knaust et al., 1998).
with other findings as to the role of calcium signaling in the response of Sequencing of lbp genes across multiple genera, including
marine phytoplankton to environmental cues (Jingwen et al., 2006; Alexandrium, Protoceratium, Gonyaulax, and Ceratocorys, revealed a very
Verret et al., 2010) as well as bioluminescence (von Dassow and diverse gene family comprised of several gene types (Valiadi and
Latz, 2002). Genomic analyses have found that conserved channel types Iglesias-Rodriguez, 2014). Phylogenetic analysis showed that, for mul-
likely to be involved in calcium signaling are widely distributed among tiple genera, the evolution of the lbp gene differed from the species’
algal species, including dinoflagellates (Verret et al., 2010). RNA-Seq general phylogenies, demonstrating the complex evolutionary history
profiling of Alexandrium species (A. fundyense) upon exposure to para- of dinoflagellate bioluminescence (Valiadi and Iglesias-
site-associated (Amoebophrya) waterborne cues showed differential Rodriguez, 2014). In N. scintillons, it occurs as a fusion with LCF
regulation of signal-transduction-related genes, including serine/ (Liu and Hastings, 2007). An additional LBP has since been identified in
threonine kinases and genes involved in calcium, mitogen-activated N. scintillans and shown to be actively transcribed in addition to the one
protein kinase and Ras signaling (Lu et al., 2016). Copepodamides have attached to LCF, supporting the notion that a significant re-structuring
also been shown to stimulate domoic acid production in the toxic of the system has taken place in this species (Valiadi and Iglesias-
diatom Pseudo-nitzschia seriata (Selander et al., 2019). RNA-sequencing Rodriguez, 2014). Collectively, the evolutionary analysis of the biolu-
results indicate P. seriata transduce copepod cues via G protein path- minescent system in multiple dinoflagellate species illustrates the
ways (Hardardottir et al., 2019). Could dinoflagellates utilize similar adaptive value of this trait.
pathways in response to the copepodamides? Recent comprehensive The light produced by toxic species is dim in comparison to that of
transcriptomic analyses identified G protein-coupled receptors (GPCRs) many of the non-toxic species. All toxic, bioluminescent species for
in 81 diatom and 36 dinoflagellate transcriptomes (Mojib and which sequence data are available possess a separate LBP in addition to
Kubanek, 2020). The GPCRs of both groups were highly divergent from the LCF (Table 1). It is worth noting that lbp gene sequences have not
known metazoan GPCRs and, unlike other eukaryotes that express been detected in the gonyaucaloids Ceratium and Fragilidium nor in
multiple classes of GPCRs, both were found to express only a single Protoperidinium and Pyrocystis (Colepicolo et al., 1993; Valiadi and
major class with greatest similarity to class C and rhodopsin‐like Iglesias-Rodriguez, 2014). Protoperidinium and Pyrocystis are both large,
GPCRs, respectively. Based on the transcriptome data, dinoflagellate bright-emitting, non-toxic cells; Ceratium and Fragilidium are also non-

21
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

toxic. Additionally, the lbp gene sequences of A. monilatum and G. spi- nucleotide sequences; this is further supported by analysis of sxtA and
nifera, both toxic, displayed increased diversity from those of other sxtG in multiple strains of G. catenatum. Although highly conserved,
Gonyaulacales. This correlation between LBP evolution and toxicity is a sxtA4 phylogeny mirrors that of species phylogeny (Mendoza-Flores
link that requires further investigation. et al., 2018; Murray et al., 2015), indicating that the gene may have
evolved in a common dinoflagellate ancestor of G. catenatum, P. baha-
7.2. Saxitoxin mense and Alexandrium and a secondary loss of these genes occurred in
non-toxic Alexandrium species (Murray et al., 2015). These collective
Three theories have been proposed as to the origin of STX produc- findings with sxtA1, sxtA4, and sxtG support the theory that the ability
tion in dinoflagellates: convergent evolution, horizontal gene transfer, for STX synthesis has been secondarily lost in some species, with sxtA
or autonomous production by associated bacteria (Orr et al., 2013b). being lost independently on multiple occasions (Orr et al., 2013b).
STX biosynthesis genes were first identified in several genera of fresh- Selection (positive or negative) analysis of sxtA1 and sxtG showed
water cyanobacteria; 14 genes were shared among the genera, leading that the clade of sxtA1 putatively associated with STX synthesis is under
to designation of 14 “core” genes (Murray et al., 2011a; Wiese et al., negative selection, indicating pressure to constrain function
2010). Due to the complexity of their genomes, identification of these (Murray et al., 2015). This differs from the evolution of the cyano-
homologs in dinoflagellates took several more years to achieve. Tran- bacterial sxt gene cluster, in which positive selection was found to play
scriptomic sequencing of toxin-producing Alexandrium strains resulted a role (Murray et al., 2011a). Similarly, sxtG also showed negative se-
in the discovery of nuclear-encoded sxt homologs in dinoflagellates lection, indicating functional constraint for this gene as well. The pre-
(Hackett et al., 2013; Stuken et al., 2011). SxtA transcripts were found sence of multiple paralogs in different clades of both sxtA1 and sxtG
to be closely related to a clade of cyanobacteria sxtA sequences, and indicate these genes may perform different functions, as both display
possessed the same domain structure (Stuken et al., 2011). However, positive selection in clades associated with a lack of toxin production
unlike their bacterial homologs, dinoflagellate transcripts contained a facilitating the development of different functions (Murray et al.,
higher GC content, occurred in multiple copies, and displayed the ty- 2015).
pical dinoflagellate spliced-leader sequences and eukaryotic polyA-tails In general, negative selection is the selective removal of (seemingly)
(Stuken et al., 2011) confirming their dinoflagellate origin and pro- deleterious genes, and removal of these genes can be achieved on the
viding strong evidence of autonomous STX production by dino- population genetics level, with as little as a single point mutation.
flagellates (Orr et al., 2013b). Subsequent characterization of sxtG, the Linking the findings of potential negative selection of sxt genes with
second “core” gene, confirmed the previous results, as transcripts also genetic diversity, how then can this be translated into toxic and non-
possessed spliced-leader sequences and poly-A tails (Orr et al., 2013a). toxic HAB populations occurring both spatially and temporally in the
While the findings of dinoflagellate STX biosynthesis homologs lessened same general geographic area (as discussed in the section on Genetic
the support for the co-culture bacteria theory it is intriguing to note that Diversity)? If the toxin serves to protect cells from predation pressure,
half of the core cyanobacterial sxt genes have their origin from pro- one would image all cells in a population would retain sxtA4. Here, the
teobacterial genomes, and half of the dinoflagellate sxt homologs also concept of bioluminescence as Batesian mimicry may come into play: if
have a putative proteobacterial origin. While no proteobacteria have all cells in the population are bioluminescent, sxtA4 may be lost in a
been shown to produce STX, many strains are found in close association portion of the sub-population over time, as the dinoflagellate predator
with STX-producing dinoflagellate strains (Orr et al., 2013b). learns to associate the flash with toxicity, thereby avoiding all biolu-
Based on the initial findings of dinoflagellate-specific sxt homologs, minescent (and at that point, toxic) cells. While this theory requires
it was proposed that a horizontal gene transfer (HGT) event between additional genetic data on sub-populations from the same geographical
ancestral STX-producing bacteria and dinoflagellates occurred (Stuken area in order to develop appropriately, it presents a viable and testable
et al., 2011). Additional sequence analysis has led to a more refined hypothesis as to potential links between bioluminescence and toxicity
theory: STX evolution in dinoflagellates was initiated via a massive in HAB species based on the findings of negative selection for sxt genes.
gene transfer event from cyanobacteria into an ancestor of Alexandrium
and Pyrodinium, who then acquired additional genes from multiple 8. Conclusions
prokaryotic sources. The sxt genes were then extensively modified, with
some homologs lost or replaced. The pathway was lost in some lineages, Both bioluminescence and toxin production have the potential to
and transferred to Gymnodinium via an endosymbiotic gene transfer. greatly influence HAB initiation and persistence. Some of the most
Genes and domains were re-shuffled, resulting in loss of sxtA4 in some notorious HABs consist of species with the capacity for both traits, and
species (Orr et al., 2013b). it is this combination for which many knowledge gaps remain, at both
The presence and high sequence conservation of the sxtA4 domain the ecological and genetic levels. The function, and thus adaptive value,
was confirmed in 40 strains (35 from Alexandrium species, a single of the flash have been long debated. Multiple hypotheses have been
Pyrodinium, and four Gymnodinium) of STX-producing species, and ab- proposed, all rooted in the premise that bioluminescence acts as an anti-
sent in non-toxin producing species from these genera (Murray et al., grazing strategy and this relief from grazing pressure plays an im-
2015). Screening for the stxA1 domain found it was widely distributed portant role in bloom formation. The same holds for toxin production;
in non-STX producing dinoflagellates and occurred as three paralogs, there are multiple hypotheses, with one of the main being that the toxin
with one of the paralogs seemingly closely associated with STX synth- functions as a defense from grazers. Combining these traits leads to the
esis (Murray et al., 2015). SxtG was also found to be generally specific hypothesis that the flash functions as an aposematic signal, and that
for STX-producing strains (Murray et al., 2015). While detected in Batesian mimicry, which often conjures up images of monarch butter-
Alexandrium species not known to synthesize STX, its presence was not flies, may well be occurring within single-celled dinoflagellates as well.
detected in non-STX-producing species outside the Alexandrium genus, The genetic machinery of bioluminescence has been defined.
P. bahamense, or G. catenatum (Hackett et al., 2013; Orr et al., 2013a). Sequencing of lcf and lbp genes across genera has shown a complex
Phylogenetic analysis of sxtA1 and sxtG indicates that the evolution evolutionary history, supporting, at the genomic level, the adaptive
of these genes has been complex, with duplication events involved in value of this trait. However, several areas remain unexplored that may
the evolution of both sxtA and sxtG (Murray et al., 2015). Analysis of provide additional insights into the eco-evolutionary role of dino-
sxtG in multiple strains of Alexandrium along with a G. catenatum strain flagellate bioluminescence, especially with regards to HAB species.
indicate multiple independent acquisitions and losses of this gene These include: the mechanism behind the general trend of LBP presence
(Orr et al., 2013a). Widespread horizontal gene transfer does not ap- in toxic, dim-emitting species and lack of LBP in non-toxic, bright-
pear to occur for sxtA4 based on analysis of Alexandrium sxtA4 emitting species; and the effects of key amino acid differences within

22
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

the LCF catalytic domain in toxic species. The core genetic machinery in Magadan, S., Maneiro, I., Riveiro, I., Iglesias, P., 2006. Feeding strategies of the co-
STX biosynthesis has also recently been determined, and with it the pepod Acartia clausi on single and mixed diets of toxic and non-toxic strains of the
dinoflagellate Alexandrium minutum. Mar. Ecol. Prog. Ser. 316, 115–125.
finding that the sxtA4 gene appears specific for toxin production. Beedessee, G., Hisata, K., Roy, M.C., Satoh, N., Shoguchi, E., 2015. Multifunctional
However, the identification of genes involved in both these traits were polyketide synthase genes identified by genomic survey of the symbiotic dino-
derived from lab-based experiments with clonal isolates. A large gap in flagellate, Symbiodinium minutum. BMC Genom. 16, 11.
Bergkvist, J., Selander, E., Pavia, H., 2008. Induction of toxin production in dino-
our understanding of both these traits is their distribution among nat- flagellates: the grazer makes a difference. Oecologia 156 (1), 147–154.
ural bloom populations, especially with regards to toxin production. Biggley, W.H., Swift, E., Buchanan, R.J., Seliger, H.H., 1969. Stimulable and spontaneous
When considering the ecological function of toxin production, one bioluminescence in marine dinoflagellates, Pyrodinium bahamense, Gonyaulax poly-
edra, and Pyrocystis lunula. J. Gen. Physiol. 54 (1P1), 96–122.
feature that is often overlooked is the variability of this trait within sub- Bodager, D., 2002. Outbreak of saxitoxin illness following consumption of Florida puf-
populations of a natural bloom of the species. It remains unknown ferfish. Fl. J. Environ. Health 179, 9–13.
whether all the cells possess the genes necessary for toxin production, Boldt, L., Yellowlees, D., Leggat, W., 2012. Hyperdiversity of genes encoding integral
light-harvesting proteins in the dinoflagellate Symbiodinium sp. PLoS One 7 (10),
or only a portion of the sub-population. Extending that reasoning leads
e47456.
to further questions as to how the genes are maintained within that Borkow, G., Gabbay, J., 2005. Copper as a biocidal tool. Curr. Med. Chem. 12 (18),
population, as both toxic and non-toxic strains have been isolated from 2163–2175.
the same geographic area. Can toxic cells become non-toxic, and vice Boyer, G.L., Sullivan, J.J., Andersen, R.J., Harrison, P.J., Taylor, F.J.R., 1987. Effects of
nutrient limitation on toxin production and composition in the marine dinoflagellate
versa? This can further be extended to bioluminescence, for just as there Protogonyaulax tamarensis. Mar. Biol. 96, 123–128.
exist toxic and non-toxic strains within a species, so too are there light Brand, L.E., Campbell, L., Bresnan, E., 2012. Karenia: the biology and ecology of a toxic
and dark strains. The frequency and potential for both toxin production genus. Harmful Algae 14, 156–178.
Brandenburg, K.M., Wohlrab, S., John, U., Kremp, A., Jerney, J., Krock, B., Van de Waal,
and bioluminescence among natural HAB sub-populations remains D.B., 2018. Intraspecific trait variation and trade-offs within and across populations
unknown yet is an area that needs to be explored in order to gain of a toxic dinoflagellate. Ecol. Lett. 21 (10), 1561–1571.
greater insights into the genetic and environmental mechanisms driving Bricelj, V.M., Connell, L., Konoki, K., Macquarrie, S.P., Scheuer, T., Catterall, W.A.,
Trainer, V.L., 2005. Sodium channel mutation leading to saxitoxin resistance in clams
HAB evolution. increases risk of PSP. Nature 434 (7034), 763–767.
Burkenroad, M.D., 1943. A possible function of bioluminescence. J. Mar. Res. 5, 161–164.
Declaration of Competing Interest Buskey, E., Mann, C.G., Swift, E., 1987. Photophobic responses of calanoid copepods:
possible adaptive value. J. Plank. Res. 9, 857–870.
Buskey, E., Mills, L., Swift, E., 1983. The effects of dinoflagellate bioluminescence on the
The authors declare that they have no known competing financial swimming behavior of a marine copepod. Limnol. Oceanogr. 28 (3), 575–579.
Buskey, E.J., Stearns, D.E., 1991. The effects of starvation on bioluminescence potential
interests or personal relationships that could have appeared to influ-
and egg release of the copepod Metridia longa. J. Plankton Res. 13 (4), 885–893.
ence the work reported in this paper. Buskey, E.J., Swift, E., 1983. Behavioral responses of the coastal copepod Acartia hud-
sonica (Pinhey) to simulated dinoflagellate bioluminescence. J. Exp. Mar. Biol. Ecol.
References 72, 43–58.
Buskey, E.J., Swift, E., 1985. Behavorial responses of oceanic zooplankton to simulated
bioluminescence. Biol. Bull. 168, 263–275.
Abrahams, M.V., Townsend, L.D., 1993. Bioluminescence in dinoflagellates: a test of the Catterall, W.A., 1985. The voltage sensitive sodium channel: a receptor for multiple
burglar alarm hypothesis. Ecology 74, 258–260. neurotoxins. In: Anderson, W., Baden (Eds.), Toxic Dinoflagellates. Elsevier Science
Akbar, M.A., Mohd Yusof, N.Y., Tahir, N.I., Ahmad, A., Usup, G., Sahrani, F.K., Bunawan, Publishing Co, Inc., pp. 329–342.
H., 2020. Biosynthesis of saxitoxin in marine dinoflagellates: an omics perspective. Cembella, A.D., 1998. Ecophysiology and metabolism of paralytic shellfish toxins in
Mar. Drugs 18 (2), 103. marine microalgae. In: Anderson, D.M., Cembella, A.D., Hallegraeff, G. (Eds.),
Alfonso, A., Vieytes, M.R., Botana, L.M., 2016. Yessotoxin, a promising therapeutic tool. Physiological Ecology of Harmful Algal Blooms. Springer-Verlag, Berlin, pp.
Mar. Drugs 1 (14), 30. 381–403.
Almandoz, G.O., Montoya, N.G., Hernando, M.P., Benavides, H.R., Carignan, M.O., Cembella, A.D., 2003. Chemical ecology of eukaryotic microalgae in marine ecosystems.
Ferrario, M.E., 2014. Toxic strains of the Alexandrium ostenfeldii complex in southern Phycologia 42, 420–447.
South America (Beagle Channel, Argentina). Harmful Algae 37, 100–109. Chakraborty, S., Pancic, M., Andersen, K.H., Kiorboe, T., 2019. The cost of toxin pro-
Alpermann, T.J., Tillmann, U., Beszteri, B., Cembella, A.D., John, U., 2010. Phenotypic duction in phytoplankton: the case of PST producing dinoflagellates. ISME J 13 (1),
variation and genotypic diversity in a planktonic population of the toxigenic marine 64–75.
dinoflagellate Alexandrium tamarense (Dinophyceae). J. Phycol. 46, 18–32. Chen, A.K., Latz, M.I., Sobolewski, P., Frangos, J.A., 2007. Evidence for the role of G-
Alvarez, G., Uribe, E., Regueiro, J., Blanco, J., Fraga, S., 2016. Gonyaulax taylorii, a new proteins in flow stimulation of dinoflagellate bioluminescence. Am. J. Physiol. –
yessotoxins-producer dinoflagellate species from Chilean waters. Harmful Algae 58, Regul. Integr. Comp. Physiol. 292 (5), R2020–R2027.
8–15. Chen, L., Zhang, H., Finiguerra, M., Bobkov, Y., Bouchard, C., Avery, D.E., Anderson,
Anderson, D.M., Alpermann, T.J., Cembella, A.D., Collos, Y., Masseret, E., Montresor, M., P.A.V., Lin, S., Dam, H.G., 2015. A novel mutation from gene splicing of a voltage-
2012a. The globally distributed genus Alexandrium: multifaceted roles in marine gated sodium channel in a marine copepod and its potential effect on channel
ecosystems and impacts on human health. Harmful Algae 14, 10–35. function. J. Exp. Mar. Biol. Ecol. 469, 131–142.
Anderson, D.M., Cembella, A.D., Hallegraeff, G.M., 2012b. Progress in understanding Colepicolo, P., Roenneberg, T., Morse, D., Taylor, W.R., Hastings, J.W., 1993. Circadian
harmful algal blooms: paradigm shifts and new technologies for research, monitoring, regulation of bioluminescence in the dinoflagellate Pyrocystis lunula. J. Phycol. 29,
and management. Annu. Rev. Mar. Sci. 4, 143–176. 173–179.
Anderson, D.M., Cheng, T.P.-O., 1988. Intracellular localization of saxitoxins in the di- Colin, S., Dam, H., 2005. Testing of resistance of pelagic marine copepods to a toxic
nofllagellate Gonyaulax tamarensis. J. Phycol 24 (1), 17–22. dinoflagellate. Evol. Ecol. 18, 355–377.
Anderson, D.M., Kulis, D.M., Sullivan, J.J., Hall, S., 1990a. Toxin composition variations Colin, S., Dam, H.G., 2004. Testing for resistance of pelagic marine copepods to a toxic
in one isolate of the dinoflagellate Alexandrium fundyense. Toxicon 28 (8), 885–893. dinoflagellate. Evol. Ecol. 18, 355–377.
Anderson, D.M., Kulis, D.M., Sullivan, J.J., Hall, S., Lee, C., 1990b. Dynamics and phy- Colin, S.P., Dam, H.G., 2002. Latitudinal differentiation in the effects of the toxic dino-
siology of saxitoxin production by the dinoflagellates Alexandrium spp. Mar. Biol. 104 flagellate Alexandrium spp. on the feeding and reproduction of populations of the
(3), 511–524. copepod Acartia hudsonica. Harmful Algae 1, 113–125.
Armstrong, M., Kudela, R., 2006. Evaluation of California isolates of Lingulodinium poly- Colin, S.P., Dam, H.G., 2007. Comparison of the functional and numerical responses of
edrum for the production of yessotoxin. Afr. J. Mar. Sci. 28, 399–401. resistant versus non-resistant populations of the copepod Acartia hudsonica fed the
Azanza, R.V., Miranda, L.N., 2001. Phytoplankton composition and Pyrodinium baha- toxic dinoflagellate Alexandrium tamarense. Harmful Algae 6, 875–882.
mense toxic blooms in Manila Bay, Philippines. J. Shellfish Res. 20 (3), 1251–1255. Cusick, K.D., Boyer, G.L., Wilhelm, S.W., Sayler, G.S., 2009. Transcriptional profiling of
Azanza, R.V., Taylor, F.J.R., 2001. Are pyrodinium blooms in the Southeast Asian region Saccharomyces cerevisiae upon exposure to saxitoxin. Environ. Sci. Technol. 43 (15),
recurring and spreading? A view at the end of the millennium. Ambio 30 (6), 6039–6045.
356–364. Cusick, K.D., Minkin, S.C., Dodani, S.C., Chang, C.J., Wilhelm, S.W., Sayler, G.S., 2012.
Bachvaroff, T.R., Place, A.R., 2008. From stop to start: tandem gene arrangement, copy Inhibition of copper uptake in yeast reveals the copper transporter Ctr1p As a po-
number and trans-splicing sites in the dinoflagellate Amphidinium carterae. PLoS One tential molecular target of saxitoxin. Environ. Sci. Technol. 46 (5), 2959–2966.
3 (8), e2929. Cusick, K.D., Sayler, G.S., 2013. An overview on the marine neurotoxin saxitoxin: ge-
Bagøien, E., Miranda, A., Reguera, B., Franco, J.M., 1996. Effects of two paralytic shellfish netics, molecular targets, methods of detection, and ecological functions. Mar. Drugs
toxin producing dinoflagellates on the pelagic harpacticoid copepod Euterpina acuti- 11, 991–1018.
frons. Mar. Biol. 126, 361–369. Cusick, K.D., Wetzel, R.K., Minkin, J., S.C., Dodani, S.C., Wilhelm, S.W., Sayler, G.S.,
Banse, K., 1982. Cell volumes, maximal growth rates of unicellular algae and ciliates, and 2013. Paralytic shellfish toxins inhibit copper uptake in Chlamydomonas reinhardtii.
the role of ciliates in the marine pelagial. Limnol. Oceanogr. 27 (6), 1059–1071. Environ. Toxicol. Chem. 32, 1388–1395.
Barreiro, A., Guisande, C., Frangopulos, M., Gonzalez-Fernandez, A., Munoz, S., Perez, D., Cusick, K.D., Widder, E.A., 2014. Intenisty differences in bioluminescent dinoflagellates

23
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

impact foraging efficiency in a nocturnal predator. Bull. Mar. Sci. 90, 797–811. in cyanobacteria and dinoflagellates. Mol. Biol. Evol. 30, 70–78.
Cusick, K.D., Wilhelm, S.W., Hargraves, P.E., Sayler, G.S., 2016. Single-cell PCR of the Haddock, S.H.D., Moline, M.A., Case, J.F., 2010. Bioluminescence in the Sea. Ann. Rev.
luciferase conserved catalytic domain reveals a unique cluster in the toxic biolumi- Mar. Sci. 2, 443–493.
nescent dinoflagellate Pyrodinium bahamense. Aquat. Biol. 25, 139–150. Hallegraeff, G.M., Blackburn, S.I., Doblin, M.A., Bolch, C.J.S., 2012. Global toxicology,
Czyż, A., Plata, K., Węgrzyn, G., 2003. Stimulation of DNA repair as an evolutionary drive ecophysiology and population relationships of the chainforming PST dinoflagellate
for bacterial luminescence. Luminescence 18, 140–144. Gymnodinium catenatum. Harmful Algae 14, 130–143.
Czyz, A., Wrobel, B., Wegrzyn, G., 2000. Vibrio harveyi bioluminescence plays a role in Hanley, K.A., Widder, E.A., 2017. Bioluminescence in dinoflagellates: evidence that the
stimulation of DNA repair. Microbiology 146, 283–288. adaptive value of bioluminescence in dinoflagellates is concentration dependent.
da Costa, R.M., Franco, J., Cacho, E., Fernandez, F., 2005. Toxin content and toxic effects Photochem. Photobiol. 93 (2), 519–530.
of the dinoflagellate Gyrodinium corsicum (Paulmier) on the ingestion and survival Harada, T., Oshima, Y., Kamiya, H., Yasumoto, T., 1982. Confirmation of paralytic
rates of the copepods Acartia grani and Euterpina acutifrons. J. Exp. Mar. Biol. Ecol. shellfish toxins in the dinoflagellate Pyrodinium bahamense var. compressa and bi-
322 (2), 177–183. valves in Palau. Bull. Jpn. Soc. Sci. Fish. 48 (6), 821–825.
de Cock, R., Matthysen, E., 1999. Aposematism and bioluminescence: experimental evi- Hardardottir, S., Wohlrab, S., Hjort, D.M., Krock, B., Nielsen, T.G., John, U., Lundholm,
dence from glow-worm larvae (Coleoptera: lampyridae). Evol. Ecol. 13, 619–639. N., 2019. Transcriptomic responses to grazing reveal the metabolic pathway leading
de Cock, R., Matthysen, E., 2002. Glow-worm larvae bioluminescence (Coleoptera: lam- to the biosynthesis of domoic acid and highlight different defense strategies in dia-
pyridae) operates as an aposemtic signal upon toads (Bufo bufo). Behav. Ecol. 14, toms. BMC Mol. Biol. 20, 14.
103–108. Harju, K., Koskela, H., Kremp, A., Suikkanen, S., de la Iglesia, P., Miles, C.O., Krock, B.,
DeSa, R., Hastings, J.W., 1968. The characterization of scintillons. Bioluminescent par- Vanninen, P., 2016. Identification of gymnodimine D and presence of gymnodimine
ticles from the marine dinoflagellate, Gonyaulax polyedra. J. Gen. Physiol. 51, variants in the dinoflagellate Alexandrium ostenfeldii from the Baltic Sea. Toxicon 112
105–122. (Supplement C), 68–76.
Dominguez, H.J., Paz, B., Daranas, A.H., Norte, M., Franco, J.M., Fernandez, J.J., 2010. Hastings, J.W., 1978. Bacterial and dinoflagellate luminescent systems. In: Herring, P.J.
Dinoflagellate polyether within the yessotoxin, pectenotoxin, and okadaic acid toxin (Ed.), Bioluminescence in Action. Academic Press, London, pp. 129–170.
groups: characterization, analysis, and human health implications. Toxicon 56, Hastings, J.W., 1983. Biological diversity, chemical mechanisms, and the evolutionary
191–217. origins of bioluminescent systems. J. Mol. Evol. 19, 309–321.
Doucette, G.J., Anderson, D.M., 1993. Intracellular distribution of saxitoxin in Holmes, M.J., Lewis, R.J., Jones, A., Hoy, A.W., 1995. Cooliatoxin, the first toxin from
Alexandrium fundyense. In: Smayda, T.J., Shimizu, Y. (Eds.), Toxic Phytoplankton Coolia monotis (Dinophyceae). Nat. Toxins 3, 355–362.
Blooms in the Sea. Elsevier Science Publishers, pp. 863–868. Hoppenrath, M., Chomérat, N., Horiguchi, T., Schweikert, M., Nagahama, Y., Murray, S.,
Draisci, R., Ferretti, E., Palleschi, L., Marchiafava, C., Poletti, R., Milandri, A., Ceredi, A., 2013. Taxonomy and phylogeny of the benthic Prorocentrum species
Pompei, M., 1999. High levels of yessotoxin in mussels and presence of yessotoxin (Dinophyceae)—a proposal and review. Harmful Algae 27, 1–28.
and homoyessotoxin in dinoflagellates of the Adriatic Sea. Toxicon 37 (8), Howard, M.D.A., Silver, M., Kudela, R.M., 2008. Yessotoxin detected in mussel (Mytilus
1187–1193. californicus) and phytoplankton samples from the U.S.west coast. Harmful Algae 7,
Eckert, R., 1965. Bioelectric control of bioluminescence in the dinoflagellate Noctiluca. II. 646–652.
Asynchronous flash initiation by a propagated triggering potential. Science 147, Howard, M.D.A., Smith, G.J., Kudela, R.M., 2009. Phylogenetic relationships of yesso-
1142–1145. toxin-producing dinoflagellates, based on large subunit and internal transcribed
Eckert, R., Sibaoka, T., 1968. The flash-triggering action potential of the luminescent spacer ribosomal DNA domains. Appl. Environ. Microbiol. 75, 54–63.
dinoflagellate Noctiluca. J. Gen. Physiol. 52, 258–282. Iwamoto, M., Sumino, A., Shimada, E., Kinoshita, M., Matsumori, N., Oiki, S., 2017.
Esaias, W.E., Curl, H.C., 1972. Effect of dinoflagellate bioluminescence on copepod in- Channel formation and membrane deformation via sterol-aided polymorphism of
gestion rates. Limnol. Oceanogr. 17 (6), 901–905. amphidinol 3. Sci. Rep. 7, 10.
Finiguerra, M., Avery, D.E., Dam, H.G., 2014. No evidence for induction or selection of Jamison, T.F., Vilotijevic, I., 2007. Epoxide-opening cascades promoted by water. Science
mutant sodium channel expression in the copepod Acartia husdsonica challenged with 317, 1189–1192.
the toxic dinoflagellate Alexandrium fundyense. Ecol. Evol. 4 (17), 3470–3481. Janouškovec, J., Gavelis, G.S., Burki, F., Dinh, D., Bachvaroff, T.R., Gornik, S.G., Bright,
Finiguerra, M., Avery, D.E., Dam, H.G., 2015. Determining the advantages, costs, and K.J., Imanian, B., Strom, S.L., Delwiche, C.F., Waller, R.F., Fensome, R.A., Leander,
trade-offs of a novel sodium channel mutation in the copepod Acartia hudsonica to B.S., Rohwer, F.L., Saldarriaga, J.F., 2017. Major transitions in dinoflagellate evo-
Paralytic Shellfish Toxins (PST). PLoS One 10 (6), e0130097. lution unveiled by phylotranscriptomics. Proc. Natl. Acad. Sci. 114 (2), E171–E180.
Fleisher, K.J., Case, J.F., 1995. Cephalopod predation facilitated by dinoflagellate lumi- Jarne, P., Lagoda, P.J.L., 1996. Microsatellites, from molecules to populations and back.
nescence. Biol. Bull. 189 (3), 263–271. Trends Ecol. Evol. 11 (10), 424–429.
Fogel, M., Hastings, J.W., 1972. Bioluminescence: mechanism and mode of control of Jin, K., Klima, J.C., Deane, G., Dale Stokes, M., Latz, M.I., 2013. Pharmacological in-
scintillon activity. Proc. Natl. Acad. Sci. 69 (3), 690–693. vestigation of the bioluminescence signaling pathway of the dinoflagellate
Fraga, S., Rodríguez, F., Caillaud, A., Diogène, J., Raho, N., Zapata, M., 2011. Lingulodinium polyedrum: evidence for the role of stretch-activated ion channels. J.
Gambierdiscus excentricus sp. nov. (Dinophyceae), a benthic toxic dinoflagellate from Phycol. 49 (4), 733–745.
the Canary Islands (NE Atlantic Ocean). Harmful Algae 11, 10–22. Jingwen, L., Nianzhi, J., Huinnong, C., 2006. Cell cycle and cell signal transduction in
Frangoulos, M., Guisande, C., Maneiro, I., Riveiro, I., Franco, J., 2000. Short-term and marine phytoplankton. Prog. Nat. Sci. 16, 671–678.
long-term effects of the toxic dinoflagellate Alexandrium minutum on the copepod John, E.H., Flynn, K.J., 2000. Growth dynamics and toxicity of Alexandrium fundyense
Acartia clausi. Mar. Ecol. Prog. Ser. 203, 161–169. (Dinophyceae): the effect of changing N:p supply ratios on internal toxin and nutrient
Fritz, L., Morse, D., Hastings, J.W., 1990. The circadian bioluminescence rhythm of levels. Eur. J. Phycol. 35 (1), 11–23.
Gonyaulax is related to daily variations in the number of light emitting organelles. J. Johnson, C.H., Inoue, S., Flint, A., Hastings, J.W., 1985. Compartmentalization of algal
Cell Sci. 95, 321–328. bioluminescence - autofluorescence of bioluminescent particles in the dinoflagellate
Gacutan, R.Q., Tabbu, M.Y., Aujero, E.J., Icatlo, F., 1985. Paralytic shellfish poisoning Gonyaulax as studied with image-intensified video microscopy and flow cytometry. J.
due to Pyrodinium bahamense var compressa in Mati, Davao-Oriental. Philippines. Mar. Cell. Biol. 100 (5), 1435–1446.
Biol. 87 (3), 223–227. Kellmann, R., Michali, T.K., Neilan, B.A., 2008a. Identification of a saxitoxin biosynthesis
Glibert, P.M., Burkholder, J.M., Kana, T.M., 2012. Recent insights about relationships gene with a history of frequent horizontal gene transfers. J. Mol. Evol. 67 (5),
between nutrient availability, forms, and stoichiometry, and the distribution, eco- 526–538.
physiology, and food web effects of pelagic and benthic Prorocentrum species. Harmful Kellmann, R., Mihali, T.K., Jeon, Y.J., Pickford, R., Pomati, F., Neilan, B.A., 2008b.
Algae 14, 231–259. Biosynthetic intermediate analysis and functional homology reveal a saxitoxin gene
Gokhale, R.S., Dipika, T., 2001. Biochemistry of polyketides. In: Rehm, H.J., Reed, G. cluster in cyanobacteria. Appl. Environ. Microbiol. 74 (13), 4044–4053.
(Eds.), Biotechnology, 2nd ed. Wiley-VHC Verlag, Weinheim, Germany, pp. 341–372. Kellmann, R., Neilan, B.A., 2007. Biochemical characterization of paralytic shellfish toxin
Gomez, F., 2005. A list of free-living dinoflagellate species in the world's oceans. Acta Bot. biosynthesis in vitro. J. Phycol. 43 (3), 497–508.
Croat. 64, 129–212. Kellmann, R., Stuken, A., Orr, R.J.S., Svendsen, H.M., Jakobsen, K.S., 2010. Biosynthesis
Grebner, W., Berglund, E.C., Berggren, F., Eklund, J., Haroadottir, S., Andersson, M.X., and molecular genetics of polyketides in marine dinoflagellates. Mar. Drugs 8 (4),
Selander, E., 2019. Induction of defensive traits in marine plankton-new copepoda- 1011–1048.
mide structures. Limnol. Oceanogr. 64 (2), 820–831. Kimura, K., Okuda, S., Nakayama, K., Shikata, T., Takahashi, F., Yamaguchi, H., Skamoto,
Gribble, K.E., Keafer, B.A., Quilliam, M.A., Cembella, A.D., Kulis, D.M., Manahan, A., S., Yamaguchi, M., Tomaru, Y., 2015. RNA sequencing revealed numerous polyketide
Anderson, D.M., 2005. Distribution and toxicity of Alexandrium ostenfeldii synthase genes in the harmful dinoflagellate Karenia mikimotoi. PLoS One 10 (11), 16.
(Dinophyceae) in the Gulf of Maine, USA. Deep-Sea Res. II: Top. Stud. Oceanogr. 52, Knaust, R., Urbig, T., Li, L., Taylor, W., Hastings, J.W., 1998. The circadian rhythm of
2745–2763. bioluminescence in Pyrocystis is not due to differences in the amount of luciferase: a
Griffins, J.E., Park, G., Dam, H.G., 2019. Relative importance of nitrogen sources, algal comparative study of three bioluminescent dinoflagellates. J. Phycol. 34, 167–172.
alarm cues and grazer exposure to toxin production of the marine dinoflagellate Kohli, G.S., Campbell, K., John, U., Smith, K.F., Fraga, S., Rhodes, L.L., Murray, S.A.,
Alexandrium catenella. Harmful Algae 84, 181–187. 2017. Role of modular polyketide synthases in the production of polyether ladder
Grober, M.S., 1988a. Brittle-star bioluminescence functions as an aposematic signal to compounds in ciguatoxin-producing Gambierdiscus polynesiensis and G. excentricus
deter crustacean predators. Anim. Behav. 36, 493–501. (Dinophyceae). J. Euk. Microbiol. 64 (5), 691–706.
Grober, M.S., 1988b. Response of tropical reef fauna to brittle-star luminescence Kohli, G.S., John, U., Figueroa, R.I., Rhodes, L.L., Harwood, D.T., Groth, M., Bolch, C.J.S.,
(Echinodermata: ophiuroidea). J. Exp. Mar. Biol. Ecol. 115, 157–168. Murray, S.A., 2015. Polyketide synthesis genes associated with toxin production in
Hackett, J.D., Anderson, D.M., Erdner, D.L., Bhattacharya, D., 2004. Dinoflagellates: a two species of Gambierdiscus (Dinophyceae). BMC Genom. 16 (1), 410.
remarkable evolutionary experiment. Am. J. Bot. 91 (10), 1523–1534. Kremp, A., Tahvanainen, P., Litaker, W., Krock, B., Suikkanen, S., Leaw, C.-.P., Tomas, C.,
Hackett, J.D., Wisecarver, J.H., Brosnahan, M.L., Kulis, D.M., Anderson, D.M., 2014. Phylogenetic relationships, morphological variation and toxin patterns in the
Bhattacharya, D., Plumley, F.G., Erdner, D.L., 2013. Evolution of saxitoxin synthesis Alexandrium ostenfeldii (Dinophyceae) complex: implications for species boundaries

24
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

and identities. J. Phycol. 50, 81–100. toxigenic dinoflagellate Azadinium spinosum, with emphasis on polykeitde synthase
Landsberg, J.H., Hall, S., Johannessen, J.N., White, K.D., Conrad, S.M., Abbott, J.P., genes. BMC Genom. 16 (1), 27.
Flewelling, L.J., Richardson, R.W., Dickey, R.W., Jester, E.L.E., Etheridge, S.M., Mihali, T.K., Carmichael, W.W., Neilan, B.A., 2011. A putative gene cluster from a
Deeds, J.R., Van Dolah, F.M., Leighfield, T.A., Zou, Y.L., Beaudry, C.G., Benner, R.A., Lyngbya wollei bloom that encodes paralytic shellfish toxin biosynthesis. PLoS One 6
Rogers, P.L., Scott, P.S., Kawabata, K., Wolny, J.L., Steidinger, K.A., 2006. Saxitoxin (2), e14657.
puffer fish poisoning in the United States, with the first report of Pyrodinium baha- Mihali, T.K., Kellmann, R., Neilan, B.A., 2009. Characterisation of the paralytic shellfish
mense as the putative toxin source. Environ. Health Persp. 114 (10), 1502–1507. toxin biosynthesis gene clusters in Anabaena circinalis AWQC131C and
Latz, M.I., Bovard, M., VanDelinder, V., Segre, E., Rohr, J., Groisman, A., 2008. Aphanizomenon sp. NH-5. BMC Biochem. 10, 8.
Bioluminescent response of individual dinoflagellate cells to hydrodynamic stress Mojib, N., Kubanek, J., 2020. Comparative transcriptomics supports the presence of G
measured with millisecond resolution in a microfluidic device. J. Exp. Biol. 211 (17), protein-coupled receptor-based signaling in unicellular marine eukaryotes. Limnol.
2865–2875. Oceanogr. 65 (4), 762–774.
Latz, M.I., Jeong, H.J., 1996. Effect of red tide dinoflagellate diet and cannibalism on the Molgo, J., Marchot, P., Araoz, R., Benoit, E., Iorga, B.I., Zakarian, A., Taylor, P., Bourne,
bioluminescence of the heterotrophic dinoflagellates Protoperidinium spp. Mar. Ecol. Y., Servent, D., 2017. Cyclic imine toxins from dinoflagellates: a growing family of
Prog. Ser. 132, 275–285. potent antagonists of the nicotinic acetylcholine receptors. J. Neurochem. 142 (Suppl
Latz, M.I., Nauen, J.C., Rohr, J., 2004. Bioluminescence response of four species of di- 2), 41–51.
noflagellates to fully developed pipe flow. J. Plankton Res. 26, 1529–1546. Montojo, U.M., Sakamoto, S., Cayme, M.F., Gatdula, N.C., Furio, E.F., Relox, J.J.R.,
Le Tortorec, A.H., Tahvanainen, P., Kremp, A., Simis, S.G.H., 2016. Diversity of luciferase Shigeru, S., Fukuyo, Y., Kodama, M., 2006. Remarkable difference in accumulation of
sequences and bioluminescence production in Baltic Sea Alexandrium ostenfeldii. Eur. paralytic shellfish poisoning toxins among bivalve species exposed to Pyrodinium
J. Phycol. 51 (3), 317–327. bahamense var. compressum bloom in Masinloc bay, Philippines. Toxicon 48 (1),
Lee, T.C.H., Fong, F.L.Y., Ho, K.C., Lee, F.W.F., 2016. The mechanism of Diarrhetic 85–92.
Shellfish Poisoning toxin production in Prorocentrum spp.: physiological and mole- Morquecho, L., 2019. Pyrodinium bahamense one the most significant harmful dino-
cular perspectives. Toxins 8 (10), 23. flagellate in Mexico. Front. Mar. Sci. 6 (1), 1–8.
Lewitus, A.J., Horner, R.A., Caron, D.A., Garcia-Mendoza, E., Hickey, B.M., Hunter, M., Murray, S.A., Diwan, R., Orr, R.J.S., Kohli, G.S., John, U., 2015. Gene duplication, loss,
Huppert, D.D., Kudela, R.M., Langlois, G.W., Largier, J.L., Lessard, E.J., RaLonde, R., and selection in the evolution of saxitoxin biosynthesis in alveolates. Mol. Phylo.
Jack Rensel, J.E., Strutton, P.G., Trainer, V.L., Tweddle, J.F., 2012. Harmful algal Evol. 92, 165–180.
blooms along the North American west coast region: history, trends, causes, and Murray, S.A., Hoppenrath, M., Orr, R.J.S., Bolch, C., John, U., Diwan, R., Yauwenas, R.,
impacts. Harmful Algae 19, 133–159. Harwood, T., de Salas, M., Neilan, B., Hallegraeff, G., 2014. Alexandrium diversa-
Li, H., Miao, J., Cui, F., Li, G., 2008. Surfactant promotion of the inhibitory effects of porum sp. nov., a new non-saxitoxin producing species: phylogeny, morphology and
cupric glutamate on the dinoflagellate Alexandrium. J. Phycol. 44, 1364–1371. sxtA genes. Harmful Algae 31, 54–65.
Lim, P.-.T., Leaw, C.-.P., Usup, G., Kobiyama, A., Koike, K., Ogata, T., 2006. Effects of Murray, S.A., Mihali, T.K., Neilan, B.A., 2011a. Extraordinary conservation, gene loss,
light and temperature on growth, nitrate uptake, and toxin production of two tropical and positive selection in the evolution of an ancient neurotoxin. Mol. Biol. Evol. 28
dinoflagellates: alexandrium tamiyavanichii and Alexandrium minutum (3), 1173–1182.
(Dinophyceae). J. Phycol. 42 (4), 786–799. Murray, S.A., Wiese, M., Stuken, A., Brett, S., Kellmann, R., Hallegraeff, G., Neilan, B.A.,
Lim, P.-.T., Ogata, T., 2005. Salinity effect on growth and toxin production of four tropical 2011b. SxtA-based quantitative molecular assay to identify saxitoxin-producing
Alexandrium species (Dinophyceae). Toxicon 45 (6), 699–710. harmful algal blooms in marine waters. Appl. Environ. Microbiol. 77 (19),
Lindstrom, J., Grebner, W., Rigby, K., Selander, E., 2017a. Effects of predator lipids on 7050–7057.
dinoflagellate defence mechanisms - increased bioluminescence capacity. Sci. Rep. Nakanishi, K., 1985. The chemistry of brevetoxins: a review. Toxicon 23, 473–479.
7, 9. Nascimento, S.M., Mendes, M.C.Q., Menezes, M., Rodriguez, F., Alves-de-Souza, C.,
Lindstrom, J.B., Pierce, N.T., Latz, M.I., 2017b. Role of TRP channels in dinoflagellate Branco, S., Riobo, P., Franco, J., Nunes, J.M.C., Huk, M., Morris, S., Fraga, S., 2017.
mechanotransduction. Biol. Bull. 233 (2), 151–167. Morphology and phylogeny of Prorocentrum caipirignum sp. nov. (Dinophyceae), a
Lindström, L., Alatalo, R.V., Mappes, J., 1997. Imperfect Batesian mimicry – the effects of new tropical toxic benthic dinoflagellate. Harmful Algae 70, 73–89.
the frequency and the distastefulness of the model. Proc. R. Soc. Lond. B 264 (1379), Nicolas, M.T., Morse, D., Bassot, J.M., Hastings, J.W., 1991. Colocalization of luciferin
149–153. binding-protein and luciferase to the scintillons of Gonyaulax polyedra revealed by
Liu, L., Hastings, J.W., 2006. Novel and rapidly diverging intergenic sequences between double immunolabeling after fast-freeze fixation. Protoplasma 160 (2–3), 159–166.
tandem repeats of the luciferase genes in seven dinoflagellate species. J. Phycol. 42 Nosenko, T., Bhattacharya, D., 2007. Horizontal gene transfer in chromalveolates. BMC
(1), 96–103. Evol. Biol. 7 (1), 173.
Liu, L., Hastings, J.W., 2007. Two different domains of the luciferase gene in the het- Nuzzo, G., Cutignano, A., Sardo, A., Fontana, A., 2014. Antifungal amphidinol 18 and its
erotrophic dinoflagellate Noctiluca scintillans occur as two separate genes in photo- 7-sulfate derivative from the marine dinoflagellate Amphidinium carterae. J. Natl.
synthetic species. Proc. Natl. Acad. Sci. 104 (3), 696–701. Prod. 77 (6), 1524–1527.
Liu, L., Wilson, T., Hastings, J.W., 2004. Molecular evolution of dinoflagellate luciferases, Oikawa, H., Satomi, M., Watabe, S., Yano, Y., 2005. Accumulation and depuration rates of
enzymes with three catalytic domains in a single polypeptide. Proc. Natl Acad. Sci. paralytic shellfish poisoning toxins in the shore crab Telmessus acutidens by feeding
USA 101, 16555–16560. toxic mussels under laboratory controlled conditions. Toxicon 45, 163–169.
Llewellyn, L., Negri, A., Robertson, A., 2006. Paralytic shellfish toxins in tropical oceans. Okamoto, O.K., Liu, L., Robertson, D.L., Hastings, J.W., 2001. Members of a dinoflagellate
Toxin Rev. 25 (2), 159–196. luciferase gene family differ in synonymous substitution rates. Biochemistry 40,
Lu, Y., Wohlrab, S., Groth, M., Glöckner, G., Guillou, L., John, U., 2016. Transcriptomic 15862–15868.
profiling of Alexandrium fundyense during physical interaction with or exposure to Orr, R.J.S., Murray, S.A., Stüken, A., Rhodes, L., Jakobsen, K.S., 2012. When naked be-
chemical signals from the parasite Amoebophrya. Mol. Ecol. 25 (6), 1294–1307. came armored: an eight-gene phylogeny reveals monophyletic origin of theca in di-
Luo, Z., Zhang, H., Krock, B., Lu, S., Yang, W., Gu, H., 2017. Morphology, molecular noflagellates. PLoS One 7 (11), e50004.
phylogeny and okadaic acid production of epibenthic Prorocentrum (Dinophyceae) Orr, R.J.S., Stuken, A., Murray, S.A., Jakobsen, K.S., 2013a. Evolutionary acquisition and
species from the northern South China Sea. Algal Res. 22, 14–30. loss of saxitoxin biosynthesis in dinoflagellates: the second "core" gene, sxtG. Appl.
Luo, Z.H., Krock, B., Mertens, K.N., Price, A.M., Turner, R.E., Rabalais, N.N., Gu, H.F., Environ. Microbiol. 79 (7), 2128–2136.
2016. Morphology, molecular phylogeny and azaspiracid profile of Azadinium po- Orr, R.J.S., Stüken, A., Murray, S.A., Jakobsen, K.S., 2013b. Evolution and distribution of
porum (Dinophyceae) from the Gulf of Mexico. Harmful Algae 55, 56–65. saxitoxin biosynthesis in dinoflagellates. Mar. Drugs 11 (8), 2814–2828.
MacKenzie, L., White, D., Oshima, Y., Kapa, J., 1996. The resting cyst and toxicity of Pančić, M., Kiørboe, T., 2018. Phytoplankton defence mechanisms: traits and trade-offs.
Alexandrium ostenfeldii (Dinophyceae) in New Zealand. Phycologia 35, 148–155. Biol. Rev. 93 (2), 1269–1303.
Marcinko, C.L.J., Painter, S.C., Martin, A.P., Allen, J.T., 2013. A review of the mea- Parkhill, J.-.P., Cembella, A.D., 1999. Effects of salinity, light, and inorganic nitrogen on
surement and modelling of dinoflagellate bioluminescence. Prog. Oceanogr. 109, growth and toxigenicity of the marine dinoflagellate Alexandrium tamarense from
117–129. northeastern Canada. J. Plank. Res. 21, 939–955.
Marek, P., Papaj, D., Yeager, J., Molina, S., Moore, W., 2011. Bioluminescent aposema- Parsons, M.L., Aligizaki, K., Bottein, M.-Y.D., Fraga, S., Morton, S.L., Penna, A., Rhodes,
tism in millipedes. Curr. Biol. 21 (18), R680–R681. L., 2012. Gambierdiscus and Ostreopsis: reassessment of the state of knowledge of their
Martini, S., Haddock, S.H.D., 2017. Quantification of bioluminescence from the surface to taxonomy, geography, ecophysiology, and toxicology. Harmful Algae 14, 107–129.
the deep sea demonstrates its predominance as an ecological trait. Sci Rep 7, 11. Paul, G.K., Matsumori, N., Murata, M., Tachibana, K., 1995. Isolation and chemical
Menden-Deuer, S., Montalbano, A.L., 2015. Bloom formation potential in the harmful structire of amphidonl-2, a potent hemolytic compound derived from marine dino-
dinoflagellate Akashiwo sanguinea: clues from movement behaviors and growth flagellate Ampidinium klebsii. Tetra Lett. 36 (35), 6279–6282.
characteristics. Harmful Algae 47, 75–85. Pawlowiez, R., Morey, J.S., Darius, H.T., Chinain, M., Van Dolah, F.M., 2014.
Mendoza-Flores, A., Leyva-Valencia, I., Band-Schmidt, C.J., Galindo-Sánchez, C.E., Transcriptome sequencing reveals single domain Type I-like polyketide synthases in
Bustillos-Guzmán, J.J., 2018. Identification of the gene sxtA (domains sxtA1 and the toxic dinoflagellate Gambierdiscus polynesiensis. Harmful Algae 36, 29–37.
sxtA4) in Mexican strains of Gymnodinium catenatum (Dinophyceae) and their evo- Paz, B., Daranas, A.H., Norte, M., Riobo, P., Franco, J.M., Fernandez, J.J., 2008.
lution. Front. Mar. Sci. 5, 289. Yessotoxins, a group of marine polyether toxins: an overview. Mar. Drugs 6 (2),
Meng, Y.H., Van Wagoner, R.M., Misner, I., Tomas, C., Wright, J.L.C., 2010. Structure and 73–102.
Biosynthesis of Amphidinol 17, a Hemolytic Compound from Amphidinium carterae. Paz, B., Riobo, P., Fernandez, A.L., Fraga, S., Franco, J.M., 2004. Production and release
J. Nat. Prod. 73 (3), 409–415. of yessotoxins by the dinoflagellates Protoceratium reticulatum and Lingulodinium
Mensinger, A.F., Case, J.F., 1992. Dinoflagellate luminescence increases susceptibility of polyedrum in culture. Toxicon 44 (3), 251–258.
zooplankton to teleost predation. Mar. Biol. 112, 207–210. Peter, C., Krock, B., Cembella, A., 2018. Effects of salinity variation on growth and yes-
Meyer, J.M., Rödelsperger, C., Eichholz, K., Tillmann, U., Cembella, A., McGaughran, A., sotoxin composition in the marine dinoflagellate Lingulodinium polyedra from a
John, U., 2015. Transcriptomic characterisation and genomic glimps into the Skagerrak fjord system (western Sweden). Harmful Algae 78, 9–17.

25
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Pettay, D.T., LaJeunesse, T.C., 2013. Long-range dispersal and high-latitude environments Bioluminescent organisms and bioluminescence measurements in the North Atlantic
influence the population structure of a "stress-tolerant" dinoflagellate endosymbiont. Ocean near latitude 59.5°N, longitude 21°W. J. Geophys. Res. 100 (C4), 6527–6547.
PLoS One 8 (11), 12. Teegarden, G.J., 1999. Copepod grazing selection and particle discrimination on the basis
Phlips, E.J., Badylak, S., Christman, M., Wolny, J., Brame, J., Garland, J., Hall, L., Hart, J., of PSP toxin content. MEPS 181, 163–176.
Landsberg, J., Lasi, M., Lockwood, J., Paperno, R., Scheidt, D., Staples, A., Steidinger, Teegarden, G.J., Cembella, A.D., 1996. Grazing of toxic dinoflagellates, Alexandrium spp.
K., 2011. Scales of temporal and spatial variability in the distribution of harmful by adult copepods of coastal Maine: implications for the fate of paralytic shellfish
algae species in the Indian River Lagoon, Florida, USA. Harmful Algae 10, 277–290. toxins in marine food webs. J. Exp. Mar. Biol. Ecol. 196, 145–176.
Place, A.R., Bowers, H.A., Bachvaroff, T.R., Adolf, J.E., Deeds, J.R., Sheng, J., 2012. Teegarden, G.J., Cembella, A.D., Capuano, C.L., Barron, S.H., Durbin, E.G., 2003.
Karlodinium veneficum—the little dinoflagellate with a big bite. Harmful Algae 14, Phycotoxin accumulation in zooplankton feeding on Alexandrium fundyense–vector or
179–195. sink? J. Plankton Res. 25, 429–443.
Prevett, A., Lindstrom, J., Xu, J., Karlson, B., Selander, E., 2019. Grazer-induced biolu- Tillmann, U., Elbrächter, M., Krock, B., John, U., Cembella, A., 2009. Azadinium spinosum
minescence gives dinoflagellates a competitive edge. Curr. Biol. 29 (12), R564–r565. gen. et sp. nov. (Dinophyceae) identified as a primary producer of azaspiracid toxins.
Rademacher, C., Masepohl, B., 2012. Copper-responsive gene regulation in bacteria. Eur. J. Phycol. 44 (1), 63–79.
Microbiology 158 (Pt 10), 2451–2464. Tillmann, U., John, U., 2002. Toxic effects of Alexandrium spp. on heterotrophic dino-
Ramstad, H., Hovgaard, P., Yasumoto, T., Larsen, S., Aune, T., 2001. Monthly variations flagellates: an allelochemical defence mechanism independent of PSP-toxin content.
in diarrhetic toxins and yessotoxin in shellfish from coast to the inner part of the Mar. Ecol. Prog. Ser. 230, 47–59.
Sognefjord, Norway. Toxicon 39, 1035–1043. Tillmann, U., Kremp, A., Tahvanainen, P., Krock, B., 2014. Characterization of spirolide
Reguera, B., Riobo, P., Rodriguez, F., Diaz, P.A., Pizarro, G., Paz, B., Franco, J.M., Blanco, producing Alexandrium ostenfeldii (Dinophyceae) from the western Arctic. Harmful
J., 2014. Dinophysis toxins: causative organisms, distribution and fate in shellfish. Algae 39, 259–270.
Mar. Drugs 12 (1), 394–461. Tomas, C.R., van Wagoner, R., Tatters, A.O., White, K.D., Hall, S., Wright, J.L.C., 2012.
Reguera, B., Velo-Suárez, L., Raine, R., Park, M.G., 2012. Harmful Dinophysis species: a Alexandrium peruvianum (Balech and Mendiola) Balech and Tangen a new toxic
review. Harmful Algae 14, 87–106. species for coastal North Carolina. Harmful Algae 17, 54–63.
Rhodes, L., McNabb, P., de Salas, M., Briggs, L., Beuzenberg, V., Gladstone, M., 2005. Topalov, G., Kishi, Y., 2001. Chlorophyll catabolism leading to the skeleton of dino-
Yessotoxin production by Gonyaulax spinifera. Harmful Algae 5, 148–155. flagellate and Krill luciferins: hypothesis and model studies. Angew. Chem. Int. Ed.
Richlen, M.L., Erdner, D.L., McCauley, L.A.R., Libera, K., Anderson, D.M., 2012. Extensive Engl. 40, 3892–3894.
genetic diversity and rapid population differentiation during blooms of Alexandrium Touzet, N., Franco, J.M., Raine, R., 2007. Characterization of nontoxic and toxin-pro-
fundyense (Dinophyceae) in an isolated salt pond on Cape Cod, MA, USA. Ecol. Evol. ducing strains of Alexandrium minutum (Dinophyceae) in Irish Coastal Waters. Appl.
2 (10), 2588–2599. Environ. Microbiol. 73 (10), 3333–3342.
Robineau, B., Gagné, J.A., Fortier, L., Cembella, A.D., 1991. Potential impact of a toxic Tubaro, A., Dell'Ovo, V., Sosa, S., Florio, C., 2010. Yessotoxins: a toxicological overview.
dinoflagellate (Alexandrium excavatum) bloom on survival of fish and crustacean Toxicon 56, 163–172.
larvae. Mar. Biol. 108 (2), 293–301. Underwood, T.J., Tallamy, D.W., Pesek, J.D., 1997. Bioluminescence in firefly larvae: a
Rodriguez, J.D., Haq, S., Bachvaroff, T., Nowak, K.F., Nowak, S.J., Morgan, D., Cherny, test of the aposematic display hypothesis (Coleoptera: lampyridae). J. Ins. Behav. 10,
V.V., Sapp, M.M., Bernstein, S., Bolt, A., DeCoursey, T.E., Place, A.R., Smith, S.M.E., 365–370.
2017. Identification of a vacuolar proton channel that triggers the bioluminescent Usup, G., Ahmada, A., Matsuoka, K., Lim, P.T., Leaw, C.P., 2012. Biology, ecology and
flash in dinoflagellates. PLoS One 12 (2), 24. bloom dynamics of the toxic marine dinoflagellate Pyrodinium bahamense. Harmful
Rynearson, T.A., Virginia Armbrust, E., 2004. Genetic differentiation among populations Algae 14, 301–312.
of the planktonic marine diatim Ditylum brightwellii (Bacillariophyceae). J. Phycol. Usup, G., Kulis, D.M., Anderson, D.M., 1994. Growth and toxin production of the toxic
40 (1), 34–43. dinoflagellate Pyrodinium bahamense var. compressum in laboratory cultures. Nat.
Saravanan, V., Godhe, A., 2010. Genetic heterogeneity and physiological variation among Toxins 2 (5), 254–262.
seasonally separated clones of Skeletonema marinoi (Bacillariophyceae) in the Gullmar Valiadi, M., de Rond, T., Amorim, A., Gittins, J.R., Gubili, C., Moore, B.S., Iglesias-
Fjord, Sweden. Eur. J. Phycol. 45 (2), 177–190. Rodriguez, M.D., Latz, M.I., 2019. Molecular and biochemical basis for the loss of
Satake, M., MacKenzie, L., Yasumoto, T., 1997. Identification of Protoceratium reticulatum bioluminescence in the dinoflagellate Noctiluca scintillans along the west coast of the
as the biogenetic origin of yessotoxin. Nat. Toxins 5 (4), 164–167. U.S.A. Limnol. Oceanogr. 64 (6), 2709–2724.
Schantz, E., 1971. The dinoflagellate poisons. In: Kadis, S., Ciegler, A., Ajl, S.J. (Eds.), Valiadi, M., Iglesias-Rodriguez, M.D., 2014. Diversity of the luciferin binding protein gene
Microbial Toxins. Academic Press, New York, pp. 3–26. in bioluminescent dinoflagellates – insights from a new gene in Noctiluca scintillans
Schmidt, R.J., Gooch, V.D., Loeblich III, A.R., Hastings, J.W., 1978. Comparative study of and sequences from gonyaulacoid genera. J. Euk. Microbiol. 61 (2), 134–145.
luminescent and nonluminescent strains of Gonyaulax excavate (Pyrrhophyta). J. Valiadi, M., Painter, S.C., Allen, J.T., Balch, W.M., Iglesias-Rodriquez, D.M., 2014.
Phycol. 14 (1), 5–9. Molecular detection of bioluminescent dinoflagellates in surface waters of the pata-
Schmitter, R., Njus, D., Siulzman, F.M., Gooch, V., Hastings, J.W., 1976. Dinoflagellate gonian shelf during early Austral summer 2008. PLoS One 9, e98849.
bioluminescence: a comparative study of in vitro components. J. Cell. Physiol. 87, Van de Waal, D.B., Smith, V.H., Declerck, S.A., Stam, E.C., Elser, J.J., 2014.
123–134. Stoichiometric regulation of phytoplankton toxins. Ecol. Lett. 17 (6), 736–742.
Schultz, L.W., Liu, L., Cegielshi, M., Hastings, J.W., 2005. Crystal structure fo a pH- Verma, A., Barua, A., Ruvindy, R., Savela, H., Ajani, P.A., Murray, S.A., 2019. The Genetic
regulated luciferase catalyzing the bioluminescent oxidation of an open tetrapyrrole. basis of toxin biosynthesis in dinoflagellates. Microorganisms 7 (8), 29.
Proc. Natl Acad. Sci. USA 102, 1378–1383. Verret, F., Wheeler, G., Taylor, A.R., Farnham, G., Brownlee, C., 2010. Calcium channels
Selander, E., Berglund, E.C., Engstrom, P., Berggren, F., Eklund, J., Hardardottir, S., in photosynthetic eukaryotes: implications for evolution of calcium-based signalling.
Lundholm, N., Grebner, W., Andersson, M.X., 2019. Copepods drive large-scale trait- New Phytol. 187 (1), 23–43.
mediated effects in marine plankton. Sci. Adv. 5 (2), 6. von Dassow, P., Latz, M.I., 2002. The role of Ca(2+) in stimulated bioluminescence of the
Selander, E., Kubanek, J., Hamberg, M., Andersson, M.X., Cervin, G., Pavia, H., 2015. dinoflagellate Lingulodinium polyedrum. J. Exp. Biol. 205 (Pt 19), 2971–2986.
Predator lipids induce paralytic shellfish toxins in bloom-forming algae. Proc. Natl. Waldbauer, G.P., Sternbury, J.G., 1987. Experimental field demonstration that two
Acad. Sci. 112 (20), 6395–6400. aposematic butterfly color patterns do not confer protection against birds in Northern
Selander, E., Thor, P., Toth, G., Pavia, H., 2006. Copepods induce paralytic shellfish toxin Michigan. Am. Midl. Nat. 118, 145–152.
production in marine dinoflagellates. Proc. R. Soc. B – Biol. Sci. 273 (1594), Wang, D.-.Z., Li, C., Zhang, Y., Wang, Y.-.Y., He, Z.-.P., Lin, L., Hong, H.-.S., 2012.
1673–1680. Quantitative proteomic analysis of differentially expressed proteins in the toxicity-
Seliger, H.H., Biggley, W.H., Swift, E., 1969. Absolute values of photon emission from the lost mutant of Alexandrium catenella (Dinophyceae) in the exponential phase. J.
marine dinoflagellate Pyrodinium bahamense, Gonyaulax polyedra, and Pyrocystis lu- Proteomics 75 (18), 5564–5577.
nula. Photobiol. Photochem. 10, 227–232. White, A.W., 1986. High toxin content in the dinoflagellate Gonyaulax excavata in nature.
Shen, B., Wong, C.-.O., Lau, O.-.C., Woo, T., Bai, S., Huang, Y., Yao, X., 2015. Plasma Toxicon 24 (6), 605–610.
membrane mechanical stress activates TRPC5 channels. PLoS One 10 (4), e0122227. White, H.H., 1979. Effects of dinoflagellate bioluminescence on the ingestion rates of
Shimizu, Y., 1993. Microalgal metabolites. Chem. Rev. 93 (5), 1685–1698. herbivorous zooplankton. J. Exp. Mar. Biol. Ecol. 36, 217–224.
Shimizu, Y., 1996. Microalgal metabolites: a new perspective. Annu. Rev. Microbiol. 50, Whittaker, K.A., Rignanese, D.R., Olson, R.J., Rynearson, T.A., 2012. Molecular sub-
431–465. division of the marine diatom Thalassiosira rotulain relation to geographic distribu-
Slamovits, C.H., Keeling, P.J., 2008. Widespread recycling of processed cDNAs in dino- tion, genome size, and physiology. BMC Evol. Biol. 12 (1), 209.
flagellates. Curr. Biol. 18 (13), R550–R552. Widder, E.A., 2010. Bioluminescence in the ocean: origins of biological, chemical, and
Smith, S.M.E., Morgan, D., Musset, B., Cherny, V.V., Place, A.R., Hastings, J.W., ecological diversity. Science 328, 704–708.
DeCoursey, T.E., 2011. Voltage-gated proton channel in a dinoflagellate. Proc. Natl. Widder, E.A., Case, J.F., 1981a. Bioluminescence excitation in a dinoflagellate. In:
Acad. Sci. 108 (44), 18162–18167. Nealson, K.H. (Ed.), Bioluminescence Current Perspectives. Burgess Publishing,
Stoecker, D.K., Thessen, A.E., Gustafson, D.E., 2008. “Windows of opportunity” for di- Minneapolis, MN, pp. 125–132.
noflagellate blooms: reduced microzooplankton net growth coupled to eutrophica- Widder, E.A., Case, J.F., 1981b. Two flash forms in the bioluminescent dinoflagellate,
tion. Harmful Algae 8, 158–166. Pyrocystis fusiformis. J. Comp. Physiol. 143, 43–52.
Stuken, A., Orr, R.J.S., Kellmann, R., Murray, S.A., Neilan, B.A., Jakobsen, K.S., 2011. Widder, E.A., Case, J.F., 1982. Distribution of subcellular bioluminescent sources in a
Discovery of nuclear-encoded genes for the neurotoxin saxitoxin in dinoflagellates. dinoflagellate, Pyrocystis fusiformis. Biol. Bull. 162, 423–448.
PLoS One 6 (5), e20096. Widder, E.A., Case, J.F., Bernstein, S.A., MacIntyre, S., Lowenstine, M.R., Bowlby, M.R.,
Swift, E., Biggley, W.H., Seliger, H.H., 1973. Species of oceanic dinoflagellates in the Cook, D.P., 1993. A new large volume bioluminescence bathyphotometer with de-
genera Dissodinium and Pyrocystis: interclonal and interspecific comparisons of the fined turbulence excitation. Deep-Sea Res. I 40, 607–627.
color and photon yield of bioluminescence. J. Phycol. 9, 420–426. Widder, E.A., Johnsen, S., Bernstein, S.A., Case, J.F., Neilson, D.J., 1999. Thin layers of
Swift, E., Sullivan, J., Batchelder, H., Van Keuren, J., Vaillancourt, R., Bidigare, R., 1995. bioluminescent copepods found at density discontinuities in the water column. Mar.

26
K.D. Cusick and E.A. Widder Harmful Algae 98 (2020) 101850

Biol. 134 (3), 429–437. Wyatt, T., Jenkinson, I.R., 1997. Notes on Alexandrium population dynamics. J. Plankton
Widder, E.A., Latz, M.I., Case, J.F., 1983. Marine bioluminescence spectra measured with Res. 19 (5), 551–575.
an optical multichannel detection system. Biol. Bull. 165 (3), 791–810. Yang, I., Selander, E., Pavia, H., John, U., 2011. Grazer-induced toxin formation in di-
Wiese, M., D'Agostino, P.M., Mihali, T.K., Moffitt, M.C., Neilan, B.A., 2010. Neurotoxic noflagellates: a transcriptomic model study. Eur. J. Phycol. 46 (1), 66–73.
alkaloids: saxitoxin and its analogs. Mar. Drugs 8, 2185–2211. Yebra, D.M., Kiil, S., Dam-Johansen, K., 2004. Antifouling technology—past, present and
Wilson, T., Hastings, J.W., 1998. Bioluminescence. Ann. Rev. Cell Dev. Biol. 14, 197–230. future steps towards efficient and environmentally friendly antifouling coatings.
Wink, M., 2003. Evolution of secondary metabolites from an ecological and molecular Prog. Organic Coat. 50 (2), 75–104.
phylogenetic perspective. Phytochemistry 64, 3–19. Zhang, H., Hou, Y.B., Miranda, L., Campbell, D.A., Sturm, N.R., Gaasterland, T., Lin, S.J.,
Wisecaver, J.H., Brosnahan, M.L., Hackett, J.D., 2013. Horizontal gene transfer is a sig- 2007. Spliced leader RNA trans-splicing in dinoflagellates. Proc. Natl. Acad. Sci. USA
nificant driver of gene innovation in dinoflagellates. Genome Biol. Evo.l 5 (12), 104 (11), 4618–4623.
2368–2381. Zhang, S.F., Zhang, Y., Lin, L., Wang, D.Z., 2018. iTRAQ-based quantitative proteomic
Wohlrab, S., Iversen, M.H., John, U., 2010. A molecular and co-evolutionary context for analysis of a toxigenic dinoflagellate alexandrium catenella at different stages of
grazer induced toxin production in Alexandrium tamarense. PLoS One 5 (11), e15039. toxin biosynthesis during the cell cycle. Mar. Drugs 16 (12), 491.
Wolfe, G.V., 2000. The chemical defense ecology of marine unicellular plankton: con- Zimmer, R.K., Ferrer, R.P., 2007. Neuroecology, chemical defense, and the keystone
straints, mechanisms, and impacts. Biol. Bull. 198, 225–244. species concept. Biol. Bull. 213 (3), 208–225.
Wu, C., Akimoto, H., Ohmiya, Y., 2003. Tracer studies on dinoflagellate luciferin with Zingone, A., Oksfeldt Enevoldsen, H., 2000. The diversity of harmful algal blooms: a
[15N]-glycine and [15N]-glutamic acid in the dinoflagellate Pyrocystis lunula. challenge for science and management. Ocean Coast. Manag. 43 (8), 725–748.
Tetrahedron Lett. 44, 1263–1266.

27

You might also like