<<

J. Phycol. 49, 1118–1127 (2013) © Published 2013. This article is a U.S. Government work and is in the public in the U.S.A. DOI: 10.1111/jpy.12120

SUBCELLULAR LOCALIZATION OF POLYKETIDE SYNTHASES AND FATTY ACID SYNTHASE ACTIVITY1

Frances M. Van Dolah,2 Mackenzie L. Zippay Marine Biotoxins Program, NOAA Center for Coastal Environmental Health and Biomolecular Research, Charleston, South Carolina 29412, USA Marine Biomedical and Environmental Sciences, Medical University of South Carolina, Charleston, South Carolina 29412, USA Laura Pezzolesi Interdepartmental Research Centre for Environmental Science (CIRSA), University of Bologna, Ravenna 48123, Italy Kathleen S. Rein Department of Chemistry and Biochemistry, Florida International University, Miami, Florida 33199, USA Jillian G. Johnson Marine Biotoxins Program, NOAA Center for Coastal Environmental Health and Biomolecular Research, Charleston, South Carolina 29412, USA Marine Biomedical and Environmental Sciences, Medical University of South Carolina, Charleston, South Carolina 29412, USA Jeanine S. Morey, Zhihong Wang Marine Biotoxins Program, NOAA Center for Coastal Environmental Health and Biomolecular Research, Charleston, South Carolina 29412, USA and Rossella Pistocchi Interdepartmental Research Centre for Environmental Science (CIRSA), University of Bologna, Ravenna 48123, Italy

Dinoflagellates are prolific producers of polyketide related to fatty acid synthases (FAS), we sought to secondary metabolites. Dinoflagellate polyketide determine if fatty acid biosynthesis colocalizes with synthases (PKSs) have sequence similarity to Type I either chloroplast or cytosolic PKSs. [3H]acetate PKSs, megasynthases that encode all catalytic domains labeling showed fatty acids are synthesized in the on a single polypeptide. However, in dinoflagellate cytosol, with little incorporation in chloroplasts, PKSs identified to date, each catalytic domain resides consistent with a Type I FAS system. However, on a separate transcript, suggesting multiprotein although 29 sequences in a K. brevis expressed sequence tag database have similarity (BLASTx e- complexes similar to Type II PKSs. Here, we provide À evidence through coimmunoprecipitation that value <10 10) to PKSs, no transcripts for either Type I single-domain ketosynthase and ketoreductase (cytosolic) or Type II (chloroplast) FAS are present. interact, suggesting a predicted multiprotein Further characterization of the FAS complexes may complex. In brevis (C.C. Davis) Gert Hansen help to elucidate the functions of the PKS & Ø. Moestrup, previously observed chloroplast identified in dinoflagellates. localization of PKSs suggested that brevetoxin Key index words: brevetoxin; Coolia monotis; dinofla- biosynthesis may take place in the chloroplast. Here, gellates; fatty acid synthase; Karenia brevis; Ostreopsis we report that PKSs are present in both cytosol and cf. ovata; palytoxin; polyketide synthase chloroplast. Furthermore, brevetoxin is not present in isolated chloroplasts, raising the question of what List of abbreviations: ACP, acyl carrier ; chloroplast-localized PKS enzymes might be doing. AT, acyl transferase; CHAPS, 3[(3-cholamidopropyl) Antibodies to K. brevis PKSs recognize cytosolic and dimethylammonio]-propanesulfonic acid; CIB, chloroplast proteins in Ostreopsis cf. ovata Fukuyo, chloroplast isolation buffer; DAPI, 4′,6-diamidin- and Coolia monotis Meunier, which produce different 2-phenylindole; DH, dehydratase; DPM, distintegra- suites of polyketide toxins, suggesting that these PKSs tions per minute; EDTA, Ethylenediaminetetraacetic may share common pathways. Since PKSs are closely acid; EST, expressed sequence tag; FAME, fatty acid methyl ester; FAS, fatty acid synthase; IP, 1 immunoprecipitation; KR, ketoreductase; KS, ketoa- Received 22 May 2013. Accepted 25 August 2013. – 2Author for correspondence: e-mail [email protected]. cyl synthase; LC-MS, liquid chromatography mass Editorial Responsibility: C. Bowler (Associate Editor) spectrometry; LOD, limit of detection; MRM, multi-

1118 DINOFLAGELLATE POLYKETIDE SYNTHASES 1119 ple reaction monitoring; PBS, phosphate-buffered tain multiple modules on a single polypeptide, each saline; PKS, polyketide synthase; SIM, selected ion of which contains all active site domains needed to monitoring; TBS, tris-buffered saline; TBST, tris- carry out one chain extension, where each module buffered saline with tween; TE, thioesterase used only once during polyketide assembly. In con- trast, bacterial Type II PKSs are organized as com- plexes of smaller proteins where each catalytic Marine dinoflagellates produce some of the most domain is located on a separate peptide, a structure potent toxins on earth, and are responsible for similar to Type II FAS present in bacteria and chlo- more than 60,000 intoxication incidents per year on roplasts of higher plants and some algae. Type III a worldwide basis (Van Dolah 2000). The majority PKSs are small, structurally divergent proteins of dinoflagellate toxins that adversely affect human typified by plant chalcone synthases, and melanin health are polyether compounds, synthesized by producing PKSs in bacteria (Hopwood 1997, Funa complex enzymes known as PKS (Shimizu 2003). et al. 1999, Khosla et al. 1999, Gross et al. 2006). Polyketides are synthesized in a manner analogous PKSs have a patchy distribution among protist lin- to fatty acid biosynthesis through the sequential eages for which genomic sequence data are available addition of carboxylic acid building blocks: the con- (John et al. 2008): type I PKSs are present in chlo- densation reaction between carboxylic acids is per- rophytes (Chlamydomonas reinhardtii and Ostreococcus formed by a b-KS domain, whereupon the b-keto spp.) and the haptophyte Emiliania huxleyi, but are group may be fully or partially reduced by successive absent from stramenopiles (Thallassiosira pseudonan- activities of KR, DH, and enoyl reductase domains na, Phaeodactylum tricornutim, Phytophthora spp.) and following each chain elongation step (Staunton and Excavata (Trypanosoma spp., Leishmania major, Naegle- Weissman 2001). The growing carbon chain resides ria gruberi, Monosiga brevicollis). Transcriptome on a phosphopantetheine “arm” on the ACP that sequencing identified Type I PKSs in Prymnesio- “swings” it into close proximity of the catalytic sites, phytes, Chrysochromulina polylepis (John et al. 2010), while an AT brings additional “extender” carboxylic and Prymnesium parvum (Freitag et al. 2011). Among acid units to be added to the growing chain. The the closest relatives to dinoflagellates, Type I PKSs full-length polyketide is then released from the PKS are absent from the , thermophile, complex by a TE, and post-PKS modifications create but are variably present in the (pres- the final polyketide structure. ent in spp., , and PKSs are structurally and functionally similar to tenella, but absent from FASs and likely evolved from an ancestral condens- falciparum and parva). The Type I PKS in ing that served to make cellular functions Cryptosporidium parvum is a 40 kb, intronless gene more efficient (Hopwood 1997). The addition of an that encodes for a single 13,000 amino acid ACP and AT produced a primitive PKS, while fur- modular protein (Zhu et al. 2002). ther additions of a reductive cycle converted the Because of the large, complex genomes of dino- primitive PKS to an FAS. The fully reductive FAS flagellates (e.g. 1 9 1011 bp in K. brevis), the avail- became fixed in primary metabolism as fatty acids ability of genomic sequences has been limited to became essential components of the . PKSs sub- date. In the amphidinolide producer ,a sequently diverged, retaining variable numbers of predicted 5,625 bp PKS was found, containing sev- reductive catalytic domains to produce a variety of eral PKS domains (KS, AT, DH, KR, ACP, and TE; endproducts. Recent studies in plants suggest that Kubota et al. 2006). However, many unexplained the duplication of genes involved in primary metab- frame shifts and gaps were present in the sequence. olism is the primary source of new genetic material Similarly, Bachvaroff and Place (2008) found for secondary metabolism. Following duplication, numerous large introns within a genomic sequence even minor sequence changes in one copy can alter for a PKS KR domain protein in Amphidinium carte- substrate specificity, leading to new secondary rae that far exceeds the size of its coding region. metabolism products (Ober 2005). Since secondary Degenerate PCR identified partial sequences con- metabolites are not essential for survival, these taining type I PKS KS domains in seven dinoflagel- genes are more tolerant of mutations than their late , including two encoded by K. brevis counterparts in primary metabolism, further pro- (Snyder et al. 2003, 2005). Using a cDNA library moting the diversity observed in secondary metabo- screening approach, Monroe and Van Dolah (2008) lites (Weng et al. 2012). identified K. brevis transcripts with sequence similar- PKSs are typically categorized as Types I–III, ity to type I PKSs. However, each transcript con- although these divisions are becoming less clear as tained only a single catalytic domain, a structure more PKSs are characterized. Type I are large pro- more similar to type II PKSs. Homologous single- teins with multiple active sites on a single polypep- domain Type I-like PKS transcripts have now been tide. Iterative Type I PKSs, found in fungi, are reported in Alexandrium ostenfeldii (Eicholz et al. structurally similar to vertebrate (Type I) FAS, in 2012) and in the nontoxic Heterocapsa circularisqu- that both are multidomain proteins used repeatedly ama (Salcedo et al. 2012), and are present in public for each chain elongation. Modular type I PKSs con- databases of other dinoflagellate transcriptomes. 1120 FRANCES M. VAN DOLAH ET AL.

Á The K. brevis transcripts predicted KS and KR used in place of EDTA Na2 and FeCl3 6H2O. Cultures were maintained under illumination from cool white light at a domain proteins of between 50 and 100 kDa. À À photon flux density of 50–65 lmol photons Á m 2 Á s 1. Employing peptide polyclonal antibodies developed = to the in silico translated KS and KR sequences, Chloroplast isolation. For all species, 2 L of culture (n 3) was harvested at late log phase for chloroplast isolation. In Monroe et al. (2010) confirmed their predicted parallel, 1 L (n = 3) was harvested for whole cell protein sizes and found altered expression in a nontoxic iso- extraction. For K. brevis, cells were harvested by centrifuga- late of K. brevis suggesting their involvement in tion at 600g for 10 min, while the other species were centri- brevetoxin biosynthesis. Furthermore, their appar- fuged at 2,400g for 20 min. Cell pellets harvested for whole ent localization to the chloroplast suggested that cell extraction were stored at À80°C until extraction. brevetoxin biosynthesis may take place in that Chloroplasts were isolated by a modification of the method of Laatsch et al. (2004). Cell pellets were resuspended in CIB cellular compartment, consistent with previously consisting of 50 mM Tris, pH 7.5, 1 mM MgCl2,10mM reported chloroplast-associated okadaic acid (Zhou EDTA, 0.25% polyvinyl pyrrolidone, and 0.4 M mannitol and and Fritz 1994) and the thylakoid association of homogenized in a Dounce homogenizer on ice until chlorop- lasts free of nuclei were apparent by epifluorescent micro- microcystin (Kaebernick et al. 2000). Nonetheless, À evidence for their function in toxin biosynthesis scopic inspection of a DAPI (0.1 lg Á mL 1; Molecular remains limited to correlation. Probes, Carlsbad, CA, USA) stained sample. The homogenate was then layered onto a step gradient consisting of 4 mL The dinoflagellates included in the current study 2.2 M sucrose and 12 mL 1.6 M sucrose, then centrifuged at produce unique polyketide toxins. Brevetoxins, pro- 14,000g for 30 min. The plastid band, visible by its dark pig- duced by K. brevis, are ladder-like all-trans-fused mentation was removed using transfer pipets and an aliquot cyclic polyethers derived from two backbones was stained with DAPI and inspected by epifluorescence containing 49 (brevetoxin-A) or 50 (brevetoxin-B) microscopy to ensure the plastid band was free of nuclei or carbon atoms (Shimizu 2003). K. brevis also pro- unbroken cells. The chloroplast-containing fractions were duces several other polyether compounds, including diluted 1:6 in CIB, then harvested by centrifugation at 12,000g for 10 min at 4°C. The chloroplast pellets were either hemibrevetoxinB, brevisamide, brevesin, brevenal, frozen at À80°C or extracted immediately for protein as and tamulamides A and B (Baden et al. 2005, described below. Truxal et al. 2012). Ostreopsis cf. ovata is also a Protein isolation. Total protein was isolated from K. brevis, â prolific producer of secondary metabolites including O. cf. ovata, C. monotis cell pellets using TRI Reagent palytoxin (Ciminiello et al. 2010) and structurally (Molecular Research Center, Cincinnati, OH, USA) according related ovatoxins (Ciminiello et al. 2012, Pezzolesi to manufacturer’s instructions with some modifications. In brief, whole cells (or chloroplast pellets) were resuspended in et al. 2012). Coolia monotis produces cooliatoxin, a â 1 mL (or 0.5 mL) of TRI Reagent . K. brevis cells were lysed polyketide structurally related to yessotoxin and with by pipetting or gentle vortexing, whereas O. cf. ovata, C. mon- a molecular weight identical to monosulfated yesso- otis were sonicated on ice with short pulses of 15–20 s for a toxin, although its complete structure is still unchar- total of 2 min. Following addition of 0.2-mL chloroform (or acterized (Holmes et al. 1995). 0.1 mL), phases were separated by centrifugation (12,000g This study builds upon the previous insights into for 15 min at 4°C). RNA was discarded while the DNA was dinoflagellate PKSs gained in K. brevis. Employing precipitated with 100% ethanol and pelleted after centrifuga- tion at 5,000g for 5 min at 4°C. RNA and DNA fractions were the antibodies developed by Monroe et al. (2010), removed according to the manufacturer’s protocol. Proteins we sought to better understand where KS and KR were precipitated from the supernatant using acetone, domain proteins reside within the dinoflagellate washed 39 with 0.3 M guanidine hydrochloride in 95% etha- and whether they interact. We next used the K. bre- nol plus 2.5% glycerol, washed with 100% ethanol, and resus- vis KS and KR antibodies to investigate if PKSs in pended in protein sample buffer (30 mM Tris, pH 8.5, 2 M thiourea, 7 M urea, and 4% (w/v) CHAPS. Protein was ali- two other polyether-producing dinoflagellate species À ° similarly possess single-domain proteins. Finally, we quoted and stored at 20 C. Protein concentrations were determined using a Bradford (Bio-Rad, Hercules, CA, investigated potential additional roles for these USA) in a 96-well plate format (BD Biosciences, Franklin PKSs through the subcellular localization of fatty Lakes, NJ, USA). The colorimetric results were read at 595- acid synthesis activity and the analysis of putative nm wavelength on a FLUOstar Optima Microplate Fluorome- PKS and FAS sequences present in a K. brevis EST ter (BMG Labtech, Cary, NC, USA). database. Western blotting. Western blotting of whole cell, cytosol, or chloroplast protein (10 lg per lane) was carried out on NuPAGE Novex 4%–12% Bis-Tris gels (InvitrogenTM, Carls- MATERIALS AND METHODS bad, CA, USA). Following separation at 200 V for 45 min, the proteins were electrophoretically transferred onto prehydrat- Algal cultures. K. brevis (C.C. Davis) G. Hansen & ed polyvinylidene fluoride membrane (0.45 lm; Millipore, Ø. Moestrup (Wilson isolate), O. cf. ovata Fukuyo (toxic Bedford, MA, USA). Transfer was carried out at 100 V for 1 h strain OOAB0801), and C. monotis Meunier from the Adriatic with a transfer buffer containing 19 Tris/Glycine/sodium Sea (strain CMAB0801; toxicity uncharacterized) were used in dodecyl sulfate and 20% methanol (Bio-Rad). Complete this study. All isolates were grown in batch cultures in auto- transfer of proteins in the 30–100 kDa region was assessed by claved, 20-lm filtered seawater at a salinity of 36 obtained staining with Ponceau stain (Sigma-AldrichTM, St. Louis, MO, from Florida Institute of Technology field station in Vero USA). Following transfer, the membrane was blocked for 1 h Beach, FL, USA at 25 Æ 1°C on a 16:8 h light:dark cycle. Sea- at room temperature with 5% nonfat dried milk in TBS for water was enriched with modified f/2 medium with 0.01 mM O. cf. ovata and C. monotis, or TBS with 0.1% Tween-20 selenous acid (final concentration) and ferric sequestrene (TBST) for K. brevis. The membranes were then probed with DINOFLAGELLATE POLYKETIDE SYNTHASES 1121 polyclonal peptide antibodies against in silico translated K. bre- and –B); open A-ring PbTx-1, -2, -3, -7 (also called open A- vis KS (KB2006) and KR (KB5299) domain proteins (Monroe ring brevetoxin-A, open A-ring brevetoxin-B, open A-ring di- et al. 2010; Fig. S1 in the Supporting Information) or poly- hydrobrevetoxin-B, open A-ring dihydrobrevetoxin-A); and clonal anti-rubisco large subunit (RbcL; Cat# ABIN93539; open A-ring oxidized PbTx-1 and-2 (also called open A-ring Agrisera AB, Vannas, Sweden) overnight at 4°C in 5% nonfat oxidized brevetoxin-A and –B). Using SIM, the LOD was À À milk in TBST. For K. brevis, the following dilutions were used: ~1.2 ng Á mL 1 for PbTx-1, 0.8 ng Á mL 1 for PbTx-2, and À KS 1:7500 and KR 1:2500. For O. cf. ovata and C. monotis the 1.2 ng Á mL 1 for PbTx-3 with 10-lL injection; using MRM À following dilutions were used: KS 1:5000 and KR 1:500. Anti- method, LOD was ~0.9 ng Á mL 1 for PbTx-3 and À body specificity was determined by incubating the antibody 2ngÁ mL 1 for PbTx-1 and -2 with 5-lL injection. SIM Data with 100-fold excess, w/w, of the peptide used to generate were used for toxin semiquantitation as not all toxin congen- the antibody for 1 h prior to membrane immunoblotting at ers were available with toxin standards and MRM data were the same dilutions indicated above. Anti- (RbcL) was used at used for further toxin identity confirmation. Concentrations 1:500,000. The blots were washed three times for 10 min with of all congeners were summed and the summed values nor- TBST, incubated at room temperature for 1 h in secondary malized on a per-cell basis and expressed as combined pg antibody (ECL Anti- IgG Horseradish Peroxidase; brevetoxins per cell. Paired t-test was used to determine sig- NA934V; GE Healthcare, Piscataway, NJ, USA) that was nificant differences (P < 0.05) in brevetoxin concentrations diluted 1:2,000 in blocking solution, and then washed two between whole cell and chloroplast extracts. times for 10 min in TBST and a final wash in TBS for 5 min. Chlorophyll analysis. Chlorophyll a concentration was mea- Western blots were developed using West Pico chemilumines- sured on subsamples of the whole cultures and isolated chlo- cent substrate (Pierce, Rockford, IL, USA) and immunoreac- roplasts used for brevetoxin analysis to estimate the tive proteins were visualized on CL-XPosureTM film (Thermo chloroplast recovery. Chlorophyll analysis was performed by Scientific, Waltham, MA, USA). The developed film was collecting cells (50 mL of algal culture or 1/10 of chloroplast scanned and analyzed with a digital imaging software package, sample obtained from the subcellular fractionation) by centri- AlphaEase FC Version 3.1.2 (Alpha Innotech, San Leandro, fugation as above. Supernatants were removed and pellets CA, USA). stored at À80°C. For the extraction, 90% acetone was added Immunoprecipitation. Using an agarose support matrix to the samples and the extracts stored for 2–4hatÀ20°C. â (Pierce Co-IP kit #26149), the isolation of KR-KS protein The extracts were then centrifuged (4°C) and filtered. The complexes from K. brevis lysate was performed with the fol- samples were analyzed on an HP-1050 liquid chromatography lowing modifications to the manufacturer’s instructions equipped with photodiode array and fluorescence detectors (Thermo Scientific, Rockford, IL, USA): (i) the volume of as described by DiTullio and Geesey (2002). The gradient AminoLink Plus Coupling Resin was doubled to 100 lL, (ii) elution program was a modification of the ammonium ace- 80 micrograms of affinity-purified antibody was coupled to tate pairing method (Wright et al. 1991). The coefficient of the column for 90 min, (iii) 209 modified Dulbecco’s PBS variation from replicate standard injections was <3%. Peak buffer was added instead of the Lysis/Wash buffer during the spectra from each eluted peak were compared to the stored co-IP steps to minimize disruption by detergents, and (iv) library spectra to confirm peak identity and determine rela- À protease inhibitor (P8215; Sigma-AldrichTM) was added to the tive peak purity. Quantification in lg Á mL 1 was converted extracted lysate to minimize degradation. All IP steps were into pg per cell. performed at 4°C. A final volume of 40 lL was collected for Fatty acid synthesis assay. Log-phase cultures were harvested the elution and loaded onto a gel for immunodetection as by centrifugation (600g, 10 min) and the pellets were homog- described for Western blot (see above). enized on ice in 2 mL of FAS buffer (100 mM Tris-HCl, Brevetoxin analysis. Two liters of K. brevis Wilson isolate 2 mM EDTA, 1 mM MgCl2, 2 mM dithiothreitol, 0.4 mM (n = 6) was grown to midlog phase, from which one liter was mannitol, pH 7.8), per liter of culture, using a glass Dounce used for whole cell toxin extraction and 1 L was used for homogenizer (Cole-Parmer, Vernon Hills, IL, USA). The chloroplast brevetoxin extraction. Five-milliliter aliquots from homogenate was then centrifuged at 2500g for 10 min, and each L culture were fixed in 2% glutaraldehyde and analyzed the resulting “crude chloroplast” pellet was resuspended and for cell concentration using a Multisizer 3 (Beckman Coulter, subjected to sucrose gradient centrifugation as described Fullerton, CA, USA). Cells were then harvested by centrifuga- above for chloroplast isolation. The supernatant from the tion at 600g for 10 min and 1 mL of 100% methanol was 2500g centrifugation was subjected to centrifugation at added to extract brevetoxins. Similarly, 1 mL of methanol 100,000g for 1 h to produce a purified cytosol fraction. was added to pelleted chloroplasts after isolation using the Whole-cell homogenates, purified cytosol, and isolated chlo- procedure above. All samples were then filtered through a roplasts were transferred to 15-mL falcon centrifuge tubes À 0.22-lm Nanosep MF centrifuge filter with a bioinert mem- and [3H] acetate (4 lCi Á mL 1; PerkinElmer, San Jose, CA, À brane (Pall Life Sciences, Ann Arbor, MI, USA). LC-MS was USA), Na NADH (0.4 mg Á mL 1), Na NADPH Á 4H O 2 À 4 À 2 performed on an Agilent 1100 HPLC (Agilent, Palo Alto, CA, (0.4 mg Á mL 1), and malonyl-CoA (0.2 mg Á mL 1) were USA) coupled to an Applied Biosystems/MDS Sciex 4000 Q added. The mixture was incubated for 2–24 h at 25°C with TRAP hybrid triple quadruple/linear ion trap mass spectrom- continuous rotation, then saponified with the addition of eter equipped with a Turbo VTM source (AB ACIEX, previ- 2 mL 4 M NaOH and stored at 4°C overnight. After acidifi- ously as Applied Biosystems, Foster City, CA, USA). LC cation with concentrated HCl (pH <2, as indicated by pH separation was performed on a Luna C8 (2) (2.0 9 150 mm, paper), the assay mixture was extracted with hexane (1:1). 5 lm; Phenomenex, Torrence, CA, USA), with a stepped lin- Tubes were briefly centrifuged (1000g for 5 min), the hexane ear gradient of water (A)/acetonitrile (B) containing 0.1% layer retained, and the extraction repeated. The resulting formic acid: hold 50% B for 2 min, 80% B at 22 min, 95% B lipid extracts were dried under nitrogen, leaving a light col- at 23–29 min, then return to 50% B at 30 min, and hold ored residue. FAMEs were generated by either by the addi- 7 min to regenerate column. The detection of brevetoxin tion of diazomethane or 2-mL boron trifluoride (BF3, 10%) congeners was achieved by SIM or MRM in positive ion in methanol (Supelco, Belefonte, PA, USA) followed by mode. Brevetoxin congeners analyzed included PbTx-1, -2, -3, incubation at 100°C for 1 h. FAMEs were extracted in 1 mL -7, and -9 (also called brevetoxin-A, brevetoxin-B, dihydrob- hexane, washed with 3-mL saturated NaCl, and dried under revetoxin-B, dihydrobrevetoxin-A, tetrahydrobrevetoxin-B); nitrogen. A mixed fatty acid standard (Sigma-Aldrich) oxidized PbTx-1 and -2 (also called oxidized brevetoxin-A containing cis-9-oleate, arachidonate, behenate, linoleate, 1122 FRANCES M. VAN DOLAH ET AL. linolenate, myristate, palmitate, stearate, lignocerate, and A α-KS α-KR docosahexaenoic acid, was similarly esterified and used as both standard and carrier for silver nitrate thin layer chroma- 120 tography (TLC), by which FAMEs were separated according 100 KS to their degree of unsaturation. Silica gel Si250F TLC plates (J.T. Baker) were soaked in 10% silver nitrate (w/v) in ace- 80 tonitrile for 20 min and oven-dried 1 h at 80°C before use. l Dried K. brevis residues were redissolved in 40 L hexane 60 and 20 lL of mixed FAME standard was added as carrier. 50 Samples and standard were spotted on the prepared TLC Heavy plates and separated using a mixture of benzene: hexane Chain (1:1, v/v). TLC plates were stained with dichlorofluorescein 40 (0.2% in 95% ethanol) and visualized under UV light. Using the FAME standards/carrier to identify mobility, zones representing saturated, mono-, di-, and polyunsatu- rated fatty acids were scraped into scintillation vials, B α-KR α-KS and scintillation cocktail was added (Insta-Fluor Plus liquid 100 scintillation cocktail for organic samples, PerkinElmer, Waltham, MA, USA). Radioactivity incorporated into each fatty acid zone was quantified by liquid scintillation 80 counting. 60 Cerulenin inhibition assays. Cerulenin, an inhibitor of de novo fatty acid synthase activity, was used as an independent 50 Heavy confirmation of the incorporation of [3H] acetate into fatty Chain acids. Assays were conducted in the same manner as 40 described above, with or without addition of 250 lg of cerule- KR nin (Sigma Aldrich), and incubation on ice for 15 min prior to the addition of radiolabelled [3H]-acetate. Further process- ing of the cerulenin-containing assays was carried out as above. FIG. 1. Immunoprecipitation of ketoacyl synthase (KS) and ketoreductase (KR) containing protein complexes from Karenia brevis. (A) Western blots probed with anti-KS antibody: left lane – RESULTS whole cell lysate identifying the expected 100 kDa KS band; center lane – proteins immunoprecipitated with anti-KS antibody KS and KR domain proteins coprecipitate. The single shows the expected 100 kDa KS band; right lane – protein immu- domains present in K. brevis Type I-like PKS pro- noprecipitated with anti-KR antibody includes the 100 kDa KS teins suggest that they might function in multipro- band. (B) Western blots probed with anti-KR antibody: left lane whole cell lysate identifying the expected ~40 kDa KR band; cen- tein complexes analogous to Type II PKS. To ter lane – proteins immunoprecipitated with anti-KR antibody address this question we carried out IP of whole cell shows the expected ~40 kDa KR band; right lane protein immu- lysates of K. brevis with either the anti-KS or the noprecipitated with anti-KS antibody includes the ~40 kDa KR anti-KR antibody, then probed the immunoprecipi- band. tated protein by Western blotting with anti-KS to determine the presence of the KS domain (Fig. 1A) or anti-KR to determine the presence of the KR domain (Fig. 1B). Immunoprecipitation with either the anti-KS or anti-KR was able to recover the other protein from a whole-cell protein lysate, suggesting that they may interact in Type II-like multiprotein PKS complexes. KS domain proteins are present in both chloroplast and cytosol. Monroe et al. (2010) demonstrated that KS 100 kDa both KS and KR domain proteins are present in the chloroplasts of K. brevis. However, the anti-KS anti- 50 kDa body was found to be more immunoreactive with KR whole cell lysates than with isolated chloroplasts, 40 kDa suggesting the KS protein(s) may reside in more than one compartment. Here, we probed Western blots of whole cell lysates, purified cytosol, and iso- RbcL 50 kDa lated chloroplasts with the previously reported anti- KS and anti-KR antibodies. Anti-KS reactive bands at ~100 kDa are present in both purified cytosol and FIG. 2. Western blotting demonstrates that ketoacyl synthase chloroplast (Fig. 2). Two protein bands immunore- domain proteins are present in both chloroplast and cytosol of active with the anti-KR antibody, at ~41 and Karenia brevis. Ketoreductase (KR) domain proteins recognized by ~ this anti-KR antibody are present only in isolated chloroplasts. 48 kDa, are present only in the chloroplast Integrity of the chloroplasts is demonstrated by the absence of (Fig. 2). The 48 kDa band is more prominent in rubisco from purified cytosol. DINOFLAGELLATE POLYKETIDE SYNTHASES 1123 isolated chloroplasts, but is typically the minor band chloroplast or that the mature molecule is not pres- in whole-cell protein extracts. The presence of the ent in the chloroplast. highly expressed, plastid localized RbcL protein in Anti-K. brevis PKS antibodies recognize proteins in the isolated chloroplasts and its absence from puri- O. cf. ovata and C. monotis fied cytosol indicates that the preparation did not Western blotting of whole cell lysates and sucrose result in significant chloroplast breakage. gradient purified chloroplasts from both O. cf. ovata Brevetoxin is not located in the chloroplast. The chlo- and C. monotis identified ~98 kDa bands immunore- roplast localization of PKS proteins in K. brevis ini- active with the anti-K. brevis KS antibody (Fig. 4A). tially suggested that brevetoxin may be synthesized This molecular weight is similar to that found in in this compartment. To date, however, the intracel- K. brevis. The anti-KR antibody was cross-reactive with lular localization of brevetoxin has not been demon- proteins of about 90 kDa in the whole cell lysates of strated. To determine if brevetoxin is localized to both C. monotis and O. cf. ovata, substantially larger the chloroplast, we isolated intact chloroplasts from than the ~41–48 kDa bands found in K. brevis 2-L cultures of K. brevis (n = 6) by sucrose gradient (Fig. 4B). However, proteins cross-reactive with the centrifugation. In parallel, whole cells from 2-L cul- anti-KR antibody (at either ~90 or ~40–50 kDa) are tures of K. brevis (n = 6) were harvested by centrifu- absent from the chloroplasts. Both the KS and KR- gation. Both chloroplast and the whole cell pellets like bands are specific for the epitopes on K. brevis were extracted in methanol and brevetoxin was ana- KS and KR proteins (Fig. S1), as these bands disap- lyzed by LC-MS/MS. The total brevetoxin concentra- pear when the anti-KS or anti-KR antibodies are tion in the whole cell extracts of the Wilson isolate preincubated with an excess of the peptide À was 6.3 Æ 1.65 pg Á cell 1, within the normal range (Fig. S2 in the Supporting Information). of concentrations reported for this isolate (Fig. 3). Subcellular localization of fatty acid synthesis in K. bre- Chloroplasts isolated from an equal volume of cells vis. Given the similarity between PKS and FAS contained only 0.01 Æ 0.001 pg per cell. Chloro- sequences, the absence of brevetoxin from the plast recovery was 29.8 Æ 4.9%, as determined by chloroplast, and the presence of similar proteins in measuring chlorophyll a in whole cell lysates nonbrevetoxin producers, it is possible that these (4.9 Æ 0.96 pg per cell) and isolated chloroplasts PKS enzymes might instead be involved in fatty acid (1.45 Æ 0.41 pg per cell) from the same cultures. synthesis. The subcellular localization of fatty acid Thus, after correcting for chloroplast recovery, the biosynthesis has not been demonstrated in K. brevis, measured brevetoxin concentration in isolated chlo- or to our knowledge, in any photosynthetic dinofla- roplasts was 0.04 Æ 0.018 pg per cell, two orders of gellate. To address this question, we investigated the magnitude lower than the whole cell concentration incorporation of [3H] acetate into fatty acids in (Fig. 3). Similar results were obtained when mea- whole cell lysates, isolated chloroplasts, and purified sured by ELISA (data not shown). Since both LC- cytosol in the presence of substrate malonyl-CoA MS/MS and ELISA methods recognize the ladder and cofactors Na2NADH, Na4NADPH. Following polyether structures of mature brevetoxin molecules saponification and esterification, fatty acids were (or fragmentation thereof), these results suggest separated according to their degree of unsaturation either that brevetoxin is not synthesized in the on silver nitrate impregnated silica gel TLC plates. Bands representing saturated, monoene, diene, and polyunsaturated FAMEs were scraped from the 10.00 Whole Cell 8.00 Chloroplast 6.00 -1 4.00

. 2.00 0.10 * 0.08 0.06

pg Brevetoxin cell 0.04 0.02 0.00 Whole Cell Chloroplast

FIG. 3. Brevetoxin concentrations in pg∙per cell in whole cell FIG. 4. (A) Proteins recognized by the anti-Karenia brevis ketoa- extracts (white-open bar) and chloroplast extracts (gray-filled cyl synthase antibody are expressed in both whole cell and chlo- bar) of Karenia brevis isolated at midlog phase. Data are roplast protein of Ostreopsis cf ovata and Coolia monotis. (B) mean Æ SD, corrected for chloroplast recovery. Asterisk (*) signi- Protein bands recognized by the anti-K. brevis ketoreductase anti- fies statistical differences between whole cell and chloroplast body occur in whole cell lysates at ~90 kDa in both C. monotis (paired t-test; P = 0.002). and O. cf. ovata, but are absent from chloroplasts. 1124 FRANCES M. VAN DOLAH ET AL.

À plates and the distribution of incorporated [3H] was identify 29 as PKSs (BLASTx e-value <10 10; determined by scintillation counting. [3H] acetate Table S1 in the Supporting Information), 28 of incorporation into fatty acids in each of the cellular which have top blast hits to the K. brevis PKSs previ- fractions tested is summarized in Table 1A. The ously characterized by Monroe and Van Dolah purified cytosol (7730 DPM) accounted for almost (2008) or other dinoflagellate single-domain PKSs. 80% of the total [3H] acetate incorporated into fatty The remaining contig, with homology to Chlamydo- acids by whole cell lysates from 4.5 L of culture monas, contains the conserved domain smart00823, (10036 DPM). Isolated chloroplasts from twice as representing the phosphopantetheine attachment much starting material (9 L of culture) incorpo- site present in ACP. Seven contigs are annotated by rated only 336 DPM, representing 3.3% of the tri- Blast2GO as Type I FAS. However, the top (or sec- tium incorporated by 4.5 L whole cell homogenates. ondary) blast hits for four of the seven are previ- Even assuming ~25%–30% recovery of chloroplasts ously identified K. brevis PKS KS domain proteins. from the sucrose density gradient purification pro- These proteins contain the conserved domain cedure, based on chlorophyll a measurements cd00833 present in PKSs, rather than a conserved (described above), this would account for only ~6% domain specific for FAS. The remaining two contigs of the total [3H] incorporation into fatty acids. All have low similarity to Toxoplasma or Ostreococcus FAS, fractions showed a similar distribution of fatty acids, but these sequences contain only the phospho pant- in which 93%–94% of the label was associated with otheneine binding region of ACPs (smart00823). saturated fatty acids, ~4% PUFAs, and the remain- Therefore, it appears that among 23,000 contigs, no der monoenes and dienes (Table 1B). Given the Type I FAS KS domains exist that are unique from low incorporation by chloroplasts and profile similar the PKSs. In addition, no Type II FASs are present, to that in the cytosol, the incorporation observed in which is consistent with the absence of chloroplast- the chloroplast preparations may represent activity localized fatty acid synthesis. found in contaminating cytosol, rather than the chloroplasts themselves. The incorporation of [3H] acetate into fatty acids was inhibited by cerulenin, DISCUSSION an inhibitor of de novo FAS activity (Table 1C). Our current knowledge of the biosynthetic No fatty acid synthases are present in an extensive machinery responsible for dinoflagellate polyketide K. brevis EST library. The colocalization of PKS pro- toxin biosynthesis is rudimentary. Dinoflagellate teins with fatty acid synthase activity in the cytosol genes with homology to Type I PKSs have been prompted us to revisit our EST database (www.mari identified in a variety of both toxin-producing and negenomics.org) for contigs annotated as FASs to non-toxic dinoflagellate species through high- better characterize FASs in K. brevis. Among 23,000 throughput transcriptome sequencing projects unique contigs, automated Blast2GO annotations (Lidie et al. 2005, Bachvaroff and Place 2008, Murray et al. 2011, Salcedo et al. 2012, Eicholz et al. 3 TABLE 1. [ H] acetate incorporation into fatty acids by 2012). Single-domain PKS proteins, first described whole cell lysates, purified cytosol, and isolated chlorop- in K. brevis (Monroe and Van Dolah 2008, Monroe lasts. Data are from a representative experiment, with a et al. 2010), have been confirmed in Alexandrium minimum of three experiments performed for each cellu- ostenfeldii (Eicholz et al. 2012), O. cf. ovata, and lar fraction. C. monotis (this study) by Western blotting. However, the function of these proteins in toxin biosynthesis Gradient purified Whole cell lysate Purified cytosol chloroplasts has not been demonstrated. Although recent investi- gations show modest increases in brevetoxin cellular (A) Total 3H acetate incorporation into fatty acids (DPM) Saturated 9,354 7,310 312 quotas under nutrient stress (Hardison et al. 2012, Monoene 146 86 0 2013), consistent with a secondary metabolite, the Diene 122 48 8 PKSs are among a large set of transcripts downregu- PUFA 414 286 16 lated during stationary phase, which is consistent Total 10,036 7,730 336 with a role in primary, rather than secondary metab- (B) Percentage of 3H acetate incorporated into fatty acids of olism (Johnson et al. 2012). There were no differ- increasing desaturation ences observed in PKS transcript levels expressed in Saturated 93.2 94.6 92.9 a toxic isolate of K. brevis as compared to a nontoxic Monoene 1.5 1.1 0.0 Diene 1.2 0.6 2.4 isolate when analyzed by microarray or qPCR. How- PUFA 4.1 3.7 4.8 ever, the expression of KS proteins was lower in the non-toxic isolate than the toxic isolate, providing a 3 (C) Effect of cerulenin on incorporation of H acetate into clue that they may be involved in brevetoxin biosyn- fatty acids (DPM) thesis (Monroe et al. 2010). To gain further insight Activity (DPM) % of control into the function of these PKS proteins, the current study focused on their subcellular localization in Whole cell 7,234 100 K. brevis and in two distantly related dinoflagellate Whole cell + inhibitor 268 3.7 species, as well as the location of brevetoxin and DINOFLAGELLATE POLYKETIDE SYNTHASES 1125 fatty acid synthesis, an alternative or additional role endoplasmic reticulum for maturation into complex these proteins may play in dinoflagellates. lipids (Koo et al. 2004). On the other hand, the Mining the currently available K. brevis EST data- cyclization of polyketides is thought to occur con- base revealed 41 contigs whose top blast hit was PKS currently with their elongation (Keating and Walsh or FAS. All contigs that contained the KS conserved 1999). If this is the case for dinoflagellate polyke- domain (cd00833) had top blast hits to one of the tides, then the precursors could not be synthesized previously described K. brevis PKSs (Monroe and in the chloroplast. What might the chloroplast-local- Van Dolah 2008), regardless of their annotation as ized PKSs then be doing? One possibility is that PKS or FAS by Blast2GO. Three of these contigs these proteins are involved in fatty acid synthesis. contain the epitope against which the anti-KS anti- To begin to address this question we sought to body was made (Table S1 and Fig. S2), which is determine if fatty acid synthesis colocalizes with found toward the C-terminal end of the KS protein these PKSs. (KB2006), outside of the conserved KS domain. Fatty acid synthesis takes place in the cytosol of Therefore, it is possible that the anti-KS antibody animals and fungi by Type I FAS multidomain pro- recognizes different proteins in the chloroplast and teins (Wakil 1989). In contrast, fatty acid biosynthe- cytosol. In contrast, the anti-KR antibody cross- sis takes place in the chloroplast of plants by Type reacted only with chloroplast protein. From the II FAS complexes (Walker and Harwood 1985). Western blot data we cannot resolve if the anti-KR Like the higher plants, fatty acid synthesis is carried reactive bands at 41 and 48 kDa represent different out by Type II FAS in the chromist algae, cryptomo- proteins or reflect posttranslational modification in nads, heterokonts, and haptophytes, that share with the same protein. The 41 kDa band has similar dinoflagellates plastids of red algal origin (Ryall mobility to the band immunoprecipitated by the et al. 2003). Among the nearest relatives to dinofla- anti-KR antibody (Fig. 1). The nondenaturing lysis gellates, the apicomplexans Toxoplasma and Plasmo- conditions used for IP may account for the absence dium also contain Type II FASs (Waller et al. 1998). of the 48 kDa band seen under denaturing condi- These species possess relict plastids where fatty acid tions of the Western blot, for example, if the epi- synthesis likely takes place. In contrast, the apicom- tope is unavailable in the native 48 kDa protein due plexan Cryptosporidium parvum, which lacks a relict to differences in folding. Only one of four contigs plastid, possesses both a large Type I FAS (Zhu with homology to the KR domain protein (KB5299) et al. 2000) as well as a closely related Type I PKS possesses the epitope used to make the anti-KR anti- (Zhu et al. 2002). Thus, predictions based on near- body (Table S1 and Fig. S1). Therefore, it is likely est neighbors to the dinoflagellates may be unpro- that there are multiple KR proteins, and our cur- ductive. In the heterotrophic dinoflagellate rent results do not exclude the possibility of cyto- Crypthecodinium cohnii, a cytosolic FAS has been solic KR domains. The co-IP data indicate that characterized as a 410 kDa protein consisting of proteins recognized by these antibodies are in fact two subunits (Sonnenborn and Kunau 1982). This capable of interacting, since either antibody was enzyme produces only saturated fatty acids, primar- capable of coprecipitating the other protein. This ily palmitate (C16:0), consistent with a Type I FAS. supports the hypothesis that these single-domain Based on 13C labeling studies of C. cohnii cultures, Type I-like PKS proteins form multiprotein com- de Swaaf et al. (2003) proposed that three separate plexes like Type II PKS. What remains to be clari- systems are responsible for fatty acid synthesis: cyto- fied is what the PKSs in the chloroplast and cytosol solic synthesis of saturated fatty acids; the conver- might be doing. sion of saturated fatty acids to monounsaturated The absence of brevetoxin from the chloroplast fatty acids; and the de novo synthesis of 22:6 with- challenges the notion that brevetoxin is synthesized out monounsaturated fatty acid intermediates. in that compartment, although it is possible that Although C. cohnii is heterotrophic, it has been the linear polyketide precursors to brevetoxin are shown since that study to possess a remnant chloro- synthesized there and exported for further process- plast (Sanchez-Puerta et al. 2007), which could be ing to produce the cyclized, mature compound. the location of one of these processes. The hapto- Since neither our LC-MS/MS analysis nor the ELISA phyte Pavlova lutheri appears to have two distinct would recognize the linear precursors to brevetox- enzyme systems that produce polyunsaturated fatty ins, they could be present and undetected. Recent acids, one in the plastid and one extra-chloroplastic studies in O. cf. ovata using anti-palytoxin antibodies (Guiheneuf et al. 2011). The [3H] acetate incorpo- revealed the cytoplasmic localization of ovatoxins, ration by cell fractions in the current study indi- while chloroplasts were negative (Honsell et al. cates that K. brevis similarly carries out the synthesis 2011). No information is available on the localiza- of saturated fatty acids in the cytosol. The absence tion of cooliatoxin. Nonetheless, cytoplasmic locali- of Type I FASs that are unique from the PKSs, zation of the mature polyketide does not preclude together with the identification of KS domain pro- its synthesis in the chloroplast. Precedence for this tein(s) in the cytosol by Western blotting, raises the is found in higher plants, which synthesize fatty possibility that this synthesis may be carried out by acids in the chloroplast and export them to the protein complexes containing PKS-like KS domains. 1126 FRANCES M. VAN DOLAH ET AL.

Polyketide synthases have been shown to produce Eicholz, K., Beszteri, B. & John, U. 2012. Putative monofunctional de novo polyunsaturated fatty acids, independently Type I polyketide synthase units: a dinoflagellate-specific feature? PloSOne 7:e48624. of FAS, elongases or desaturases, in the marine bac- Freitag, M., Beszteri, S., Vogel, H. & John, U. 2011. Effects of teria Shewenella as well as the marine protist Schizo- physiological shock treatments on toxicity and polyketide chytrium, a member of the stramenopile lineage synthase in Prymnesium parvum (Prymnesio- (Metz et al. 2001). In Schizochytrium, this activity was phyceae) Eur. J. Phycol. 46:193–201. found in the cytosol (100 kg supernatant). In K. bre- Funa, N., Ohnishi, Y., Fujii, I., Shibuya, M., Ebizuka, Y. & Hori- ~ 3 nouchi, S. 1999. A new pathway for polyketide synthesis in vis, 4% of [ H] acetate was incorporated into PU- microorganisms. Nature 400:897–9. FAs. Whether these are derived through Gross, F., Luniak, N., Perlova, O., Gaitatzis, N., Jenke-Kodama, elongation/desaturation dependent processes or H., Gerth, K., Gottschalk, D., Dittmann, E. & Muller,€ R. independent of them cannot be resolved with cur- 2006. Bacterial type III polyketide synthases: phylogenetic 3 analysis and potential for the production of novel secondary rent data set. The absence of significant [ H] ace- metabolites by heterologous expression in pseudomonads. tate incorporation into fatty acids by chloroplasts, Arch. Microbiol. 185:28–38. coupled with the absence of Type II FAS sequences Guiheneuf, F., Ulmann, L., Termblin, G. & Mimount, V. 2011. in the transcriptome suggests that chloroplast-local- Light dependent utilization of two radiolabeled carbon ized PKS complexes carry out some other function. sources, sodium biocarbonate and sodium acetate, and relationships with long chain polyunsaturated fatty acidy Whether either the chloroplast or cytosolic com- synthesis in the microalga Pavlova lutheri (Haptophyta). Eur. plexes carry out polyketide toxin biosynthesis in J. Phycol. 46:143–52. K. brevis, O. cf. ovata,orC. monotis is currently Hardison, R. D., Sunda, W. G., Litaker, R. W., Shea, D. & Tester, undetermined. P. A. 2012. Nitrogen limitation increases brevetoxins in Karenia brevis (): implications for bloom toxicity. J. Phycol. 48:844–58. This work was funded by NOAA/NOS/NCCOS project Hardison, D. R., Sunda, W. G., Shea, D. & Litaker, R. W. 2013. #02E0025. MLZ was supported by a NOAA Oceans and Increased toxicity of Karenia brevis during phosphate limited Human Health Initiative postdoctoral traineeship through growth: ecological and evolutionary implications. PLOSOne 8: the Medical University of South Carolina. LP was supported e58545. by a fellowship from the Marco Polo funding programme Holmes, M. J., Lewis, J. L., Jones, A. & Wong Hoy, A. W. 1995. through The University of Bologna. Contributions from KR Cooliatoxin, the first toxin from Coolia monotis (Dinophy- were supported by NIEHS ARCH grant S11 ES0 11181 while ceae). Nat. Toxins 3:355–62. on sabbatical leave from FIU. We thank Jack DiTullio and Honsell, G., De Bortoli, M., Boscolo, S., Dell’Aversano, C., - Tyler Cyronak for assistance with chlorophyll analysis. tocchi, C., Fontanive, G., Penna, A. et al. 2011. Harmful dinoflagellate Ostreopsis cf. ovata Fukuyo: detection of ovatoxins in field samples and cell immunolocalization NOAA DISCLAIMER using antipalytoxin antibodies. Environ. Science Technol. 45:7051–9. This publication does not constitute an endorsement of Hopwood, D. A. 1997. Genetic ontributions to understanding – any commercial product or intend to be an opinion beyond polyketide synthases. Chem. Rev. 97:2465 98. scientific or other results obtained by the National Oceanic John, U., Beszteri, B., Derelle, E., Van de Peer, Y., Read, B., and Atmospheric Administration (NOAA). No reference shall Moreau, H. & Cembella, A. 2008. Novel insights into evolution of protistan polyketide synthases through phyloge- be made to NOAA, or this publication furnished by NOAA, – to any advertising or sales promotion which would indicate or nomic analysis. Protist 159:21 30. John, U., Beszteria, S., Glockner, G., Singh, R., Medlin, L. & Cem- imply that NOAA recommends or endorses any proprietary bella, A. 2010. Genomic characterisation of the ichthyotoxic product mentioned herein, or which has as its purpose an prymnesiophyte Chrysochromulina polylepis, and the expres- interest to cause the advertised product to be used or pur- sion of polyketide synthase genes in synchronized cultures. chased because of this publication. Eur. J. Phycol. 45:215–29. Johnson, J. G., Morey, J. S., Beal, M., Ryan, J. & Van Dolah, F. M. Bachvaroff, T. & Place, A. 2008. From stop to start: tangem gene 2012. Transcriptome remodeling associated with chronologi- rearrangement, copy number and trans-splicing in the dino- cal aging in the dinoflagellate, Karenia brevis. Mar. Genom. – flagellate, Amphidinium carterae. PloSOne 3:e2929. 5:15 25. Baden, D. G., Bourdelais, A. J., Jacocks, H., Michelizza, S. & Naar, J. Kaebernick, M., Neilan, B. A., Borner, T. & Dittmann, E. 2000. 2005. Natural and derivative brevetoxins: historical back- Light and the transcriptional response of the microcystin – ground, mutiplicity, and effects. Environ. Health Perspect. biosynthetic genes. Appl. Environ. Microbiol. 66:3387 92. 113:621–5. Keating, T. A. & Walsh, C. T. 1999. Initiation, elongation, and ter- Ciminiello, P., Dell’Aversano, C., Dello Iacovo, E., Fattorusso, E., mination in polyketide and polyketide antibiotic biosynthe- – Forino, M., Grauso, L., Tartaglione, L., Guerrini, F., Pezzole- sis. Curr. Opin. Chem. Biol. 3:598 606. si, L., Pistocchi, R. & Vanucci, S. 2012. Isolation and struc- Khosla, C., Gokhale, R. S., Jacobsen, J. R. & Cane, D. E. 1999. ture elucidation of ovatoxin-a, the major toxin produced by Tolerance and specificity of polyketide synthases. Annu. Rev. – Ostreopsis ovata. J. Am. Chem. Soc. 134:1869–75. Biochem. 68:219 53. Ciminiello, P., Dell’Aversano, C., Dello Iacovo, E., Fattorusso, E., Koo, A. J. K., Ohlrogge, J. B. & Pollard, M. 2004. On the export – Forino, M., Grauso, L., Tartaglione, L., Guerrini, F. & Pistoc- of fatty acids from the chloroplast. J. Biol. Chem. 279:16101 chi, R. 2010. Complex palytoxin-like profile of Ostreopsis ovat- 10. a. Identification of four new ovatoxins by high-resolution Kubota, T., Iinuma, Y. & Kobayashi, J. 2006. Cloning of polyke- liquid chromatography/mass spectrometry. Rapid Commun. tide synthase genes from amphidinolide-producing, dinofla- – Mass Spectrom. 24:2735–44. gellate Amphidinium sp. Biol. Pharm. Bull. 29:1314 8. DiTullio, G. R. & Geesey, M. E. 2002. Photosynthetic pigments in Laatsch, T., Zauner, S., Stoebe-Maier, B., Kowallik, K. V. & Maier, marine algae and bacteria. In Bitton, G. [Ed.] The Encyclope- U. G. 2004. Plastid-derived single gene minicircles of the dia of Environmental Microbiology. Wiley-Interscience, Hobo- dinoflagellate horridum are localized in the – ken, NJ, pp. 2453–70. nucleus. Mol. Biol. Evol. 21:1318 22. DINOFLAGELLATE POLYKETIDE SYNTHASES 1127

Lidie, K. B., Ryan, J. C., Barbier, M. & Van Dolah, F. M. 2005. Walker, K. A. & Harwood, J. L. 1985. Localization of chloroplastic Gene expression in Florida red tide dinoflagellate Karenia fatty acid synthesis de novo in the stroma. Biochem. J. brevis: analysis of an expressed sequence tag library and 226:551–6. development of DNA microarray. Mar. Biotechnol. 7:481–93. Waller, R. F., Keeling, P. J., Donald, R. G. & Al, E. 1998. Nuclear Metz, J. G., Roessler, P., Facciotti, D. & Al, E. 2001. Production of encoded proteins target to the plastid in Toxoplasma gondii polyunsaturated fatty acids by polyketide synthases in both and Plasmodium flaciparum. PNAS 95:12352–7. prokaryotes and . Science 293:290–3. Weng, J.-K., Phillipe, R. N. & Noel, J. P. 2012. The rise of Monroe, E. A., Johnson, J. G., Wang, Z. H., Pierce, R. K. & chemodiversity in plants. Science 336:1667–70. Van Dolah, F. M. 2010. Characterization and expression of Wright, S. W., Jeffrey, S. W., Mantoura, R. F. C., Llewellyn, C. A., nuclear-encoded polyketide synthases in the brevetoxin- Bjornland, T., Repeta, D. & Welschmeyer, N. 1991. Improved producing dinoflagellate Karenia brevis. J. Phycol. 46:541–52. HPLC method for the analysis of chlorophylls and carote- Monroe, E. A. & Van Dolah, F. M. 2008. The toxic dinoflagellate noids from marine phytoplankton. Mar. Ecol. Prog. Ser. Karenia brevis encodes novel type I-like polyketide synthases 77:183–96. containing discrete catalytic domains. Protist 159:471–82. Zhou, J. & Fritz, L. 1994. Okadaic acid antibody localizes to chlo- Murray, S. A., Garby, T., Hoppenrath, M. & Neilan, B. A. 2011. roplasts in the DSP-toxin-producing dinoflagellates Prorocen- Genetic diversity, morphological uniformity, and polyketide trum lima and Prorocentrum maculosum. Phycologia 33:455–61. production in dinoflagellates (Amphidinium, Dinoflagellata). Zhu, G., LaGier, M. J., Stejskal, F., Millership, J. J., Cai, X. M. & PloSOne 7:e38253. Keithly, J. S. 2002. Cryptosporidium parvum: the first protist Ober, D. 2005. Seeing double: gene duplication and diversification known to encode a putative polyketide synthase. Gene in plant secondary metabolism. Trends Plant Sci. 10:444–9. 298:79–89. Pezzolesi, L., Guerrini, F., Ciminiello, P., Dell’Aversano, C., Dello Zhu, G., Marchewka, M. J., Woods, K. M., Upton, S. J. & Keithly, Iacovo, E., Fattorusso, E., Forino, M., Tartaglione, L. & J. S. 2000. Molecular analysis of a Type I fatty acid synthase Pistocchi, R. 2012. Influence of temperature and salinity on in Cryptosporidium parvum. Mol. Biochem. Parasitol. 105:253–60. Ostreopsis cf. ovata growth and evaluation of toxin content through HR LC-MS and bioloigcal assays. Water Res. 46:82–92. Ryall, K., Harper, J. Y. & Keeling, P. J. 2003. Plastid-dervived type – II fatty acid iosynthetic enzymes in chromists. Gene 313:139 Supporting Information 48. Salcedo, T., Upadhyay, R. J., Nagasaki, K. & Bhattacharya, D. Additional Supporting Information may be 2012. Dozens of toxin-related genes are expressed in a non- toxic strain of the dinoflagellate Heterocapsa circularisquama. found in the online version of this article at the Mol. Biol. Evol. 29:1503–6. publisher’s web site: Sanchez-Puerta, M. V., Lippmeier, J. C., Apt, K. E. & Delwiche, C. F. 2007. Plastid genes in a non-photosynthetic dinoflagellate. Figure S1. K. brevis PKS protein sequences and Protist 158:105–17. location of peptides used for antibody production Shimizu, Y. 2003. Microalgal metabolites. Curr. Opin. Microbiol. (squares). A. KS domain protein KB2006, B. KR 6:236–43. domain protein KB5299. Snyder, R. V., Gibbs, P. D. L., Palacios, A., Abiy, L., Dickey, R., Lopez, J. V. & Rein, K. S. 2003. Polyketide synthase genes Figure S2. Western blot analysis of PKS proteins from marine dinoflagellates. Mar. Biotechnol. 5:1–12. in O. cf. ovata (O.o.) and C. monotis (C.m.) using: Snyder, R. V., Guerrero, M. A., Sinigalliano, C. D., Winshell, J., Perez, R., Lopez, J. V. & Rein, K. S. 2005. Localization of (A.) peptide polyclonal anti-KS against K. brevis polyketide synthase encoding genes to the toxic dinoflagel- KS domain protein KB2006 (left panel) and pep- late Karenia brevis. Phytochemistry 66:1767–80. tide-blocked anti-KS (right panel), demonstrating Sonnenborn, U. & Kunau, W.-H. 1982. Purification and proper- specificity of the ~98 kDa band; (B) anti-KR ties of the fatty acid synthetase complex from the marine against K. brevis KR domain protein KB2599 (left dinoflagellate, Crypthecodinium cohnii. Biochim. Biophys. Acta 712:523–34. panel) and peptide-blocked anti-KR (right panel) Staunton, J. & Weissman, K. J. 2001. Polyketide biosynthesis: a demonstrating specificity of the ~90 kDa band. millennium review. Nat. Prod. Rep. 18:380–416. de Swaaf, M. E., de Rijk, T. C., van der Meer, P., Eggink, G. & Table S1. K. brevis sequences annotated as PKS Sijtsma, L. 2003. Analysis of docosahexaenoic acid biosynthe- or FAS among a database of 65,000 ESTs that sis in Crypthecodinium cohnii by C-13 labelling and desaturase assemble into 23,000 contigs. KB2006, KB6736, inhibitor experiments. J. Biotechnol. 103:21–9. KB6380, KB4285, and KB1008 are K. brevis KS Truxal, L. T., Bourdelais, A. J., Jacocks, H., Abraham, W. M. & Baden, D. G. 2012. Characterization of tamulamides A and domain proteins previously described by Monroe B, polyethers isolated from the marine dinoflagellate Karenia and Van Dolah (2008). KB5299 is a KR domain brevis. J. Nat. Prod. 73:536–40. protein. Van Dolah, F. M. 2000. Marine algal toxins: origins, health effects, and their increased occurrence. Environ. Health Perspect. 108:133–41. Wakil, S. J. 1989. Fatty acid synthase, a proficient multifunctional enzyme. Biochemistry 28:4523–30.