12.07.2015 Views

Current Status and Historical Trends of Brown Tide and Red Tide ...

Current Status and Historical Trends of Brown Tide and Red Tide ...

Current Status and Historical Trends of Brown Tide and Red Tide ...

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Current</strong> <strong>Status</strong> <strong>and</strong> <strong>Historical</strong> <strong>Trends</strong><strong>of</strong> <strong>Brown</strong> <strong>Tide</strong> <strong>and</strong> <strong>Red</strong> <strong>Tide</strong>Phytoplankton Blooms in the CorpusChristi Bay National EstuaryProgram Study AreaCorpus Christi Bay National Estuary ProgramCCBNEP-07 • January 1996


This project has been funded in part by the United States EnvironmentalProtection Agency under assistance agreement #CE-9963-01-2 to the TexasNatural Resource Conservation Commission. The contents <strong>of</strong> this documentdo not necessarily represent the views <strong>of</strong> the United States EnvironmentalProtection Agency or the Texas Natural Resource ConservationCommission, nor do the contents <strong>of</strong> this document necessarily constitute theviews or policy <strong>of</strong> the Corpus Christi Bay National Estuary ProgramManagement Conference or its members. The information presented isintended to provide background information, including the pr<strong>of</strong>essionalopinion <strong>of</strong> the authors, for the Management Conference deliberations whiledrafting <strong>of</strong>ficial policy in the Comprehensive Conservation <strong>and</strong>Management Plan (CCMP). The mention <strong>of</strong> trade names or commercialproducts does not in any way constitute an endorsement or recommendationfor use.


<strong>Current</strong> <strong>Status</strong> <strong>and</strong> <strong>Historical</strong> <strong>Trends</strong> <strong>of</strong><strong>Brown</strong> <strong>Tide</strong> <strong>and</strong> <strong>Red</strong> <strong>Tide</strong> Phytoplankton Blooms in theCorpus Christi Bay National Estuary ProgramStudy AreaDr. Edward J. BuskeyPrincipal InvestigatorContributors:Mr. Scott StewartMr. Jay PetersonMr. Christopher CollumbMarine Science InstituteThe University <strong>of</strong> Texas at AustinP.O. Box 1267Port Aransas, Texas 78373512/749-6794512/749-6777 FAXPublication CCBNEP-07January 1996


Policy CommitteeCommissioner John BakerPolicy Committee ChairTexas Natural ResourceConservation CommissionMr. Ray AllenCoastal CitizenThe Honorable Vilma LunaTexas RepresentativeThe Honorable Josephine MillerCounty Judge, San Patricio CountyThe Honorable Mary RhodesMayor, City <strong>of</strong> Corpus ChristiMs. Jane SaginawPolicy Committee Vice-ChairRegional Administrator, EPA Region 6Commissioner John ClymerTexas Parks <strong>and</strong> Wildlife DepartmentCommissioner Garry MauroTexas General L<strong>and</strong> OfficeMr. Bernard PaulsonCoastal CitizenThe Honorable Carlos TruanTexas SenatorManagement CommitteeMr. Dean Robbins, Co-Chair Mr. William H. Hathaway, Co-ChairLocal Governments Advisory CommitteeMr. James Dodson, Chair Commissioner Gordon Porter, Vice-ChairScientific/Technical Advisory CommitteeDr. Terry Whitledge, Chair Dr. Wes Tunnell, Vice-ChairCitizens Advisory CommitteeMr. William Goldston, Co-Chair Mr. John Hendricks, Co-ChairFinancial Planning Advisory CommitteeDr. Joe Mosely, ChairProgram DirectorMr. Richard VolkTAMU-CC Campus Box 290 • 6300 Ocean Drive • Corpus Christi, TX 78412512/985-6767 • FAX 512/985-6301 • CCBNEP home page: //www.sci.tamucc.edu/ccbnepiii


Barry R. McBee, ChairmanR. B. Ralph Marquez, CommissionerJohn M. Baker, CommissionerDan Pearson, Executive DirectorAuthorization for use or reproduction <strong>of</strong> any original material contained inthis publication, i.e., not obtained from other sources, is freely granted. TheCommission would appreciate acknowledgment.Published <strong>and</strong> distributedby theTexas Natural Resource Conservation CommissionPost Office Box 13087Austin, Texas 78711-3087The TNRCC is an equal opportunity/affirmative action employer. The agency does not allow discrimination on the basis <strong>of</strong> race, color,religion, national origin, sex, disability, age, sexual orientation or veteran status. In compliance with the Americans with Disabilities Act,this document may be requested in alternate formats by contacting the TNRCC at (512) 239-0010, Fax 239-0055 or 1-800-RELAY-TX(TDD), or by writing P.O. Box 13087, Austin, TX 78711- 3087.iv


CORPUS CHRISTI BAY NATIONAL ESTUARY PROGRAMThe Corpus Christi Bay National Estuary Program (CCBNEP) is a four-year,community based effort to identify the problems facing the bays <strong>and</strong> estuaries <strong>of</strong> theCoastal Bend, <strong>and</strong> to develop a long-range, Comprehensive Conservation <strong>and</strong>Management Plan. The Program's fundamental purpose is to protect, restore, or enhancethe quality <strong>of</strong> water, sediments, <strong>and</strong> living resources found within the 600 square mileestuarine portion <strong>of</strong> the study area.The Coastal Bend bay system is one <strong>of</strong> 28 estuaries that have been designated as anEstuary <strong>of</strong> National Significance under a program established by the United StatesCongress through the Water Quality Act <strong>of</strong> 1987. This bay system was so designated in1992 because <strong>of</strong> its benefits to Texas <strong>and</strong> the nation. For example:• Corpus Christi Bay is the gateway to the nation's sixth largest port, <strong>and</strong> home to thethird largest refinery <strong>and</strong> petrochemical complex. The Port generates over $1 billion<strong>of</strong> revenue for related businesses, more than $60 million in state <strong>and</strong> local taxes, <strong>and</strong>more than 31,000 jobs for Coastal Bend residents.• The bays <strong>and</strong> estuaries are famous for their recreational <strong>and</strong> commercial fisheriesproduction. A study by Texas Agricultural Experiment Station in 1987 found thatthese industries, along with other recreational activities, contributed nearly $760million to the local economy, with a statewide impact <strong>of</strong> $1.3 billion, that year.• Of the approximately 100 estuaries around the nation, the Coastal Bend ranks fourthin agricultural acreage. Row crops -- cotton, sorghum, <strong>and</strong> corn -- <strong>and</strong> livestockgenerated $480 million in 1994 with a statewide economic impact <strong>of</strong> $1.6 billion.• There are over 2600 documented species <strong>of</strong> plants <strong>and</strong> animals in the Coastal Bend,including several species that are classified as endangered or threatened. Over 400bird species live in or pass through the region every year, making the Coastal Bendone <strong>of</strong> the premier bird watching spots in the world.The CCBNEP is gathering new <strong>and</strong> historical data to underst<strong>and</strong> environmental status<strong>and</strong> trends in the bay ecosystem, determine sources <strong>of</strong> pollution, causes <strong>of</strong> habitatdeclines <strong>and</strong> risks to human health, <strong>and</strong> to identify specific management actions to beimplemented over the course <strong>of</strong> several years. The 'priority issues' under investigationinclude:• altered freshwater inflow • degradation <strong>of</strong> water quality• declines in living resources • altered estuarine circulation• loss <strong>of</strong> wetl<strong>and</strong>s <strong>and</strong> other habitats • selected public health issues• bay debrisThe COASTAL BEND BAYS PLAN that will result from these efforts will be thebeginning <strong>of</strong> a well-coordinated <strong>and</strong> goal-directed future for this regional resource.v


STUDY AREA DESCRIPTIONThe CCBNEP study area includes three <strong>of</strong> the seven major estuary systems <strong>of</strong> the TexasGulf Coast. These estuaries, the Aransas, Corpus Christi, <strong>and</strong> Upper Laguna Madre areshallow <strong>and</strong> biologically productive. Although connected, the estuaries arebiogeographically distinct <strong>and</strong> increase in salinity from north to south. The LagunaMadre is unusual in being only one <strong>of</strong> three hypersaline lagoon systems in the world.The study area is bounded on its eastern edge by a series <strong>of</strong> barrier isl<strong>and</strong>s, including theworld's longest -- Padre Isl<strong>and</strong>.Recognizing that successful management <strong>of</strong> coastal waters requires an ecosystemsapproach <strong>and</strong> careful consideration <strong>of</strong> all sources <strong>of</strong> pollutants, the CCBNEP study areaincludes the 12 counties <strong>of</strong> the Coastal Bend: Refugio, Aransas, Nueces, San Patricio,Kleberg, Kenedy, Bee, Live Oak, McMullen, Duval, Jim Wells, <strong>and</strong> Brooks.This region is part <strong>of</strong> the Gulf Coast <strong>and</strong> South Texas Plain, which are characterized bygently sloping plains. Soils are generally clay to s<strong>and</strong>y loams. There are three majorrivers (Aransas, Mission, <strong>and</strong> Nueces), few natural lakes, <strong>and</strong> two reservoirs (LakeCorpus Christi <strong>and</strong> Choke Canyon Reservoir) in the region. The natural vegetation is amixture <strong>of</strong> coastal prairie <strong>and</strong> mesquite chaparral savanna. L<strong>and</strong> use is largely devoted torangel<strong>and</strong> (61%), with cropl<strong>and</strong> <strong>and</strong> pasturel<strong>and</strong> (27%) <strong>and</strong> other mixed uses (12%).The region is semi-arid with a subtropical climate (average annual rainfall varies from 25to 38 inches, <strong>and</strong> is highly variable from year to year). Summers are hot <strong>and</strong> humid,while winters are generally mild with occasional freezes. Hurricanes <strong>and</strong> tropical stormsperiodically affect the region.On the following page is a regional map showing the three bay systems that comprise theCCBNEP study area.vi


PrefaceThe following report has been an attempt to bring together all available information onharmful algal blooms, commonly referred to as brown <strong>and</strong> red tides, in the CorpusChristi Bay National Estuary Program (CCBNEP) area. Since the number <strong>of</strong> publishedstudies on this subject is very small, especially if only studies from within the CCBNEParea are included, unpublished data <strong>and</strong> data from non-peer reviewed technical reportsare also presented. This is especially true in the case <strong>of</strong> the brown tide report, sincemany studies are still ongoing <strong>and</strong> much essential data has not yet been written up forpublication. Unpublished data is clearly identified as such, <strong>and</strong> cited as a personalcommunication from the scientist providing the data. I want to thank the many scientistswho shared unpublished data with us for this report. Many scientists are reluctant toshare unpublished data, but few conclusions could have been drawn on the brown tidebased on the paucity <strong>of</strong> published data. I have attempted to check with those providingdata to ensure that I have interpreted them correctly, but any errors in interpretation inthis report are only my own.The organization <strong>of</strong> the red tide report is different from that <strong>of</strong> the brown tide report,since there were relatively few published reports or unpublished data from theCCBNEP study area because <strong>of</strong> the rarity <strong>of</strong> these events. In addition to presentinginformation available on blooms in this area, relevant generalities are presented basedon the more frequent occurrences <strong>of</strong> these blooms in other locations, especially thewest coast <strong>of</strong> Florida.Dr. Edward J. BuskeyProject LeaderNovember 28, 1995Port Aransas, Texasix


TABLE OF CONTENTSPagePREFACE..............................................................................................................xiEXECUTIVE SUMMARY........................................................................................1BROWN TIDEI. INTRODUCTION TO THE TEXAS BROWN TIDE...........................................3II. HISTORICAL TRENDS OF TEXAS BROWN TIDES.......................................4Causative Species......................................................................................4Comparison to the New Engl<strong>and</strong> <strong>Brown</strong> <strong>Tide</strong>.............................................5III. AVAILABLE DATA...........................................................................................7Environmental Conditions During <strong>and</strong> Immediately Precedingthe <strong>Brown</strong> <strong>Tide</strong>.................................................................................7<strong>Brown</strong> <strong>Tide</strong> Biomass.................................................................................20Spatial Distributions.................................................................................24Environmental Impact...............................................................................24IV. IDENTIFICATION OF PROBABLE CAUSES.............................................. 54V. IDENTIFICATION OF DATA AND INFORMATION GAPS...........................55VI. RECOMMENDATIONS FOR PRIORITY RESEARCH ANDMONITORING EFFORTS.............................................................................56Research..................................................................................................56Monitoring.................................................................................................57VII. LITERATURE CITED....................................................................................59RED TIDEVIII. INTRODUCTION TO HARMFUL RED TIDESIN TEXAS WATERS.....................................................................................66IX. HISTORICAL TRENDS OF TEXAS RED TIDE.............................................67Frequency................................................................................................67Duration....................................................................................................70Environmental Effects...............................................................................71Possible Causes.......................................................................................76X. AVAILABLE DATA.........................................................................................79Causative Species....................................................................................79Cell Concentrations/Other Biomass Estimates.........................................80Spatial <strong>and</strong> Temporal Bloom Distributions...............................................82Fish Kills...................................................................................................82Shellfish Beds...........................................................................................83Gulf <strong>of</strong> Mexico Circulation........................................................................84XI. POTENTIAL STATUS....................................................................................85x


TABLE OF CONTENTSPageXII. IDENTIFICATION OF DATA AND INFORMATION GAPS...........................87XIII. LITERATURE CITED...................................................................................89XIV. MATERIALS AND METHODS: UNPUBLISHED DATA...............................96XV. ANNOTATED BIBLIOGRAPHY....................................................................98XVI. ACKNOWLEDGMENTS............................................................................172xi


LIST OF TABLESBROWN TIDETablePage1. Comparison <strong>of</strong> <strong>Brown</strong> <strong>Tide</strong>s.............................................................................5RED TIDE1. Shellfish Toxicity............................................................................................692. Aerial Survey <strong>of</strong> G. breve...............................................................................823. Fish Kills.........................................................................................................834. Shellfish Fishery Closures.............................................................................84xii


LIST OF FIGURESBROWN TIDEFigurePage1. Map <strong>of</strong> the Estuaries <strong>of</strong> the Texas Coast........................................................32. Map <strong>of</strong> Upper Laguna Madre <strong>and</strong> Baffin Bay..................................................83. Map <strong>of</strong> the Transects Taken in the Nueces Esturine System.........................84. Distribution <strong>of</strong> <strong>Brown</strong> <strong>Tide</strong> in Baffin Bay..........................................................95. Water Temperature (Laguna Madre <strong>and</strong> East Flats)....................................106. Water Temperature <strong>and</strong> Salinity (Laguna Madre <strong>and</strong> Baffin Bay)................117. Water Temperature (CPL Plant)...................................................................128. Secchi Depths (Nueces Estuary <strong>and</strong> Laguna Madre)...................................149. Secchi Depths (Laguna Madre <strong>and</strong> Baffin Bay)............................................1410. PAR (UTMSI, East Flats <strong>and</strong> Laguna Madre)..............................................1511. Dissolved Oxygen (CPL Plant).....................................................................1612. Ammonium Concentrations..........................................................................1713. Nitrate Concentrations.................................................................................1714. Nitrite Concentrations..................................................................................1815. Phosphate Concentrations...........................................................................1916. Silica Concentrations...................................................................................2017. <strong>Brown</strong> <strong>Tide</strong> Cell Counts (1990-1991)...........................................................2118. <strong>Brown</strong> <strong>Tide</strong> Cell Counts (1994-1995)...........................................................2219. Chlorophyll a Concentration (Transects).....................................................2220. Chlorophyll a Concentration (Stations A, B, C, <strong>and</strong> D)...............................2321. Primary Production for Laguna Madre.........................................................2422. Seagrass Shoot Density..............................................................................2623. Seagrass Shoot Biomass.............................................................................2724. Seagrass Root/Rhizome Biomass...............................................................2825. Ratio <strong>of</strong> Root/Rhizome to Shoot Biomass....................................................2926. Area <strong>of</strong> Seagrass Loss.................................................................................3027. Seagrass Biomass at Various Water Depths...............................................3128. Mean Seagrass Biomass.............................................................................3129. Mesozooplankton Abundance <strong>and</strong> Ctenophore Volumes............................3330. Acartia tonsa Prosome Lengths...................................................................3331. Acartia tonsa Egg Release Rates................................................................3432. Acartia tonsa Gut Pigment Concentration....................................................3433. Microzooplankton <strong>and</strong> <strong>Brown</strong> <strong>Tide</strong> Concentration.......................................3534. Microzooplankton grazing............................................................................3635. Growth Rate <strong>of</strong> Strombidinopsis...................................................................3736. Growth Rate <strong>of</strong> Fabrea salina......................................................................3837. Growth Rate <strong>of</strong> Euplotes sp.........................................................................38xiii


LIST OF FIGURESFigurePage38. Growth Rate <strong>of</strong> Noctiluca scintillans.............................................................3939. Growth Rate <strong>of</strong> Oxyhrris marina...................................................................3940. Growth rate <strong>of</strong> Brachionus plicatilus.............................................................4041. Survival <strong>of</strong> Acartia tonsa nauplii...................................................................4142. Survival <strong>of</strong> Acartia tonsa nauplii...................................................................4143. Egg Release Rate <strong>of</strong> Acartia tonsa..............................................................4244. Survival <strong>of</strong> Spotted Seatrout........................................................................4545. Survival <strong>of</strong> <strong>Red</strong> Drum...................................................................................4646. Feeding Rates <strong>of</strong> Spotted Seatrout..............................................................4647. Spotted Seatrout Larval Densities...............................................................4748. Spotted Seatrout Larval Densities...............................................................4849. Black Drum Larval Densities........................................................................4850. Bay Anchovy Larval Densities......................................................................4951. Macrobenthos Biomass................................................................................5052. Macrobenthos Abundance...........................................................................5053. Macrobenthos Diversity................................................................................5154. Game Fish Abundance................................................................................5255. Forage Fish Abundance...............................................................................53RED TIDE1. Gymnodinium breve Blooms........................................................................692. G. breve, Temperature, <strong>and</strong> Salinity............................................................803. G. breve Concentrations..............................................................................81xiv


Executive SummaryHarmful algal blooms appear to be increasingly common phenomena on aworldwide scale. These blooms can be considered "harmful" either because <strong>of</strong> theirpotential threat to human health through the consumption <strong>of</strong> contaminated seafood, asin the case <strong>of</strong> many toxic phytoplankton blooms, or through the changes in speciesabundances <strong>and</strong> distributions (<strong>of</strong>ten including species <strong>of</strong> commercial value), as in thecase <strong>of</strong> recent "brown tide" blooms. In the Corpus Christi Bay National Estuary Programarea, harmful algal blooms have been relatively rare; there appear to have been fewerthan a dozen documented red tide blooms over the entire Texas coast <strong>and</strong> only onedocumented (although extremely persistent) brown tide bloom. This may be due in partto blooms going undocumented, especially in the years prior to World War Two whenthe cause <strong>and</strong> effect relationship between toxic din<strong>of</strong>lagellate blooms <strong>and</strong> fish <strong>and</strong>other marine life kills had not been established. In any event, harmful algal blooms arequite rare events <strong>and</strong> there is little evidence to suggest that they are increasing infrequency locally.Since January <strong>of</strong> 1990 the Laguna Madre has been experiencing a persistentbloom <strong>of</strong> small phytoplankton species generally referred to as the "brown tide". Thebrown tide began in January 1990 in Baffin Bay in an ecosystem that was alreadydisrupted by persistent high salinities that reduced the populations <strong>of</strong> planktonic <strong>and</strong>benthic grazers. Two severe freezes in December <strong>of</strong> 1989 caused massive fish kills.The decomposition <strong>of</strong> these fish released a large nutrient pulse that was sufficient t<strong>of</strong>uel the initial bloom <strong>of</strong> brown tide. This nearly monospecific bloom has been caused byhigh densities <strong>of</strong> a small (4-5 µm diameter) algal species that was previouslyundescribed. The proposed name for this new species is Aureomonas lagunensis.<strong>Brown</strong> tide cell concentrations have ranged from 0.5 - 6 million cells/ml throughout thecourse <strong>of</strong> the bloom, <strong>and</strong> chlorophyll a concentrations have reached as high as 120µg/l with a mean <strong>of</strong> about 60 µg/l.The brown tide has reduced the clarity <strong>of</strong> waters <strong>of</strong> the Laguna Madre, shadingout sea grass beds <strong>and</strong> disrupting sports-fishing activities. The biomass <strong>of</strong> roots <strong>and</strong>rhizomes in the seagrass beds has decreased dramatically in the past two years <strong>of</strong> thebrown tide, indicating the seagrasses are using up their energetic reserves. It isestimated that over 9 square kilometers <strong>of</strong> seagrass beds had been lost by the end <strong>of</strong>1994. The nearly 200 square miles <strong>of</strong> seagrass beds in Laguna Madre represent thelargest undisturbed seagrass habitat on the Texas coast <strong>and</strong> these seagrass beds arean important nursery habitat for fish <strong>and</strong> an essential winter food resource to migratingwaterfowl.Zooplankton are the major consumers <strong>of</strong> phytoplankton in most marinesystems, <strong>and</strong> it is puzzling that zooplankton populations have not increased during thisalgal bloom. Mesozooplankton populations declined at the beginning <strong>of</strong> the bloom <strong>and</strong>have remained low in areas impacted by the brown tide. The size <strong>of</strong> adult copepodswas lower <strong>and</strong> egg production rates were reduced in brown tide affected areas,1


indicating that brown tide was a poor food for these grazers. The brown tide has alsohad a dramatic affect on the benthic organisms <strong>of</strong> Laguna Madre; abundance, biomass<strong>and</strong> diversity <strong>of</strong> benthic fauna have all decreased.The extended brown tide bloom has had no apparent effect on populations <strong>of</strong>adult fish <strong>and</strong> shellfish. This is in contrast to the apparent effects <strong>of</strong> brown tide on larvalfish populations. Both laboratory <strong>and</strong> field studies suggest that the brown tide may betoxic to newly hatched larval fish <strong>and</strong> that larval fish populations are reduced in areasseverely impacted by the brown tide.Perhaps the most important unanswered question is: why has the brown tidepersisted for so long? Part <strong>of</strong> the answer lies in long turnover times for waters inLaguna Madre <strong>and</strong> Baffin Bay. The lack <strong>of</strong> freshwater inflow <strong>and</strong> restricted exchangewith the Gulf <strong>of</strong> Mexico make it difficult to disperse an algal bloom once it becomesestablished. It also seems likely that this algal species may be competitively superior toother species in the harsh conditions <strong>of</strong> the Laguna Madre. It is also possible that thisspecies suppresses other algal species through allelopathic effects. It also appears tohave negative effects on populations <strong>of</strong> grazers that might help bring the bloom undercontrol.Apart from brown tides, red tides comprise the other common harmful algalbloom in Texas coastal waters. Two species <strong>of</strong> din<strong>of</strong>lagellates are responsible for toxicred tides in Texas: Gymnodinium breve <strong>and</strong> Alex<strong>and</strong>rium monilata. The toxins <strong>of</strong> boththese din<strong>of</strong>lagellate species can cause extensive mortality in fish <strong>and</strong> invertebratepopulations. Only G. breve blooms have been reported to cause human healthproblems through neurotoxic shellfish poisoning <strong>and</strong> respiratory irritation associatedtoxin containing aerosols. <strong>Red</strong> tides have been infrequent events on the Texas coastcompared to other areas <strong>of</strong> the Gulf <strong>of</strong> Mexico, such as the West coast <strong>of</strong> Florida.There have been only 4 documented G. breve <strong>and</strong> 7 A. monilata red tide blooms.Although occasionally the economic impacts <strong>of</strong> red tides are very severe, suchas in the 1986 red tide, they appear to have no long term affects on the ecosystem, asmay be the case with the brown tide. Fish populations, <strong>and</strong> tourism, soon return tonormally following the infrequent blooms, <strong>and</strong> no basic changes in marine policy seemwarrented to deal with them. Should toxic blooms in Texas increase in frequency <strong>and</strong>/orseverity, however, monitoring efforts should be considered to provide warning <strong>of</strong> bloominitiation <strong>and</strong> transport along the Texas coast.2


II. <strong>Historical</strong> trends <strong>of</strong> Texas brown tidesSince the algal species responsible for the Texas brown tide bloom has not beenpreviously described taxonomically, it is difficult to know with certainty if there havebeen previous occurrences <strong>of</strong> the Texas brown tide. Certainly algal blooms haveoccurred in the past in the Laguna Madre, as in any body <strong>of</strong> water. Long time residents<strong>of</strong> south Texas recall periods <strong>of</strong> "brown water" in the Laguna Madre, but it is impossibleto know if the causative organism was the same one present in the current brown tide.Gunter (1945) cites reports <strong>of</strong> “reddish colored ‘bad’ water”, observed by pilots for theTexas Game, Fish <strong>and</strong> Oyster Commission, in Laguna Madre, <strong>and</strong> Simmions (1957)refers to red waters in the Laguna Madre as possibly being caused by: 1) dissolvediron 2) decaying vegetation 3) phytoplankton or 4) clay deposits. None <strong>of</strong> the previousreports seem to compare to the present bloom in either duration or severity.Examination <strong>of</strong> the stable carbon isotope ratios <strong>of</strong> organic carbon from sedimentcores collected in Baffin Bay suggest that the relative importance <strong>of</strong> seagrasses,macrophytic algae <strong>and</strong> phytoplankton may have shifted for decade long periods overthe last several thous<strong>and</strong> years (Parker <strong>and</strong> Scalan, personal communication). Thisalone is not pro<strong>of</strong> that the brown tide organism dominated the Laguna Madre at anypoint in the past, but it does suggest that the relative importance <strong>of</strong> phytoplankton <strong>and</strong>seagrasses as primary producers in Baffin Bay have been different than the patternexisting in historical periods before the brown tide. To determine if the brown tide werethe organism responsible for phytoplankton dominance during these periods,biomarkers specific to the brown tide alga would be needed, <strong>and</strong> cores could beexamined for these compounds. This would require study <strong>of</strong> the diagenesis <strong>of</strong> browntide derived organic material.A. Causative speciesThe organism responsible for this bloom is a small (4-5 µm diameter) phytoplanktonspecies which has not yet been formally classified. Attempts to formally describe thisspecies have recently been completed <strong>and</strong> a species description has been submittedfor publication. The proposed name for this new species is Aureomonas lagunensis (H.DeYoe, personal communication). The Texas "brown tide" alga is devoid <strong>of</strong> obviousexternal diagnostic features when observed by optical microscopy, but examination <strong>of</strong>ultrastructure by transmission electron microscopy <strong>and</strong> <strong>of</strong> photosynthetic pigments byhigh performance liquid chromatography reveal that it is similar in morphology <strong>and</strong>pigments to Aureococcus anophagefferens (Sieburth et al., 1988; Stockwell et al.,1993) which has been responsible for recurrent blooms in Narragansett Bay <strong>and</strong> LongIsl<strong>and</strong> bays since 1985 (Cosper et al., 1987). Recent molecular data indicate that bothspecies belong to a newly recognized class Pelagophyceae (Anderson et al., 1993),but the two species appear to differ enough from one another to warrant placing themin separate genera (DeYoe <strong>and</strong> Suttle, 1995).4


Organism:AureococcusanophageffernsTexas "browntide"Cell size (µm) 2 -3 4 - 5PigmentsType III Chrysophyte Type III ChrysophyteExternal Polysaccaride layer + +Cell wall None NoneViral inclusions + +DMS/cell 0.13pg 0.76pgPolyclonal antibodyfor Aureococcus + -Bloom CharacteristicsCell density (cells x 10 6 /ml) 0.5 - 2.6 0.5 - 5.0Chlorophyll a (µg/L) 18 20 - 140C 14 uptake (mgC/m 2 /h) 200 - 400 230 - 840Growth rates(doublings/day) 3 0.3 - 0.9Carbon:Chlorophyll ? 380:1Maximum duration 6 months 72 monthsReoccurence frequency 3 - 5 years ?Deleterious effects Yes* Yes*** Mussels, clams, scallops, eelgrass, zooplankton** Zooplankton, larval fish, seagrassTable 1. Comparison <strong>of</strong> the Texas "brown tide" organism with Aureococcusanophagefferens (adapted from Table in Stockwell et al., 1993).B. Comparison to the East Coast brown tideAlgal blooms <strong>of</strong> Aureococcus anophagefferens, a small (2 µm diameter)pelagiophyte, are also referred to as the East Coast brown tide. These blooms haveoccurred in several bays in the Northeastern United States including Narragansett Bay,RI, the Peconic-Gardiners Bay system, the south shore <strong>of</strong> Long Isl<strong>and</strong>, NY <strong>and</strong>Barnegat Bay, NJ (Lonsdale et al., in prep). These brown tides are restricted toshallow, vertically well-mixed waters, <strong>and</strong> have not occurred in Long Isl<strong>and</strong> Sound(Olsen, 1989; Nuzzi <strong>and</strong> Waters, 1989). The first bloom occurred in the summer <strong>of</strong>1985, lasting for up to six months in some locations. Widespread blooms recurred in1986 <strong>and</strong> 1987, <strong>and</strong> on a much more localized scale <strong>and</strong> <strong>of</strong> shorter duration in recentyears. During blooms <strong>of</strong> the East Coast brown tide, total phytoplankton biomass <strong>and</strong>primary production rates are similar to those in non-bloom years (Cosper et al., 1989;Dennison et al., 1989). The demise <strong>of</strong> brown tide blooms is <strong>of</strong>ten associated withreplacement <strong>of</strong> the brown tide with other pico- <strong>and</strong> nanoplankton species present in lowconcentrations during the bloom (Sieburth <strong>and</strong> Johnson, 1989; Keller <strong>and</strong> Rice, 1989).There is no evidence that A. anophagefferens blooms are initiated from cysts, since low5


numbers <strong>of</strong> cells are found in affected waters throughout the year (Lonsdale et al., inprep).The exact causes <strong>of</strong> blooms <strong>of</strong> the East Coast brown tide remain unknown.Aureococcus has the ability to adapt to decreasing light levels as a bloom progresses<strong>and</strong> to maintain high growth efficiency under nutrient limiting conditions (Milligan,1992). A. anophagefferens is also tolerant <strong>of</strong> a wide range <strong>of</strong> temperature <strong>and</strong> salinityconditions. It grows best at higher temperatures found in East Coast waters in summer(20-25 o C) but can also grow at a slower rate at temperatures as low as 5 o C, allowingseed populations to be maintained through the winter (Cosper et al., 1989b).Physical factors that might contribute to the East Coast brown tide include longresidence times <strong>of</strong> bay waters <strong>and</strong> reduced rainfall (Cosper et al., 1989b). Thereappears to be a correlation between brown tide distribution <strong>and</strong> areas <strong>of</strong> elevatedsalinity. Growth rates <strong>of</strong> Aureococcus anophagefferens are faster at 30 ppt than 28 pptsalinity, <strong>and</strong> growth can be maintained at salinities between 22 <strong>and</strong> 35 ppt (Cosper etal., 1989b).The biomass <strong>of</strong> the East Coast brown tide cells was inversely related toconcentrations <strong>of</strong> dissolved inorganic nitrogen in both field measurements <strong>and</strong>mesocosm-based nutrient enrichment studies (Keller <strong>and</strong> Rice, 1989). <strong>Brown</strong> tideblooms were <strong>of</strong> shorter duration <strong>and</strong> lower biomass in nutrient enriched mesocosmsthan in controls. This suggests that the East Coast brown tide blooms are not causedby macronutrient loading, <strong>and</strong> in fact the persistence <strong>of</strong> East Coast brown tide bloomsmay be due in part to its ability to grow at nutrient concentrations previouslydemonstrated to limit the growth <strong>of</strong> diatoms (Keller <strong>and</strong> Rice, 1989). Aureococcus hasgreater ability for heterotrophic take up <strong>of</strong> dissolved organic compounds compared toother species <strong>of</strong> microalgae (Dzurica et al., 1989). Trace metals <strong>and</strong> their chelatorshave been shown to have an important effect on the growth rate <strong>of</strong> Aureococcus inculture. Both iron <strong>and</strong> selenium additions caused faster growth rates, <strong>and</strong> the use <strong>of</strong>citric acid as a chelator resulted in faster growth rates than with the chelators EDTA(ethylene-diminetetraacetic acid) or NTA (nitrilotriacetic acid) (Cosper et al., 1993).Increased use <strong>of</strong> citric acid instead <strong>of</strong> phosphates in some detergents <strong>and</strong> increasedmunicipal use <strong>of</strong> deep groundwater with higher iron concentrations may havecontributed to the brown tide blooms (Cosper et al., 1993). No evidence <strong>of</strong> allelopathiceffects <strong>of</strong> chemicals released by the brown tide on the growth <strong>of</strong> other phytoplanktonspecies has been found (Cosper et al., 1989b).There is less information available on the factors leading to the termination <strong>of</strong>blooms <strong>of</strong> Aureococcus anophagefferens. Little is known about the role <strong>of</strong>microzooplankton grazing in bloom dynamics <strong>of</strong> the East Coast brown tide (Lonsdale etal., in prep). Sieburth et al. (1988) noted the occurrence <strong>of</strong> viral particles within A.anophagefferens cells during the 1985 bloom in Narragansett Bay, but the role <strong>of</strong>viruses in the bloom dynamics <strong>of</strong> this organism remains unknown.6


Perhaps the most important impact <strong>of</strong> the East Coast brown tide has been itsdetrimental effects on commercially important species <strong>of</strong> benthic macr<strong>of</strong>auna, such asscallops <strong>and</strong> mussels, <strong>and</strong> on eelgrass beds. Populations <strong>of</strong> the mussel Mytilus edulisin Narragansett Bay experienced mass mortalities <strong>of</strong> 30 to 100% (Tracey, 1988). InPeconic Bay, New York, adult bay scallops Argopecten irradians suffered heavymortality (Wenczel, 1987) <strong>and</strong> losses in mussel weight reducing harvest (Bricelj et al.,1987) resulting in economic losses to the scallop fishery estimated at two million dollarsper year (Kahn <strong>and</strong> Rockel, 1988). No data has been collected on the effects <strong>of</strong> thebrown tide on other benthic invertebrates (Lonsdale et al., in prep).Early outbreaks <strong>of</strong> the East Coast brown tide also caused severe reduction inlight penetration during the main growing season in shallow seagrass beds, resulting ina significant reduction in leaf biomass <strong>and</strong> confinement <strong>of</strong> eelgrass populations toshallow water (Dennison et al., 1989). Eelgrass provides an important nursery habitatfor many marine species, including bay scallops, <strong>and</strong> the losses <strong>of</strong> seagrass habitatmay have contributed to the lower populations <strong>of</strong> scallops during brown tide outbreaks.III. Available dataA. Environmental conditions during <strong>and</strong> immediately preceding the brown tide1. Initiation <strong>of</strong> the bloomScientists at the Marine Science Institute <strong>of</strong> the University <strong>of</strong> Texas (UTMSI)were carrying out a multi-investigator study <strong>of</strong> the Laguna Madre during the time whenthe brown tide began. Most experimental work was performed at two stations overseagrass beds in the upper Laguna Madre designated stations A <strong>and</strong> B, <strong>and</strong> at twostations in Baffin Bay over muddy bottoms, designated stations C <strong>and</strong> D (Figure 2).This study also included a monthly hydrographic survey at up to 44 stations in theupper Laguna Madre where basic hydrographic data were collected including seawatertemperature, salinity, pH <strong>and</strong> oxygen concentration. Seawater samples were alsocollected for chemical analysis <strong>of</strong> nutrient concentrations, including nitrate, nitrite,ammonium, phosphate <strong>and</strong> silicate. Water was also collected for determination <strong>of</strong>phytoplankton biomass estimated as chlorophyll a concentration. At a subset <strong>of</strong>stations in the upper reaches <strong>of</strong> Baffin Bay, whole water samples were preserved inLugol's iodine preservative for possible future analysis <strong>of</strong> the effects <strong>of</strong> high salinitieson phytoplankton <strong>and</strong> microzooplankton species composition. Secchi depth <strong>and</strong>chlorophyll a readings were taken from various transects in the Nueces <strong>and</strong> CorpusChristi Bays prior to the onset <strong>of</strong> the brown tide by Texas Water Development Board(Figure 3). UTMSI scientists first became aware <strong>of</strong> the presence <strong>of</strong> the brown tide inJune <strong>of</strong> 1990, when brown waters were encountered throughout upper Laguna Madrewith extremely high chlorophyll concentrations (Figure 4). The unusual nature <strong>of</strong> thebrown tide bloom was not realized for several months. Previously unexaminedpreserved whole water samples from Baffin Bay were then examined for the presence7


This hypothesis was rejected when a brown tide bloom developed in Copano Bay in thesummer <strong>of</strong> 1991 in waters <strong>of</strong>


Temperature <strong>and</strong> Salinity for Laguna Madre70Temp (0-35 0 C) <strong>and</strong> Salinity (30-70 ppt)605040302010Sta. A(Temp)Sta. A(Sal'n)Sta. B(Temp)Sta. B(Sal'n)Sta. C(Temp)Sta. C(Sal'n)Sta. D(Temp)Sta. D(Sal'n)0Mar'89May'89July'89Sept'89Nov'89Jan'90Mar'90May'90July'90Sept'90Nov'90Jan'91Mar'91May'91July'91DateFigure 6. Water temperature <strong>and</strong> salinity at four locations in the upper Laguna Madre<strong>and</strong> Baffin Bay from March 1989 until August 1991. Data from Dr. E.J. Buskey,University <strong>of</strong> Texas Marine Science Institute.Water temperatures in the Laguna Madre can vary by 5 o C or more over shortperiods <strong>of</strong> time due to changes in cloud cover <strong>and</strong> wind conditions <strong>and</strong> the shallowbody <strong>of</strong> water involved. In the late fall, winter <strong>and</strong> early spring, the passage <strong>of</strong> coldfronts can cause even wider fluctuations in the temperature <strong>of</strong> the Laguna Madre over ashort period <strong>of</strong> time. Daily temperatures at the intake to the Central Power <strong>and</strong> Lightelectrical generating station on the shore <strong>of</strong> Laguna Madre during part <strong>of</strong> 1993 <strong>and</strong>1994 are shown in Figure 7.11


Plot <strong>of</strong> Temperatures measured at CPL intake '93-'94353025Temp ( 0 C)201510506/30/93 8/19/93 10/8/93 11/27/93 1/16/94 3/7/94 4/26/94 6/15/94 8/4/94 9/23/94 11/12/94 1/1/95DateFigure 7. Water temperatures at the intake for the Central Power <strong>and</strong> Light electricalgenerating plant in Corpus Christi. Water for cooling the power plant is withdrawn fromthe Laguna Madre <strong>and</strong> returned to Oso Bay. Data provided by the GCCA/CPL redfishhatchery.3. Residence Time <strong>of</strong> Water in BaysWater in coastal bays is in a constant state <strong>of</strong> flux. Fresh water is added throughrainfall, groundwater <strong>and</strong> rivers. Freshwater can leave coastal bays directly throughevaporation or indirectly as less saline waters flow out <strong>of</strong> the bays into the Gulf <strong>of</strong>Mexico. Waters from the Gulf <strong>of</strong> Mexico are also mixed into coastal bays by way <strong>of</strong> tidalexchange. The residence times <strong>of</strong> waters in Texas coastal bays depends mainly on theamount <strong>of</strong> freshwater inflow <strong>and</strong> the tidal exchange from surrounding areas. Shormann(1992) estimated the residence times <strong>of</strong> three Texas bays in the CCBNEP area, BaffinBay, Nueces Bay <strong>and</strong> Copano Bay, in order to look into its possible effects on thepersistence <strong>of</strong> brown tide blooms in these three areas. He used the equationsdescribed in Armstrong (1982) to estimate residence time based on freshwater inflow<strong>and</strong> salinity data, <strong>and</strong> the equations described in Smith (1985) to estimate the effects <strong>of</strong>tidal flushing.Shormann (1992) assumed that tidal flushing <strong>of</strong> Baffin Bay was negligiblebecause <strong>of</strong> the bay's isolation from any passes to the Gulf <strong>of</strong> Mexico. Based onfreshwater inflow data, a residence time <strong>of</strong> approximately 12.5 years was calculated. A12


second residence time was calculated taking into account the losses <strong>of</strong> water throughevaporation. Based on these data, a residence time <strong>of</strong> approximately one year wascalculated. However, it should be noted that this turnover time is based on waterleaving by evaporation, <strong>and</strong> planktonic organisms such as the brown tide wouldactually be concentrated in Baffin Bay rather than be diluted out. In reality, asevaporation occurs, water becomes more dense, so some <strong>of</strong> this dense water probablyflows out along the bottom <strong>of</strong> Baffin Bay <strong>and</strong> is replaced by less saline water from theupper Laguna Madre. However, the upper Laguna Madre is also heavily impacted bythe brown tide, <strong>and</strong> this exchange does little to dilute the brown tide.Shormann (1992) also calculated residence times for Nueces Bay <strong>of</strong> 25 days,<strong>and</strong> for Copano Bay <strong>of</strong> 214 days. These calculated residence times correspondapproximately to the relative persistence <strong>of</strong> episodes <strong>of</strong> brown tide in each <strong>of</strong> thesebays. The brown tide spread into Nueces Bay on at least one occasion in August 1992,but it only persisted for a few weeks. Examination <strong>of</strong> brown tide cells using transmissionelectron microscopy showed the presence <strong>of</strong> what appeared to be viral particles withincells, suggesting that termination <strong>of</strong> this short lived bloom may have been enhanced bypathogens (Stockwell, pers. comm.).The brown tide has also spread into Copano Bay,<strong>and</strong> has persisted there for periods <strong>of</strong> several months. In Baffin Bay, with a calculatedresidence time <strong>of</strong> a year or longer, the brown tide has persisted for over five years, <strong>and</strong>has not abated since the bloom began.4. Light transmissionLight transmission through seawater is an important factor, along withtemperature <strong>and</strong> nutrient availability, determining the amount <strong>of</strong> primary production thatcan occur in a body <strong>of</strong> water. A brown tide event, typically with a cell concentration <strong>of</strong>one million cells per ml <strong>of</strong> seawater, can severely reduce the penetration <strong>of</strong> lightthrough seawater. Light transmission through seawater is largely affected by thepresence <strong>of</strong> dissolved <strong>and</strong> suspended materials that both absorb <strong>and</strong> scatter light. Asphotosynthetic organisms, brown tide algal cells absorb light in order tophotosynthesize; as small particles suspended in the water column they also scatterlarge amounts <strong>of</strong> light. One <strong>of</strong> the simplest ways to quantify the transparency <strong>of</strong> naturalwaters is to measure the greatest depth below the surface at which a small white disk,called a Secchi disk, can still be seen by an observer above the water's surface. Theyearly average Secchi depth, the depth at which the disk can no longer be seen, forseveral transects along Nueces Bay, Corpus Christi Bay <strong>and</strong> the upper Laguna Madre,prior to the brown tide, is shown in Figure 8. There is a general increase in watertransparency from average Secchi depths <strong>of</strong> ca. 50 cm in Nueces Bay to 80 - 100 cmas you approach the Laguna Madre. A somewhat incomplete record <strong>of</strong> Secchi depthsat four stations in the Laguna Madre <strong>and</strong> Baffin Bay is shown in Figure 9. Before thebrown tide began, Secchi depths were as great as 1.5m (the depth <strong>of</strong> the water column)in the Laguna Madre. During the most severe parts <strong>of</strong> the brown tide, Secchi depthswere <strong>of</strong>ten less than 0.5 m. Water clarity <strong>of</strong>ten improves for short periods <strong>of</strong> time in the13


winter, when cold fronts displace the brown tide laden water with relatively clear waterfrom Corpus Christi Bay.Sechi depths for Nueces estuary <strong>and</strong> upper Laguna Madre1.80Depth (m)1.601.401.201.000.800.600.400.200.0085 86 87 88 89Year475364122127142147Figure 8. Yearly average Secchi depth data at various transects throughout Nuecesestuary (47 & 53) <strong>and</strong> the upper Laguna Madre (64-147). Data provided by the TexasWater Development Board.Secchi depths for Stations A, B, C, D1.61.41.2Depth (m)10.80.6STA ASTA BSTA CSTA D0.40.204/2/89 7/11/89 10/19/89 1/27/90 5/7/90 8/15/90 11/23/90 3/3/91 6/11/91 9/19/91 12/28/91DateFigure 9. Monthly Secchi depth measurements taken at four stations in Laguna Madre<strong>and</strong> Baffin Bay. Data provided by Dr. Terry Whitledge, University <strong>of</strong> Texas MarineScience Institute. Station locations are shown in Fig. 2.14


In spite <strong>of</strong> the warm temperatures <strong>and</strong> high organic loads in the water columnduring the brown tide event, there have been no major fish kills reported in theCCBNEP that can be traced to anoxic events. This is due primarily to the shallow, wellmixed water column, which does not allow water near the bottom to be isolated from thesurface for long periods <strong>of</strong> time. Dissolved oxygen levels typically remain above 2 mg/l,even during warm summer months with high brown tide concentrations (Figure 11).D.O. levels at the CPL intake '93-'94109876D.O. (mg/L)5432106/30/93 8/19/93 10/8/93 11/27/93 1/16/94 3/7/94 4/26/94 6/15/94 8/4/94 9/23/94 11/12/94 1/1/95DateFigure 11. Morning dissolved oxygen levels measured in seawater at the intake to theCentral Power <strong>and</strong> Light electrical generating station. Samples taken betwee 4:00 <strong>and</strong>5:00 AM. Data provided by GCCA/CPL redfish hatchery.6. NutrientsPrior to the bloom <strong>of</strong> the Texas brown tide alga, concentrations <strong>of</strong> dissolvedinorganic nitrogen (DIN) in Baffin Bay <strong>and</strong> Upper Laguna Madre increased from


NH 4 concentrations for Laguna Madre3025uM201510STA ASTA BSTA CSTA D5003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91DateFigure 12. Ammonium concentrations in the upper Laguna Madre <strong>and</strong> Baffin Bay, fromMarch 1989 through December 1991. The 1990 winter / spring maxima was ascribedto the freeze induced fish <strong>and</strong> invertebrate die <strong>of</strong>f. Unpublished data from Dr. T.E.Whitledge, University <strong>of</strong> Texas Marine Science Institute.NO 3 concentrations in Laguna Madre1210uM8642003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91STA ASTA BSTA CSTA DDateFigure 13. Nitrate concentrations in the upper Laguna Madre <strong>and</strong> Baffin Bay, fromMarch 1989 through December 1991. Unpublished data from Dr. T.E. Whitledge,University <strong>of</strong> Texas Marine Science Institute.17


NO 2 concentrations for Laguna Madre2.52.0uM1.51.0STA ASTA BSTA CSTA D0.50.003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91DateFigure 14. Nitrite concentrations in the upper Laguna Madre <strong>and</strong> Baffin Bay, fromMarch 1989 through December 1991. Unpublished data from Dr. T.E. Whitledge.,University <strong>of</strong> Texas Marine Science Institute.Of this DIN increase, 60-95 % was in the form <strong>of</strong> NH 4+(Stockwell et al., 1993,Whitledge 1993, unpublished data). This large influx <strong>of</strong> ammonium was beneficial tothe Texas brown tide alga since, unlike most other algae, it is unable to use NO 3 - as anitrogen source. Aureococcus anophagefferens, the algal species associated with thenortheastern brown tides, is able to take up <strong>and</strong> reduce NO 3 - , NO 2 - , <strong>and</strong> NH 4 + . Evidencesuggests that the Texas brown tide species is able to take up all three inorganic forms<strong>of</strong> nitrogen, but is unable to reduce NO 3 - due to a lack <strong>of</strong> a functional nitrate reductase(DeYoe <strong>and</strong> Suttle, 1994).The large input <strong>of</strong> nitrogen is most <strong>of</strong>ten ascribed to two severe winter freezeswhich induced large fish kills <strong>and</strong> a salinity / temperature induced invertebrate die <strong>of</strong>f(Stockwell et al., 1993, Whitledge, 1993, DeYoe <strong>and</strong> Suttle, 1994). An estimated965,000 fish died in Baffin Bay <strong>and</strong> Upper Laguna Madre. This fish kill could havereleased approximately 2.01 x 10 7 g <strong>of</strong> N into the system (DeYoe <strong>and</strong> Suttle, 1994). Inthe year prior to the bloom, the invertebrate population decreased by over 90%releasing an estimated 0.48 g <strong>of</strong> N m -2 into Baffin Bay <strong>and</strong> 6.29 g <strong>of</strong> N m -2 into upperLaguna Madre (Montagna, personal communication). DeYoe <strong>and</strong> Suttle (1994)estimate that this increase in N would be sufficient to produce 6x10 12 cells m -3 , anamount larger than was actually observed during the bloom.This pulse <strong>of</strong> ammonium has been ascribed as a factor leading to the initiation <strong>of</strong>the Texas brown tide algal bloom. However, high concentrations <strong>of</strong> DIN might not benecessary to maintain the bloom. A study <strong>of</strong> the brown tide organism, Aureococcusanophagefferens, in Narragansett Bay has shown that the northeastern brown tidebloom was maintained during periods <strong>of</strong> extremely low DIN levels (< 1.0 µM) (Keller18


<strong>and</strong> Rice, 1989). Since Aureococcus is a picoplankton (cell size 2-3 µm), its highsurface area to volume ratio could allow it to take up a greater portion <strong>of</strong> the availablenutrients than the other, larger, phytoplankton (Joint, 1986). Keller <strong>and</strong> Rice (1989)found that the brown tide population diminished <strong>and</strong> other phytoplankton populationsrecovered when nutrient levels increased in bloom waters. In Texas, after the initialpre-bloom pulse, inorganic nutrient levels remained low for ca. 12 months (DIN< 3µM<strong>and</strong> PO 4 - ca. 1µM). However, the Texas brown tide bloom has successfully persistedthrough several periods <strong>of</strong> nutrient enrichment in Laguna Madre (Whitledge, 1993 <strong>and</strong>unpubl. data, Dunton, unpubl. data).Phosphorus <strong>and</strong> silica are rarely thought to be limiting in a coastal environment(Boynton et al., 1982); as a result almost no studies have been done on the effects <strong>of</strong>these nutrients on the Texas brown tide alga. A small increase <strong>of</strong> PO 4 - was observedprior to the start <strong>of</strong> the brown tide bloom (Figure 15). Phosphate levels increased fromca. 1µM to 3 µM (Whitledge, unpubl. data). It is unlikely that this pulse <strong>of</strong> phosphatewould have much effect since, phosphorous is limiting to the Texas brown tide algaonly when an unnaturally large (ca. 100µM) pulse <strong>of</strong> nitrogen is added to the system(Whitledge, pers. comm.). No studies have been done on the effect <strong>of</strong> silica levels onthe Texas brown tide alga. While the brown tide organism does not have aphysiological requirement for silica, it is conceivable that sufficient levels could allowdiatoms to compete with the alga. A drop in silica levels prior to the bloom (Figure 16)helps support this idea.PO4 concentration for Laguna Madre76uM54321003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91STA ASTA BSTA CSTA DDateFigure 15. Phosphate concentrations in the upper Laguna Madre <strong>and</strong> Baffin Bay, fromMarch 1989 through December 1991. Unpublished data from Dr. T.E. Whitledge,University <strong>of</strong> Texas Marine Science Institute.19


SiO 4 concentrations for Laguna Madre140120uM1008060STA ASTA BSTA CSTA D4020003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91DateFigure 16. Silica concentrations in the upper Laguna Madre <strong>and</strong> Baffin Bay from March1989 through December 1991. Unpublished data from Dr. T.E. Whitledge, University <strong>of</strong>Texas Marine Science Institute.B. <strong>Brown</strong> tide biomassOne way to estimate the severity <strong>of</strong> the brown tide bloom at any specific place ortime is to collect a water sample <strong>and</strong> count the number <strong>of</strong> brown tide cells present in avolume <strong>of</strong> water. Since most marine phytoplankton populations are quite diverse, it canbe a very difficult <strong>and</strong> time consuming task to try to identify all the species <strong>of</strong>phytoplankton present in a water sample <strong>and</strong> determine their abundance throughquantitative microscopic analysis <strong>of</strong> samples. A more common practice is toconcentrate plankton samples on filters <strong>and</strong> measure the concentration <strong>of</strong>photosynthetic pigments characteristic <strong>of</strong> all phytoplankton, such as chlorophyll a. Theamount <strong>of</strong> chlorophyll a present in a volume <strong>of</strong> water is then assumed to beproportional to the total biomass <strong>of</strong> phytoplankton present. The pigment composition <strong>of</strong>phytoplankton varies both with species <strong>and</strong> environmental conditions, so the ratio <strong>of</strong>chlorophyll a to total biomass can vary considerably from sample to sample. In the case<strong>of</strong> a nearly monospecific phytoplankton bloom such as the brown tide, most <strong>of</strong> thepigment is produced by a single species, so some <strong>of</strong> the variability is reduced <strong>and</strong>chlorophyll concentrations can be quite useful for comparing the concentrations <strong>of</strong>brown tide within different areas <strong>of</strong> the bloom <strong>and</strong> at different times <strong>of</strong> the year.However, since all phytoplankton species contain chlorophyll a, this parameterprovides no information about the proportion <strong>of</strong> the phytoplankton biomass contributedby the brown tide compared to other species. It can provide no indication <strong>of</strong> whenbrown tide is present or absent in the water.20


It is also possible to enumerate the concentration <strong>of</strong> brown tide cellsmicroscopically. The cells <strong>of</strong> the brown tide alga are small (4-5 µm in diameter) <strong>and</strong>very abundant in areas obviously affected by the brown tide. However, cells <strong>of</strong> thebrown tide alga are morphologically indistinct, <strong>and</strong> there are no external characteristicsthat can be used for positive microscopic identification.1. Cell countsIn the spring <strong>of</strong> 1990 the first "bloom" level abundances <strong>of</strong> the brown tide algawere observed in Laguna Madre. Prior to this time, the brown tide alga had not beenobserved in this area. Due to its indistinct appearance, this does not mean it was notpresent in low numbers <strong>and</strong> simply not recognized in phytoplankton samples. As figure17 shows, brown tide levels quickly increased from their initial observed levels <strong>of</strong> ca.500,000 cells/ml in June 1990 to over 2 x 10 6 cells/ml in July 1990. Populationabundances decreased from this summer high during the winter months but remainedat bloom levels (ca. 1 x10 6 cells/ml). Cell counts once again increased during thesummer months. Unlike A. anophagefferens, the Texas brown tide is able to maintainhigh abundance levels through the winter months. Thus, instead <strong>of</strong> a bloom which lastsfor several months, Texas has experienced a bloom <strong>of</strong> brown tide alga which has lastedover 5 years <strong>and</strong> still has an abundance over 500,000 cells/ml (Figure 18).<strong>Brown</strong> <strong>Tide</strong> Counts for Laguna Madre400000035000003000000Cells/ml250000020000001500000STA ASTA BSTA CSTA D10000005000000Oct-89 Jan-90 May-90 Aug-90 Nov-90 Mar-91 Jun-91 Sep-91DateFigure 17. <strong>Brown</strong> tide cell counts for 4 locations in upper Laguna Madre <strong>and</strong> Baffin Bayfrom December 1990 through August 1991. Buskey <strong>and</strong> Stockwell (1993) <strong>and</strong>unpublished data <strong>of</strong> Dr. E.J. Buskey, The University <strong>of</strong> Texas Marine Science Institute.21


<strong>Brown</strong> <strong>Tide</strong> Counts Upper Laguna Madre5Cells/ml x10^64.543.532.521.510.5011/12/94 1/1/95 2/20/95 4/11/95 5/31/95 7/20/95DateMouth Corpus ChristiBayStation AStation CLaguna Madre South<strong>of</strong> Baffin BayFigure 18. <strong>Brown</strong> tide counts for 4 locations in upper Laguna Madre <strong>and</strong> Baffin Bayfrom November 1994 through June 1995. Unpublished data <strong>of</strong> Dr. Roy Lehman, TexasA&M University - Corpus Christi, Center for Coastal Studies.2. Chlorophyll aFrom 1983 through 1989, prior to the initiation <strong>of</strong> the brown tide bloom, thechlorophyll a maxima never exceeded 17 µg/L (Figure 19). After the bloom, chlorophylla concentrations, during peaks, have exceeded 40µg/L (Figure 20).CHlorophyll a Concentrations for Nueces Bay <strong>and</strong> Laguna Madre1816Chl a (ug / L)1412108647536412212714214718342083 84 85 86 87 88 89YearFigure 19. Yearly average chlorophyll a concentrations averaged along cross baytransects running from Nueces Bay through the upper Laguna Madre. (See Figure 3 forthe location <strong>of</strong> these transects.) Data provided by the Texas Water DevelopmentBoard.22


Chlorophyll Data From STEPS60A5040µg Chla /L30STA ASTA C20100Jan-89 Jun-89 Nov-89 Apr-90 Sep-90 Feb-91 Jul-91 Dec-91 May-92 Sep-92 Feb-93 Jul-93DateChl concentrations for Laguna Madre7060Bµg/ml5040302010003/8906/8908/8910/8912/8902/9004/9006/9008/9010/9012/9002/9104/9106/9108/9110/9112/91STA ASTA BSTA CSTA DDateFigure 20. (A) Chlorophyll a data from stations A <strong>and</strong> C in the upper Laguna Madre <strong>and</strong>Baffin Bay, respectively, from March 1989 through February 1993. Unpublished datafrom Dr. Dean Stockwell, University <strong>of</strong> Texas Marine Science Institute.(B) Surface chlorophyll a concentrations (µg/l) from four stations in the upper LagunaMadre from March 1989 until December 1991. Unpublished data <strong>of</strong> Dr. TerryWhitledge, University <strong>of</strong> Texas Marine Science Institute.Primary production (gC m -2 d -1 ) before the brown tide averaged 1.2 gC m -2 d -1 ,while after the brown tide the average increased to 2.0 gC m -2 d -1 (Figure 21). Both23


D. Environmental impact1. SeagrassesOne <strong>of</strong> the greatest potential impacts <strong>of</strong> the brown tide on the Laguna Madreecosystem is the loss <strong>of</strong> seagrass due to reduced light penetration. <strong>Red</strong>uctions in waterclarity are thought to have been responsible for large scale losses in seagrass habitatin the past (Kenworthy <strong>and</strong> Haunert, 1991), <strong>and</strong> the depth limit <strong>of</strong> seagrass meadows is<strong>of</strong>ten assumed to be constrained by the depth <strong>of</strong> light penetration required forphotosynthesis (Dennison, 1987; Duarte, 1991). The brown tide has caused areduction in light penetration which increases with depth <strong>of</strong> the water column, so thegreatest impact <strong>of</strong> the brown tide should be on seagrasses at the greatest depths.A study by Dr. Ken Dunton at the University <strong>of</strong> Texas Marine Science Institutecompared seagrass growth characteristics at sites in the Laguna Madre near station Awhich has been heavily impacted by the brown tide <strong>and</strong> a site in the East Flats <strong>of</strong>Corpus Christi Bay that has been generally free <strong>of</strong> the brown tide. Seagrass shootdensity can be quite variable at both sites, particularly during the period from 1989 untilthe summer <strong>of</strong> 1992 (Figure 22). Following the summer <strong>of</strong> 1992, shoot density at thesite in the East Flats <strong>of</strong> Corpus Christi Bay remained largely unchanged, whereas shootdensity showed a steady decline during the same period at the site in Laguna Madre,decreasing from 8-10 thous<strong>and</strong> shoots per square meter in the summer <strong>of</strong> 1992 to lessthan 4,000 shoots per square meter by the end <strong>of</strong> 1994 (Figure 22). Seagrass arealshoot biomass, measured in grams dry weight per square meter also showsconsiderable variability, but a seasonal pattern <strong>of</strong> high shoot biomass in late summer isobserved in each year except 1991 at the East Flats site, whereas summer biomasspeaks appear to be dampened out after 1991 at the Laguna Madre site (Figure 23).However, above ground shoot density <strong>and</strong> biomass measures may not necessarily bethe best indicators <strong>of</strong> stress induced by reduced light intensities because seagrassesstore reserve energy in their rhizomes, <strong>and</strong> can use these reserves to produce aboveground biomass during periods <strong>of</strong> reduced light. The biomass <strong>of</strong> roots <strong>and</strong> rhizomeshas been declining steadily since 1992 at the Laguna Madre site, compared to a morevariable pattern at the East Flats site (Figure 25). The ratio <strong>of</strong> root <strong>and</strong> rhizomebiomass to shoot biomass dropped from a high <strong>of</strong> 16 in Laguna Madre before thebrown tide began, <strong>and</strong> then fell to a value <strong>of</strong> 2 to 4 during the brown tide bloom, whereit has stabilized. The range <strong>of</strong> ratios observed at the East Flats site, unlike LagunaMadre, varied with a relatively consistent seasonal pattern from 1 to 3 (Figure 25).25


Station A1400012000Density (Shoots m -2 )10000800060004000200009/88 4/89 10/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateEast Flats1400012000Density (Shoots m -2 )100008000600040002000010/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateFigure 22. Seagrass shoot density (shoots/m 2 ) at a location in upper Laguna Madre(top) <strong>and</strong> a location in East Flats <strong>of</strong> Corpus Christi Bay (bottom). Data from Dr. KenDunton, University <strong>of</strong> Texas Marine Science Institute.26


Station A300Areal Shoot Biomass (gdw m-2)2502001501005001/89 7/89 2/90 8/90 3/91 9/91 4/92 11/92 5/93 12/93 6/94DateEast Flats300Areal Shoot Biomass (gdw m-2)25020015010050010/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateFigure 23. Seagrass areal shoot biomass (gdw/m 2 ) at a location in upper Laguna Madre(top) <strong>and</strong> a location in East Flats <strong>of</strong> Corpus Christi Bay (bottom). Data from Dr. KenDunton, University <strong>of</strong> Texas Marine Science Institute.27


Station A600Root/Rhizome Biomass (gdw m -2 )50040030020010009/88 4/89 10/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateEast Flats600Root/Rhizome Biomass (gdw m -2 )500400300200100010/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateFigure 24. Seagrass root/rhizome biomass (gdw/m 2 ) at a location in upper LagunaMadre (top) <strong>and</strong> a location in East Flats <strong>of</strong> Corpus Christi Bay (bottom). Data from Dr.Ken Dunton, University <strong>of</strong> Texas Marine Science Institute.28


Station A2018Root/Rhizome:Shoot Ratio16141210864209/88 4/89 10/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateEast Flats2018Root/Rhizome:Shoot Ratio161412108642010/89 5/90 11/90 6/91 12/91 7/92 1/93 8/93 3/94 9/94 4/95DateFigure 25. Ratio <strong>of</strong> root/rhizome biomass to shoot biomass at a location in upperLaguna Madre (top) <strong>and</strong> a location in East Flats <strong>of</strong> Corpus Christi Bay (bottom). Datafrom Dr. Ken Dunton, University <strong>of</strong> Texas Marine Science Institute.Dr. Christopher P. Onuf (submitted), reports on changes in seagrass distribution<strong>and</strong> biomass due to light attenuation <strong>of</strong> the Texas brown tide. Comparisons <strong>of</strong> pre- <strong>and</strong>post-brown tide studies were made to give an estimate on the amount <strong>of</strong> sea grasses29


It is thought that light attenuation due to the brown tide is responsible for theseagrass losses. Underwater photosynthetically available radiation (PAR) wasmeasured at 1 min intervals at three locations (Figure 26). Secchi depths were taken atapproximately monthly intervals from March 1992 to November 1994 at stations 100mfrom the eastern <strong>and</strong> western shores <strong>of</strong> the Laguna <strong>and</strong> at the midpoint on east-westtransects at 27 o 20’, 22’, 26’ 30’, 38’ <strong>and</strong> 40’. These Secchi depths were related to PARby determining light vs. depth pr<strong>of</strong>iles <strong>and</strong> computing attenuation coefficients todetermine the proportion <strong>of</strong> surface light reaching the bottom. Light attenuationincreases from north to south which is consistent with the higher biomass seen at thegreatest depths in the northern portion <strong>of</strong> the study prior to 1993 (Figure 28). Thecompensation point for Halodule has been estimated at 12 - 20% <strong>of</strong> surface irradiation(Kenworthty, et al. 1991, Onuf, 1991). Using an average compensation point <strong>of</strong> 15%surface irradiance, Onuf predicts that 23% <strong>of</strong> the study area is too deep to supportHalodule while the brown tide persists. Prior to the brown tide only 6% <strong>of</strong> LagunaMadre was too deep .Although 17% <strong>of</strong> the 1988 meadow has continually received light below thecompensation level, only 6% <strong>of</strong> the meadow has been lost. Onuf (1995) proposes thatHalodule is “able to survive periods <strong>of</strong> inadequate light by reducing the number <strong>of</strong>shoots it maintains <strong>and</strong> supporting the reduced dem<strong>and</strong>s by metabolizing thecarbohydrate reserves stored in below ground tissues.” The reduction in biomass priorto loss in distribution supports this hypothesis. As the brown tide persists the coverageshould continue to decrease (as it did from 1993 to 1994). Since seagrasses preventresuspension <strong>of</strong> sediment, the loss <strong>of</strong> the seagrass could increase the light attenuation<strong>of</strong> suspended particles, leading to even greater seagrass losses. These losses couldhave a major effect on species, such as the <strong>Red</strong>head ducks (Anthya americana) whichdepend on Halodule wrightii for food <strong>and</strong> shelter.2. Planktonic GrazersMesozooplankton populations, dominated by the copepod Acartia tonsa, weregenerally abundant before the onset <strong>of</strong> the brown tide, although there was considerablespatial <strong>and</strong> temporal variability in population density (Figure 29). In Baffin Bay,mesozooplankton densities ranged from 1,500 to 23,500 organisms per cubic meterbefore the onset <strong>of</strong> the brown tide. After the beginning <strong>of</strong> the brown tide,mesozooplankton densities remained below 8,000 per cubic meter through July 1991.The ctenophore Mnemiopsis mccradyi was abundant in Baffin Bay in March <strong>and</strong> April1990, just prior to the spread <strong>of</strong> the brown tide.The size <strong>of</strong> adult female Acartia tonsa, measured as prosome length, is usually afunction <strong>of</strong> the temperature <strong>and</strong> salinity <strong>of</strong> the waters the organisms grow in. Highertemperatures <strong>and</strong> higher salinities result in smaller copepods. In Baffin Bay, whereconditions were extremely hypersaline, both temperature <strong>and</strong> salinity affected the size<strong>of</strong> A. tonsa. In Laguna Madre, where conditions were less hypersaline, temperature hadthe most important influence on the size <strong>of</strong> A. tonsa before the brown tide, with32


copepods reaching their maximum size in the cool winter months (ca. 0.75 mmprosome length) <strong>and</strong> their minimum size in hot summer months (ca. 0.55 mm prosomelength) (Figure 30). During the brown tide, A. tonsa remained small throughout thecooler winter months (0.6 mm prosome length).Mesozooplankton(x1000 m-3)504540353025201510504035302520151050Mar-89Jun-89Sep-89Dec-89Mar-90Ctenophore Volume(ml/m3)Jun-90Oct-90Jan-91Apr-91Jul-91DateMesozooplanktonCtenophoresFigure 29. Mesozooplankton abundance <strong>and</strong> ctenophore displacement volume in BaffinBay before <strong>and</strong> during the brown tide. Adapted from Buskey <strong>and</strong> Stockwell (1993).350.930Temperature ( o C)2520151050.80.70.6Prosome Length (mm)00.5Sep-88 Apr-89 Oct-89 May-90 Nov-90 Jun-91 Dec-91Month (1989 -1991)Surface TempProsome lengthFigure 30. Prosome lengths <strong>of</strong> adult female Acartia tonsa at station A in Laguna Madrebefore <strong>and</strong> during the brown tide bloom, <strong>and</strong> temperature data from the same station.Before the brown tide, the size <strong>of</strong> copepods was a function mainly <strong>of</strong> watertemperature; after the bloom copepods remained small throughout the year, indicatingpoor nutrition. Adapted from Buskey <strong>and</strong> Stockwell (1993).33


Egg release rates by adult female Acartia tonsa ranged from 20-60 eggs perfemale per day in Baffin Bay before the brown tide began (Figure 31). After the browntide began, the egg release rate was consistently below 5 eggs per female per day.706050403020100<strong>Brown</strong> <strong>Tide</strong> BeganMar-89Jul-89Nov-89Eggs/Female/DayMar-90Jul-90Nov-90Mar-91Month (1989-1991)Figure 31. Egg release rates <strong>of</strong> adult female Acartia tonsa before <strong>and</strong> during the browntide, based on 48 hour field incubations. Adapted from Buskey <strong>and</strong> Stockwell (1993).Gut pigment contents can be used to estimate grazing rates <strong>of</strong>mesozooplankton. Before the brown tide began, gut pigment contents for adult femaleAcartia tonsa ranged from 2 -18 ng pigment per copepod. Gut pigment contents were atthe detection limit <strong>of</strong> our analytical technique (< 1 ng chl a /copepod) following thebrown tide, even though surface water chlorophyll concentrations were 2-10x higher(Figure 32).Gut Pigments (ng/copepod)181614121086420Before <strong>Brown</strong> tideAfter <strong>Brown</strong> <strong>Tide</strong>0 20 40 60 80Surface Chl a (ug/l)Figure 32. Daytime gut pigment (chlorophyll a plus phaeopigments) concentrations foradult female Acartia tonsa compared to phytoplankton biomass (chlorophyll a) before<strong>and</strong> during the brown tide. High gut pigment concentrations indicate active feeding onphytoplankton. Adapted from Buskey <strong>and</strong> Stockwell (1993).34


The observed changes in zooplankton populations could have been due to thenutritional inadequacy <strong>of</strong> the brown tide alga to support zooplankton growth <strong>and</strong>reproduction, or there may be a substance produced by the algae that interferes withzooplankton feeding. Lower egg laying rates were reported for Acartia tonsa fromNarragansett Bay, RI, when fed phytoplankton collected in the field during anAureococcus bloom (Durbin <strong>and</strong> Durbin, 1989). Increases in body size, condition factor(a ratio <strong>of</strong> weight to a unit <strong>of</strong> length), egg laying rate <strong>and</strong> gut pigments were reportedwhen field collected Acartia tonsa were fed cultured algae in their preferred size rangeduring laboratory experiments.Microzooplankton populations consisting mainly <strong>of</strong> ciliates, copepod nauplii <strong>and</strong>rotifers, were generally abundant before the brown tide began, with populationsgenerally ranging between 20-200 per ml. After the brown tide began, themicrozooplankton populations remained below 30 per ml (Figure 33). Microzooplanktongrazed ca. 85-98% <strong>of</strong> the phytoplankton st<strong>and</strong>ing stock per day before the brown tide,but < 3% per day during the brown tide (Figure 34). The inability <strong>of</strong> microzooplanktongrazers to control populations <strong>of</strong> the brown tide alga may result from the production <strong>of</strong>substances that inhibit the grazing <strong>and</strong>/or growth <strong>of</strong> microzooplankton. Anotherpossible explanation might be that the mesozooplankton exerted additional predationpressure on microzooplankton populations during the brown tide, due to the lack <strong>of</strong>phytoplankton food in their preferred size range.1202.5Microzooplankton/ml10080604020021.510.50Dec-88Apr-89Jul-89Oct-89Jan-90May-90Aug-90Nov-90Mar-91<strong>Brown</strong> <strong>Tide</strong> (cells/ml x 10 6 )Jun-91Sep-91Month (1989-1991)Microzooplankton<strong>Brown</strong> <strong>Tide</strong>Figure 33. Microzooplankton abundance <strong>and</strong> brown tide cell concentrations at StationC in Baffin Bay before <strong>and</strong> during the brown tide. Adapted from Buskey <strong>and</strong> Stockwell(1993).35


Chl a (mg/m 3 )4035302520151050<strong>Brown</strong> <strong>Tide</strong> BeganMar-89May-89Jul-89Sep-89Sep-89Jan-90Mar-90May-90Jul-90Sep-90Nov-90Month (1989-1990)Initial St<strong>and</strong>ing Crop Amount GrazedFigure 34. Microzooplankton grazing <strong>of</strong> phytoplankton st<strong>and</strong>ing stock before <strong>and</strong> duringthe brown tide, based on field incubation using the dilution technique <strong>of</strong> L<strong>and</strong>ry <strong>and</strong>Hassett (1982). Adapted from Buskey <strong>and</strong> Stockwell (1993)Laboratory StudiesLaboratory studies <strong>of</strong> the effects <strong>of</strong> the Texas brown tide alga on zooplanktonhave been carried out by Dr. Edward J. Buskey <strong>of</strong> the University <strong>of</strong> Texas MarineScience Institute. The ciliate Strombidinopsis sp. showed a typical numerical response<strong>of</strong> increasing specific growth rate with increases in food concentration until maximumgrowth rate is reached, when the food <strong>of</strong>fered was Pyrenomonas salina. The thresholdconcentration for growth appeared to be ca. 0.1 mg C l -1 , <strong>and</strong> a maximum specificgrowth rate <strong>of</strong> 0.96 d -1 was achieved at a food concentration <strong>of</strong> 1 mg C l -1 (Figure 35).When this same species was fed the Texas brown tide alga, a different numericalresponse was found. All specific growth rates were negative, indicating mortality ratherthan growth on this food source. Since specific growth rates became more negative asfood concentrations <strong>of</strong> brown tide increase (i.e. death rate increased), it appears thatthe brown tide is toxic to this species.36


Specific Growth Rate (day -1 )1.510.50-0.5-1-1.50 1 2 3 4 5Phytoplankton Concentration (mg C/l)Fed Pyrenomonas Fed <strong>Brown</strong> <strong>Tide</strong>Figure 35. Specific growth rate (d -1 ) <strong>of</strong> the ciliate Strombidinopsis sp. on variousconcentrations <strong>of</strong> the Texas brown tide alga <strong>and</strong> on Pyrenomonas salina. Error barsrepresent the st<strong>and</strong>ard error <strong>of</strong> the slope <strong>of</strong> line for the natural logarithm <strong>of</strong> the number<strong>of</strong> cells regressed against time, used to determine specific growth rates. Adapted fromBuskey <strong>and</strong> Hyatt (1995).In contrast, the ciliates Fabrea salina <strong>and</strong> Euplotes sp. grew well on the browntide. Maximum specific growth rate for F. salina <strong>of</strong> 0.52 d -1 was achieved at 1 mg C l -1<strong>of</strong> brown tide. However, the numerical response <strong>of</strong> F. salina to various concentrations<strong>of</strong> the brown tide alga shows a decreasing growth rate at food concentrations above 1mg C l -1 , falling to 0.25 d -1 at 5 mg C l -1 (Figure 36). In contrast, F.salina fedIsochrysis galbana showed a more typical numerical response curve, with growth rateremaining high at higher food concentrations (maximum specific growth rate <strong>of</strong> 0.71 d -1at 5 mg C l -1. ) The ciliate Euplotes sp. showed a similar numerical response when fedthe brown tide alga. Maximum specific growth rates <strong>of</strong> 0.5 d -1 were observed at abrown tide concentration <strong>of</strong> 1 mg C l -1 , but specific growth rates fell to 0.29 at 5 mg C l -1 (Figure 37). Growth <strong>of</strong> Euplotes sp. was higher when fed I. galbana at 2.5 or 5 mgcarbon l -1 than when fed similar concentrations <strong>of</strong> the brown tide.The Texas brown tide alga did not support the growth <strong>of</strong> the heterotrophicdin<strong>of</strong>lagellate Noctiluca scintillans. However, there is no evidence that the brown tidealga is highly toxic to N. scintillans, since growth (death) rate was only -0.04 d -1 at 5 mgC l -1 <strong>of</strong> brown tide. Specific growth rates <strong>of</strong> N. scintillans on Thalassiosira sp. rangedfrom 0.38 d -1 for 0.1 mg carbon l -1 to 0.6 d -1 for 1.0 mg carbon l -1 (Figure 38). When amixture <strong>of</strong> Thalassiosira sp. <strong>and</strong> the brown tide alga was <strong>of</strong>fered as food, the specificgrowth rate ranged from 0.22 d -1 to 0.38 d -1 . Even though total food concentration wastwice as high when equal amounts <strong>of</strong> brown tide <strong>and</strong> Thalassiosira sp. were <strong>of</strong>fered asfood, growth rates were always lower than when N. scintillans were <strong>of</strong>feredThalassiosira sp. alone.37


Specific Growth Rate (day -1 )0.800.700.600.500.400.300.200.100.000.00 1.00 2.00 3.00 4.00 5.00Phtyoplankton Concentration (mg C/l)Fed <strong>Brown</strong> <strong>Tide</strong>Fed IsochrysisFigure 36. Specific growth rate (d -1 ) <strong>of</strong> the ciliate Fabrea salina on variousconcentrations <strong>of</strong> the Texas brown tide alga <strong>and</strong> on Isochrysis galbana. Adapted fromBuskey <strong>and</strong> Hyatt (1995).0.6Specific Growth Rate (day -1 )0.50.40.30.20.100 1 2 3 4 5Phytoplankton Concentration (mg C/l)Fed <strong>Brown</strong> <strong>Tide</strong>Fed IsochrysisFigure 37. Specific growth rates (d -1 ) <strong>of</strong> the ciliate Euplotes sp. on variousconcentrations <strong>of</strong> the Texas brown tide alga <strong>and</strong> on Isochrysis galbana. Adapted fromBuskey <strong>and</strong> Hyatt (1995).38


Specific Growth Rate (Day -1 )0.70.60.50.40.30.20.10-0.1-0.20 1 2 3 4 5Phytoplankton Concentration (mg C/l)Thalassiosira <strong>Brown</strong> <strong>Tide</strong> BothFigure 38. Specific growth rates (d -1 ) <strong>of</strong> the heterotrophic din<strong>of</strong>lagellate Noctilucascintillans on various concentrations <strong>of</strong> the Texas brown tide alga, Thalassiosira sp. orequal concentrations <strong>of</strong> both species (twice the total food concentration). Adapted fromBuskey <strong>and</strong> Hyatt (1995).The small heterotrophic din<strong>of</strong>lagellate Oxyhrris marina did not grow on browntide at food concentrations below 0.1 mg C l -1 , but grew well on higher concentration <strong>of</strong>this species (Figure 39). Maximum specific growth rates <strong>of</strong> 0.55 d -1 were observed at aconcentration <strong>of</strong> 5 mg C l -1 . However, O. marina grew faster on low concentrations <strong>of</strong>Isochrysis galbana, <strong>and</strong> reached maximum growth at 0.5 mg C l -1 . Growth rates <strong>of</strong> O.marina were higher when fed equivalent concentrations <strong>of</strong> I. galbana than when fed thebrown tide alga at concentrations below 5 mg C l -1 (Figure 39).Specific Growth (day -1 )10.80.60.40.20-0.2-0.40 1 2 3 4 5Phytoplankton Concentration (mg C/l)<strong>Brown</strong> <strong>Tide</strong>IsochrysisFigure 39. Specific growth rates (d -1 ) <strong>of</strong> the heterotrophic din<strong>of</strong>lagellate Oxyhrrismarina on the Texas brown tide alga <strong>and</strong> on Isochrysis galbana. Adapted from Buskey<strong>and</strong> Hyatt (1995).39


The Texas brown tide alga did not support the growth <strong>of</strong> the rotifer Brachionusplicatilus at any food concentration. Specific growth (death) rates were similar to thosefor rotifers that were starved over the same 3 day period (ca. -0.15 d -1 ). At a brown tideconcentration <strong>of</strong> 5 mg C l -1 , specific growth (death) rate decreased to -0.39. When therotifers were fed Isochrysis galbana, specific growth rates ranged from 0.19 to 0.61 d -1(Figure 40). However, when B. plicatilus was fed an equal amount <strong>of</strong> both I. galbana<strong>and</strong> the Texas brown tide alga (yielding twice the total phytoplankton concentration), B.plicatilus showed little or no growth.Specific Growth Rate (day -1 )0.80.60.40.20-0.2-0.4-0.60 1 2 3 4 5Phytoplankton Concentration (mg C/l)Fed Isochrysis Fed <strong>Brown</strong> <strong>Tide</strong> Fed BothFigure 40. Specific growth rates (d -1 ) <strong>of</strong> the rotifer Brachionus plicatilus on the Texasbrown tide alga, Isochrysis galbana <strong>and</strong> equal concentrations <strong>of</strong> each (twice the totalfood concentration). Adapted from Buskey <strong>and</strong> Hyatt (1995).When groups <strong>of</strong> 36 Acartia tonsa nauplii were raised in 3 ml cell wells in tissueculture plates, the group held without any food exhibited daily mortality <strong>and</strong> were alldead by the end <strong>of</strong> a five day period. For a similar group <strong>of</strong> nauplii held in 3 ml <strong>of</strong>seawater containing brown tide at a concentration <strong>of</strong> 5 mg C l -1 , there was extensivemortality on the second day, <strong>and</strong> all nauplii were dead by day 3. For groups <strong>of</strong> naupliiraised on 5 mg C l -1 <strong>of</strong> Isochrysis galbana, Pyrenomonas salina or a combination <strong>of</strong> thetwo foods, over 85% <strong>of</strong> the nauplii were still alive at the end <strong>of</strong> the 5 day period (Figure41). Since the brown tide appeared to settle to the bottom <strong>of</strong> the cell wells over thecourse <strong>of</strong> the experiment, subsequent experiments were done with groups <strong>of</strong> 50 naupliiin 50 ml tissue culture flasks rotated at 2 rpm to keep the algae in suspension.Replicate experiments were run at 2, 3.5 <strong>and</strong> 5 mg C l -1 . Survival <strong>of</strong> A. tonsa naupliiranged from 1% to 6% on the brown tide alone, from 70 to 83% on I. galbana alone <strong>and</strong>from 41 to 55% on a combination <strong>of</strong> equal amounts <strong>of</strong> each food (twice the foodconcentration) (Figure 42).40


Percent Survival10090807060504030201000 1 2 3 4 5Time (days)Starved <strong>Brown</strong> <strong>Tide</strong> IsochrysisPyrenomomos Iso & PryenFigure 41. Survival <strong>of</strong> groups <strong>of</strong> 24 nauplii <strong>of</strong> the copepod Acartia tonsa fed 5 mg C l -1<strong>of</strong> the Texas brown tide alga, Isochrysis galbana, Pyrenomonas salina, a combination<strong>of</strong> I. galbana <strong>and</strong> P. salina or starved over a 96 hour period. Adapted from Buskey <strong>and</strong>Hyatt (1995).0.9Proportion <strong>of</strong> Nauplii Surviving0.80.70.60.50.40.30.20.102 3.5 5Food Concentration (mg C/l)Isochrysis <strong>Brown</strong> <strong>Tide</strong> BothFigure 42. Survival <strong>of</strong> groups <strong>of</strong> 50 nauplii <strong>of</strong> the copepod Acartia tonsa fed Isochrysisgalbana, the Texas brown tide alga or a combination <strong>of</strong> both (twice the total foodconcentration) at food concentrations <strong>of</strong> 2, 3.5 <strong>and</strong> 5 mg C l -1 over a 96 hour period.Each bar is the mean value <strong>of</strong> two replicate experiments. Adapted from Buskey <strong>and</strong>Hyatt (1995).41


Egg release rates for adult female Acartia tonsa fed the Texas brown tide alga ata food concentration <strong>of</strong> 1.5 mg C l -1 was 3.4 + 2.3 (mean + 1 SD, n=6) eggs per femaleper day, which was not significantly different from the egg release rates <strong>of</strong> 1.7 + 0.7 forA. tonsa that had been starved over the same 48 h period (t-test, p = 0.05). A. tonsafemales fed similarly sized small phytoplankton species showed intermediate eggrelease rates <strong>of</strong> 9.3 + 2.6 eggs per female per day when fed 1.5 mg C l -1 <strong>of</strong> Isochrysisgalbana (5 µm diameter) <strong>and</strong> 13.1 + 3.5 eggs per female per day when fed 1.5 mg C l -1<strong>of</strong> Emiliania huxleyi (4 µm diameter). Highest egg release rates <strong>of</strong> 25.9 + 7.2 eggs perfemale per day were measured for A. tonsa fed 1.5 mg C l -1 <strong>of</strong> the diatom Thalassiosirasp. When A. tonsa females were <strong>of</strong>fered a combination <strong>of</strong> 1.5 mg C l -1 each <strong>of</strong>Thalassiosira sp. <strong>and</strong> the Texas brown tide alga (3 mg C l -1 total), egg release rateswere 26.9 + 3.6 eggs per female per day, which is not significantly different from therelease rate when the copepods were fed Thalassiosira sp. alone (t-test, α = 0.05)(Figure 43).3530Eggs/Female/Day252015105Starved<strong>Brown</strong> <strong>Tide</strong>IsochrysisEmilianiaThalassiosiraThal & <strong>Brown</strong> <strong>Tide</strong>0Figure 43. Mean egg release rates (eggs female -1 day -1 ) <strong>of</strong> adult female Acartia tonsafed 1.5 mg C l -1 <strong>of</strong> the Texas brown tide alga, Isochrysis galbana, Emiliania huxleyi,Thalassiosira sp., a combination <strong>of</strong> Thalassiosira sp. <strong>and</strong> brown tide, or starved. Eachbar represents the mean (+ SD) <strong>of</strong> eight replicate experiments. Adapted from Buskey<strong>and</strong> Hyatt (1995).The Texas brown tide alga appears to be a poor food for a variety <strong>of</strong>zooplankton species. It supports no growth <strong>of</strong> the ciliate Strombidinopsis sp., theheterotrophic din<strong>of</strong>lagellate Noctiluca scintillans or the rotifer Brachionus plicatilus. It isnot unusual to find zooplankton that can not be cultured on a particular species <strong>of</strong>phytoplankton; some can only capture particles in a relatively narrow size range(Fenchel, 1980). However N. scintillans can be grown on a wide range <strong>of</strong> phytoplanktonspecies, including species <strong>of</strong> similar size (Buskey, 1995) <strong>and</strong> B. plicatilus is an easy to42


culture organism widely used as a food for larval fish. Based on the results <strong>of</strong> thisstudy, there is evidence that the brown tide may be directly toxic to some species <strong>of</strong>zooplankton, at cell concentrations similar to those found in nature. For the ciliateStrobidinopsis sp. (Figure 35) <strong>and</strong> for the rotifer B. plicatilus (Figure 40), mortality ratesincrease with increasing brown tide concentration. For one experiment with Acartiatonsa nauplii, mortality was faster in the presence <strong>of</strong> the brown tide than when no foodwas <strong>of</strong>fered (Figure 41). Additional evidence for toxicity <strong>of</strong> the brown tide to somespecies <strong>of</strong> zooplankton comes from the decrease in survival <strong>of</strong> A. tonsa nauplii whenboth a suitable food (Isochrysis galbana) <strong>and</strong> the brown tide are <strong>of</strong>fered together(Figure 42). In addition, when both I. galbana <strong>and</strong> the brown tide are <strong>of</strong>fered together,growth <strong>of</strong> the heterotrophic din<strong>of</strong>lagellate N. scintillans <strong>and</strong> the rotifer B. plicatilus areinhibited (Figures 38 & 40). In contrast, there is little evidence that the related speciesAureococcus anophagefferens is toxic to microzooplankton. No evidence was found forchanges in protozoan grazing rate or for suppression <strong>of</strong> growth in protozoans fed A.anophagefferens, nor was there any evidence that A. anophagefferens caused areduction in protozoan populations in nature (Caron et al., 1989).The brown tide alga appears to be a poor food item for Acartia tonsa, thedominant mesozooplankton in the Laguna Madre. Egg release rates <strong>of</strong> adult femalesfed the brown tide were not significantly different from those held without food over thesame time interval. This may have been due in part to the small size <strong>of</strong> the brown tidecells (4-5 µm diameter), which is outside the optimum size range for particle capture byA. tonsa (Berggreen et al., 1988). However, A. tonsa females produced an intermediatenumber <strong>of</strong> eggs on two similarly small sized algal species, indicating that size alonewas not the problem. There was no evidence <strong>of</strong> brown tide toxicity to adult female A.tonsa at 1.5 mg C l -1 , since there was no direct mortality to adults <strong>and</strong> egg release wasnot lowered with the combination <strong>of</strong> Thalassiosira <strong>and</strong> brown tide. Lower egg releaserates are reported for A. tonsa fed picoalgae during an A. anophagefferens bloom inNarragansett Bay, (Durbin <strong>and</strong> Durbin, 1989). In addition to lowering the egg releaserates <strong>of</strong> adult females, the presence <strong>of</strong> the brown tide resulted in lower survival <strong>of</strong> A.tonsa nauplii, suggesting toxic effects. The Texas brown tide alga has also been shownto be toxic to first feeding red drum <strong>and</strong> spotted sea trout larvae (G.J. Holt, personalcommunication) but it does not appear to have adverse affect on adult fish populations.Field evidence also supports the concept that the Texas brown tide alga is apoor food for zooplankton <strong>and</strong> disrupts trophic transfer in the planktonic food web.Mesozooplankton abundance (mainly Acartia tonsa) was lower in the Laguna Madreafter the brown tide began than in the preceding year, <strong>and</strong> adult female A. tonsa weresmaller <strong>and</strong> produced fewer eggs in field incubations than before the brown tide began(Buskey <strong>and</strong> Stockwell, 1993). Microzooplankton abundances were also lower after thebrown tide began, <strong>and</strong> microzooplankton community grazing rates <strong>of</strong> phytoplanktonst<strong>and</strong>ing stock were reduced from ca. 95% to less than 5% during the brown tide(Buskey <strong>and</strong> Stockwell, 1993). It is still difficult to underst<strong>and</strong> why species <strong>of</strong>microzooplankton capable <strong>of</strong> growing on the brown tide have not flourished <strong>and</strong> helpedbring the brown tide under control. It is possible that A. tonsa may be exerting43


additional predation pressure on microzooplankton populations during the brown tide,due to the reduction <strong>of</strong> other species <strong>of</strong> phytoplankton in their preferred size range. It iswell documented that A. tonsa also feed on microzooplankton (reviewed in Pierce <strong>and</strong>Turner, 1992), so it is possible that they are holding microzooplankton populationsbelow a level where they can exert sufficient grazing pressure to help control the browntide.The related brown tide species, Aureococcus anophagefferens, has beendemonstrated to inhibit feeding <strong>and</strong> cause mass mortality <strong>of</strong> the mussel Mytilus edulis(Tracey, 1988). Bricelj <strong>and</strong> Kuenstner (1989) concluded that this mortality was due totoxicity <strong>and</strong> not to small size or nutritional inadequacy <strong>of</strong> this phytoplankton species.Laboratory studies also indicate that A. anophagefferens reduces growth <strong>and</strong> causeshigh mortality <strong>of</strong> bay scallop larvae (Gallager et al., 1989). A. anophagefferens wasshown to inhibit the cilliary activity <strong>of</strong> isolated gills <strong>of</strong> some bivalve species such asMercenaria mercenaria <strong>and</strong> Mytilus edulis but not others that were affected by browntide in nature such as Argopecten irradians (Gainey <strong>and</strong> Shumway, 1991). In contrast,the Texas brown tide alga is readily consumed by the dwarf surfclam Mulinia lateraliswithout adverse affects (Montagna et al., 1993), <strong>and</strong> there is no evidence that theTexas brown tide alga is toxic to adults <strong>of</strong> other species <strong>of</strong> invertebrates.It seems likely that the Texas brown tide alga may produce a chemical thatinhibits grazing or growth <strong>of</strong> microzooplankton, <strong>and</strong>/or may act as an allelopathic agentto reduce competition from other phytoplankton species. For example, bothAureococcus anophagefferens <strong>and</strong> the Texas brown tide contain high concentrations <strong>of</strong>dimethylsulfoniopropionate (DMSP) which is a precursor to dimethylsulfide (DMS) <strong>and</strong>acrylic acid (Keller et al., 1989; Stockwell et al., 1993). The role <strong>of</strong> DMSP in grazerinhibition is unclear, however. For example, Phaeocystis pouchetii, which also producesa large amount <strong>of</strong> DMSP (Keller et al., 1989) appears to be consumed by a wide variety<strong>of</strong> zooplankton (Admiraal <strong>and</strong> Venekamp, 1986; Huntley et al., 1987), whereasChrysochromulina polylepis, which produces DMSP reduces growth <strong>and</strong> feeding rates<strong>of</strong> the tintinnid Favella ehrenbergii (Carlsson et al., 1990). The polysaccharide-likelayer on the surface <strong>of</strong> A. anophagefferens contains a bioactive compound responsiblefor the reduction in cilliary beat frequency in bivalve gills (Gainey <strong>and</strong> Shumway, 1991),but no similar compounds have yet been identified in the Texas brown tide alga.Many species <strong>of</strong> harmful <strong>and</strong> nuisance algae are toxic to a variety <strong>of</strong> marineorganisms. Most <strong>of</strong> the toxins associated with harmful algal blooms were first noticedbecause <strong>of</strong> the extensive fish kills they caused or for the human health risk associatedwith consumption <strong>of</strong> contaminated seafood. It is difficult to underst<strong>and</strong> why algalspecies would evolve toxins that were specifically aimed at humans or fish species thatdo not directly consume these algal species. It is possible that in some cases thesetoxins might be substances that have evolved for some other physiological function inthe cell, which coincidentally happen to be toxic to human or marine life. In the cases <strong>of</strong>Aureococcus anophagefferens <strong>and</strong> the Texas brown tide alga, it appears as if toxicsubstances may be targeted at benthic <strong>and</strong> planktonic grazers that feed on these44


species <strong>of</strong> phytoplankton, <strong>and</strong> although there are no direct threats to human healthfrom these species, they may have a pr<strong>of</strong>ound effect on the structure <strong>and</strong> function <strong>of</strong>the ecosystems in which they reside.3. Larval Fish<strong>Red</strong> drum (Sciaenops ocellatus) <strong>and</strong> spotted seatrout(Cynoscion nebulosus)eggs <strong>and</strong> larvae were tested for survival, growth <strong>and</strong> feeding success in the presence<strong>of</strong> brown tide by Dr. Joan Holt <strong>of</strong> the University <strong>of</strong> Texas Marine Science Institute. Testswere carried out both in the laboratory <strong>and</strong> in situ in the Texas Parks <strong>and</strong> Wildlife fishproduction ponds at the GCCA-CPL Marine Development Center. Eggs from laboratoryspawns placed in brown tide concentrations <strong>of</strong> 1 to 1.6 million cells per ml showedsignificantly reduced hatch rate <strong>and</strong> 2 or 3 day survival, compared to controls inseawater (Figure 44). Only 20% <strong>of</strong> spotted seatrout survived to day 3 post hatch. <strong>Red</strong>drum eggs also showed significantly reduced hatch rates <strong>and</strong> reduced survival in bothlaboratory studies <strong>and</strong> in in situ studies at the GCCA-CPL Marine Development Center(Figure 45) although survival rates <strong>of</strong> controls were 100% in the laboratory <strong>and</strong> only20% on day two in the ponds. Eggs did considerably worse when placed in the pondscompared to the laboratory, probably due to high pH values, caused by the effects <strong>of</strong>high photosynthetic rates on the carbonate cycle. Values ranged from 8.9 to 9.6 inponds with brown tide blooms, compared to normal seawater pH <strong>of</strong> 8.2. Studies areunderway to determine the effect <strong>of</strong> increasing pH on red drum egg <strong>and</strong> larval survival.Spotted Sea Trout% Survival1009080706050403020100EggDay1Day2Day3Control<strong>Brown</strong> tideFigure 44. Survival <strong>of</strong> spotted seatrout eggs <strong>and</strong> early larvae in the presence <strong>and</strong>absence <strong>of</strong> brown tide algae at a concentration <strong>of</strong> 1.6 million cells per ml. Unpublisheddata from Dr. G. Joan Holt, University <strong>of</strong> Texas Marine Science Institute.45


<strong>Red</strong> Drum% Survival1009080706050403020100Egg Day 1 Day 2Control (field) <strong>Brown</strong> <strong>Tide</strong> (field) Control (lab) <strong>Brown</strong> <strong>Tide</strong> (lab)Figure 45. Survival <strong>of</strong> red drum eggs <strong>and</strong> larvae in the presence <strong>and</strong> absence <strong>of</strong> browntide in both field <strong>and</strong> laboratory studies. Unpublished data from Dr. G. Joan Holt,University <strong>of</strong> Texas Marine Science Institute.Feeding studies were carried out with larvae from eggs hatched in normalseawater. Spotted seatrout larvae fed at significantly reduced rates in brown tide atages 5 to 7 days post hatch but not at day 14 (Figure 46).Feeding Rates: Spotted Seatrout10.000Prey Eaten/Hour8.0006.0004.0002.000control<strong>Brown</strong> <strong>Tide</strong>0.000Day 5 Day 7 Day 14Figure 46. Feeding rates <strong>of</strong> larval spotted seatrout on rotifers at 5, 7 <strong>and</strong> 14 days posthatch,in brown tide water compared to bloom-free water. Significantly lower feedingrates were observed in the presence <strong>of</strong> the brown tide for 5 <strong>and</strong> 7 day old larvae.Unpublished data from Dr. G. Joan Holt, University <strong>of</strong> Texas Marine Science Institute.46


These results are particularly important for spotted seatrout that spawn in theLaguna Madre where high concentrations <strong>of</strong> brown tide occur, <strong>and</strong> to the red drumproduction ponds at the GCCA-CPL Marine Development Center. The overall effects <strong>of</strong>brown tide on larvae can be interpreted to show that eggs spawned in water with highbrown tide counts will not survive to first feeding <strong>and</strong> larvae that encounter high browntide concentrations in the first two weeks <strong>of</strong> life will suffer high mortality <strong>and</strong> reducedfeeding <strong>and</strong> growth rates.Field data collected by Mr. Scott Holt <strong>of</strong> the University <strong>of</strong> Texas Marine ScienceInstitute demonstrated that spotted seatrout larvae had reduced densities, averagedfrom 7 sites in the upper Laguna Madre (Figure 47), during the brown tide in 1992-1993 compared to collections from four months in 1989, before the brown tide began.Similar data from a site in the Port Mansfield channel (lower panel) where there was nomeasurable brown tide are quite variable, but there is no obvious reduction in larvaldensity between the same two time intervals.(b)Port Mannsfield Channel (East Cut)25(a)Laguna Madre Stations454035Density (No./ 100 m 3 )2015105Density (No./100m 3 )302520151005MarMayJulSepMayAugSepPre- <strong>Brown</strong> <strong>Tide</strong> | Post <strong>Brown</strong> <strong>Tide</strong>(1989) (92 - 93)JulSep0MarMayJulSepMayAugSepJulPre <strong>Brown</strong> <strong>Tide</strong> | Post <strong>Brown</strong> tide(1989) (92 - 93)SepFigure 47. A comparison <strong>of</strong> spotted seatrout larval densities in two areas <strong>of</strong> the LagunaMadre for pre- <strong>and</strong> post brown tide periods. Panel a shows monthly means for sevensites in the Laguna Madre, all <strong>of</strong> which were impacted by the brown tide. Panel b showssimilar data for a site in the Port Mansfield channel where there was no measurablebrown tide. Unpublished data <strong>of</strong> Scott Holt, University <strong>of</strong> Texas Marine ScienceInstitute.When larval density <strong>of</strong> spotted seatrout is plotted against brown tide cell counts,for samples taken in the upper Laguna Madre, there appears to be a trend <strong>of</strong>decreasing larval density with increasing brown tide density (Figure 48). When the47


density <strong>of</strong> black drum larvae is plotted against phytoplankton biomass measured aschlorophyll a concentration (Figure 49) there is a distinct pattern <strong>of</strong> lower larval densityat higher chlorophyll concentrations, which are areas most highly impacted by thebrown tide. In contrast, high chlorophyll concentrations seemed to have little impact onthe density <strong>of</strong> larval anchovy (Figure 50).Spotted Seatrout Larvae60Larval Density (No./100m 3 )504030201000 100 200 300<strong>Brown</strong> <strong>Tide</strong> (cellsx 10 4 /ml)Figure 48. Variations in density <strong>of</strong> larval spotted seatrout with brown tide cell counts forcollections from the Laguna Madre. Unpublished data <strong>of</strong> Scott Holt, University <strong>of</strong> TexasMarine Science Institute.Black Drum Larvae120Larval Density (No./100m 3 )1008060402000 10 20 30 40 50Chlorophyll (µg/l)Figure 49. Variations in density <strong>of</strong> black drum larvae with changes in phytoplanktonbiomass measured as chlorophyll a concentration, for collections in Laguna Madre.Areas <strong>of</strong> high phytoplankton biomass represent brown tide affected areas. Unpublisheddata <strong>of</strong> Scott Holt, University <strong>of</strong> Texas Marine Science Institute.48


Larval AnchovyLarval Density (No./100m 3 )90080070060050040030020010000 20 40 60 80Chlorophyll (µg/l)Figure 50. Variations in density <strong>of</strong> larval bay anchovy with changes in phytoplanktonbiomass measured as chlorophyll a concentration, for collections in Laguna Madre.Areas <strong>of</strong> high phytoplankton biomass represent brown tide affected areas. Unpublisheddata <strong>of</strong> Scott Holt, University <strong>of</strong> Texas Marine Science Institute.4. BenthosThe benthic organisms that live in the Laguna Madre include a large number <strong>of</strong>filter feeding invertebrates that could potentially have an important grazing impact onthe brown tide, given the shallow depth <strong>of</strong> the water column. Also since strongseabreezes from the Gulf <strong>of</strong> Mexico generally keep the water column well mixed,stratification <strong>of</strong> the water column generally does not occur. Thus, layering which couldisolate phytoplankton populations in surface water from potential grazers in the benthosdoes not occur.Biomass <strong>of</strong> the macrobenthos (defined as benthic organisms > 0.5 mm) isgenerally higher in the seagrass beds <strong>of</strong> Laguna Madre than in the mud-bottomhabitats <strong>of</strong> Baffin Bay (Figure 51). There was a decline in benthic biomass from >80 gdry weight per square meter to 40 species per 35 cm 2 before the brown tide began, <strong>and</strong> has remained at40 species or below since the brown tide began through 1992 (Figure 53). Diversity <strong>of</strong>species in Baffin Bay has remained below 10 species per 35 cm 2 both before <strong>and</strong>during the brown tide (Figure 53).49


Macr<strong>of</strong>auna BiomassBaffin Bay <strong>and</strong> Laguna Madre (all stations)Laguna Madre Biomass(g/m 2 )1009080706050403020100876543210Mar-89Jul-89Dec-89Apr-90Oct-90Feb-91Apr-91Jun-91Aug-91Oct-91Dec-91Apr-92Baffin Bay Biomass (g/m 2 )Oct-92Laguna MadreBaffin BayFigure 51. Macrobenthos biomass in Baffin Bay <strong>and</strong> the Laguna Madre from March1989 through October 1992. Unpublished data provided by Dr. Paul Montagna,University <strong>of</strong> Texas Marine Science Institute.Macr<strong>of</strong>auna AbundanceBaffin Bay <strong>and</strong> Laguna Madre(all stations)180000160000Abundance (#/m 2 )140000120000100000800006000040000200000Mar-89Jul-89Dec-89Apr-90Oct-90Feb-91Apr-91Jun-91Aug-91Oct-91Dec-91Apr-92Oct-92Baffin BayLaguna MadreFigure 52. Macrobenthos abundance in Baffin Bay <strong>and</strong> the Laguna Madre from March1989 through October 1992. Unpublished data provided by Dr. Paul Montagna,University <strong>of</strong> Texas Marine Science Institute.50


60Macr<strong>of</strong>auna DiversityBaffin Bay <strong>and</strong> Laguna Madre (all stations)12Laguna Madre Diversity (#Species)50403020101086420Mar-89Jul-89Dec-89Apr-90Oct-90Feb-91Apr-91Jun-91Aug-91Oct-91Apr-92Oct-92Baffin Bay Diversity(# Species)0Baffin BayLaguna MadreFigure 53. Macrobenthos diversity in Baffin Bay <strong>and</strong> the Laguna Madre from March1989 through October 1992. Unpublished data provided by Dr. Paul Montagna,University <strong>of</strong> Texas Marine Science Institute.The brown tide appears to have caused a reduction in macrobenthos biomass<strong>and</strong> abundance, <strong>and</strong> a reduction in species diversity in seagrass habitats. To the extentthat this reduction is in populations <strong>of</strong> filter feeding invertebrates (as opposed todeposit <strong>and</strong> detritus feeders), this reduction in benthos may have reduced grazingpressure on the brown tide alga (Montagna et al., 1993). The filter feeding molluskMulinia lateralis nearly disappeared from Baffin Bay during 1990 <strong>and</strong> 1991 (Montagnaet al., 1993). Montagna et al. (1993) found that adult Mulinia lateralis consumed thebrown tide alga as well as three cultured species <strong>of</strong> phytoplankton. It was furtherestimated that at its peak density before the brown tide began (800 per square meter),its maximum clearance rate <strong>of</strong> 10 ml per hour would allow this population <strong>of</strong> clams toclear a 1.2 m water column in 150 hours or about 6-7 days. However, since the browntide can double its population in 1-2 days under good growing conditions (Stockwell etal., 1993), even under ideal conditions these benthic filter feeders can not controlbrown tide populations.5. Adult fish <strong>and</strong> shellfish populations.The extended brown tide bloom had no apparent prolonged effect onpopulations <strong>of</strong> adult fish <strong>and</strong> shellfish . Game fish populations (red drum, black drum,<strong>and</strong> spotted seatrout) were sampled by gill net in the spring <strong>and</strong> fall (Figure 54). Thereis no apparent decline in the populations <strong>of</strong> these important recreational species; in factthere appears to be a substantial increase in black drum since the initiation <strong>of</strong> thebrown tide. Small forage fish (bay anchovy <strong>and</strong> striped mullet) <strong>and</strong> commerciallyimportant shellfish (brown shrimp <strong>and</strong> blue crab) populations were sampled by bag51


seine during the appropriate season (Figure 55). No prolonged decline in abundancehas occurred since the initiation <strong>of</strong> the brown tide. Each population did experience atemporary decline in abundance close to the time <strong>of</strong> initiation <strong>of</strong> the brown tide. Thisdecrease in numbers may represent the large fish kill caused by the December 1989freeze, <strong>and</strong>/or the effects <strong>of</strong> the extended drought in 1989.ANumber/hr54.543.532.521.510.508284Spring Gill Net Data(15 April-15 June)8688Year909294<strong>Red</strong> Drum Spotted Seatrout Black DrumBNumber/hr54.543.532.521.510.508284Fall Gill Net Data(15 Sept.-15 Nov.)8688Year909294<strong>Red</strong> Drum Spotted Seatrout Black DrumFigure 54. Seasonal average <strong>of</strong> three game fish species (red drum, spotted seatrout<strong>and</strong> black drum) based on spring (a) <strong>and</strong> fall (b) gill net sets at several r<strong>and</strong>omlyselected locations within the Laguna Madre. Data provided by Larry McEachron <strong>and</strong>Kyle Spiller, Texas Parks <strong>and</strong> Wildlife Department.52


ABay AnchovyBStripped Mullet250600Number/hectare20015010050Number/hectare50040030020010000798183858789919395798183858789919395YearYearCNumber/hectare160014001200100080060040020007981<strong>Brown</strong> Shrimp83858789Year919395DNumber/hectare250200150100500798183Blue Crab858789Year919395Figure 55. Seasonal average catch based on bag seine data at several r<strong>and</strong>omlychosen sites within the Laguna Madre. Collection seasons for each species were: A.bay anchovy March-June, B. striped mullet May-June, C. brown shrimp April-July, D.blue crab March-June. Data provided by Larry McEachron <strong>and</strong> Kyle Spiller, TexasParks <strong>and</strong> Wildlife Department.The lack <strong>of</strong> adverse effects <strong>of</strong> brown tide on adult fish populations is in sharpcontrast to its apparent effects on larval fish populations. In a nearly enclosed systemsuch as the Laguna Madre, it might be expected that if larval fish abundance declined,adult fish populations would soon decline as well. Several possible explanations existfor this apparent paradox, but there is at present no way to determine which, if any, <strong>of</strong>the possible explanations is most plausible. For example, the brown tide may haveenhanced the survival <strong>of</strong> post-larval, juvenile fish by reducing their susceptibility tovisual predation by larger fish in the reduced visibility conditions associated with thebrown tide. Another contributing factor may have been reduced sports fishing pressurein brown tide affected areas, due to reduced success experienced by many anglers inthe low visibility waters. In addition, there have been no severe freezes affecting theLaguna Madre since the brown tide began, <strong>and</strong> increased fish populations may reflectrecovery <strong>of</strong> populations from the severe freezes in the 1980's. A ban on commercial netfisheries since the early 1980’s also could lead to the increase in adult population. It isalso possible that additional fish may have migrated into the Laguna Madre fromadjacent habitats not affected by the brown tide.53


IV. Identification <strong>of</strong> Probable CausesLaguna Madre is vulnerable to ecosystem disruption due to its lack <strong>of</strong> consistentfreshwater inflow, making hypersaline conditions likely during periods <strong>of</strong> extendeddrought. This propensity for extreme hypersaline conditions has been reducedsomewhat in recent decades by the construction <strong>of</strong> the Gulf Intracoastal Waterway.This channel allows, at least, a minimum amount <strong>of</strong> water exchange with the Gulf <strong>of</strong>Mexico.The long turnover times estimated for waters in Baffin Bay <strong>and</strong> Laguna Madreclearly contribute to the persistence <strong>of</strong> the brown tide bloom. The time required toexchange all the water contained within the Laguna Madre with adjacent areas in theGulf <strong>of</strong> Mexico <strong>and</strong> Corpus Christi Bay is on the order <strong>of</strong> one year. This makes itdifficult to disperse <strong>and</strong> dilute an algal bloom for a species that can easily double itspopulation in a matter <strong>of</strong> days if sufficient nutrients are available. There has been muchpublic discussion <strong>of</strong> the possibility <strong>of</strong> eliminating the brown tide through increasing thecirculation <strong>of</strong> the Laguna Madre by raising the John F. Kennedy Causeway or byopening new passes between the Gulf <strong>of</strong> Mexico <strong>and</strong> the Laguna Madre. In order forthese engineering projects to have an impact on the brown tide, they would have tosubstantially reduce the residence time <strong>of</strong> water in the Laguna Madre, from one year toa matter <strong>of</strong> weeks, as is the case in other Texas bays where the brown tide hasbloomed but not persisted (Shormann, 1992). Recent studies <strong>of</strong> the impact <strong>of</strong> raisingthe Kennedy causeway for the Texas Department <strong>of</strong> Transportation by Shiner, Moseley<strong>and</strong> Associates indicate that the cost <strong>of</strong> raising the Kennedy causeway through most <strong>of</strong>its length would be on the order <strong>of</strong> $50 million dollars, <strong>and</strong> that this major engineeringproject would have a minimal impact on the turnover time <strong>of</strong> water in the upper LagunaMadre. The total amount <strong>of</strong> water exchange may increase slightly, but the major impactwould be to spread this exchange over a wider area, compared to the two channels thatnow h<strong>and</strong>le most <strong>of</strong> the exchange. Other suggested projects, such as reopeningYarborough Pass would also be expensive, initially <strong>and</strong> in terms <strong>of</strong> maintenance costs,<strong>and</strong> there is little evidence that they would increase circulation enough to eliminatedthe brown tide. By comparison, the lower Laguna Madre, which has better circulationthan the upper Laguna Madre due to major passes at Port Isabel to the south <strong>and</strong> PortMansfield to the north, still has persistent blooms <strong>of</strong> the brown tide. It would beinteresting to determine if the brown tide in lower Laguna Madre would persist withoutthe major pulse <strong>of</strong> brown tide pushed south each winter by the north winds <strong>of</strong> strongcold fronts.Laguna Madre fish populations are vulnerable to extensive die-<strong>of</strong>fs during thepassage <strong>of</strong> severe cold fronts. Due to a shallow (1.2 m) average depth (Armstrong,1987), few deep water refuges <strong>and</strong> no direct passes to the Gulf <strong>of</strong> Mexico, extreme coldfronts, with below freezing temperatures, during winter can cause dramatic drops inwater temperature resulting in mass mortality <strong>of</strong> fish <strong>and</strong> benthic invertebrates (DeYoe<strong>and</strong> Suttle, 1994). It has been calculated that the nutrients released from the fish <strong>and</strong>54


invertebrate die-<strong>of</strong>f in the freezes <strong>of</strong> December 1989 would have provided sufficientnutrients to explain the initiation <strong>of</strong> the brown tide bloom in January 1990. However,there has been much public discussion <strong>of</strong> the possible role <strong>of</strong> agricultural run<strong>of</strong>f fromthe King Ranch in the initiation <strong>and</strong> maintenance <strong>of</strong> the brown tide. A large amount <strong>of</strong>l<strong>and</strong> adjacent to the Laguna Madre is alleged to have been changed from grazing toagricultural use prior to the initiation <strong>of</strong> the brown tide, <strong>and</strong> concerns have been raisedthat run<strong>of</strong>f from this agricultural l<strong>and</strong> may have contributed to the initiation <strong>of</strong> the bloom,<strong>and</strong> may be a continuing source <strong>of</strong> nutrients helping to maintain the bloom. Thispossibility has not yet been explored, but is currently under investigation by the TexasAgricultural Experiment Station <strong>and</strong> Corpus Christi Bay National Estuary Program.It is clear from studies <strong>of</strong> changes in the Laguna Madre ecosystem before <strong>and</strong>during the initiation <strong>of</strong> the brown tide, that changes in biotic components <strong>of</strong> the LagunaMadre ecosystem also indirectly contributed to the initiation <strong>of</strong> the brown tide. A period<strong>of</strong> extended drought in 1989 had raised the salinity in the upper reaches <strong>of</strong> Baffin Bayto >60 ppt. This is nearly twice the salinity <strong>of</strong> normal seawater, <strong>and</strong> many marinespecies cannot survive under these high salinity conditions. Population <strong>of</strong> zooplankton<strong>and</strong> benthic organisms, the major feeders on water column algae, were declining priorto the onset <strong>of</strong> the brown tide. With depressed populations <strong>of</strong> planktonic <strong>and</strong> benthicgrazers, it became much easier for a major bloom such as the brown tide to get started,since the normal biological controls on algal growth were impaired. Once the bloombecame established, the algae was capable <strong>of</strong> growing faster than the grazers couldremove it.There are indications that the brown tide alga is toxic to some species <strong>of</strong>planktonic grazers, <strong>and</strong> this resultant reduction in grazing pressure may help the browntide bloom persist. In particular, it appears that brown tide concentrations above athreshold concentration <strong>of</strong> approximately one million cells per ml are either directlytoxic to some species <strong>of</strong> zooplankton or inhibit their growth (Buskey <strong>and</strong> Hyatt, 1995).Additional research on the potentially toxic effects <strong>of</strong> brown tide on marine organisms,including their planktonic <strong>and</strong> benthic grazers <strong>and</strong> larval fish is needed.V. Identification <strong>of</strong> Data <strong>and</strong> Information GapsCompared to the body <strong>of</strong> information available on blooms <strong>of</strong> the East Coastbrown tide alga Aureococcus anophagefferens, very little research has been performedon the Texas brown tide alga to date. Initial strong support <strong>of</strong> research efforts by theTexas Higher Education Coordinating Board (THECB) have provided a wealth <strong>of</strong>information about conditions before <strong>and</strong> during the initiation <strong>of</strong> the Texas brown tide.This is unique among studies <strong>of</strong> harmful algal bloom phenomena; usually studies donot begin until the bloom is already under way <strong>and</strong> recognized as a potential problem.The studies funded by the THECB also clearly documented the initial changes in theLaguna Madre ecosystem caused by the bloom, but much remains to be learned.55


In particular, it is clear that very little is known about the autecology <strong>of</strong> the algaitself. For scientists, one clear indication <strong>of</strong> the lack <strong>of</strong> interest in this algal species itselfis that after over five <strong>and</strong> one half years <strong>of</strong> algal bloom, this species has still not beentaxonomically described. It is an embarrassment to everyone working with the browntide that we still can only refer to it as the "brown tide alga". A taxonomic description <strong>of</strong>this species is critical to scientific communication, so that scientists can potentiallyidentify this species from other parts <strong>of</strong> the world. Fortunately, a taxonomic descriptionwill soon be published (DeYoe, personal communication).Most studies <strong>of</strong> the brown tide alga have been field studies <strong>of</strong> biomassdistributions <strong>and</strong> primary production rates (Stockwell <strong>and</strong> Whitledge, unpublished), <strong>and</strong>the only studies <strong>of</strong> the alga under laboratory conditions are those <strong>of</strong> DeYoe <strong>and</strong> Suttle(1994) who provided important insights into the nutrient requirements <strong>of</strong> this species,<strong>and</strong> DeYoe et al. (1995) that examined gene sequences to examine the taxonomicrelationship <strong>of</strong> this species to other algal taxa. Additional isolates <strong>of</strong> this alga need tobe brought into culture, <strong>and</strong> a systematic study <strong>of</strong> the factors affecting its survival <strong>and</strong>growth needs to be undertaken. Since this species, like the East Coast brown tideAureococcus anophagefferens is not readily cultured under conditions suitable formany other phytoplankton species (e.g. it does not grow well in f/2 media) a study <strong>of</strong>micronutrient requirements, including trace metals <strong>and</strong> vitamins could provide importantinsights into the conditions that favor the growth <strong>of</strong> this species.VI. Recommendations for priority research <strong>and</strong> monitoring effortsA. Research1. To study <strong>and</strong> define the autecology <strong>of</strong> the brown tide alga in terms <strong>of</strong> the physical<strong>and</strong> chemical factors that affect its growth. Specifically to look at the effects <strong>of</strong>temperature, salinity, light intensity, macro- <strong>and</strong> micronutrients on the growth rate <strong>of</strong> thebrown tide.2. To identify possible toxins <strong>and</strong>/or inhibitory chemicals produced by the brown tideorganism, their affects on the growth <strong>of</strong> other phytoplankton species (possibleallelopathic effects) <strong>and</strong> on the survival, growth <strong>and</strong> feeding <strong>of</strong> potential grazers on thebrown tide, including both planktonic (micro- <strong>and</strong> mesozooplankton) <strong>and</strong> benthicgrazers. It should be emphasized that potential toxic effects should be examined on alllife history stages <strong>of</strong> grazers, including the planktonic larvae <strong>of</strong> benthic macr<strong>of</strong>auna.Further study <strong>of</strong> the potential toxic effects <strong>of</strong> the brown tide on larval fish is alsoneeded.3. A method for positively identifying the Texas brown tide alga needs to be developed,such as a fluorescently labeled polyclonal antibody marker with high specificity for thebrown tide, or a fluorescently labeled genetic marker, specific to a particular uniquegenetic sequence in the DNA or RNA <strong>of</strong> the brown tide alga. Using these methods, thebrown tide alga could be positively identified at low cell concentrations to determine if it56


is a common component <strong>of</strong> phytoplankton communities throughout the coastal zone <strong>of</strong>Texas <strong>and</strong> other areas <strong>of</strong> the Gulf <strong>of</strong> Mexico. Such a marker would also be useful foridentifying brown tide cells in the stomach contents <strong>of</strong> grazers from protozoa tomacr<strong>of</strong>auna, to help determine the extent to which natural grazer populations feed onthe brown tide.4. To investigate the use <strong>of</strong> various biological agents to help control brown tidepopulations, including planktonic <strong>and</strong> benthic grazers <strong>and</strong> possible pathogens such asviruses specific to the brown tide alga. It is doubtful if these agents could be used toeliminate the brown tide during periods <strong>of</strong> its peak abundance <strong>and</strong> extent, but they maybe useful for breaking up seed populations during periods <strong>of</strong> low natural abundance <strong>of</strong>the brown tide <strong>and</strong> could also prove useful as control agents in mariculture facilitieswhere the area that needs to be treated is small.5. To investigate the role <strong>of</strong> non-point source pollutants, including fertilizer <strong>and</strong>pesticide run<strong>of</strong>f from agricultural areas on the growth <strong>and</strong> maintenance <strong>of</strong> brown tideblooms <strong>and</strong> associated plankton populations.6. <strong>Historical</strong> studies <strong>of</strong> the Laguna Madre should be pursued to determine the historicalimportance <strong>of</strong> seagrass-based primary production versus phytoplankton based primaryproduction in the Laguna Madre over the past several thous<strong>and</strong> years. If this can betied in to data on paleosalinity <strong>and</strong> paleotemperature, as well as information on theoccurrence <strong>of</strong> major hurricanes (via the s<strong>and</strong> deposits they leave in the sedimentrecord) the importance <strong>of</strong> circulation with the Gulf <strong>of</strong> Mexico on the dominance <strong>of</strong>seagrass or phytoplankton can be determined. If biomarkers specific to diageniccompounds associated with the brown tide alga can be identified, it may also bepossible to determine if the brown tide alga has bloomed for extended periods in thepast.B. Monitoring1. Continued monitoring <strong>of</strong> the areal distribution <strong>of</strong> seagrass beds in the Laguna Madremust be made so that the impact <strong>of</strong> the brown tide on the Laguna Madre ecosystem canbe assessed.2. Continued monitoring <strong>of</strong> the effects <strong>of</strong> the brown tide on the structure <strong>and</strong> function <strong>of</strong>the Laguna Madre ecosystem, especially finfish populations, will provide us with abasis for cost/benefit analysis for potential management strategies for dealing with thebrown tide.3. Continued monitoring <strong>of</strong> the distribution <strong>and</strong> abundance <strong>of</strong> the brown tide willincrease our knowledge <strong>of</strong> the factors influencing the persistence <strong>of</strong> this bloom.57


4. A monitoring plan should be designed to quantify non-point source pollutants,including pesticides <strong>and</strong> fertilizers, <strong>and</strong> help identify their sources.58


VII. Literature CitedAdmiraal, W., Venekamp, L.A.H. (1986). Significance <strong>of</strong> tintinnid grazing during blooms<strong>of</strong> Phaeocystis pouchetii (Haptophyceae) in Dutch coastal waters. Neth. J. Sea. Res.20: 61-66.Andersen, R.A., Saunders, G.W., Paskind, M.P., Sexton, J.P. (1993). Ultrastructure<strong>and</strong> 18s r RNA gene sequence for Pelagomonas calceolata gen. et sp. nov. <strong>and</strong> thedescription <strong>of</strong> a new algal class, the Pelagophyceae classis nov. J. Phycol. 29: 701-715.Armstrong, N.E. (1982). Responses <strong>of</strong> Texas estuaries to freshwater inflow. pp. 103-120. In: V.S. Kennedy (ed) Estuarine Comparisons. Academic Press, New York.Armstrong, N.E. (1987). The ecology <strong>of</strong> open - bay bottoms <strong>of</strong> Texas: a communitypr<strong>of</strong>ile. U.S. Fish. Wildl. Serv. Biol. Rep. 85 (7.12): 104 pp.Berggreen, U., Hansen, B., Kiorboe, T. (1988). Food size spectra, ingestion <strong>and</strong> growth<strong>of</strong> the copepod Acartia tonsa during development: implications for determination <strong>of</strong>copepod production. Mar. Biol. 99: 341-352.Bricelj, V.M., Epp, J., Malouf, R.E. (1987). Intraspecific variation in reproductive <strong>and</strong>somatic growth cycles <strong>of</strong> bay scallops Argopecten irradians. Mar. Ecol. Prog. Ser. 36:123-137.Bricelj, V.M., Kuenstner, S.H. (1989). Effects <strong>of</strong> the "brown tide" on the feeding,physiology <strong>and</strong> growth <strong>of</strong> bay scallops <strong>and</strong> mussels. In: Cosper, E.M., Bricelj, V.M.,Carpenter, E.J. (eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong> Estuarine Studies 35,Springer-Verlag, Berlin. p. 491-509.Buskey, E.J. (1995). Growth <strong>and</strong> bioluminescence <strong>of</strong> Noctiluca scintillans on varyingalgal diets. J. Plank. Res. 17: 29-40.Buskey, E.J., Hyatt, C.J. (1995). Effects <strong>of</strong> the Texas (USA) ‘brown tide’ alga onplanktonic grazers. Mar. Ecol.Prog.Ser. 126: 285-292.Buskey, E.J., Stockwell, D.A. (1993). Effects <strong>of</strong> a persistent "brown tide" onzooplankton populations in the Laguna Madre <strong>of</strong> South Texas. In Smayda, T.J.,Shimizu, Y. (eds.), Toxic Phytoplankton Blooms in the Sea. Proc. 5th Int. Conf. ToxicMarine Phytopl. Amsterdam, Elsevier. p. 659-666.Carlsson, P., Graneli, E., Olsson, P. (1990). Grazer elimination through poisoning: one<strong>of</strong> the mechanisms behind Chrysochromulina polylepis blooms? In: Graneli, E.,Sundstrom, B., Edler, L., Anderson, D.M. (eds). Toxic Marine Phytoplankton. Elsevier,New York, p. 116-122.59


Caron, D.A., Lim, E., Kunze, H., Cosper, E.M., Anderson, D.M. (1989). Trophicinteractions between nano- <strong>and</strong> microzooplankton <strong>and</strong> the "brown tide". In: Cosper,E.M., Bricelj, V.M., Carpenter, E.J. (eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong>Estuarine Studies 35, Springer-Verlag, Berlin. p. 265-294.Cosper, E.M., Carpenter, E.J., Cotrell, M. (1989). Primary productivity <strong>and</strong> growthdynamics <strong>of</strong> the "brown tide" in Long Isl<strong>and</strong> embayments. In: Cosper, E.M., Bricelj,V.M., Carpenter, E.J. (eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong> EstuarineStudies 35, Springer-Verlag, Berlin. p. 139-158.Cosper, E.M., Dennison, W.C., Carpenter, E.J., Bricelj, V.M., Mitchell, J.G., Kuenstner,S.H., Colflesh, D., Dewey, M. (1987). Recurrent <strong>and</strong> persistent brown tide bloomsperturb a coastal marine ecosystem. Estuaries 10: 284-290.Cosper, E.M., Dennison, W.C., Milligan, A., Carpenter, E.J., Lee, C. Holzapfel, J.,Milanese, L. (1989b). An examination <strong>of</strong> environmental factors important to initiating<strong>and</strong> sustaining brown tide blooms. In: Cosper, E.M., Bricelj, V.M., Carpenter, E.J. (eds).Novel Phytoplankton Blooms. Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag,Berlin. p. 317-340.Cosper, E.M., Garry, R.T., Milligan A., Doall, M.H. (1993). Iron, selenium <strong>and</strong> citric acidare critical to the growth <strong>of</strong> the brown tide microalga, Aureococcus anophagefferens. In:Toxic Phytoplankton Blooms in the Sea. T.J. Smayda <strong>and</strong> Y. Shimizu (eds.), Proc. 5thInt. Conf. Toxic Marine Phytopl. Amsterdam, Elsevier, pp 667-673.Dagg, M.J., Whitledge, T.E. (1991). Concentrations <strong>of</strong> copepod nauplii associated withthe nutrient-rich plume <strong>of</strong> the Mississippi River. Cont. Shelf Res. 11: 1409-1423Daily, J.A., Kana, J.C., McEachron, L.W. (1991). <strong>Trends</strong> in selective abundance <strong>and</strong>size <strong>of</strong> selected finfishes <strong>and</strong> shellfishes along the Texas coast: November 1975-December 1990. Tex. Parks Wildl. Dept. Manag. Data Ser. 74: 1 - 128.Dennison, W.C. (1987). Effects <strong>of</strong> light on seagrass photosynthesis, growth <strong>and</strong> depthdistribution. Aquat. Bot. 27: 3-14.Dennison, W.C., Marshall, G.J., Wig<strong>and</strong>, W. (1989). Effect <strong>of</strong> "brown tide" shading oneelgrass (Zostera marina) distributions. In: Cosper, E.M., Bricelj, V.M., Carpenter, E.J.(eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 675-692.DeYoe, H.R., Suttle, C.A. (1994). The inability <strong>of</strong> the Texas "brown tide" alga to usenitrate <strong>and</strong> the role <strong>of</strong> nitrogen in the initiation <strong>of</strong> a persistent bloom <strong>of</strong> this organism. J.Phycol. 30: 800-806.60


DeYoe, H.R., Chan, A.M., Suttle, C.A. (1995). Phylogeny <strong>of</strong> Aureococcusanophagefferens <strong>and</strong> a morphologically similar bloom forming algae from Texas asdetermined by 18s rRNA gene sequence analysis. J. Phycol. 31: 413-418.Duarte, C.M. (1991). Seagrass depth limits. Aquat. Bot. 40: 363-377.Dunton, K.H. (1994). Seasonal growth <strong>and</strong> biomass <strong>of</strong> the subtropical seagrassHalodule wrightii in relation to continuous measurements <strong>of</strong> underwater irradiance.Mar. Bio. 120: 479-489.Durbin, A.G., Durbin, E.G. (1989). Effect <strong>of</strong> the "brown tide" on feeding, size <strong>and</strong> egglaying rate <strong>of</strong> adult female Acartia tonsa. In: Cosper, E.M., Bricelj, V.M., Carpenter, E.J.(eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 625-646.Dzurica, S., Lee, C., Cosper, E.M., Carpenter, E.J. (1989). Role <strong>of</strong> environmentalvariables, specifically organic compounds <strong>and</strong> micronutrients, in the growth <strong>of</strong> theChrysophyte Aureococcus anophagefferens. In: Cosper, E.M., Bricelj, V.M., Carpenter,E.J. (eds). Novel Phytoplankton Blooms. Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 229-252.Fenchel, T. (1980). Suspension feeding in ciliated protozoa: functional response <strong>and</strong>particle size selection. Microb. Ecol. 6: 1-11.Flint, R.W. (1984) Phytoplankton production in the Corpus Christi Bay estuary. Contr.Mar. Sci. 27: 65-83.Gainey, L.F., Jr.,Shumway, S.E. (1991). The physiological effect <strong>of</strong> Aureococcusanophagefferens ("brown tide") on the lateral cilia <strong>of</strong> bivalve mollusks. Biol. Bull. 181:298-306.Gallager, S.M., Stoecker, D.K., Bricelj, V.M. (1989). Effects <strong>of</strong> the brown tide alga ongrowth, feeding physiology <strong>and</strong> locomotory behavior <strong>of</strong> scallop larvae (Argopectenirradians). In: Novel phytoplankton Blooms, E.M. Cosper, V.M. Bricelj <strong>and</strong> E.J.Carpenter (eds). Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 511-541.Graneli, E., Olsson, P., Carlsson, P., Sundstrom, B., Lindahl, O. (1989). From anoxia t<strong>of</strong>ish poisoning: the last ten years <strong>of</strong> phytoplankton blooms in Swedish marine waters. In:Novel phytoplankton Blooms, E.M. Cosper, V.M. Bricelj <strong>and</strong> E.J. Carpenter (eds).Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 407-428.Gunter, G. (1945) sime characteristics <strong>of</strong> ocean waters <strong>and</strong> Laguna Madre. TexasGame <strong>and</strong> Fish 3: 7, 19, 21,22.61


Holm-Hansen, O., Lorenzen, C.J., Holmes, R., Strickl<strong>and</strong>, J.D.H. (1965). Fluormetricdetermination <strong>of</strong> chlorophyll. J. Cons. Perm. Int. Explor. Mer. 30: 3-15.Holt, S.A., Pratt, C.M., Arnold C.R. (1994). Evaluation <strong>of</strong> larval red drum releases forstock enhancement. Annual Proceedings <strong>of</strong> the Texas Chapter, American FisheriesSociety. 16: 16-21.Huntley, M., T<strong>and</strong>e, K., Eilertsen, H.C. (1987). On the trophic fate <strong>of</strong> Phaeocystispouchetii Hariot. II. Grazing rates <strong>of</strong> Calanus hyperboreous. J. Exp. Mar. Biol. Ecol.110: 197-212.Joint, I.R. (1986). Physiological ecology <strong>of</strong> picoplankton in various oceanic provincesIn: Photosynthetic picoplankton. Platt, T. (ed). Can. Bull. Fish. Aquat. Sci. 214:287-309.Kahn, J. <strong>and</strong> Rockel, M. (1988). Measuring the economic effects <strong>of</strong> brown tides. J.Shellfish Res. 7: 677-682.Keller, A.A. <strong>and</strong> Rice, R.L. (1989). Effects <strong>of</strong> nutrient enrichment on natural populations<strong>of</strong> the brown tide phytoplankton Aureococcus anophagefferens (Chrysophyceae). J.Phycol. 25:636-646.Keller, M.D., Bellows, W.K., Guillard, R.R.L. (1989). Dimethylsulfide production <strong>and</strong>marine phytoplankton: an additional impact <strong>of</strong> unusual blooms. In: Novel phytoplanktonBlooms, E.M. Cosper, V.M. Bricelj <strong>and</strong> E.J. Carpenter (eds). Coastal <strong>and</strong> EstuarineStudies 35, Springer-Verlag, Berlin. p. 511-541.Kenworthy, W.J., Fonseca,M.S., Dipiero, S.J. (1991). Defining the ecological lightcompensation point for seagrasses Halodule wrightii <strong>and</strong> Syringodium filiforme fromlong - term submarine light regime monitoring in the southern Indian River. In: The lightrequirements <strong>of</strong> seagrasses: proceedings <strong>of</strong> a workshop to examine the capability <strong>of</strong>water quality criteria, st<strong>and</strong>ards <strong>and</strong> monitoring programs to protect seagrasses,Kenworthy, W.J., Haunert, D. (eds). NOAA Technical Memor<strong>and</strong>um NMFS-SEFC-287,p. 106-113.Kenworthy, W.J., Haunert, D.E. (1991). The light requirements <strong>of</strong> seagrasses:proceedings <strong>of</strong> a workshop to examine the capability <strong>of</strong> water quality criteria, st<strong>and</strong>ards<strong>and</strong> monitoring programs to protect seagrasses. NOAA Technical Memor<strong>and</strong>um NMFS-SEFC-287.L<strong>and</strong>ry, M.R., Hassett, R.P. (1982). Estimating the grazing impact <strong>of</strong> marine microzooplankton.Mar. Biol. 67: 283.Lonsdale, D.J., Bricelj, M. <strong>and</strong> Taylor, G. (in prep). Aureococcus anophagefferens:causes <strong>and</strong> ecological consequences <strong>of</strong> "<strong>Brown</strong> <strong>Tide</strong>".62


Milligan, A.J. (1992). An investigation <strong>of</strong> factors contributing to blooms <strong>of</strong> the "browntide" Aureococcus anophagefferens (Chysophyceae) under nutrient saturated, lightlimited conditions. M.S. Thesis, State University <strong>of</strong> New York at Stony Brook, 84pp.Montagna, P.A., Stockwell, D.A., Kalke, R.D. (1993). Dwarf surfclam Mulinia lateralis(Say, 1822) populations <strong>and</strong> feeding during the Texas brown tide event. J. ShellfishRes. 12: 433-442.Nuzzi, R., Waters, RM. (1989). The spatial <strong>and</strong> temporal distribution <strong>of</strong> "brown tide" ineastern Long Isl<strong>and</strong> Sound. In: Novel phytoplankton Blooms, E.M. Cosper, V.M. Bricelj<strong>and</strong> E.J. Carpenter (eds). Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p.117-138.Odum, H.T. <strong>and</strong> Wilson, R. (1962) Futher studies on reaeration <strong>and</strong> metabolism <strong>of</strong>Texas bays, 1958-1960. Publ. Inst. Mar. Sci. Univ. Texas. 8: 23-55.Olsen, P.S. (1989). Development <strong>and</strong> distribution <strong>of</strong> a brown-water algal bloom inBarneget Bay, New Jersey with perspective on resources <strong>and</strong> other red tides in theregion. In: Novel phytoplankton Blooms, E.M. Cosper, V.M. Bricelj <strong>and</strong> E.J. Carpenter(eds). Coastal <strong>and</strong> Estuarine Studies 35, Springer-Verlag, Berlin. p. 189-212.Onuf, C.P. (1995). Seagrass responses to long-term light reduction by brown tide inupper Laguna Madre, Texas: Distribution <strong>and</strong> biomass patterns. Mar. Ecol. Prog. Ser.(Submitted).Onuf, C.P. (1991). Light requirements <strong>of</strong> Halodule wrightii, Syringodium filiforme, <strong>and</strong>Halophila engelmanni in a heterogeneous <strong>and</strong> variable environment inferred from long -term monitoring. In: The light requirements <strong>of</strong> seagrasses: proceedings <strong>of</strong> a workshopto examine the capability <strong>of</strong> water quality criteria, st<strong>and</strong>ards <strong>and</strong> monitoring programsto protect seagrasses, Kenworthy, W.J., Haunert, D. (eds). U.S. Department <strong>of</strong>Commerce, National Oceanic <strong>and</strong> Atmospheric Administration, National MarineFisheries, NOAA Technical Memor<strong>and</strong>um MMFS-SEFC-287, p. 95-105.Pierce, R.W., Turner, J.T. (1992). Ecology <strong>of</strong> planktonic ciliates in marine food webs.Rev. Aquat. Sci. 67: 1710-1714.Quammen, M.L., Onuf, C.P. (1993). Laguna Madre: seagrass changes continuedecades after salinity reduction. Estuaries 16: 303-311.Shormann, D.E. (1992). The effects <strong>of</strong> freshwater inflow <strong>and</strong> hydrography on thedistribution <strong>of</strong> brown tide in South Texas. M.A. Thesis, Department <strong>of</strong> Marine Science,University <strong>of</strong> Texas at Austin,, 112 pp.Sieburth, J. McN., Johnson, P.W. (1989). Picoplankton ultrastructure: A decade <strong>of</strong>preparation for the brown tide alga, Aureococcus anophagefferens. In: Novel63


phytoplankton Blooms, E.M. Cosper, V.M. Bricelj <strong>and</strong> E.J. Carpenter (eds). Coastal <strong>and</strong>Estuarine Studies 35, Springer-Verlag, Berlin. p. 1-22.Sieburth, J. McN., Johnson, P.W., Hargraves, P.E. (1988). Ultrastructure <strong>and</strong> ecology<strong>of</strong> Aureococcus anophagefferens gen. et sp. nov. (Chrysophyceae); the dominantpicoplankter during a bloom in Narragansett Bay, Rhode Isl<strong>and</strong>, summer 1985. J.Phycol. 24: 416-425.Simmons, E.G. (1957). An ecological survey <strong>of</strong> the upper Laguna Madre <strong>of</strong> Texas.Publ. Inst. Nar. Sci. Univ. Tex. 4: 156-200.Smayda, T.J., Villareal, T.A. (1989). The 1985 "brown tide" <strong>and</strong> the open phytoplanktonniche in Narragansett Bay during summer. In: Novel Phytoplankton Blooms, E.M.Cosper, V.M. Bricelj <strong>and</strong> E.J. Carpenter (eds). Coastal <strong>and</strong> Estuarine Studies 35,Springer-Verlag, Berlin. p. 159-188.Smith, N.P. (1985). Numerical simulation <strong>of</strong> bay-shelf exchanges with a onedimensionalmodel. Contr. Mar. Sci. 28: 1-15.Stockwell, D.A. (1989). Nitrogen Processes Study (NIPS): Effects <strong>of</strong> freshwater inflowon the primary production <strong>of</strong> a texas Coastal Bay System. Technical Report No. TR89-010, (UT Marine Science Institute, Port Aransas, Texas) 180 pp.Stockwell, D.A., Buskey, E.J., Whitledge, T.E. (1993). Studies on conditions conduciveto the development <strong>and</strong> maintenance <strong>of</strong> a persistent "brown tide" in Laguna Madre,Texas. In: Toxic Phytoplankton Blooms in the Sea. T.J. Smayda <strong>and</strong> Y. Shimizu (eds.),Proc. 5th Int. Conf. Toxic Marine Phytopl. Amsterdam, Elsevier. p. 693-698.Tracey, G.A. (1988). Feeding reduction, reproductive failure, <strong>and</strong> mortality in Mytilusedulis during the 1985 "brown tide" in Narragansett Bay, Rhode Isl<strong>and</strong>. Mar. Ecol. Prog.Ser. 50: 73-81.Virnstein, R.W. (1982) Leaf growth rate <strong>of</strong> the seagrass Halodule wrightiiphotographically measured in situ. Aquat. Bot. 12: 209-218.Wenczel, P. (1987). Proposed bay scallop reseeding activities for 1987. Reportsubmitted to New York State Development Corporation, New York, 28 pp.Whitledge, T.E., Malloy, S.C., Patton, C.J., Wirick, C.D. (1981). Automated nutrientanalyses in sewater. Format Report 51398, Brookhaven National Laboratory, Upton,NewYork, 216pp.Whitledge, T.E. (1989). Nitrogen Processes Study (NIPS): Nutrient distributions <strong>and</strong>dynamics in Lavaca, San Antonio <strong>and</strong> Nueces/Corpus Christi Bays in relation to64


freshwater inflow. Technical Report No. tr/89-007, (UT Marine Science Institute PortAransas, Texas) 211pp.Whitledge, T.E, (1993). The nutrient <strong>and</strong> hydrographic conditions prevailing in LagunaMadre, Texas before <strong>and</strong> during a brown tide bloom. In: Toxic Phytoplankton Blooms inthe Sea. T.J. Smayda <strong>and</strong> Y. Shimizu (eds.), Proc. 5th Int. Conf. Toxic Marine Phytopl.Amsterdam, Elsevier. p. 711-716.65


VIII. Introduction to harmful red tides in Texas watersApart from brown tides, red tides comprise the other common harmful algal bloom inTexas coastal waters. <strong>Red</strong> tides are caused by blooms <strong>of</strong> din<strong>of</strong>lagellates that at highdensities can produce colors from yellow to reddish-brown in the water. The cells areattracted to light <strong>and</strong> actively swim toward the surface where they may be concentratedto high densities by wind, currents <strong>and</strong> tides (Tester <strong>and</strong> Fowler, 1990). No more thantwenty din<strong>of</strong>lagellate species are thought to be toxic (Steidinger, 1979), providingsources <strong>of</strong> poisonous compounds that during a bloom that can cause mass mortalities<strong>of</strong> marine organisms <strong>and</strong> can lead to human health problems when contaminatedseafoods are consumed or aerosol toxins are inhaled. Toxic red tide blooms occurthroughout the world but are relatively rare in Texas coastal waters. So far, there havebeen two species <strong>of</strong> din<strong>of</strong>lagellates responsible for toxic red tides in Texas: theunarmored Gymnodinium breve (formerly Ptychodiscus brevis) <strong>and</strong> Alex<strong>and</strong>rium(formerly Gonyaulax) monilata, an armored, chain-forming species. Typically,Gymnodinium breve first blooms in the Gulf <strong>of</strong> Mexico at least several miles <strong>of</strong>f thecoast. <strong>Current</strong>s may move these blooms to shore <strong>and</strong>/or into coastal bays, <strong>and</strong> bloomconcentrations can persist from one week to several months. Blooms may be confinedto a particular bay or estuary (typical <strong>of</strong> A. monilata) or may spread to cover a massivearea <strong>of</strong> coastal waters <strong>and</strong> embayments (as with G. breve). The toxins in both <strong>of</strong> thesedin<strong>of</strong>lagellate species can cause extensive mortality in fish <strong>and</strong> invertebrates, but inTexas, only Gymnodinium breve red tides have been reported to cause human healthproblems in the forms <strong>of</strong> temporary respiratory irritation from aerosol toxin <strong>and</strong>neurotoxic shellfish poisoning (NSP). The toxin commonly becomes an aerosol whencells are ruptured by wave action, as in heavy surf, with toxin or toxin-laced cellfragments carried into the air with water vapor <strong>and</strong>/or minute salt particles. NSP resultsfrom consuming raw or cooked oysters, clams or mussels contaminated with what iscommonly called "brevetoxin." The symptoms, including nausea, dizziness <strong>and</strong>tingling in the extremities, go away in a few days; no deaths have been reported due toNSP (Steidinger, 1983).Both Alex<strong>and</strong>rium monilata <strong>and</strong> Gymnodinium breve undergo sexual as well asasexual reproduction. The possibility that blooms initiate in these red tide species frombenthic resting cysts known as hypnozygotes is likely, though the factors that triggerexcystment <strong>of</strong> the hypnozygotes are yet unknown. The sexual cycle in A. monilatainvolves production <strong>of</strong> two motile, isogamous, haploid gametes from each singlehaploid asexual cell. The fusion <strong>of</strong> two gametes produces first a double-flagellated,diploid planozygote, then a hypnozygote; excystment releases one cell that soondivides asexually into a small chain <strong>of</strong> four, but whether meiosis re-establishes haploidybefore or after excystment is not known (Walker <strong>and</strong> Steidinger, 1979). The G. brevesexual cycle is similar. Haploid cells form isogamous gametes which, upon pairing <strong>and</strong>fusion, produce diploid planozygotes (Walker, 1982). In the same study, Walker wasunable to induce hypnozygotic cyst formation from the planozygotes, <strong>and</strong> to date,neither laboratory cultures nor sediment analyses have documented the existence <strong>of</strong> G.breve hypnozygotes (K. A. Steidinger, pers. comm.).66


IX. <strong>Historical</strong> trends <strong>of</strong> Texas red tides<strong>Red</strong> tides appear to be infrequent events in the CCBNEP area. Preliminaryinvestigations indicate that only four Gymnodinium breve <strong>and</strong> six or seven Alex<strong>and</strong>riummonilata blooms have been documented along the entire Texas coast since the majorred tide <strong>of</strong> 1935 (Snider, 1987; the ambiguity regarding the number <strong>of</strong> A. monilata redtides stems from the non-published sources <strong>of</strong> information available to Snider).Whereas G. breve blooms have been relatively well-studied, given the extent <strong>of</strong> theirnegative impact on fish <strong>and</strong> shellfish in Texas <strong>and</strong> their frequent recurrence <strong>of</strong>f WestFlorida, A. monilata blooms, which have not contaminated shellfish, have not been sowell-documented. Before the apparent 1935 red tide, the relationship between fish kills<strong>and</strong> din<strong>of</strong>lagellate blooms was not completely understood, making the causes <strong>of</strong> fishkills difficult to determine prior to that time.A. Frequency1. Offshore blooms <strong>of</strong> Gymnodinium breve in TexasThe mass mortality <strong>of</strong> fishes from the end <strong>of</strong> June to mid-September <strong>of</strong> 1935along the South Texas coast was not documented as linked to a din<strong>of</strong>lagellate bloom(Lund, 1936; cf. Steidinger <strong>and</strong> Joyce, 1973), but, given Lund's report <strong>of</strong> an associatedaerosol irritant ("irritating 'gas'") in the absence <strong>of</strong> a positive species identification, thisbloom was almost certainly due to G. breve (Snider, 1987). Newspaper reports at thetime indicate that local scientists varied in their opinion <strong>of</strong> the cause <strong>of</strong> the fish kill. Dr.E.J. Lund <strong>of</strong> the University <strong>of</strong> Texas blamed it on low salinity waters along the coast,while Dr. C.J. Reed <strong>of</strong> Texas A&I University blamed it on a bloom <strong>of</strong> diatoms whichwas clogging the gills <strong>of</strong> fish <strong>and</strong> causing them to suffocate. Local fishermensuspected that canisters <strong>of</strong> poison gas left over from the First World War were finallydecomposing <strong>and</strong> releasing their deadly fumes. Others suspected release <strong>of</strong> gasesassociated with volcanic activity (Corpus Christi Caller times, Houston Chronicle,Houston Post, <strong>and</strong> San Antonio Evening News).Whereas two to three other suspected red tide blooms occurred in Texas waterswithin the two decades after 1935 (Gunter, 1951), the next major bloom occurred inSeptember <strong>of</strong> 1955. As indicated by fish mortality, the bloom ranged from 18 milesnorth <strong>of</strong> Port Isabel to the coastal waters <strong>of</strong> the adjacent Mexican coastal state <strong>of</strong>Tamaulipas. Highest reported densities <strong>of</strong> the causative species, G. breve, werefound 36 miles south <strong>of</strong> the Rio Gr<strong>and</strong>e River mouth at >22,000 cells/ml (Wilson <strong>and</strong>Ray, 1956). Wilson <strong>and</strong> Ray (1956) <strong>of</strong>fer no indication that the bloom originated inTexas waters <strong>and</strong> proceeded south or vice versa, though their short publication doesimply a Texas origin.The red tide <strong>of</strong> 1974 is the poorest documented <strong>of</strong> the four major bloom eventsin South Texas. The Port Isabel/South Padre Press on 10 October 1974 reported thatthe bloom was sighted south <strong>and</strong> east <strong>of</strong> <strong>Brown</strong>sville/Matamoros some 150 miles <strong>of</strong>f67


the coast, causing a massive fish kill that eventually littered Mexican beaches, butgave no indication <strong>of</strong> its duration. The causative organism was not identified, but maywell have been Gymnodinium breve.This major Gymnodinium breve bloom lasting from the end <strong>of</strong> August to lateOctober <strong>of</strong> 1986 is arguably the worst yet experienced along the Texas coast witheffects stretching from its inception near Galveston Isl<strong>and</strong> to Mexico. Along MustangIsl<strong>and</strong> Beach alone, investigators estimated 100,000 dead fish per linear mile over 14miles <strong>of</strong> beach; the Rockport harbor boat basin produced water samples with up to 1.1million G. breve cells/ml, concentrations almost certainly due to physical factors suchas wind-driven currents (Trebatoski, 1988).2. Inshore blooms <strong>of</strong> Gymnodinium breve in TexasThere has been one prominent G. breve bloom in inshore waters. On 16December 1990, the <strong>Brown</strong>sville Herald newspaper began a series <strong>of</strong> reports on watersampling by health <strong>and</strong> marine <strong>of</strong>ficials in the <strong>Brown</strong>sville Ship Channel in response toa previous fish kill in November, <strong>and</strong> though the newspaper was not specific, thesuspected causative organism was identified by Dr. Eleanor Cox at Texas A & M asGymnodinium breve (Roy Lehman, pers. comm.). While the ship channel remainedclosed to shellfishing, G. breve concentrations apparently dropped below detectionlevel shortly before the Herald's final report on this red tide on 12 April 1991.According to records from the Shellfish Sanitation Division <strong>of</strong> the Texas Department <strong>of</strong>Health, there have been problem concentrations <strong>of</strong> G. breve at other times <strong>and</strong> placesforcing closure <strong>of</strong> shellfish beds, yet most <strong>of</strong> these blooms seem to have escaped widernotice. For instance, from 25 August to 2 September 1990, Aransas <strong>and</strong> Copano Bayswere closed to shellfish harvesting due to a short-lived G. breve bloom (see Table 4 inSection III. E.). Table 4 highlights some <strong>of</strong> the more prominent commercial harvestingareas affected by the closure <strong>and</strong> reopening <strong>of</strong> their shellfish beds. Though the lowfrequency <strong>of</strong> major <strong>of</strong>fshore blooms suggests that conditions favorable for initiation donot commonly occur in South Texas waters, the many closures <strong>of</strong> shellfish beds by theTexas Department <strong>of</strong> Health imply that G. breve blooms occur with greater frequencythan is commonly realized (G. Heideman, pers. comm.). Table 1 <strong>of</strong>fers further supportfor more frequent blooms by listing selected dates on which shellfish meats fromvarious Texas bays caused the death <strong>of</strong> one or more mice used in the brevetoxinbioassay. The dates in Table 1 were selected because they do not coincide with thedurations <strong>of</strong> the well-documented G. breve red tides discussed throughout this report<strong>and</strong> because they may represent a higher frequency <strong>of</strong> bloom concentrations largelyunnoticed except by the Shellfish Sanitation Division <strong>of</strong> the Texas Department <strong>of</strong>Health.68


Table 1. Selected Texas bays with prominent commercial shellfish beds matched with dates on whichsamples <strong>of</strong> shellfish meats proved fatal to one or more mice in the brevetoxin bioassay used by the TexasDepartment <strong>of</strong> Health. (Data courtesy <strong>of</strong> G. Heideman, Texas Department <strong>of</strong> Health.)Matagorda EspirituSantoSan Antonio Copano CorpusChristiNueces South3/3/876/24/878/14/873/4/87 2/15/873/4/874/1/87 4/2/878/17/874/2/878/17/871/27/888/18/883/6/913. Comparative frequencies <strong>of</strong> G. breve bloomsSimply determining the average frequency <strong>of</strong> Gymnodinium breve red tides fromthe four major <strong>of</strong>fshore blooms gives a rough mean occurrence <strong>of</strong> one major bloom <strong>of</strong>fthe Texas coast every 17 years with a st<strong>and</strong>ard deviation <strong>of</strong> 3 to 4 years. Whencompared to a recent detailed list <strong>of</strong> apparent G. breve red tides <strong>of</strong>f the southern WestFlorida coast, blooms or near blooms were reported in a total <strong>of</strong> 65 months from 1975-1991 with 46% <strong>of</strong> those reports occurring in September, October <strong>and</strong> November(Abdelghani, 1994; Figure 1).No. <strong>of</strong> Occurrences Recorded121086420MayJunJulAugSepOctNovDecJanFebMarAprFIGURE 1. Cumulative recorded occurrences per month <strong>of</strong> Gymnodiniumbreve blooms from 1975-1991 in West Florida(Data from Abdelghani, 1994)What Abdelghani's (1994) unpublished draft report does not indicate is whether anymortality was associated with any report <strong>of</strong> higher than normal G. breve concentrationsin that span <strong>of</strong> time; shellfish harvesting areas were closed on at least severaloccasions. Rounsefell <strong>and</strong> Nelson (1966) found records <strong>of</strong> 17 toxic red tides onFlorida's west coast between 1844 <strong>and</strong> 1960; the adjective "toxic" may imply marinemortality. Joyce <strong>and</strong> Roberts (1975) state that about 30 red tides occurred in Floridawaters between 1844 <strong>and</strong> 1974, an average <strong>of</strong> one every four to five years, but theymake no mention <strong>of</strong> mortality <strong>and</strong> can only attribute those since 1946 to G. breve. Onemay easily calculate that about 13 red tides occurred <strong>of</strong>f West Florida between 1960<strong>and</strong> 1974, but whether that comparatively high rate is due to an increase in bloomevents or improved detection is debatable (Hallegraeff, 1993; Anderson, 1994). For69


further comparison, Gunter (1951) reported that red tide-induced mortalities <strong>of</strong> marinelife along the West Florida coast tended to occur once every 10 years on average withdelays <strong>of</strong> up to 30 years, far more similar to the Texas experience.4. Alex<strong>and</strong>rium monilata blooms in TexasDocumentation <strong>of</strong> Alex<strong>and</strong>rium monilata blooms in Texas is less available thanthat for G. breve. Connell <strong>and</strong> Cross (1950) discussed the correlation <strong>of</strong> mass mortality<strong>of</strong> fish with a 1949 episode <strong>of</strong> red tide in Offats Bayou near Galveston. Such summerfish kills had been documented in the bayou from 1936 to 1941 by Gordon Gunter, <strong>and</strong>local fishermen, who claim that summer mortality had been an almost annualoccurrence since as early as 1929. The organism Connell <strong>and</strong> Cross (1950) describedoccurred in chains <strong>and</strong> was similar but not identical to Gonyaulax catenella. Usingsamples mainly from the Indian <strong>and</strong> Banana Rivers on Florida's east coast, Howell(1953) described a small din<strong>of</strong>lagellate species known to form chains <strong>of</strong> up to 40 cellsin length, <strong>and</strong> in early September <strong>of</strong> 1952, Howell found identical chain-formingdin<strong>of</strong>lagellates in a sample taken from Offatts Bayou. Whereas there was some debateover the proper generic classification at the time, what Connell <strong>and</strong> Cross (1950) <strong>and</strong>Howell (1953) described as a Gonyaulax species is now known as Alex<strong>and</strong>riummonilata. There are reports <strong>of</strong> "annual" blooms in the East Lagoon <strong>of</strong> Galveston Isl<strong>and</strong>(Marvin, 1965; Proctor, 1965; Proctor, 1966; Ray <strong>and</strong> Aldrich, 1967) <strong>and</strong> twoconsecutive blooms <strong>of</strong>fshore <strong>of</strong> Galveston in 1971-72 (Wardle, Ray <strong>and</strong> Aldrich, 1975).Since the 1971-72 blooms, A. monilata has not reappeared in bloom concentrations inthe Galveston area (W. J. Wardle, pers. comm.), though local blooms have occurred invarious parts <strong>of</strong> South Texas. Jensen <strong>and</strong> Bowman (1975) state that a colleague at theTexas Water Quality Board had documented an A. monilata bloom <strong>and</strong> fish kill in theViola Turning Basin <strong>of</strong> the Corpus Christi Inner Harbor in the fall <strong>of</strong> 1972. An A.monilata bloom reappeared in the Viola Turning Basin in the summer <strong>of</strong> 1975, firstdetected by routine monitoring at what became the maximum recorded concentration <strong>of</strong>5.14 million cells/liter (Jensen <strong>and</strong> Bowman, 1975).B. Duration1. Gymnodinium breveThe G. breve red tide <strong>of</strong> 1935 was detectable from 30 June to 13 August, thefirst report coming from Padre Isl<strong>and</strong> <strong>and</strong> the last from the coast at the northern tip <strong>of</strong>Matagorda Bay, a total <strong>of</strong> 45 days (Lund, 1936). The apparent 1955 G. breve red tidelasted for more than 12 days in September, but Wilson <strong>and</strong> Ray (1956) do not specifywhen the fish kill was first noticed nor when fish mortality ceased <strong>of</strong>f the TamaulipasCoast <strong>of</strong> Mexico. No duration could be determined for the 1974 red tide 150 miles<strong>of</strong>fshore <strong>and</strong> southeast <strong>of</strong> <strong>Brown</strong>sville. The <strong>Brown</strong>sville Inner Harbor bloom <strong>of</strong> G. brevereported by Jensen <strong>and</strong> Bowman (1975) began to recede after about two months butlingered for perhaps two more months. The Coast Guard first noted the 1986-87 redtide as what appeared to be an oil slick washing ashore at the southern tip <strong>of</strong>70


Galveston Isl<strong>and</strong> on 27 August 1986; the subsequent large-scale event ceased 57 dayslater (Trebatoski, 1988). Trebatoski (1988) also noted that a dense, isolatedrecurrence appeared on 8 January 1987 in Corpus Christi Bay but apparently did notpersist past 4 February (Texas Department <strong>of</strong> Health). In comparison, the impliedduration <strong>of</strong> reported red tide bloom concentrations (apparently all G. breve) along theWest Florida coast from 1975-1991 ranged from "brief" to approximately seven months,the latter in 1980 (Abdelghani, 1994).2. Alex<strong>and</strong>rium monilataThe 1972 A. monilata bloom in the Viola Turning Basin <strong>of</strong> the Corpus ChristiInner Harbor lasted from August to September; the 1975 resurgence <strong>of</strong> A. monilata inthe same location persisted from July to September (Jensen <strong>and</strong> Bowman, 1975).During the 1975 bloom, cell counts <strong>and</strong> the number <strong>and</strong> length <strong>of</strong> A. monilata chainsgenerally decreased throughout the inner harbor until no longer detectable on 17September, but a fish kill did occur in the turning basin on 22 August, despite the smallconcentrations <strong>of</strong> A. monilata detected there at that time (Jensen <strong>and</strong> Bowman, 1975).C. Environmental effects1. CommercialThe most immediate negative economic impact <strong>of</strong> red tide blooms is fish <strong>and</strong>invertebrate mortality. Aquaculture projects are especially susceptible to the toxins byvirtue <strong>of</strong> stock density <strong>and</strong> potentially catastrophic loss <strong>of</strong> investment (Steidinger <strong>and</strong>Vargo, 1988; Shumway, 1990). Two experimental studies with both Gymnodiniumbreve <strong>and</strong> Alex<strong>and</strong>rium monilata illustrate the probable impact <strong>of</strong> the respective toxinson organisms <strong>of</strong> commercial value in the field. Sievers (1969) exposed several kinds<strong>of</strong> marine animals to concentrations <strong>of</strong> G. breve <strong>and</strong> A. monilata cultures fromundiluted to 90% dilution over 48 hours to determine the comparative toxicity <strong>of</strong> eachin terms <strong>of</strong> mortality. Fish were quite sensitive to both toxins (especially so to G. brevetoxin), crustaceans were resistant to both, <strong>and</strong> both annelids <strong>and</strong> mollusks were moresensitive to A. monilata toxin, including the American oyster (Crassostrea virginica)<strong>and</strong> a mussel (Brachidontes recurvus). The oysters <strong>and</strong> mussels immediately failed toopen upon exposure to lower dilutions <strong>of</strong> A. monilata cultures <strong>and</strong> suffered highmortality in undiluted cultures. Ray <strong>and</strong> Aldrich (1967) indicate that such mortality mayarise simply because the bivalves cannot maintain shell closure well past 24 hours. Intheir 24-hour laboratory study, oysters remained tightly closed almost constantly whilein A. monilata cultures. Chicks were not poisoned by ingestion <strong>of</strong> the oysterhomogenate, even when the homogenate was combined with actual A. monilata cellsor prepared from oysters exposed to an in situ bloom <strong>of</strong> A. monilata. In the samestudy, three <strong>of</strong> four oysters exposed to G. breve cultures filtered <strong>and</strong> consumed cellsafter some delay in opening; five grams <strong>of</strong> the subsequent oyster homogenate forcefedto each <strong>of</strong> four chicks proved fatal to all within 24 hours.71


In 1971 <strong>and</strong> 1972, A. monilata blooms occurred not far <strong>of</strong>fshore <strong>of</strong> Galveston,producing high numbers <strong>and</strong> a wide variety <strong>of</strong> dead <strong>and</strong> dying marine organisms,including the commercially important blue crab, Callinectes sapidus (both years), thestone crab, Menippe mercenaria (both years), <strong>and</strong> the oyster, Crassostrea virginica(1972 only; Wardle et al., 1975). Though the findings <strong>of</strong> Wardle et al. (1975) indicatethat crustaceans are not immune to mortality induced by A. monilata blooms, otherdin<strong>of</strong>lagellate toxins (including brevetoxin) are not known to cause mortality orcontaminate the meat, so crabs, shrimp <strong>and</strong> lobsters are marketable even at the height<strong>of</strong> toxic blooms (Roberts et al., 1979, re brevetoxin; Shumway, 1990).Edible bivalve mollusks are another matter, <strong>and</strong> the public health implications <strong>of</strong>contaminated bivalve meats can close lucrative fisheries temporarily, on a seasonalbasis or indefinitely. Shumway (1990) reviews sources that document a worst-casescenario in Alaska. With tens <strong>of</strong> thous<strong>and</strong>s <strong>of</strong> miles <strong>of</strong> coastline <strong>and</strong> over 100 clamspecies, the Alaskan clam industry produced millions <strong>of</strong> pounds <strong>of</strong> shellfish productsuntil it was forced to close permanently shortly after World War II because <strong>of</strong> toxic redtide problems. When the Gulf Stream transported G. breve cells or cysts from WestFlorida to the Carolinas in 1987, the resultant bloom closed shellfish beds for the firsttime in North Carolina's history <strong>and</strong> caused losses <strong>of</strong> $20 million (Anderson, 1994).Efficient monitoring programs, however, enable some shellfish aquaculture efforts tocontinue safely in spite <strong>of</strong> recurrent blooms, exemplified by successful mussel culturesin the northeastern United States <strong>and</strong> in the Japanese scallop industry (Shumway,1990); this may be the only way to effectively reduce the negative economic impact <strong>of</strong>toxic din<strong>of</strong>lagellates on bivalve fisheries.The remaining problem for shellfish beds exposed to G. breve blooms isthe necessary ban on harvesting for up to two months or until tests indicate that themeat is no longer toxic (Steidinger <strong>and</strong> Ingle, 1972). In 1990, Tester <strong>and</strong> Fowlerpointed out that the U. S. Food <strong>and</strong> Drug Administration's interpretation <strong>of</strong> theAmerican Public Health Association guidelines regarding an acceptable amount <strong>of</strong>NSP in shellfish contaminated by G. breve was more conservative than that forshellfish contaminated with the potentially lethal PSP toxin (paralytic shellfishpoisoning, caused by numerous other din<strong>of</strong>lagellate species). In spite <strong>of</strong> theirrecommendation that the acceptable amount <strong>of</strong> NSP toxin in shellfish be equated withthat for PSP toxin, given the more than acceptable public health risk in doing so(Tester <strong>and</strong> Fowler, 1990), the FDA to date has not acted upon that recommendation(P. A. Tester, pers. comm.). In effect, the overly conservative restriction increases thenegative economic impact produced by G. breve blooms on the shellfish industry bykeeping shellfish beds closed for longer periods than truly necessary.Steidinger <strong>and</strong> Vargo (1988) also cite adverse impacts on tourism, real estate<strong>and</strong> seafood sales in areas suffering from G. breve red tides. Jensen (1975)discussed the economic impact as not just a problem for the affected area but insurrounding states as well because <strong>of</strong> the "halo effect." His example was the G.tamarensis (PSP) red tide <strong>of</strong> 1972 in New Engl<strong>and</strong>. The affected New Engl<strong>and</strong> states72


quickly banned harvest <strong>and</strong> shipment <strong>of</strong> their shellfish, but despite the fact that noblooms affected New York waters, many citizens <strong>of</strong> that state responded irrationally byrefusing to buy New York lobster, shellfish <strong>and</strong> finfish, avoiding Long Isl<strong>and</strong> seafoodrestaurants <strong>and</strong> causing a drop in the wholesale price for clams harvested fromunaffected areas. Jensen (1975) concludes that, if necessary an efficient means existwith which governmental agencies can quickly note the occurrence <strong>and</strong> distribution <strong>of</strong>red tide blooms, what remains is for the seafood industry <strong>and</strong> the government to stemrumors <strong>and</strong> counteract misinformation with accurate <strong>and</strong> current facts about any toxicblooms, curtailing the "halo effect" that needlessly increases the scale <strong>of</strong> the negativeeconomic impact.Both A. monilata <strong>and</strong> G. breve possess icthyotoxins that threaten game <strong>and</strong> foodfish populations in the Gulf <strong>of</strong> Mexico. Using juvenile mullet (Mugil cephalus) as theirbioassay, Gates <strong>and</strong> Wilson (1960) noted mortality within 4.5 hours in all aliquots <strong>of</strong> invitro cultures <strong>of</strong> A. monilata (some aliquots containing cells <strong>and</strong>/or cell fragments <strong>and</strong>others not) except the uninoculated control medium. Quick <strong>and</strong> Henderson (1975) didnot completely list the fifteen species they necropsied during an extended G. breve redtide from October 1973 through June 1974 along the west coast <strong>of</strong> Florida, but mullet,ladyfish <strong>and</strong> anchovies were mentioned. Enormous numbers <strong>of</strong> fish <strong>of</strong> even greatersport or commercial value were estimated to have died as a result <strong>of</strong> the 1986 G.breve red tide along the Texas coast (Texas Parks <strong>and</strong> Wildlife Department, 1986; seeTable 3.), including greater than 40,000 individuals each <strong>of</strong> red drum <strong>and</strong> southernflounder, about 80,000 spotted seatrout <strong>and</strong> more than 3.2 million Gulf menhaden.Riley et al. (1989) report that G. breve blooms can also reduce red drum recruitmentduring spawning periods; their source <strong>of</strong> red tide was the 1986 Texas bloom whichachieved average peak densities <strong>of</strong> 7000 cells/ml with concentrations <strong>of</strong> almost 5 x 10 4cells/ml in isolated patches in bays near Port Aransas throughout October. Theyshowed that G. breve caused paralysis <strong>and</strong> death in laboratory-spawned <strong>and</strong> wildcaughtred drum larvae at all concentrations above 40 cells/ml. Since both A. monilata<strong>and</strong> G. breve cause fish mortality, an extensive bloom <strong>of</strong> either can produce acorresponding loss <strong>of</strong> commercial <strong>and</strong> game species.2. Public healthShellfish contaminated with Gymnodinium breve cells or the heat stable toxin canproduce neurotoxic shellfish poisoning (NSP) in people who consume raw or cookedmeat. No paralysis occurs with NSP, but tingling <strong>of</strong> the skin (paresthesia), nausea,vomiting <strong>and</strong> some loss <strong>of</strong> voluntary muscle control (ataxia) are major symptomsoccurring within the first three hours (Abdelghani, 1994). Contact dermatitis <strong>and</strong>/orconjunctivitus may occur in people immersed in water containing brevetoxin (Hemmert,1975). An aerosol <strong>of</strong> either toxin-laced cell fragments or salt crystals released in surfzones can cause respiratory, skin <strong>and</strong> mucus membrane irritation, but the symptomscease with no lasting effects once victims leave the area (Hemmert, 1975; Steidinger<strong>and</strong> Vargo, 1988). The temporary <strong>and</strong> mild effects <strong>of</strong> the aerosol toxin are fortunatesince Pierce et al. (1990) determined that laboratory culture concentrations <strong>of</strong> G. breve73


could give rise to toxins enriched by 5 to 50 times in the aerosol relative to the culture.Though no deaths have yet been attributed to NSP (Shumway, 1990), people shouldnot consume shellfish exposed to a G. breve bloom until reliable tests are conducted toensure that the meat is no longer contaminated; this typically means a wait <strong>of</strong> one totwo months after cessation <strong>of</strong> the bloom (Steidinger <strong>and</strong> Ingle, 1972).Regarding brevetoxin itself, Baden <strong>and</strong> Thomas (1989) examined severaldifferent clones <strong>of</strong> G. breve <strong>and</strong> found significant variability in toxin content between G.breve populations experiencing different environmental conditions. Baden (pers.comm.) states that G. breve is very adept at performing the simple metabolicmodifications that create variability in the toxin pr<strong>of</strong>ile, <strong>and</strong> one may presume that thevariability is due to different metabolic states within the organism. One should not,therefore, rule out the possibility that toxins from monoclonal G. breve populations, ifsuch exist, could exhibit differential toxicity to marine animals <strong>and</strong> humans during abloom depending on the bloom environment. Also, the 1990 studies by Pierce et al.<strong>and</strong> Roszell et al. revealed that six or more possible toxin pr<strong>of</strong>iles appear individually orin combination at different times in the life <strong>of</strong> a population, whether from examination <strong>of</strong>a monoclonal culture or a natural population. Taken together, these findings mayindicate that one or more clones may exist in any given G. breve bloom population;regardless <strong>of</strong> the clonal constitution, a bloom may produce, as elements age within thepopulation, a suite <strong>of</strong> toxins that varies in (1) the composition <strong>of</strong> toxic fractions, (2) theoverall quantity <strong>of</strong> toxin released at any given point <strong>and</strong>/or (3) potency. The publichealth implications <strong>of</strong> these possibilities should be noted. Apparently no public healthrisks have yet been associated with the toxin <strong>of</strong> Alex<strong>and</strong>rium monilata.3. Threats to endangered wildlifeWhereas red tide blooms may be detrimental to a variety <strong>of</strong> endangeredspecies, very little documentation exists for adverse affects on, for example, marinemammals <strong>and</strong> sea turtles. The most likely <strong>and</strong> serious threat from red tide to suchspecies in South Texas concerns the whooping crane.The Aransas National Wildlife Refuge reported an overwintering population <strong>of</strong>133 adult <strong>and</strong> 8 juvenile whooping cranes for early 1995. Given such low numbers, aGymnodinium breve outbreak could be catastrophic for the entire flock (Tom Stehn,pers. comm.). The G. breve red tide <strong>of</strong> 1986-87 in South Texas came very close toinfiltrating whooping crane critical habitat in the Aransas National Wildlife Refuge inthe fall <strong>of</strong> 1986. Because clams, swallowed whole, comprise a small part <strong>of</strong> awhooping crane's diet in fall <strong>and</strong> early winter, brevetoxin could possibly sicken or killany cranes consuming contaminated clams. Scaup <strong>and</strong> cormorants have been knownto perish from exposure to brevetoxin in Florida, so the potential threat to the smallpopulation <strong>of</strong> endangered whooping cranes in South Texas calls for careful futuremonitoring (Stehn, 1987).4. Ecosystem74


The effects <strong>of</strong> Gymnodinium breve blooms on the ecosystem may be positive aswell as negative. Both aspects are covered quite well by Steidinger <strong>and</strong> Vargo (1988),the source for much <strong>of</strong> what follows.Obviously, negative ecosystem effects begin with marine organism mortality,whether due to toxin or oxygen depletion, whether affecting wild or culturedpopulations <strong>of</strong> marine life. Oxygen depletion by dense concentrations <strong>of</strong> either toxic ornon-toxic red tide species can be as detrimental as toxins to fisheries or aquaculture;toxic blooms can cause problems as well for local, national <strong>and</strong> internationaleconomies, all discussed above.Other ecosystem-level problems include (1) an undefined negative impact fromexcessive nutrient enrichment <strong>and</strong> bacteria concentrations in waters in whichnumerous fish carcasses decompose <strong>and</strong> (2) the influence <strong>of</strong> toxic din<strong>of</strong>lagellates onother planktonic organisms. Regarding the latter, Freeberg et al. (1979) used mediumin which G. breve had been grown as a medium for each <strong>of</strong> 28 phytoplankton speciesin axenic cultures. The medium significantly inhibited growth in 18 species, but itseffect varied within species. The population levels <strong>of</strong> several diatoms <strong>and</strong>din<strong>of</strong>lagellates <strong>and</strong> one flagellate barely increased above inoculum concentrations;two din<strong>of</strong>lagellate inocula suffered lysis. Toxin extracts totally arrested growth in eight<strong>of</strong> twelve species (four diatoms <strong>and</strong> four din<strong>of</strong>lagellates), but column chromotographycould not separate the algal inhibition component from the ichthyotoxin. Huntley et al.(1986) used thirteen species <strong>of</strong> din<strong>of</strong>lagellates as possible prey for two species <strong>of</strong>copepods in a large suite <strong>of</strong> experiments. Results <strong>of</strong> one experiment included G. breveas one <strong>of</strong> several din<strong>of</strong>lagellates consistently rejected by the copepod Calanuspacificus; ingestion <strong>of</strong> G. breve cells produced elevated heart rate <strong>and</strong> loss <strong>of</strong> motorcontrol in the copepod. The findings <strong>of</strong> Freeberg et al. (1979) <strong>and</strong> Huntley et al.(1986) counter the suggestion by Steidinger <strong>and</strong> Ingle (1972) that G. breve temporarilyaffects inshore <strong>and</strong> nearshore reef fisheries only, though Steidinger <strong>and</strong> Ingle (1972)may have had no data on smaller-scale effects.In terms <strong>of</strong> the positive impacts G. breve has on ecosystems, Steidinger <strong>and</strong>Vargo (1988) point out that G. breve, like all known toxic din<strong>of</strong>lagellates, arephotosynthetic primary producers. As such, if a consumer can tolerate the toxin <strong>and</strong>digest the cellulose cell wall, G. breve can be a high quality food source. In addition,din<strong>of</strong>lagellates in general tend to leak a variable percentage <strong>of</strong> their daily carbon in theform <strong>of</strong> amino acids <strong>and</strong> other complex biochemicals. In bloom proportions,contributions <strong>of</strong> the din<strong>of</strong>lagellate population in terms <strong>of</strong> food for consumers, dissolvedorganics <strong>and</strong> the detritus <strong>of</strong> dead <strong>and</strong> dying cells must be large. Vargo et al. (1987)<strong>of</strong>fers numerical estimates <strong>of</strong> various carbon production rates <strong>and</strong> estimates a highpotential annual carbon input for G. breve blooms <strong>of</strong>f West Florida based on severalsources <strong>of</strong> data. At the time <strong>of</strong> publication, however, Steidinger <strong>and</strong> Vargo (1988)refer to Vargo et al. (1987) <strong>and</strong> only two other largely speculative studies on thesupposed magnitude <strong>of</strong> the carbon contribution <strong>of</strong> any bloom species, but no75


subsequent studies have confirmed the significant carbon contribution postulated forG. breve blooms since 1988 (K. A. Steidinger, pers. comm.). Another possible positiveresult following the negative impact <strong>of</strong> a toxic bloom is that affected bottomcommunities are kept in states <strong>of</strong> simplified ecological interactions <strong>and</strong> increasedsystem efficiency (Steidinger <strong>and</strong> Vargo, 1988). They cite several sources <strong>of</strong> supportfor the possibility that periodic fish <strong>and</strong> invertebrate kills from toxic red tides couldmake for communities less prone to catastrophic collapse from any major perturbationover the long term because their fauna <strong>and</strong> flora never have time to develop complex,fragile interdependencies. If this is true, the scientific community, the government,fishing industries <strong>and</strong> the public should best resign themselves to short-term toxic redtide problems in exchange for long-term benefits for affected benthic <strong>and</strong> reefcommunities <strong>and</strong> the commercially valuable resources harvested from them.D. Possible causesSteidinger <strong>and</strong> Vargo (1988) used the terms "initiation," "growth" <strong>and</strong>"maintenance" to categorize the various factors that produce <strong>and</strong> promote each stagein the progress <strong>of</strong> a red tide bloom, <strong>and</strong> they are used in this section as headings fordiscussing the causative factors for each stage in a Gymnodinium breve bloom. (Theless problematical din<strong>of</strong>lagellate Alex<strong>and</strong>rium monilata has not received as muchattention in terms <strong>of</strong> published data on the factors prompting it to bloom.) Becausemultiple variables both known <strong>and</strong> unknown are at work in the marine environmentduring a bloom, any discussion <strong>of</strong> possible causes remains largely speculative. Thecauses <strong>of</strong> initiation in particular are problematical, for no proven catalyst has yet beenfound, whether a single variable or suite <strong>of</strong> variables. As a result, those factorsdiscussed below as possible initiators for G. breve blooms definitely play roles ingrowth <strong>and</strong> maintenance as well (Rounsefell <strong>and</strong> Nelson, 1966; Steidinger <strong>and</strong>Haddad, 1981).1. InitiationIn 1958, Collier suggested that a complex <strong>of</strong> biological factors causes a red tidebloom <strong>and</strong> that biologically active organic compounds are important to G. breve,including vitamin B12 <strong>and</strong> organic chelators, <strong>of</strong> which the latter can be supplemented orreplaced by sulfides that commonly occur in West Florida estuaries.On a more basic level, salinity is a principal factor in the initiation <strong>and</strong>subsequent progress <strong>of</strong> a G. breve bloom. Aldrich <strong>and</strong> Wilson (1960) determined thatthe optimal salinity range for axenic cultures <strong>of</strong> G. breve was from 27 to 37 ppt. Someorganisms survived salinities as low as 22.5 <strong>and</strong> as high as 46.0 ppt for ten weeks, butthe authors concluded that typical Gulf <strong>of</strong> Mexico salinities should not imposerestrictions on the growth <strong>of</strong> this din<strong>of</strong>lagellate, though water with salinities <strong>of</strong> 24 ppt orless should (see also Tester <strong>and</strong> Fowler, 1990). This low-salinity limitation helpsexplain why G. breve blooms commonly occur in oceanic rather than estuarine waters.76


Along with favorable salinity levels, sufficient light <strong>and</strong> an adequate carbonsource like carbon dioxide are necessary for any photosynthetic organism. Aldrich(1962) noted a correlation between red tide outbreaks involving Gymnodinium breve onthe Florida Gulf coast <strong>and</strong> extended periods <strong>of</strong> heavy rainfall <strong>and</strong> subsequent organicinput to the sea from river discharge. After unsuccessfully testing a multitude <strong>of</strong>organic substances as potential direct energy sources <strong>and</strong> discovering that theorganism did not grow in the dark in spite <strong>of</strong> growth additives, he determined that G.breve was not heterotrophic <strong>and</strong> that sunlight <strong>and</strong> carbon dioxide were its principalgrowth requirements. He suggested that vitamins, trace metals <strong>and</strong> chelators fromFlorida river waters may facilitate photoautotrophy <strong>and</strong> thus bloom formation in G.breve (cf. Collier, 1958) <strong>and</strong> should be studied further.Ten years later, Steidinger <strong>and</strong> Ingle (1972) included in their summary <strong>of</strong>information on G. breve red tides the facts that pollution does not catalyze blooms inFlorida <strong>and</strong> that din<strong>of</strong>lagellate blooms may initiate from seed populations well <strong>of</strong>fshore.Steidinger (1975b) postulated that the <strong>of</strong>fshore seed populations could be either restingpelagic cells or benthic resting cysts (hypnozygotes) from which blooms could initiate<strong>and</strong> be transported inshore with the aid <strong>of</strong> physical factors. For instance, Roberts(1979) confirmed that, in 1976, surface concentrations <strong>of</strong> G. breve increased gradually<strong>of</strong>fshore <strong>and</strong> were then transported inshore due to winds <strong>and</strong> currents. Initiation inWest Florida, therefore, may well be an <strong>of</strong>fshore phenomenon that accelerates when<strong>and</strong> if transport to depth <strong>and</strong> toward shore concentrates red tide cells in the photiczones <strong>of</strong> shallow waters (Seliger et al.,1979). That a process similar to West Floridaoccurs for the initiation <strong>of</strong> blooms <strong>of</strong>f South Texas is quite possible, particularly giventhe prevailing onshore winds in the western Gulf <strong>of</strong> Mexico.Temperature is another likely significant factor, supported in part by theapparently strong correlation between surface water temperature patterns <strong>and</strong> majorbloom initiation reported by Baldridge (1975), who claimed that simple indicatorpatterns <strong>of</strong> water temperatures from mid-January to early April could predict red tideoutbreaks twelve months in advance near Tampa Bay, Florida. He stated that suchpatterns had already shown strong correlation with five major red tides between 1957<strong>and</strong> 1974 at Egmont Key, Florida, but had no desire to claim cause-effect dependencynor to diminish the importance <strong>of</strong> other environmental conditions favoring bloominitiation. His claim <strong>of</strong> the strong predictive power <strong>of</strong> certain surface temperaturepatterns, however, has not been substantiated since 1974 (K. A. Steidinger, pers.comm.). Though Baldridge's predictive model may not be acceptable, a necessarytemperature range may still be required for initiation. Rounsefell <strong>and</strong> Nelson (1966)cite results <strong>of</strong> studies that examined temperature effects on G. breve which, whencombined, defined the survival range from 7 o -32 o C, the range <strong>of</strong> possible growth from


excystment <strong>and</strong> cited other studies on different din<strong>of</strong>lagellate species that had reportedidentical results with temperature increases.Walker (1982) has confirmed the potential for G. breve cells to formhypnozygotes as part <strong>of</strong> their sexual cycle, but hypnozygotes have not been found onthe West Florida shelf to date. The Florida program established to look for <strong>and</strong> map<strong>of</strong>fshore "seed beds" was canceled, <strong>and</strong> hypnozygotes have not yet been induced t<strong>of</strong>orm in culture (K. A. Steidinger, pers. comm.). Haddad <strong>and</strong> Carder (1979) proposethat if seed beds <strong>of</strong> G. breve hypnozygotes exist on the shelf, the cysts themselves maybe carried in suspension to depths shallower than 40 meters when the Loop <strong>Current</strong>intrudes into shelf waters <strong>and</strong> may encounter conditions conducive to excystment <strong>and</strong>growth, raising the possibility <strong>of</strong> inshore initiation as well as the previouslysubstantiated <strong>of</strong>fshore process (Steidinger, 1975b; Roberts, 1979; Seliger et al.,1979).Haddad <strong>and</strong> Carder (1979) cite numerous instances <strong>of</strong> deep Loop <strong>Current</strong> <strong>and</strong> EasternGulf Water (EGW) intrusion onto the shelf that coincided with G. breve blooms <strong>of</strong>f WestFlorida <strong>and</strong> describe a 1977 Florida bloom that occurred simultaneously with EGWupwelling nearshore. They suggest that some conditions favorable for initiation <strong>and</strong>subsequent blooms <strong>of</strong> G. breve include the decreased temperatures <strong>of</strong> the Loop<strong>Current</strong> <strong>and</strong> EGW combined with the increased light availability <strong>and</strong> increased nutrients<strong>of</strong> the nearshore shelf. Finally, Florida blooms need not arise from excystment <strong>of</strong>hypnozygotes, for Steidinger (1975a) reports the perennial presence <strong>of</strong> motile forms <strong>of</strong>unknown ploidy <strong>of</strong> G. breve at less than 1000 cells/liter in the same area <strong>of</strong> the WestFlorida shelf as the potential seed beds. Such background concentrations could alsobe responsible for the initiation <strong>of</strong> some blooms.2. GrowthAs early as 1958, Collier suggested that G. breve may produce an organicsubstance that indirectly conditions the water for its own growth in addition to makingpossible use <strong>of</strong> biologically active organic compounds, vitamin B12, sulfides <strong>and</strong>organic chelators. Since red tide outbreaks involving G. breve on the Florida Gulf coastwere correlated with extended periods <strong>of</strong> heavy rainfall <strong>and</strong> subsequent organic inputto the sea from river discharge, Aldrich (1962) suggested that vitamins, trace metals<strong>and</strong> chelators from river waters may facilitate photoautotrophy <strong>and</strong> thus bloomformation (i. e., growth) in G. breve <strong>and</strong> should be studied further. Collier et al. (1969)exp<strong>and</strong>ed Collier's 1958 study <strong>of</strong> organic <strong>and</strong> inorganic chemicals to determine thatchelated metals (e. g., EDTA-Fe), sulfide, nitrogen, phosphorus <strong>and</strong> vitamins canpositively influence the growth <strong>of</strong> G. breve. In a review, Steidinger (1975a) definedsupport for the growth <strong>of</strong> G. breve in terms <strong>of</strong> favorable levels <strong>of</strong> nutrients, growthfactors, temperature <strong>and</strong> salinity. From available data, she characterized G. breveblooms, however, as gradual increases in motile cells, not sudden increases in celldivision rates. Steidinger <strong>and</strong> Vargo (1988) cite thesis research that determined lightsaturateddivision rates in G. breve from 0.1 to 0.5/day, comparable to what isconsidered a "normal" din<strong>of</strong>lagellate division rate <strong>of</strong> < 1.0/day. If a normal doublingrate is the rule, then immense G. breve blooms are not explosive episodes <strong>of</strong> asexual78


eproduction, but even division rates <strong>of</strong> 0.5/day can produce in eight days aconcentration <strong>of</strong> 250,000 cells/liter from an initial density <strong>of</strong> 20,000/liter (Roberts,1979). Densities on the order <strong>of</strong> 10 6 /liter, however, are best attributed to physicalconcentrating factors, as discussed below.3. Maintenance/TransportBloom maintenance may be defined as the persistent presence <strong>of</strong> a bloom atconcentrations above background levels. With support from numerous theoreticaltreatments <strong>of</strong> the behavior <strong>of</strong> semi-buoyant particles, a term that also describesdin<strong>of</strong>lagellates, Collier (1958) explained that, whereas a complex <strong>of</strong> biological factorsapparently prompt a red tide bloom, physical factors such as water mass convergence,wind-driven currents <strong>and</strong> convection cells cause its mechanical concentration, astatement reiterated by Steidinger <strong>and</strong> Ingle (1972), Steidinger (1975a), Roberts (1979)<strong>and</strong> Seliger et al. (1979) (who include tidal influences). G. breve is known for itsinedibility to grazers (Huntley et al., 1986), a definite advantage for maintenance.Huntley et al. (1986) add that, not only do noxious din<strong>of</strong>lagellate species such as G.breve gain an interspecific competitive advantage in survival over edible phytoplankton,they also seem far more likely to be able to maintain bloom proportions in spite <strong>of</strong>zooplankton grazers when other factors are favorable. Several other factors alsocontribute to the maintenance <strong>of</strong> high densities <strong>of</strong> G. breve, including migration <strong>of</strong> themotile cells (Steidinger, 1975a; Steidinger <strong>and</strong> Vargo, 1988) <strong>and</strong> allelopathicsubstances released by G. breve that inhibit the growth <strong>of</strong> other phytoplankton species(Steidinger <strong>and</strong> Vargo, 1988) <strong>and</strong> may aid its own growth (Steidinger, 1983).4. DeclineIt is important to note that the same meteorologic <strong>and</strong> hydrologic forces thatconcentrate a red tide bloom can disperse it as well (Steidinger, 1983). Steidinger <strong>and</strong>Vargo (1988) list the many other ways besides advection that may diminish <strong>and</strong>terminate any algal bloom: benthic filter feeders, cell death, grazing, life cycle changes,nutrient limitation <strong>and</strong> parasitism. To this list one may add sinking <strong>and</strong> toxic naturalchemicals, the latter due to Collier's (1958) observation that copper levels in somecoastal waters <strong>of</strong> West Florida were clearly toxic to G. breve. Astronomical factors mayhelp explain the tendency for blooms to be seasonal. Aldrich (1962) noted that G.breve did not grow in the dark in spite <strong>of</strong> growth-stimulating additives; the decreasingtemperatures <strong>and</strong> photoperiod <strong>of</strong> the winter months seem to reduce or prevent thepersistence or initiation <strong>of</strong> toxic blooms (note the trend <strong>of</strong> a winter decline seen in Fig.1). Yet atypical winter weather patterns have made exceptions to this general trend,<strong>and</strong> in areas such as South Texas, mild winter weather <strong>and</strong> longer photoperiodsrelative to higher latitudes may have little preventive effect.X. Available dataA. Causative species79


1. Gymnodinium breveG. breve was positively identified as the cause <strong>of</strong> the major red tide blooms <strong>of</strong>1955 <strong>and</strong> 1986 as well as the 1990-91 bloom within the <strong>Brown</strong>sville Ship Channel.One may reasonably assume that the bloom <strong>of</strong> 1935 was due to G. breve based onreports <strong>of</strong> respiratory irritation. The cause <strong>of</strong> the 1974 red tide remains uncertain,though anything other than G. breve would be surprising.2. Alex<strong>and</strong>rium monilataIt seems very likely that the red tides in Offats Bayou in the 1930s-40s <strong>and</strong>perhaps earlier (Connell <strong>and</strong> Cross, 1950) were caused by the chain-forming, armoreddin<strong>of</strong>lagellate A. monilata, but confirmation <strong>of</strong> the species in the bayou did not comeuntil 1952 (Howell, 1953). A. monilata was identified in the East Lagoon <strong>of</strong> GalvestonIsl<strong>and</strong> in the 1960s (Ray <strong>and</strong> Aldrich, 1967) <strong>and</strong> was the cause <strong>of</strong> two blooms <strong>of</strong>fshore<strong>of</strong> Galveston in 1971 <strong>and</strong> 1972 (Wardle et al., 1975). Since then, A. monilata bloomshave only occurred farther south on the Texas coast, with two documented cases in theViola Turning Basin <strong>of</strong> the Corpus Christi Inner Harbor in 1972 <strong>and</strong> 1975 (Jensen <strong>and</strong>Bowman, 1975). There are yet unsubstantiated reports <strong>of</strong> similar blooms in the<strong>Brown</strong>sville ship channel in 1988 (Dr. Terry Allison, pers. comm).B. Cell concentrations/other biomass estimatesDuring the 1990-91 G. breve bloom in <strong>Brown</strong>sville, Tony Reisinger, MarineExtension Agent for Cameron County, collected water samples near the surface at onesite in the turning basin <strong>of</strong> the ship channel from 4 December 1990 to 23 February1991, noting G. breve concentrations along with salinity <strong>and</strong> surface temperature(Figure 2).35300000Salinity (ppt) <strong>and</strong> Surface WaterTemperature (Celsius)30252015105TemperatureSalinityCells/ml25000020000015000010000050000G. breve cells/ml0012/1/9012/5/9012/9/9012/13/9012/17/9012/21/9012/25/9012/29/901/2/911/6/911/10/91Figure 2. G. breve concentrations, salinity <strong>and</strong> surface temperature overtime at one station in the <strong>Brown</strong>sville Ship Channel Turning Basin(Unpublished data courtesy <strong>of</strong> T. Reisinger.)1/14/911/18/911/22/911/26/911/30/912/3/912/7/912/11/912/15/912/19/912/23/9180


Though the data in Figure 2 appear to reflect some influence by salinity <strong>and</strong>temperature on cell concentration (determined by hemacytometer), the association isnot significant (p = 0.23 from multiple regression), a fact not surprising with knowledge<strong>of</strong> the concentrating effect <strong>of</strong> winds <strong>and</strong> currents. For example, the lowest temperaturereading in Figure 2, 6 o C on 31 December, was due to a strong cold front producingnortherly winds, a wind direction quite contrary to the typical southeasterly winds on theSouth Texas Gulf Coast at that time <strong>of</strong> year. Northerly winds rather than thetemperature decrease most likely reduced cell concentrations in the turning basin, themost distal part <strong>of</strong> the <strong>Brown</strong>sville Inner Harbor, from the incredibly high numbers on 30December which were attributed to the concentrating effects <strong>of</strong> typical winds <strong>and</strong>currents (T. Reisinger, pers. comm.). The 1990-91 G. breve bloom in <strong>Brown</strong>svilleseems odd relative to the normal <strong>of</strong>fshore initiation <strong>and</strong> growth <strong>of</strong> conspecific blooms inWest Florida <strong>and</strong> Texas, as does the small G. breve bloom that occurred on 8 January1987 in Corpus Christi Bay, about two months after the massive 1986 event(Trebatoski, 1988). Viable cells may have remained in the Corpus Christi Bay systemto produce that brief but dense bloom, but the 1986 bloom may well have seeded<strong>Brown</strong>sville's ship channel with hypnozygotes that excysted in response to unknownstimulating factors in the channel prior to the initial fish kill in late November <strong>of</strong> 1990.This hypothesis could only be confirmed by discovery <strong>of</strong> cysts in ship channelsediments, if any cysts remain, for the 1990-91 bloom at times occupied much if not all<strong>of</strong> the ship channel <strong>and</strong> the associated Lake San Martin, where a sample with 30,000cells/ml was taken on 13 December 1990 (T. Reisinger, pers. comm.).The 1986 G. breve bloom did penetrate well into the inl<strong>and</strong> waterways <strong>of</strong> SouthTexas, passing through such entrances as the ship channel at Port Aransas, wherewater samples were taken at the pier laboratory <strong>of</strong> the University <strong>of</strong> Texas MarineScience Institute (Figure 3; Buskey, unpublished data).7.06.05.04.03.02.01.00.027-Sep29-Sep1-Oct3-OctLog G. breve cells/ml5-Oct7-Oct9-Oct11-Oct13-Oct15-Oct17-Oct19-Oct21-Oct23-Oct25-OctFigure 3. G. breve concentrations as means or singlecounts from water sampled at the UTMSI Pier Lab, PortAransas in 1986. Error bars are + log 10 st<strong>and</strong>ard errors.81


C. Spatial <strong>and</strong> temporal bloom distributionsOnly data from the 1986 G. breve bloom provide enough information on spatial<strong>and</strong> temporal distribution <strong>of</strong> any major bloom. The extent <strong>of</strong> the 1986 bloom can bedepicted in terms <strong>of</strong> coastal Texas counties with din<strong>of</strong>lagellates present in theirnearshore waters in concentrations detectable by aerial survey (Table 2).Table 2. Confirmation <strong>of</strong> G. breve presence by twelve aerial surveys <strong>of</strong> the nearshore waters <strong>and</strong> bays <strong>of</strong>coastal Texas counties from Galveston southward to the Mexican border (excluding Jackson County) inthe fall <strong>of</strong> 1986.* "X" indicates confirmed observation; "?" indicates unconfirmed presence north <strong>of</strong>Matagorda Bay. (Data from Trebatoski, 1988.)GalvestonBrazoriaSEP4SEP9SEP10SEP12SEP14Matagorda X ? X X X X X X X X X XCalhoun X X X X X X X X X X X XRefugio X X X X X X X X X XAransas X X X X X X X X X XSan Patricio X X X X XNueces X XKleberg X XKenedy X XWillacyCameronSEP16*Notes: (1) This G. breve red tide was first observed near Galveston Isl<strong>and</strong> on 27 August 1986 but was no longervisible to aerial surveys north <strong>of</strong> the border between Matagorda <strong>and</strong> Brazoria Counties by 4 September. (2) Aerialsurveys on 23 <strong>and</strong> 24 October 1986 reported no visible red tide along Texas shores.Of special interest is the fact that no red tide was visible by air only one week after themaximum extent <strong>of</strong> G. breve presence along the Texas coast. Spatial distributions mayvary widely, to some degree depending on whether the bloom is inshore or <strong>of</strong>fshore.Temporal distributions are also widely variable <strong>and</strong> may change by many orders <strong>of</strong>magnitude in relatively short periods.D. Fish killsAn estimated 2 million pounds <strong>of</strong> dead fish resulted from the 1935 red tide, forwhich the species responsible was likely G. breve (Snider, 1987). A widelydocumented fish kill, again apparently due to G. breve, littered the Gulf coast from apoint 17 miles north <strong>of</strong> Port Isabel, Texas to a 120-mile stretch in the Mexican state <strong>of</strong>Tamaulipas in September <strong>of</strong> 1955. The 1974 red tide well <strong>of</strong>fshore <strong>of</strong> <strong>Brown</strong>sville <strong>and</strong>Matamoros caused a massive fish kill that eventually littered Mexican beaches, but noestimates <strong>of</strong> the extent <strong>of</strong> fish mortality are known. The bulk <strong>of</strong> available data on fishmortality is associated with the G. breve red tide <strong>of</strong> 1986; Trebatoski (1988) includes anSEP22SEP26SEP29OCT10OCT14OCT16XX82


extensive list <strong>of</strong> all fish species that suffered mortality as identified by the Texas Parks<strong>and</strong> Wildlife Department. Table 3 summarizes the minimum numbers <strong>of</strong> individualskilled for seven common fish species <strong>of</strong> recreational or commercial importance (slightlymodified from Texas Parks <strong>and</strong> Wildlife Department, 1986).Table 3. Minimum estimates <strong>of</strong> fish killed by the G. breve red tide <strong>of</strong> 1986 in inshore <strong>and</strong> <strong>of</strong>fshoreTexas waters. (Data from Texas Parks <strong>and</strong> Wildlife Department, 1986.)Selected SpeciesArea <strong>Red</strong> Spotted Southern Atlantic Black Gulf StripedDrum Seatrout Flounder Croaker Drum Menhaden Mullet TotalBays 15,800 41,300 38,200 70,000 2,100 480,000 790,000 1,437,400Gulf 27,100 39,500 2,800 172,000 1,700 2,770,000 3,240,000 6,253,100Total 42,900 80,800 41,000 242,000 3,800 3,250,000 4,030,000 7,690,500E. Shellfish bedsAmong many other duties, the Division <strong>of</strong> Seafood Safety in the TexasDepartment <strong>of</strong> Health is responsible for determining acceptable levels <strong>of</strong> red tidecontamination in shellfish meats <strong>and</strong> closing shellfish beds when meats contain abovethresholdlevels <strong>of</strong> toxin. Since 1986, South Texas bays <strong>and</strong> inshore bodies <strong>of</strong> water,including two <strong>of</strong> those in the Corpus Christi Bay National Estuary region, have facedclosure for varied lengths <strong>of</strong> time as indicated in Table 4. Because shellfish areprincipally harvested in waters north <strong>of</strong> Port Aransas <strong>and</strong> those are the sites mostfrequently sampled by the Texas Department <strong>of</strong> Health (G. Heideman, pers. comm.),only those bodies <strong>of</strong> water from Aransas Bay <strong>and</strong> northward are listed in Table 4.Notice the duration <strong>of</strong> some closures <strong>and</strong> the dates <strong>of</strong> those shellfish beds that wereclosed well after the 1986 G. breve red tide. Continued shellfish contamination after the1986 bloom <strong>and</strong> the dates <strong>of</strong> later incidents <strong>of</strong> contamination imply that G. breve eitherblooms with a frequency <strong>and</strong>/or persists to degrees only poorly realized at present (cf.Table 1 <strong>and</strong> associated text in Section II. A.).83


Table 4. Dates <strong>of</strong> closure due to red tide (G. breve) contamination <strong>and</strong> subsequent reopening <strong>of</strong> selectedSouth Texas shellfish harvesting areas from 1986-1990 by order <strong>of</strong> the Texas Department <strong>of</strong> Health.(Data courtesy <strong>of</strong> G. Heideman, Texas Department <strong>of</strong> Health.)Area <strong>Status</strong> 1986 1987 1988 1989 1990East Matagorda CLOSED 6 SepBay REOPENED 4 DecMatagorda Bay CLOSED 6 SepREOPENED2 DecTres Palacios CLOSED 6 SepBay REOPENED 20 FebCarancahua CLOSED 6 SepBay REOPENED 20 FebLavaca Bay CLOSED 6 SepREOPENED 20 Feb (in part) 13 Feb (all)Powderhorn CLOSED 6 SepLake REOPENED 15 OctEspiritu Santo CLOSED 6 SepBay REOPENED 15 OctSan Antonio Bay CLOSED 6 SepREOPENED 14 DecMesquite Bay CLOSED 6 SepREOPENED20 FebCopano Bay* CLOSED 6 Sep 25 AugREOPENED 20 Feb 2 SepAransas Bay* CLOSED 6 Sep 10 Jan 25 AugREOPENED 1 Jan; 20 Feb 2 Sep*These bays are included in the Corpus Christi Bay National Estuary region.F. Gulf <strong>of</strong> Mexico circulationElliot (1982) presented convincing evidence <strong>of</strong> one or more anti-cyclonic eddiesor gyres that form <strong>and</strong> depart annually from the Gulf Stream's Loop <strong>Current</strong> west <strong>of</strong>Florida to migrate westward, maintaining some structural <strong>and</strong> thermal integrity for aslong as a year. H<strong>of</strong>mann <strong>and</strong> Worley (1986) later substantiated the existence <strong>of</strong> thesegyres, <strong>and</strong> satellite data have shown that the Loop <strong>Current</strong> occasionally intrudes ontothe West Florida Shelf (P. A. Tester, pers. comm.) where G. breve blooms typicallyinitiate (Steidinger, 1975b). Already documented is the apparent transport <strong>of</strong> G. breveseed populations from the West Florida Shelf in the Loop <strong>Current</strong> around the Floridapeninsula to the shores <strong>of</strong> North Carolina via the Gulf Stream, the cause <strong>of</strong> a novelbloom there in late 1987 (Tester et al., 1991).Should entrainment <strong>of</strong> seed populations occur in the Loop <strong>Current</strong> followed bygyre formation, subsequent eddy transport could carry G. breve populations to the GulfCoast <strong>of</strong> Texas <strong>and</strong> Mexico. Once in the vicinity <strong>of</strong> the Texas-Mexico Shelf, surface84


circulation patterns could work in concert with the decaying anti-cyclonic gyre tointegrate gyre waters <strong>and</strong> their plankton into the westward coastal current described byBarron <strong>and</strong> Vastano (1994). The westward coastal current results from the inferredexistence <strong>of</strong> a cyclonic gyre normally present from September through May, thesouthernmost extent <strong>of</strong> which marks a convergence <strong>of</strong> coastal currents driven by theprevailing southerly to southeasterly winds in the Western Gulf <strong>of</strong> Mexico (Cochrane<strong>and</strong> Kelly, 1986; H<strong>of</strong>mann <strong>and</strong> Worley, 1986). The prevailing winds, counteredsomewhat in winter by the northerly winds <strong>of</strong> cold fronts, produce a convergence <strong>of</strong>coastal flow patterns at a seasonally variable location somewhere between theMatagorda Peninsula <strong>and</strong> <strong>Brown</strong>sville (Watson <strong>and</strong> Behrens, 1970; Barron <strong>and</strong>Vastano, 1994). The geopotential anomaly data <strong>of</strong> Cochrane <strong>and</strong> Kelly (1986),according to Barron <strong>and</strong> Vastano (1994), can only describe a very general cycloniccurrent pattern <strong>of</strong>f the Texas-Louisiana Shelf that is subject to significant short-termvariability. That variability, however, does not preclude the possibility that both theSouth Texas coastal flow convergence <strong>and</strong> the cyclonic gyre on the Texas-LouisianaShelf could bring G. breve populations into nearshore waters as far north as Galveston(the initial site <strong>of</strong> the 1986 bloom) or along the coast <strong>of</strong> South Texas <strong>and</strong> Mexico (as inthe <strong>of</strong>fshore blooms <strong>of</strong> 1935, 1955 <strong>and</strong> 1974) whether G. breve is resident in the<strong>of</strong>fshore waters <strong>of</strong> the Texas shelf (Geesey <strong>and</strong> Tester, 1993) or transported byanticyclonic gyre from the Eastern Gulf. The same coastal current patterns (theconvergence <strong>and</strong> cyclonic gyre) could facilitate immigration <strong>of</strong> A. monilata populationsfrom the north central Gulf, as described below.XI. Potential status <strong>of</strong> red tide species in Texas coastal watersAny potential means <strong>of</strong> controlling Gymnodinium breve (<strong>and</strong> probablyAlex<strong>and</strong>rium monilata) blooms that involves the destruction <strong>of</strong> the cells may do more toexacerbate rather than prevent or diminish the toxic effect <strong>of</strong> the bloom. For example, acytolytic chemical extracted from a blue-green alga together with an effective, non-toxicdelivery medium could be used to reduce dense concentrations <strong>of</strong> G. breve (Eng-Wilmot et al., 1979a,b) but would prompt the release <strong>of</strong> brevetoxin upon lysis. Thepotential status <strong>of</strong> G. breve <strong>and</strong> A. monilata as problematic bloom species along theSouth Texas Gulf Coast, therefore, may not be subject to human control except viareduction <strong>of</strong> anthropogenic eutrophication <strong>of</strong> inshore <strong>and</strong> coastal waters, if excessnutrients actually promote toxic din<strong>of</strong>lagellate blooms (Hallegraeff, 1993).Regarding the potential status <strong>of</strong> A. monilata, Howell's (1953) work on Florida'seast coast predated documentation <strong>of</strong> A. monilata blooms <strong>and</strong> associated fish kills onthe west coast <strong>of</strong> Florida by almost twenty years (Williams <strong>and</strong> Ingle, 1972), though thefirst detection <strong>of</strong> this din<strong>of</strong>lagellate south <strong>of</strong> Fort Myers, Florida actually occurred inearly August <strong>of</strong> 1966. A fish kill soon followed, <strong>and</strong> from the first report <strong>of</strong> mortality on16 August until the end <strong>of</strong> that month, high cell counts discolored the waters <strong>and</strong>continued to kill fish between Tampa Bay <strong>and</strong> Cape Romano. A. monilataconcentrations finally fell to low levels by late September. Williams <strong>and</strong> Ingle (1972)pointed out that records <strong>of</strong> A. monilata up to that time had been known only from85


estuaries such as those in the Chesapeake Bay, on the coast <strong>of</strong> Venezuela <strong>and</strong> in theGalveston, Texas area. They also found high concentrations from less than a mile to42 miles <strong>of</strong>fshore <strong>of</strong> West Florida in late 1966, with a maximum <strong>of</strong> slightly over 1.3million cells/l at a site 400 meters out from the barrier isl<strong>and</strong>s near Sarasota on 21August. The possibility exists, therefore, for <strong>of</strong>fshore blooms <strong>of</strong> this din<strong>of</strong>lagellate inSouth Texas as well.Because A. monilata does not contaminate shellfish <strong>and</strong> may have only a slightimpact on finfish depending on the location, small blooms may be frequent in SouthTexas, escaping <strong>of</strong>fical attention <strong>and</strong> documentation. Seed populations may bearriving from the north central Gulf <strong>of</strong> Mexico, also, for A. monilata cells from awidespread bloom in September <strong>of</strong> 1980 <strong>of</strong>f Louisiana not only occupied numerousbayous <strong>and</strong> sounds but were found in large numbers in water samples from west <strong>of</strong> theMississippi River (Perry, 1980), exposing them to possible transport to Texas in thewestward coastal current resulting from the river's outflow. Several A. monilata bloomsoccurred at about that time in the north central Gulf, one in August 1979, the onementioned above <strong>and</strong> another in July 1981 (Perry <strong>and</strong> McLell<strong>and</strong>, 1981). Thoseoutbreaks showed some correlation with lowered salinity which, if causal, may promptblooms along the Texas coast should a hurricane or tropical storm decrease salinitiesin inshore <strong>and</strong> nearshore waters, perhaps especially in areas that have sufferedprevious A. monilata blooms.Discussion <strong>of</strong> the postulated trans-Gulf movement <strong>of</strong> G. breve populations inLoop <strong>Current</strong> gyres, if it occurs, should not obscure the potential for endemicpopulations <strong>of</strong> the din<strong>of</strong>lagellate in the <strong>of</strong>fshore waters <strong>of</strong> Texas to seed nearshoreblooms. Geesey <strong>and</strong> Tester (1993) report that G. breve is found throughout the Gulf <strong>of</strong>Mexico, but, until their study, its cell density in nearshore <strong>and</strong> <strong>of</strong>fshore waters <strong>of</strong>various depths was largely unknown. For twelve months beginning in March <strong>of</strong> 1990,NOAA vessels sampled 61 stations throughout the Gulf <strong>of</strong> Mexico from depths <strong>of</strong> 0 to>150 m. The density <strong>of</strong> G. breve in shallow, well-mixed waters was constant, but indeeper waters with a thermocline, the din<strong>of</strong>lagellate was more abundant near thesurface. Concentrations in central Gulf waters remained at 100 cells/ml,including two stations <strong>of</strong>fshore <strong>of</strong> Galveston, Texas. Concentrations from 1-100cells/ml occurred at coastal stations between St. Joseph Isl<strong>and</strong> <strong>and</strong> Matagorda Isl<strong>and</strong>.With the well-documented westward coastal current along the Texas coast (Barron <strong>and</strong>Vastano, 1994), concentrations <strong>of</strong> G. breve at >100 cells/ml could theoretically befound anywhere along the coastline included in the Corpus Christi Bay NationalEstuary region. The low frequency <strong>of</strong> major <strong>of</strong>fshore blooms implies that conditionsfavorable for initiation do not commonly occur in South Texas waters, but the frequentdetection <strong>of</strong> brevetoxin-contaminated shellfish <strong>and</strong> the closures <strong>of</strong> shellfish beds by theTexas Department <strong>of</strong> Health imply that red tide blooms <strong>of</strong>ten occur without beingotherwise documented (G. Heideman, pers. comm.; see Tables 1 <strong>and</strong> 4). Similarly,Steidinger's (1975) statement that 75% <strong>of</strong> G. breve blooms that begin <strong>of</strong>fshore <strong>of</strong> WestFlorida never make it to nearshore waters may also apply to <strong>of</strong>fshore waters <strong>of</strong> South86


Texas, a possibility that may only be substantiated if toxic din<strong>of</strong>lagellates become thefocus <strong>of</strong> an <strong>of</strong>fshore sampling program.Besides the unknown extent <strong>of</strong> the negative economic impact <strong>of</strong> G. breve bloomson the South Texas shellfishing industry, not to mention recreational fishing <strong>and</strong>tourism, the relatively infrequent red tide problems experienced in Texas coastal <strong>and</strong>inshore waters do not apparently require any changes in current marine policy. <strong>Red</strong>tide blooms should be understood as naturally occurring events, unless G. breve hasbeen or could be introduced by the discharge <strong>of</strong> ships ballast water containing viablecells or hypnozygotes (Hallegraeff <strong>and</strong> Bolch, 1992). Should toxic blooms in Texasincrease in frequency <strong>and</strong> severity, however, Texas marine policy should definitelychange to provide warning <strong>of</strong> bloom initiation, expansion <strong>and</strong> transport <strong>and</strong> to collectthe data necessary to define the factors involved in each step <strong>of</strong> a bloom.XII. Identification <strong>of</strong> data <strong>and</strong> information gaps1. The most obvious gap is in the lack <strong>of</strong> consistent data on red tide concentrationsduring <strong>and</strong> after a bloom (if not also before). This requires regular water sampling <strong>of</strong>established sites that ideally includes measurement <strong>of</strong> in situ temperature, salinity <strong>and</strong>perhaps winds <strong>and</strong> currents. Samples should then be analyzed with respect tonutrients <strong>and</strong> biologically active organic compounds as well as cell concentrations.2. Thorough sediment analyses, with the object <strong>of</strong> looking for hypnozygotes at thebottom <strong>of</strong> such areas as the turning basins in Corpus Christi Inner Harbor <strong>and</strong> the<strong>Brown</strong>sville Ship Channel, would provide useful <strong>and</strong> novel information on the life cycles<strong>of</strong> both A. monilata <strong>and</strong> G. breve in Texas waters regardless <strong>of</strong> the results. Similaranalyses <strong>of</strong> ship's ballast water for cysts or cells in cargo vessels coming from portsknown to have G. breve blooms (such as those on Florida's West Coast) would also beuseful (Hallegraeff <strong>and</strong> Bolch, 1992).3. Examining future <strong>of</strong>fshore G. breve blooms with continually available satelliteimagery (Gallegos, 1990) <strong>and</strong> aerial or shipboard means <strong>of</strong> determining temperature,salinity, nutrients <strong>and</strong> the bloom's surface coverage would be useful in delineatinginitiation, growth <strong>and</strong> eddy transport factors. Likewise, conducting regular sampling <strong>of</strong>the coastal current <strong>and</strong> sites <strong>of</strong> previous A. monilata blooms would be helpful butperhaps not as high a priority relative to the greater negative impact <strong>of</strong> G. breveblooms.4. Automatic notification <strong>of</strong> incidents <strong>of</strong> shellfish contamination <strong>and</strong> information transferbetween the Shellfish Sanitation Division <strong>of</strong> the Texas Department <strong>of</strong> Health <strong>and</strong>scientists in both academia <strong>and</strong> other public agencies who are interested in toxicblooms would greatly facilitate timely work with G. breve whenever concentrations arehigh enough to produce positive toxin bioassays. This notification could be as simple,inexpensive <strong>and</strong> efficient as one message sent to a st<strong>and</strong>ard list <strong>of</strong> researchers <strong>and</strong>governmental agencies. In turn, any recipient would also be able to notify the87


Department <strong>of</strong> Health <strong>and</strong> all other listed recipients <strong>of</strong> any new blooms. In fact, thebulletin board-style service <strong>of</strong> the recently established "TexCoast" on the Internet mayserve in this capacity. [For further information on "TexCoast," contact Don Hockaday(hockaday@panam.edu) or Dr. Terry Whitledge (terry@utmsi.zo.utexas.edu).]88


XIII. Literature CitedAbdelghani, A., Hartley, W. R., Esmundo, F. R., <strong>and</strong> Harris, T. F. (1994). Biological <strong>and</strong>chemical contaminants in the Gulf <strong>of</strong> Mexico <strong>and</strong> the potential impact on public health:a characterization report. Tulane University School <strong>of</strong> Public Health <strong>and</strong> TropicalMedicine, New Orleans, LA. 77 pp.Aldrich, D. V. (1962). Photoautotrophy in Gymnodinium breve Davis. Science. 37(3534): 988-990.Aldrich, D. V., Wilson, W. B. (1960). The effect <strong>of</strong> salinity on growth <strong>of</strong> Gymnodiniumbreve Davis. Biol. Bull. 119: 57-64.Anderson, D. M. (1994). <strong>Red</strong> tides. Scientific American. August, 1994.Anderson, D. M., Wall, D. (1978). Potential importance <strong>of</strong> benthic cysts <strong>of</strong> Gonyaulaxtamarensis <strong>and</strong> G. excavata in initiating toxic din<strong>of</strong>lagellate blooms. Journal <strong>of</strong>Phycology. 14 (2): 224-234.Baden, D. G., Tomas, C. R. (1989). Variations in major toxin composition for six clones<strong>of</strong> Ptychodiscus brevis. <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology.Proceedings <strong>of</strong> the First International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14,1987, in Takamatsu, Kagawa Prefecture, Japan. T. Okaichi, D. M. Anderson <strong>and</strong> T.Nemoto (eds). Elsevier Science Publishing Company, Inc., New York. p. 415-418.Baldridge, H. D. (1975). Temperature patterns in the long-range prediction <strong>of</strong> red tide inFlorida waters. In: Proceedings <strong>of</strong> The First International Conference on ToxicDin<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). The Massachusetts Science <strong>and</strong>Technology Foundation, Wakefield, MA. p. 69-79.Barron, C. N., Jr., Vastano, A. C. (1994). Satellite observations <strong>of</strong> surface circulation inthe northwestern Gulf <strong>of</strong> Mexico during March <strong>and</strong> April 1989. Continental ShelfResearch. 14 (6): 607-628.Cochrane, J. D., Kelly, F. J. (1986). Low-frequency circulation on the Texas-Louisianacontinental shelf. Journal <strong>of</strong> Geophysical Research . 91 (C9): 10,645-10,659.Collier, A. (1958). Some biochemical aspects <strong>of</strong> red tides <strong>and</strong> related oceanographicproblems. Limnol. Oceanogr. 3: 33-39.Collier, A., Wilson, W. B., Borkowski, M. (1969). Responses <strong>of</strong> Gymnodinium breveDavis to natural waters <strong>of</strong> diverse origins. J. Phycol. 5: 168-172.Connell, C. H., Cross, J. B. (1950). Mass mortality <strong>of</strong> fish associated with the protozoanGonyaulax in the Gulf <strong>of</strong> Mexico. Science. 112 (2909): 359-363.89


Elliott, B. A. (1982). Anticyclonic rings in the Gulf <strong>of</strong> Mexico. Journal <strong>of</strong> PhysicalOceanography. 12: 1292-1309.Eng-Wilmot, D. L., McCoy, L. F., Jr., Martin, D. F. (1979). Isolation <strong>and</strong> synergism <strong>of</strong> ared tide (Gymnodinium breve) cytolytic factor(s) from cultures <strong>of</strong> Gomphosphaeriaaponina. In: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second InternationalConference on Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5,1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds). Elsevier/North-Holl<strong>and</strong>, Inc. p. 355-360.Eng-Wilmot, D. L., Henningsen, B. F., Martin, D. F., Moon, R. E. (1979). Model solventsystems for delivery <strong>of</strong> compounds cytolytic towards Gymnodinium breve. In: ToxicDin<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conference on ToxicDin<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H.H. Seliger (eds). Elsevier/North-Holl<strong>and</strong>, Inc. p. 361-366.Freeberg, L. R., Marshall, A., Heyl, M. (1979). Interrelationships <strong>of</strong> Gymnodinium breve(Florida red tide) within the phytoplankton community. In: Toxic Din<strong>of</strong>lagellate Blooms,Proceedings <strong>of</strong> the Second International Conference on Toxic Din<strong>of</strong>lagellate Blooms,Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds).Elsevier/North-Holl<strong>and</strong>, Inc. p. 139-144.Gallegos, S. (1990). Evaluation <strong>of</strong> the potential <strong>of</strong> the NOAA-n AVHRR reflective datain oceanography (dissertation). Department <strong>of</strong> Oceanography, Texas A & M University,College Station, TX, USA. 189 pp.Gates, J. A., Wilson, W. B. (1960). The toxicity <strong>of</strong> Gonyaulax monilata Howell to Mugilcephalus. Limnol. Oceanogr. 5 (2): 171-174.Geesey, M., Tester, P. A. (1993). Gymnodinium breve: ubiquitous in Gulf <strong>of</strong> Mexicowaters? In: Toxic Phytoplankton Blooms in the Sea, Proceedings <strong>of</strong> the FifthInternational Conference on Toxic Marine Phytoplankton. T. J. Smayda <strong>and</strong> Y. Shimizu(eds). Elsevier p. 251-255.Gunter, G. (1951). Mass mortality <strong>and</strong> din<strong>of</strong>lagellate blooms in the Gulf <strong>of</strong> Mexico.Science 113:250-251.Haddad, K. D., Carder K. L. (1979). Oceanic intrusion: one possible initiationmechanism <strong>of</strong> red tide blooms on the west coast <strong>of</strong> Florida. In: Toxic Din<strong>of</strong>lagellateBlooms, Proceedings <strong>of</strong> the Second International Conference on Toxic Din<strong>of</strong>lagellateBlooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger(eds). Elsevier/North-Holl<strong>and</strong>, Inc. p. 269-274.Hallegraeff, G. M. (1993). A review <strong>of</strong> harmful algal blooms <strong>and</strong> their apparent globalincrease. Phycologia 32 (2): 79-99.90


Hemmert, W. H. (1975). The public health implications <strong>of</strong> Gymnodinium breve redtides, a review <strong>of</strong> the literature <strong>and</strong> recent events. In : Proceedings <strong>of</strong> The FirstInternational Conference on Toxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). TheMassachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MA. p. 489-497.H<strong>of</strong>mann, E. E., Worley, S. J. (1986). An investigation <strong>of</strong> the circulation <strong>of</strong> the Gulf <strong>of</strong>Mexico. Journal <strong>of</strong> Geophysical Research. 91(C12): 14,221-14.236.Howell, J. F. (1953). Gonyaulax monilata, sp. nov., the causative din<strong>of</strong>lagellate <strong>of</strong> a redtide on the east coast <strong>of</strong> Florida in August-September, 1951. Transactions <strong>of</strong> theAmerican Microscopical Society. 72: 153-156.Huntley, M., Sykes, P., Rohan, S., Marin, V. (1986). Chemically-mediated rejection <strong>of</strong>din<strong>of</strong>lagellate prey by the copepods Calanus pacificus <strong>and</strong> Paracalanus parvus:mechanism, occurrence <strong>and</strong> significance. Marine Ecology Progress Series. 28: 105-120.Jensen, A. C. (1975). The economic halo <strong>of</strong> a red tide. In: Proceedings <strong>of</strong> The FirstInternational Conference on Toxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). TheMassachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MA. p. 507-516.Jensen, D. A., Bowman, J. (1975). On the occurrence <strong>of</strong> the "red tide" Gonyaulaxmonilata in the Corpus Christi Inner Harbor (July-September 1975). Texas WaterQuality Board, District 12. 20 pp.Joyce, E. A., Jr., Roberts, B. S. (1975). Florida Department <strong>of</strong> Natural Resources <strong>Red</strong><strong>Tide</strong> Research Program. In: Proceedings <strong>of</strong> The First International Conference onToxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). The Massachusetts Science <strong>and</strong>Technology Foundation, Wakefield, MA. p. 95-103.Lund, E. J. (1936). Some facts relating to the occurrences <strong>of</strong> dead <strong>and</strong> dying fish on theTexas coast during June, July, <strong>and</strong> August, 1935 In: Annual Report <strong>of</strong> the Texas Game,Fish <strong>and</strong> Oyster Commission, 1934-35. Texas Game, Fish <strong>and</strong> Oyster Commission. p.47-50.Marvin, K. T. (1965). Operation <strong>and</strong> maintenance <strong>of</strong> salt-water laboratories. In: FisheryResearch: Biological Laboratory, Galveston, Fiscal Year 1964, Circular 230. UnitedStates Department <strong>of</strong> the Interior, Fish <strong>and</strong> Wildlife Service, Bureau <strong>of</strong> CommercialFisheries. p. 84-86.Perry, H.M. (1980). Din<strong>of</strong>lagellate blooms occur <strong>of</strong>f Lousiana. Coastal OceanogClimatol News 3(1): 3.91


Perry, H.M., McLell<strong>and</strong> J.A. (1981). First recorded observance <strong>of</strong> the din<strong>of</strong>lagellateProrcentrum minimum (Pavillard) Schiller 1933 in Mississippi Sound <strong>and</strong> adjacentwaters. Gulf Res Repts. 7(1): 83-85.Pierce, R. H., Henry, M. S., Pr<strong>of</strong>fitt, L. S., Hasbrouck, P. A. (1990). <strong>Red</strong> tide toxin(brevetoxin) enrichment in marine aerosol. In: Toxic Marine Phytoplankton,Proceedings <strong>of</strong> the Fourth International Conference on Toxic Marine Phytoplankton. E.Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. Anderson (eds). Elsevier p. 128-131.Proctor, R. R., Jr.(1965). Special Report: Biological Indicators in East Lagoon,Galveston Isl<strong>and</strong> In: Fishery Research: Biological Laboratory, Galveston, Fiscal Year1964, Circular 230. United States Department <strong>of</strong> the Interior, Fish <strong>and</strong> Wildlife Service,Bureau <strong>of</strong> Commercial Fisheries. p. 87-88.Proctor, R. R., Jr. (1966). Special Report: Oyster growth experiment in East Lagoon In:Annual Report <strong>of</strong> the Bureau <strong>of</strong> Commercial Fisheries Biological Laboratory, Galveston,Texas, Fiscal Year 1965, Circular 246. United States Department <strong>of</strong> the Interior, Fish<strong>and</strong> Wildlife Service, Bureau <strong>of</strong> Commercial Fisheries. p. 48-49.Quick, J. A., Jr., Henderson, G. E. (1975). Evidences <strong>of</strong> new ichtyointoxicativephenomena in Gymnodinium breve red tides. In: Proceedings <strong>of</strong> The First InternationalConference on Toxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). The MassachusettsScience <strong>and</strong> Technology Foundation, Wakefield, MA. p. 413-422.Ray, S. M., Aldrich, D. V. (1967). Ecological interactions <strong>of</strong> toxic din<strong>of</strong>lagellates <strong>and</strong>molluscs in the Gulf <strong>of</strong> Mexico. In: Animal Toxins. F. Russell <strong>and</strong> P. Saunders (eds).Pergamon Press, NY. p. 75-83.Riley, C. M., Holt, S. A., Holt, G. J., Buskey, E. J., Arnold, C. R. (1989). Mortality <strong>of</strong>larval red drum (Sciaenops ocellatus) associated with a Ptychodiscus brevis red tide.Contributions in Marine Science 31: 137-146.Roberts, B. S. (1979). Occurrence <strong>of</strong> Gymnodinium breve red tides along the west <strong>and</strong>east coasts <strong>of</strong> Florida during 1976 <strong>and</strong> 1977. In: Toxic Din<strong>of</strong>lagellate Blooms,Proceedings <strong>of</strong> the Second International Conference on Toxic Din<strong>of</strong>lagellate Blooms,Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds).Elsevier/North-Holl<strong>and</strong>, Inc. p. 199-202.Roberts, B. S., Henderson, G. E., Medlyn, R. A. (1979). The effect <strong>of</strong> Gymnodiniumbreve toxin(s) on selected mollusks <strong>and</strong> crustaceans. In: Toxic Din<strong>of</strong>lagellate Blooms,Proceedings <strong>of</strong> the Second International Conference on Toxic Din<strong>of</strong>lagellate Blooms,Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds).Elsevier/North-Holl<strong>and</strong>, Inc. p. 419-424.92


Rounsefell, G. A., Nelson, W. R. (1966). <strong>Red</strong>-tide research summarized to 1964including an annotated bibliography. United States Fish <strong>and</strong> Wildlife Service SpecialScientific Report, Fisheries No. 535, Washington, D. C. 85 pp.Roszell, L. E., Schulman, L. S., Baden, D. G. (1990). Toxin pr<strong>of</strong>iles are dependent ongrowth stages in cultured Ptychodiscus brevis. Toxic Marine Phytoplankton,Proceedings <strong>of</strong> the Fourth International Conference on Toxic Marine Phytoplankton. E.Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. Anderson (eds). Elsevier. p. 403-406.Seliger, H. H., Tyler, M. A., McKinley, K. R. (1979). Phytoplankton distributions <strong>and</strong> redtides resulting from frontal circulation patterns. In: Toxic Din<strong>of</strong>lagellate Blooms,Proceedings <strong>of</strong> the Second International Conference on Toxic Din<strong>of</strong>lagellate Blooms,Key Biscayne, Florida, Oct. 31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds).Elsevier/North-Holl<strong>and</strong>, Inc. p. 239-248.Shumway, S. E. (1990). A review <strong>of</strong> the effects <strong>of</strong> algal blooms on shellfish <strong>and</strong>aquaculture. Journal <strong>of</strong> the World Aquaculture Society. 21 (2): 65-104.Sievers, A. M. (1969). Comparative toxicity <strong>of</strong> Gonyaulax monilata <strong>and</strong> Gymnodiniumbreve to annelids, crustaceans, molluscs <strong>and</strong> a fish. Journal <strong>of</strong> Protozoology 16 (3):401-404.Snider, R. (ed.). (1987). <strong>Red</strong> tide in Texas: an explanation <strong>of</strong> the phenomenon. MarineInformation Service, Texas A&M Sea Grant College Program. 4 pp.Stehn, T. (1987). Whooping cranes during the 1986-1987 winter. Aransas NationalWildlife Refuge, U. S. Fish <strong>and</strong> Wildlife Service. 45 pp.Steidinger, K. A. (1975a). Basic factors influencing red tides. In: Proceedings <strong>of</strong> theFirst International Conference on Toxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). TheMassachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MA. p. 153-162.Steidinger, K. A. (1975b). Implications <strong>of</strong> din<strong>of</strong>lagellate life cycles on initiation <strong>of</strong>Gymnodinium breve red tides. Environmental Letters. 9 (2): 129-139.Steidinger, K. A. (1979). Collection, enumeration <strong>and</strong> identification <strong>of</strong> free-living marinedin<strong>of</strong>lagellates. In: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the SecondInternational Conference on Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct.31-Nov. 5, 1978. D. L. Taylor <strong>and</strong> H. H. Seliger (eds). Elsevier/North-Holl<strong>and</strong>, Inc. p.435-442.Steidinger, K. A. (1983). A re-evaluation <strong>of</strong> toxic din<strong>of</strong>lagellate biology <strong>and</strong> ecology In:Progress in Phycological Research, Vol. 2. Round <strong>and</strong> Chapman (eds). ElsevierScience Publishers B. V. p.147-188.93


Steidinger, K. A., Ingle, R. M. (1972). Observations on the 1971 summer red tide inTampa Bay, Florida. Environmental Letters 3 (4): 271-278.Steidinger, K. A., Joyce, E. A., Jr. (1973). Educational Series No. 17: Florida <strong>Red</strong><strong>Tide</strong>s. State <strong>of</strong> Florida Department <strong>of</strong> Natural Resources, St. Petersburg. 26 pp.Steidinger, K. A., Vargo, G. A. (1988). Marine din<strong>of</strong>lagellate blooms: dynamics <strong>and</strong>impacts. In: Algae <strong>and</strong> Human Affairs. C. A. Lembi <strong>and</strong> J. R. Waal<strong>and</strong> (eds). CambridgeUniversity Press, New York, NY. p. 373-401.Tester, P.A., Fowler, P.K. (1990). Brevetoxin contamination <strong>of</strong> Mercenaria mercenaria<strong>and</strong> Crassostrea virginica: a management issue. In: Toxic Marine Phytoplankton,Proceedings <strong>of</strong> the Fourth International Conference on Toxic Marine Phytoplankton E.Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. Anderson (eds). Elsevier. p.499-503.Tester, P. A., Stumpf, R. P., Vukovich, F. M., Fowler, P. K., Turner, J. T. (1991). Anexpatriate red tide bloom: transport, distribution, <strong>and</strong> persistence. Limnol. Oceanogr. 6(5): 1053-1061.Texas Parks <strong>and</strong> Wildlife Department. (1986). Commission Agenda Item (BriefingSession): <strong>Red</strong> <strong>Tide</strong>. 3 pp.Trebatoski, B. (1988). Observations on the 1986-1987 Texas <strong>Red</strong> <strong>Tide</strong> (Ptychodiscusbrevis), Report 88-02. Texas Water Commission. 48 pp.Vargo, G. A., Carder, K. L., Gregg, W., Shanley, E., Neil, C., Steidinger, K. A., Haddad,K. D. (1987). The potential contribution <strong>of</strong> primary production by red tides to the westFlorida shelf ecosystem. Limnol. Oceanogr. 32 (3): 762-767.Walker, L. M. (1982). Evidence for a sexual cycle in the Florida red tide din<strong>of</strong>lagellate,Ptychodiscus brevis (= Gymnodinium breve). Transactions <strong>of</strong> the AmericanMicroscopical Society 101(3): 287-293.Walker, L. M., Steidinger, K. A. (1979). Sexual reproduction in the toxic din<strong>of</strong>lagellateGonyaulax monilata. Journal <strong>of</strong> Phycology 15: 312-315.Wardle, W. J., Ray, S. M., Aldrich, A. S. (1975). Mortality <strong>of</strong> marine organismsassociated with <strong>of</strong>fshore summer blooms <strong>of</strong> the toxic din<strong>of</strong>lagellate Gonyaulax monilataHowell at Galveston, Texas In: Proceedings <strong>of</strong> the First International Conference onToxic Din<strong>of</strong>lagellate Blooms. V. R. LoCicero (ed). The Massachusetts Science <strong>and</strong>Technology Foundation, Wakefield, MA. p. 257-263.Watson, R. L., Behrens, E. W. (1970). Nearshore surface currents, southeastern TexasGulf Coast. Contributions in Marine Science 15: 133-143.94


Williams, J., Ingle, R. M. (1972). Ecological Notes on Gonyaulax monilata(Dinophyceae) Blooms Along the West Coast <strong>of</strong> Florida. In: Florida Department <strong>of</strong>Natural Resources Leaflet Series: Phytoplankton, Part 1 (Din<strong>of</strong>lagellates), No. 5.Wilson, W. B., Ray, S. M. (1956). The occurrence <strong>of</strong> Gymnodinium brevis in thewestern Gulf <strong>of</strong> Mexico. Ecology 87 (2) :388.95


XIV. Materials <strong>and</strong> Methods: Unpublished DataKen H. Dutton: Seagrass Density <strong>and</strong> Biomass.Measurements <strong>of</strong> total plant biomass were made at 2 to 3 month intervals fromfour replicate samples collected at the two stations with a 9 cm diameter coring device.Samples were thoroughly cleaned <strong>of</strong> any epiphyte material in the laboratory, separatedinto above-ground <strong>and</strong> below-ground live biomass (to calculate root:shoot ratios) <strong>and</strong>dried at 60 o C to a constant weight. Results are expressed as total biomass (g dry wtm -2 ) <strong>of</strong> all shoot, rhizome <strong>and</strong> root material collected in each core. Measurements <strong>of</strong>leaf elongation rate <strong>and</strong> shoot production were collected using the leaf-clippingtechnique described in Virnstein 1982. for Halodule wrightii (the leaves <strong>of</strong> H. wrightii aretoo thin for leaf-tagging techniques). Shoots were clipped about 2 cm above the basalsheath to permit regrowth, <strong>and</strong> cores were collected from each clipped area to measurenet growth at 3 to 6 week intervals. In the laboratory, the length <strong>of</strong> the newly formedblades was recorded for individual shoots <strong>and</strong> all new blade material was pooled fromeach <strong>of</strong> four replicate cores for determination <strong>of</strong> shoot production on a dry weight basis.the mean leaf elongation rate for each replicate core sample was based on themeasurement <strong>of</strong> 20 to 30 blades (Adapted from Dunton 1994).G. Joan Holt: Larval Feeding <strong>and</strong> SurvivalStudies were carried out in brown tide containing ponds at the GCCA hatchery inCorpus Christi Texas. Feeding studies were carried out in 20 oz soda bottles in whichthe neck had been removed <strong>and</strong> 3, 8 cm X 5 cm, panels were cut. The resultingopenings were covered with 100 µm nitex mesh nets. For laboratory studies, the bottleswere placed inside 1 liter beakers filled with either 48 µm filtered brown tide water (witha concentration <strong>of</strong> brown tide cells <strong>of</strong> 1 million ml -1 or greater) or control sea water(without brown tide). For field studies, the bottles were placed in either a 48 µm meshbag, or a control water filled plastic bag <strong>and</strong> hung from a floating rack. Five to fifteenlarvae, all the same age, were placed in each bottle, <strong>and</strong> allowed to acclimate for 1hour prior to food addition. The 5 <strong>and</strong> 7 day old larvae were fed rotifers at aconcentration <strong>of</strong> 5 ml -1 . The 14 day old larvae were fed Artemia at a concentration <strong>of</strong> 3ml -1 . After a one hour feeding time the larvae were removed <strong>and</strong> gut dissection wasperformed to count the number <strong>of</strong> prey items consumed.Hatch rates <strong>and</strong> survival studies in the field were also carried out with thebagged bottles hung from the floating rack. While the laboratory studies wereperformed in 100 ml beakers filled with the test treatments <strong>of</strong> brown tide water orcontrol seawater. To each treatment a defined number <strong>of</strong> eggs were added. The eggswere left overnight <strong>and</strong> the unhatched eggs <strong>and</strong> dead larvae were removed <strong>and</strong>counted to determine hatching rates <strong>and</strong> day 1 survival rates. Each day <strong>of</strong> theexperiment the dead larvae were removed <strong>and</strong> counted to obtain survival rates.Scott A. Holt: Larval Densities.96


Larvae were collected with five replicate ichthyoplankton tows using a 1-m net(500 µm mesh) attached to an epibenthic sled pulled through the lower 1m <strong>of</strong> the watercolumn. The volume <strong>of</strong> water filtered for each tow was measured with a mechanicalflowmeter. Samples were preserved in 5% formalin or 70% ethanol. All sciaenids in thesamples were identified to species (except Menticirrhus sp.), enumerated, <strong>and</strong>measured. Larval densities are expressed an number per 1000 m 3 <strong>of</strong> water (adaptedfrom Holt, et al. 1994).Paul A. Montagna: Benthic Biomass, Abundance, <strong>and</strong> Diversity.The micr<strong>of</strong>auna were sampled with diver held core tubes with diameters <strong>of</strong> 6.7cm. The cores were sectioned at the 0-3 cm <strong>and</strong> 3-10 cm depth intervals. Themei<strong>of</strong>auna were sampled with diver held core tubes with diameters <strong>of</strong> 1.8 cm. Thesecores were sectioned at the 0-1 cm <strong>and</strong> 1-3 cm depth intervals. Three replicates, takenwithin a 2m radius, <strong>of</strong> each sample was obtained. Samples were preserved with 5%buffered formalin, sorted, identified <strong>and</strong> counted. The macr<strong>of</strong>auna samples were usedto measure biomass. Samples were dried for 24 hr. at 55 o C <strong>and</strong> weighed. Beforedrying, mollusks were placed in 1N HCl to dissolve the carbonate shells, <strong>and</strong> washed(adapted from Montagna <strong>and</strong> Kalke 1992).Dean A. Stockwell: Chlorophyll a.See Chlorophyll a analysis methods from Terry E. Whitledge: Nutrient <strong>and</strong>Hydrographic Conditions.Terry E. Whitledge: Nutrient <strong>and</strong> Hydrographic Conditions.Samples were collected monthly from 57 stations in upper Laguna Madre <strong>and</strong>Mansfield Pass in lower Laguna Madre. A SeaBird model SBE 19 CTD pr<strong>of</strong>iler wasused to obtain salinity <strong>and</strong> temperature data with respect to depth. Discrete watersamples for salinity by refractometer, nutrients (nitrate, ammonium, nitrite, phosphate<strong>and</strong> silicate), chlorophyll a <strong>and</strong> Phaeophytin were collected near the surface by h<strong>and</strong>,<strong>and</strong> near the bottom with a Van Dorn style water sampler. The samples were collectedin clean pre- labeled polyethylene bottles <strong>and</strong> immediately chilled with ice in the dark.Secchi depths were measured on all stations.Nutrient analyses were determined on the chilled samples with a segmented flowautomated chemical analyzer using methods developed for marine <strong>and</strong> estuarinewaters (Whitledge et al. 1981, 1989) Chlorophyll <strong>and</strong> phaeopigment measurementswere determined by the fluorometric method (Holm-Hansen et al. 1965), withmodification for 60% acetone: 40% DMSO extraction at 4 o C (Dagg <strong>and</strong> Whitledge,1991, Stockwell 1989) (adapted from Whitledge 1991).97


XV. Annotated BibliographyTHE ANNOTATED BIBLIOGRAPHY OF THE"BROWN TIDE AND RED TIDE CURRENT STATUS" CONTRACT WITH THECCBNEP[All entries pertain wholly or in part to red or brown tide events or experiments; if a sourceis general, key words will reflect only sections pertinent to brown or red tide.]Author: Abdelghani, A., Hartley, W. R., Esmundo, F. R., <strong>and</strong> Harris, T. F.Date: 1994Title: Biological <strong>and</strong> chemical contaminants in the Gulf <strong>of</strong> Mexico <strong>and</strong> the potentialimpact on public health: a characterization reportPages: 77 pp.Source: Tulane University School <strong>of</strong> Public Health <strong>and</strong> Tropical Medicine, New Orleans,LAKey words: biological contaminant, Gulf <strong>of</strong> Mexico, public health, biotoxin, red tide,din<strong>of</strong>lagellate, bloom, Gymnodinium breve, neurotoxin, neurotoxic shellfishpoisoning (NSP)Summary: A small portion <strong>of</strong> this unpublished draft report describes briefly <strong>and</strong> in generalterms the negative effects <strong>of</strong> Gymnodinium breve red tide on people who ingestcontaminated shellfish or inhale the toxin as an aerosol as well as on othermembers <strong>of</strong> the marine ecosystem. A list <strong>of</strong> red tide occurrences in Florida from1975-1991 follows. No cases <strong>of</strong> NSP have been reported from any state otherthan Florida, <strong>and</strong> while red tides have occurred in Florida <strong>and</strong> Texas, none havebeen reported for Alabama, Mississippi or Louisiana.Methods: N/A (Draft report)QA/QC: N/AContact: Dr. Fred KopflerSource Inst.: Not available; contact by phone at (601) 688-2712.Author: Aldrich, D. V.Date: 1962Title: Photoautotrophy in Gymnodinium breve DavisJournal: Science 137(3534):988-990Key words: Gymnodinium breve, photoautotrophy, culture, Florida, red tide, din<strong>of</strong>lagellate,light, carbon dioxide, growth, micronutrientsSummary: Since red tide outbreaks involving Gymnodinium breve on the Florida Gulf coastwere correlated with extended periods <strong>of</strong> heavy rainfall <strong>and</strong> subsequent organicinput to the sea from river discharge, the author tested a multitude <strong>of</strong> organicsubstances as potential direct energy sources for G. breve. Light <strong>and</strong> carbondioxide, however, were the principal growth requirements. No growth occurred inthe dark in spite <strong>of</strong> additives. It is therefore unlikely that heterotrophy plays alarge role in bloom formation. Vitamins, trace metals <strong>and</strong> chelators from river98


waters may facilitate photoautotrophy <strong>and</strong> thus bloom formation in G. breve <strong>and</strong>should be studied further.Methods: See text (no divisions).QA/QC: None per se; see text.Contact: David V. AldrichSource Inst.: Bureau <strong>of</strong> Commercial Fisheries, Biological Laboratory, Galveston, Texas, USAAuthor: Aldrich, D. V., <strong>and</strong> Wilson, W. B.Date: 1960Title: The effect <strong>of</strong> salinity on growth <strong>of</strong> Gymnodinium breve DavisJournal: Biol. Bull. 119:57-64Key words: salinity, growth, Gymnodinium breve, din<strong>of</strong>lagellate, bloom, red tide, cultureSummary: Bacteria-free cultures <strong>of</strong> Gymnodinium breve experienced media with salinitiesranging from 6.3 to 46.0 ppt, growing best between 27 <strong>and</strong> 37 ppt. Someorganisms survived salinities as low as 22.5 <strong>and</strong> as high as 46.0 ppt for ten weeks;there was no indication <strong>of</strong> reduced survival at the extremes <strong>of</strong> 24.8 <strong>and</strong> 46.0 ppt.The authors conclude that typical Gulf <strong>of</strong> Mexico salinities should not imposerestrictions on the growth <strong>of</strong> this din<strong>of</strong>lagellate, though water with salinities <strong>of</strong> 24ppt or less should.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: David V. AldrichSource Inst.: Biological Laboratory, U. S. Bureau <strong>of</strong> Commercial Fisheries, Galveston, Texas,USAAuthor: Aldrich, D. V., Ray, S. M., <strong>and</strong> Wilson, W. B.Date: 1967Title: Gonyaulax monilata: population growth <strong>and</strong> development <strong>of</strong> toxicity in culturesJournal: J. Protozool. 14(4):636-639Key words: Gonyaulax monilata, growth, toxicity, culture, din<strong>of</strong>lagellate, Lebistes reticulatus,autolysisSummary: The chain-forming din<strong>of</strong>lagellate Gonyaulax monilata [now known asAlex<strong>and</strong>rium monilata] was cultured in three 8-liter <strong>and</strong> four 12-liter containers,the population densities <strong>of</strong> which were estimated weekly for 17 weeks.Populations peaked at 3-5 weeks, declined from 6-10 weeks <strong>and</strong> usually stabilizedthereafter until the final week. Populations that showed more rapid increase hadthe greatest proportion <strong>of</strong> long chains, a possible growth index, <strong>and</strong> peak toxicity(gauged by the toxic effect on the guppy, Lebistes reticulatus) occurred only whencultures had been in decline for a month, indicating autolytic release <strong>of</strong> toxin, <strong>and</strong>had no relationship to estimated cell density.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: David V. AldrichSource Inst.: Marine Laboratory, Texas A & M University, Galveston, TX 77550 USAAuthor: Anderson, D. M.99


Date: 1994Title: <strong>Red</strong> tidesJournal: Scientific American, August 1994Key words: red tide, bloom, phytoplankton, din<strong>of</strong>lagellate, cyst, shellfish poisoning, biotoxin,Gymnodinium breve, pollution, ballast waterSummary: The author presents a summary <strong>of</strong> the deleterious effects <strong>of</strong> toxic red tide blooms.Topics addressed are biology, including cyst formation <strong>and</strong> germination; toxincharacteristics <strong>and</strong> poisoning; initiation, growth <strong>and</strong> hydrologic transport <strong>of</strong> bloomspecies; the correlation between blooms <strong>and</strong> increasing coastal pollution; <strong>and</strong> novelred tide infestations following transport in the ballast water <strong>of</strong> ships.Methods: None (summary)QA/QC: N/AContact: Donald M. AndersonSource Inst.: Biology Department, Woods Hole Oceanographic Institution, Woods Hole, MA02543 USAAuthor: Anderson, D. M., <strong>and</strong> Wall, D.Date: 1978Title: Potential importance <strong>of</strong> benthic cysts <strong>of</strong> Gonyaulax tamarensis <strong>and</strong> G. excavata ininitiating toxic din<strong>of</strong>lagellate bloomsJournal: Journal <strong>of</strong> Phycology 14(2):224-234Key words: cyst, Gonyaulax [later Protogonyaulax] tamarensis, Gonyaulax excavata,din<strong>of</strong>lagellate, bloom, hypnocysts, excystment, temperature, red tide, pellicle cystSummary: Sediments from salt ponds in Cape Cod, Massachusetts yielded hypnocysts <strong>of</strong> twotoxic din<strong>of</strong>lagellate species that were later successfully germinated by temperatureincrease alone (to 16 o C) after incubation for six months at 5 o C, unaffected bynutrient or chelator concentrations nor light regime. The authors conclude thathypnocysts do seed recurrent annual blooms. The hypnocysts are evidently sexualzygotes whereas a form <strong>of</strong> asexual cyst commonly formed in laboratory cultures,termed a "pellicle cyst," is less durable, cannot overwinter in nature <strong>and</strong> most likelydoes not produce toxic blooms. Excystments <strong>of</strong> hypnocysts with increasingtemperature has been demonstrated in several other din<strong>of</strong>lagellate species [<strong>and</strong>could possibly apply to Gymnodinium breve <strong>and</strong> Alex<strong>and</strong>rium monilata].Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Donald Mark AndersonSource Inst.: Biology Department, Woods Hole Oceanographic Institution, Woods Hole, MA02543 USAAuthor: Anderson, D. M., Kulis, D. M. <strong>and</strong> Cosper, E. M.Date: 1989Title/Chap.: Immun<strong>of</strong>luorescent detection <strong>of</strong> the brown tide organism, AureococcusanophagefferensBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35100


Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 213-228Publisher: Springer-VerlagKey words: immun<strong>of</strong>luorescence, brown tide, Aureococcus anophagefferens, bloom,chrysophyte, picoplankton, antibodies, antiserumSummary: The authors report the rapid, accurate detection <strong>of</strong> the brown tide organism,Aureococcus anophagefferens, with immun<strong>of</strong>luorescent detection methods, evenwhen the phytoplankton sample is mixed with A. anophagefferens in low relativeabundance. The antiserum at effective concentrations did not cross-react with any<strong>of</strong> 46 other phytoplankton cultures from 5 algal classes, including 20 species fromthe class Chrysophycophyta; at higher antiserum concentrations, cross-reactivityoccurred with Pelagococcus subviridis, which shares some ultrastructural <strong>and</strong>possibly phylogenetic similarities. The technique may also reliably serve indetecting the presence <strong>of</strong> A. anophagefferens in variously preserved samplesseveral years old.Methods: See “Methods”QA/QC: None per se; see “Results”Contact: Donald M. AndersonSource Inst.: Biology Department, Woods Hole Oceanographic InstitutionWoods Hole, MA 02543 USAAuthor: Anderson, D. M., Keafer, B. A., Kulis, D. M., Waters, R. M. <strong>and</strong> Nuzzi, R.Date: 1993Title: An immun<strong>of</strong>luorescent survey <strong>of</strong> the brown tide chrysophyte Aureococcusanophagefferens along the northeast coast <strong>of</strong> the United StatesJournal: Journal <strong>of</strong> Plankton Research 15(5):563-580Key words: immun<strong>of</strong>luorescence, brown tide, chrysophyte, Aureococcus anophagefferensSummary: The authors conducted surveys from Portsmouth, NH to Chesapeake Bay in 1988<strong>and</strong> 1990 to determine the population distribution <strong>of</strong> the brown tide chrysophyteAureococcus anophagefferens. A species-specific immun<strong>of</strong>luorescent techniquesensitive to as few as 10-20 cells/ml revealed that water samples from almost half<strong>of</strong> the stations to the north <strong>and</strong> south <strong>of</strong> the geographic center <strong>of</strong> the brown tideblooms <strong>of</strong> 1985 (the Long Isl<strong>and</strong> area <strong>and</strong> Barnegat Bay, NJ, where highconcentrations still existed), A. anophagefferens was detected at very low cellconcentrations within a PSU salinity range <strong>of</strong> 18-32. Many <strong>of</strong> these stations, bothopen coastal <strong>and</strong> estuarine, have no history <strong>of</strong> a brown tide bloom, yet apparentlyhave the potential.Methods: See “Method.”QA/QC: None per se; see “Method.”Contact: Donald M. AndersonSource Inst.: Woods Hole Oceanographic Institution, Biology Department, Woods Hole, MA02543 USAAuthor: Baden, D. G. <strong>and</strong> Tomas, C. R.Date: 1989Title/Chap.: Variations in major toxin composition for six clones <strong>of</strong> Ptychodiscus brevis101


Book: <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology. Proceedings <strong>of</strong> theFirst International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14, 1987, inTakamatsu, Kagawa Prefecture, JapanEditor: T. Okaichi, D. M. Anderson <strong>and</strong> T. NemotoPages: 415-418Publisher: Elsevier Science Publishing Company, Inc., New YorkKey words: toxin, clone, high pressure liquid chromatography, bloom, Texas, Florida, isolate,haploid, diploidSummary: Six Ptychodiscus brevis [now Gymnodinium breve] clones examined for toxincontent included one or more isolated from single cells collected during the 1986Texas bloom, with the remainder from Florida. The oldest clone (from Florida, ca.1953) was previously characterized as diploid, whereas the more recent isolateswere noted as haploid. In the three toxin fractions analyzed with HPLC, there waswide clonal variability; each <strong>of</strong> the recent clones was significantly different fromthe oldest. Obviously, the variability in toxin content between G. brevepopulations experiencing different environmental conditions deserves furtherattention.Methods: See "Experimental."QA/QC: None per se; see "Experimental."Contact: Daniel G. BadenSource Inst.: University <strong>of</strong> Miami Rosenstiel School <strong>of</strong> Marine <strong>and</strong> Atmospheric Science,Division <strong>of</strong> Biology <strong>and</strong> Living Resources, Miami, FL 33149 USAAuthor: Baldridge, H. D.Date: 1975Title/Chap.: Temperature patterns in the long-range prediction <strong>of</strong> red tide in Florida watersBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 69-79Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: surface temperature, red tide, Florida, bloom, Gymnodinium breve, din<strong>of</strong>lagellates,toxin, bloom predictionSummary: A technique for predicting red tide blooms in the local area <strong>of</strong> the Gulf near TampaBay, Florida that depends on the strong empirical relationship between surfacewater temperature patterns <strong>and</strong> major bloom initiation is presented as a means <strong>of</strong>forecasting the likelihood <strong>of</strong> outbreaks twelve months in advance. Predictions arebased on simple indicator patterns <strong>of</strong> water temperatures from mid-January toearly April, <strong>and</strong> the author claims that such patterns have already shown strongcorrelation with five major red tides between 1957 <strong>and</strong> 1974 at Egmont Key,Florida. No claim is made <strong>of</strong> cause-effect dependency nor does the author wish todiminish the importance <strong>of</strong> other environmental conditions favoring bloominitiation.Methods: See "Methods <strong>and</strong> Results."QA/QC: None per se; see "Methods <strong>and</strong> Results."Contact: H. David Baldridge102


Source Inst.: Mote Marine Laboratory, Sarasota, Florida, USAAuthor: Barron, C. N., Jr. <strong>and</strong> Vastano, A. C.Date: 1994Title: Satellite observations <strong>of</strong> surface circulation in the northwestern Gulf <strong>of</strong> Mexicoduring March <strong>and</strong> April 1989Journal: Continental Shelf Research 14(6):607-628Key words: satellite, surface circulation, Gulf <strong>of</strong> Mexico, drifter, drogue, Texas-LouisianaShelf, current, convergence, infrared imagery, sea surface temperatureSummary: Six drift buoys drogued to 2.7 m depth in the northwestern Gulf <strong>of</strong> Mexicoprovided eight tracks over the Texas-Louisiana Shelf in March <strong>and</strong> April <strong>of</strong> 1989.Tracks indicated cross-slope <strong>and</strong> cross-shelf water movement northward towardLouisiana from the central western Gulf <strong>and</strong> a westward coastal current from theMississippi delta region to Galveston, Texas <strong>and</strong> farther south. Low-energycurrent patterns occupied the middle <strong>of</strong> the northwestern portion <strong>of</strong> the continentalshelf, <strong>and</strong> a nearshore convergence occurred between the Matagorda Peninsula <strong>and</strong>southern Padre Isl<strong>and</strong>. The effects <strong>of</strong> wind-induced currents over the shelf couldbe seen over 7 o <strong>of</strong> longitude <strong>and</strong> 3 o <strong>of</strong> latitude. Infrared satellite imageryilluminated details <strong>of</strong> the spatial scale <strong>of</strong> Gulf circulation.Methods: See "Observations <strong>and</strong> Methods."QA/QC: N/A (Physical oceanography)Contact: Charlie N. Barron, Jr.Source Inst.: Satellite Ocean Analysis Research, Department <strong>of</strong> Oceanography, Texas A & MUniversity, College Station, TX 77843 USAAuthor: Beltrami, E. J.Date: 1989Title/Ch.: <strong>Brown</strong> tide dynamics as a catastrophe modelBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent<strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong> Other Unusual Blooms, Coastal <strong>and</strong> EstuarineStudies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 307-315Publisher: Springer-VerlagKey words: brown tide, Aureococcus anophagefferens, mathematical modeling,Long Isl<strong>and</strong>Summary: Some interesting results <strong>of</strong> this model include infrequent outbreaks<strong>of</strong> brown tide occurring at irregular intervals with varying severity<strong>and</strong> persisting from two to five years, while low endemic densities<strong>of</strong> the brown tide organism persist in the intervals. Within themodel, blooms are initiated only when high salinity accompaniesambient waters sufficiently rich in nutrients to sustain a bloompopulation. While the model itself is not predictive, it may explainthe basic dynamics <strong>of</strong> brown tide outbreaks.Methods: See “The Model”QA/QC: N/A103


Contact: Edward J. BeltramiSource Inst.: Department <strong>of</strong> Applied mathematics, State University <strong>of</strong> NewYork, Stony Brook, NY 11794-3600 USAAuthor: Beltrami, E. <strong>and</strong> Cosper, E.Date: 1993Title/Ch.: Modeling the temporal dynamics <strong>of</strong> unusual bloomsBook: Toxic Phytoplankton Blooms in the Sea, Proceedings <strong>of</strong> the Fifth InternationalConference on Toxic Marine PhytoplanktonEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 731-735Publisher: ElsevierKey words: modeling, bloom, brown tide, Aureococcus anophagefferens, picoplankton,grazing pressureSummary: Concentrating on the form <strong>of</strong> bloom behavior characterized by continuous butvariable growth in cell numbers, a need for the right suite <strong>of</strong> conditions to initiatethe bloom <strong>and</strong> persistence even if conditions diminish, the authors identify <strong>and</strong>incorporate two major sets <strong>of</strong> factors into their mathmatical model <strong>of</strong> bloomdynamics. Using field <strong>and</strong> laboratory data from the brown tide blooms in LongIsl<strong>and</strong> <strong>and</strong> Narragansett Bay to calibrate the model, the external, meterorologicalfactors combined with the internal factors <strong>of</strong> trophic system dynamics producedresults that implicate low grazing pressure as the principal selective advantage toan A. anophagefferens bloom <strong>and</strong> that mimic the fluctuating cell numbers as aproduct <strong>of</strong> chaotic population dynamics.Methods: See “Mathematical Model.”QA/QC: N/AContact: Edward BeltramiSource Inst.: State University <strong>of</strong> New York, Stony Brook, New York 11794 USAAuthor: Bidigare, R. R.Date: 1989Title/Ch.: Photosynthetic pigment composition <strong>of</strong> the brown tide alga: unique chlorophyll<strong>and</strong> carotenoid derivativesBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 57-75Publisher: Springer-VerlagKey words: Ultraplankton, pigments, brown tide, Aureococcus anophagefferens,Chrysophyceae, carotenoids, Pelagococcus subviridis, chlorophyll c 3 ,,fucoxanthin, xanthophyll.Summary: The author discusses the possible use <strong>of</strong> pigments unique to ultraplanktonicchrysophytes as a means <strong>of</strong> identification, particularly with regard to Aureococcusanophagefferens, a brown tide microalga. Such pigments include xanthophylls,fucoxanthins, carotenoids <strong>and</strong> chlorophyll c.Methods: See “Methods.”104


QA/QC: None per se; see “Methods.”Contact: Robert R. BidigareSource Inst.: Geochemical & Environmental Research Group, Department <strong>of</strong>Oceanography, Texas A&M University, College Station, TX77843 USAAuthor: Bricelj, V. M., Epp, J. <strong>and</strong> Malouf, R. E.Date: 1987Title: Intraspecific variation in reproductive <strong>and</strong> somatic growth cycles <strong>of</strong> bay scallopsArgopecten irradiansJournal: Marine Ecology Progress Series 36:123-137Key words: reproductive cycle, somatic growth cycle, bay scallop, Argopecten irradians,fecundity, adductor muscle, New York, picoplankton, bloomSummary: This study examined variability in fecundity, reproductive growth cycles <strong>and</strong>somatic growth cycles in four New York populations <strong>of</strong> the bay scallopArgopecten irradians. The brown tide bloom <strong>of</strong> 1985 caused starvation <strong>and</strong>reduced adult muscle weight by 76% relative to 1984. Once the bloom waned inlate summer, surviving scallops showed a 3-fold increase in mean muscle weight inSeptember.Methods: See "Methods."QA/QC: None per se; see "Methods."Contact: V. M. BriceljSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York at Stony Brook,New York 11794 USAAuthor: Bricelj, V. M. <strong>and</strong> Kuenstner, S. H.Date: 1989Title/Ch.: Effects <strong>of</strong> the “brown tide” on the feeding physiology <strong>and</strong> growth <strong>of</strong> bay scallops<strong>and</strong> musselsBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 491-509Publisher: Springer-VerlagKey words: brown tide, bay scallop, mussel, bloom, chrysophyte, Aureococcusanophagefferens, Argopecten irradians, Mytilus edulis, bivalves, Long Isl<strong>and</strong>Summary: This study tests the small size, high density <strong>and</strong> poor digestibility mechanisms inthe interaction between A. anophagefferens <strong>and</strong> juveniles <strong>of</strong> both mussels <strong>and</strong> bayscallops in feeding trials. Results indicated neither small size, indigestibility norpoor nutritional quality as reasons for harmful effects in bivalves, but rather thechronic toxicity <strong>of</strong> A. anophagefferens induced by direct contact <strong>of</strong> brown tidecells in high densities with bivalve tissue.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: V. Monica Bricelj105


Source Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794-5000 USAAuthor: Bricelj, V. M., Fisher, N. S., Guckert, J. B. <strong>and</strong> Fu-Lin, E. C.Date: 1989Title/Ch.: Lipid composition <strong>and</strong> nutritional value <strong>of</strong> the brown tide alga AureococcusanophagefferensBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 85-100Publisher: Springer-VerlagKey words:Summary:lipids, brown tide, Aureococcus anophagefferens, bloom, chrysophyte, fatty acidsOf the two hypotheses potentially explaining the brown tide’s negative impact onbivalves, chronic toxicity or nutritional inadequacy, this study concerns the latter,analyzing the variable lipid content, lipid fractionation <strong>and</strong> fatty acid composition<strong>of</strong> Aureococcus anophagefferens in culture during exponential growth <strong>and</strong>stationary phase with respect to possible deficiencies in polyunsaturated fatty acidsessential for bivalve growth. Results indicated no comparative deficiency betweenessential fatty acids in A. anophagefferens <strong>and</strong> other algae known to be nutritiousfor bivalves, which suggests that the chronic toxicity hypothesis is more likely.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: V. Monica BriceljSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794-5000 USAAuthor: Buskey, E. J.Date: 1994Title: Impact <strong>of</strong> a persistent “brown tide” algal bloom on the Laguna Madre <strong>of</strong> SouthTexas. Final report to the Texas Higher Education Coordinating Board, AdvancedTechnology Program.Pages: 5 pp.Key words: brown tide, Pelageophycae, Laguna Madre, chrysophyte, Halodule wrightii,Acartia tonsa, Streblospio benedicti, bay anchovy, black drum, spotted sea troutSummary: This report summarizes the adverse impact <strong>of</strong> the Texas brown tide’s almost fiveyearpersistence on the flora <strong>and</strong> fauna <strong>of</strong> the Laguna Madre, including shoalgrass,copepods, microzooplankton, polychaetes <strong>and</strong> fish larvae.Methods: N/AQA/QC: N/AContact: Edward J. BuskeySource Inst.: Marine Science Institute, The University <strong>of</strong> Texas at Austin, P. O. Box 1267,Port Aransas, TX 78373 USAAuthor: Buskey, E. J., <strong>and</strong> Stockwell, D. A..Date: 1993106


Title/Ch.: Effects <strong>of</strong> a persistent “brown tide” on zooplankton populations in the LagunaMadre <strong>of</strong> South TexasBook: Toxic Phytoplankton Blooms in the SeaEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 659-666Publisher: Elsevier Science Publishers B. VKey words: brown tide, zooplankton, Laguna Madre, Texas, nanoplankton, chrysophyte,mesozooplankton, microzooplankton, Baffin Bay, Acartia tonsaSummary: Relates effects on zooplankton <strong>of</strong> a brown tide due to an unidentified Type IIInanoplanktonic chrysophyte <strong>of</strong> 4-5 µm diameter that appeared in regions <strong>of</strong> theSouth Texas coast centering around the Laguna Madre including Baffin Bay from6/90-7/91 <strong>and</strong> continuing. Cell densities ranged from 0.5-6.0 x 10 6 cells/ml,dausing sharp decline in micro- <strong>and</strong> mesozooplankton populations.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Edward J. BuskeySource Inst.: Marine Science Institute, The University <strong>of</strong> Texas at Austin, Port Aransas, TX78373 USAAuthor: Caron, D. A., Lim, E. L., Kunze, H., Cosper, E. M. <strong>and</strong> Anderson, D. M.Date: 1989Title/Ch.: Trophic interactions between nano- <strong>and</strong> microzooplankton <strong>and</strong> the “brown tide”Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms , Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 265-294Publisher: Springer-VerlagKey words: nanoplankton, microzooplankton, brown tide, bloom, grazing, Aureococcusanophagefferens, chrysophyte, protozoa, Long Isl<strong>and</strong>Summary: Examines the ability <strong>of</strong> five protozoan species to consume A. anophagefferens inlaboratory culture, since the size <strong>of</strong> the brown tide organism is ideal for protozoangrazers. Paradoxically, brown tide outbreaks have suffered no apparent reductionby protozoan grazers that typically reproduce at greater rates than those reportedfor A. anophagefferens. Results revealed that two <strong>of</strong> the five protozoan grazerswere able to consume <strong>and</strong> outgrow the A. anophagefferens cultures. The authorsspeculate that brown tide organisms could achieve bloom proportions if lowdensities <strong>of</strong> protozoan consumers were present at the time <strong>of</strong> bloom initiation <strong>and</strong>that this scenario is more likely than chemical inhibition based on the results <strong>of</strong>these experiments.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: David A. CaronSource Inst.: Biology Department, Woods Hole Oceanographic Institution, Woods Hole, MA02543 USAAuthor: Castro, L. R. <strong>and</strong> Cowen, R. K.Date: 1989107


Title/Ch.: Growth rates <strong>of</strong> bay anchovy (Anchoa mitchilli) in Great South Bay underrecurrent brown tide conditions, summers 1987 <strong>and</strong> 1988Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms , Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 663-674Publisher: Springer-VerlagKey words: bay anchovy, Anchoa mitchilli, Great South Bay, brown tide, larvae,phytoplankton, bloom, growth rate, otolithSummary: Examining the possible effects <strong>of</strong> dense brown tide on larval growth rates <strong>of</strong> bayanchovy (< 13 mm SL) in Great South Bay, Long Isl<strong>and</strong>, the authors discoveredthat neither increased turbidity nor elevated average bay surface temperatureproduced a significant change in average larval growth rates as determined byotolith diameter, though the rates were not corrected for shrinkage <strong>and</strong> werepossibly higher than those reported.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Leonardo R. CastroSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York at Stony Brook,Stony Brook, NY 11794-5000 USAAuthor: Cochrane, J. D., <strong>and</strong> Kelly, F. J.Date: 1986Title: Low-frequency circulation on the Texas-Louisiana continental shelfJournal: Journal <strong>of</strong> Geophysical Research 91(C9):10,645-10,659Key words: low-frequency circulation, continental shelf, Texas, Louisiana, coastal current,wind stress, convergence, salinity, geopotential, cyclonic gyreSummary: The authors, using several data series on coastal winds, current measurements, <strong>and</strong>distributions <strong>of</strong> surface salinity <strong>and</strong> geopotential, <strong>of</strong>fer a sketch <strong>of</strong> the lowfrequencysurface circulation above the Texas-Louisiana Shelf. They infer theexistence <strong>of</strong> a cyclonic gyre composed <strong>of</strong> a nearshore southward coastal jet, an<strong>of</strong>fshore northerly to easterly flow over the shelf break <strong>and</strong> lateral componentscompleting the circuit. The gyre's southern end marks a convergence <strong>of</strong> coastalcurrents between Port Isabel <strong>and</strong> Port Aransas. The integrity <strong>of</strong> the gyre wasmaintained in all months but July.Methods: None (Data interpretation)QA/QC: N/AContact: J. D. CochraneSource Inst.: Department <strong>of</strong> Oceanography, Texas A & M University, College Station, TX,USAAuthor: Collier, A.Date: 1958Title: Some biochemical aspects <strong>of</strong> red tides <strong>and</strong> related oceanographic problemsJournal: Limnol. Oceanogr. 3:33-39108


Key words: red tide, bloom, plankton, Prorocentrum, Gymnodinium breve, Florida, Gulf <strong>of</strong>Mexico, copper, hydrogen sulfide, din<strong>of</strong>lagellateSummary: The author suggests that a complex <strong>of</strong> biological factors causes a red tide bloom,<strong>and</strong> physical factors cause its mechnical concentration. Biologically active organiccompounds are important, <strong>and</strong> the paper discusses their possible modes <strong>of</strong> action.The armored din<strong>of</strong>lagellate Prorocentrum served as a means <strong>of</strong> estimatingdin<strong>of</strong>lagellate potential in terms <strong>of</strong> producing free organic substances in the aquaticenvironment. Regarding G. breve, the interactions <strong>of</strong> a particular bacterial colonytype <strong>and</strong> its production <strong>of</strong> vitamin B12 with the red tide organism implies that thedin<strong>of</strong>lagellate may be producing an organic substrate that indirectly conditions thewater for its own growth. West Florida waters from different sources <strong>and</strong>collection times have varied effects on G. breve growth; this paper describessulfides as suitable substitutes for organic chelators (<strong>and</strong> abundantly supplied fromWest Florida estuaries) <strong>and</strong> copper as highly toxic to the din<strong>of</strong>lagellate (in widelyvarying concentrations <strong>of</strong>f West Florida).Methods: None per se; see text.QA/QC: None per se.Contact: Albert CollierSource Inst.: U. S. Fish <strong>and</strong> Wildlife Service, Galveston, TX, USAAuthor: Collier, A., Wilson, W. B. <strong>and</strong> Borkowski, M.Date: 1969Title: Responses <strong>of</strong> Gymnodinium breve Davis to natural waters <strong>of</strong> diverse originsJournal: J. Phycol. 5:168-172Key words: Gymnodinium breve, growth, enrichment, chelator, nutrient, Florida,din<strong>of</strong>lagellate, sulfide, plankton, bloomSummary: Using river water <strong>and</strong> seawater collected from different locations in differentseasons on <strong>and</strong> <strong>of</strong>f the coast <strong>of</strong> West Florida, the authors examined subsequentgrowth <strong>of</strong> the din<strong>of</strong>lagellate Gymnodinium breve when cultured in those waterswith <strong>and</strong> without enrichment with chelated metals (e. g., EDTA-Fe), sulfide,nitrogen, phosphorus <strong>and</strong> vitamins. All additions produced enhanced growthrelative to controls in river water. For both water types, a large growth responsewas noted with waters collected during the summer <strong>of</strong> 1966, perhaps linked toriverine contributions to the seawater samples. Of all additives, sulfides, naturalchelators, nitrogen <strong>and</strong> phosphorus may best contribute to blooms. Given theunpredictability <strong>of</strong> red tide blooms, the necessity <strong>of</strong> continual water sampling isimplied in order to characterize water quality during a bloom.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Albert CollierSource Inst.: Department <strong>of</strong> Biological Science, Florida State University, Tallahassee, FL32306 USAAuthor: Connell, C. H. <strong>and</strong> Cross, J. B.Date: 1950109


Title: Mass mortality <strong>of</strong> fish associated with the protozoan Gonyaulax in the Gulf <strong>of</strong>MexicoJournal: Science 112(2909):359-363Key words: fish, mortality, Gonyaulax sp., Gulf <strong>of</strong> Mexico, Galveston, Texas, GonyaulaxcatenellaSummary: This paper discusses the correlation <strong>of</strong> mass mortality <strong>of</strong> fish with historic reports<strong>and</strong> a 1949 episode <strong>of</strong> red tide or luminescent water in a saltwater lagoon nearGalveston, Texas. In terms <strong>of</strong> (1) insufficient cells to define the sutures <strong>of</strong> thecellulose plates, (2) longer chains <strong>of</strong> cells <strong>and</strong> (3) different vectors for the toxin,the 1949 red tide din<strong>of</strong>lagellate near Galveston was similar but not identical toGonyaulax catenella, the organism responsible for tainted shellfish <strong>and</strong> resultinghuman <strong>and</strong> mammal mortality (but not fish) on the Pacific coast <strong>of</strong> North America.Small quantities <strong>of</strong> sewage pollution, wide <strong>and</strong> variable readings <strong>of</strong> dissolvedoxygen content <strong>and</strong> unusually high values for biochemical oxygen dem<strong>and</strong> in theaffected lagoon were concurrent with the occurrence <strong>of</strong> this Gonyaulax sp.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Cecil H. ConnnellSource Inst.: Dept. <strong>of</strong> Preventive Medicine, Medical Branch, Univ. <strong>of</strong> Texas, GalvestonAuthor: Cosper, E. M.Date: 1987Title: Culturing the "brown tide" algaJournal: Applied Phycology ForumKey words: culture, brown tide, algae, bloom, Long Isl<strong>and</strong>, phytoplankton, Chrysophyceae,Aureococcus anophagefferens, growth rate, growth factorsSummary: Attempts to culture the chrysophyte responsible for the algal blooms thatdevastated eelgrass <strong>and</strong> bay scallop populations in Long Isl<strong>and</strong> bays in 1985 weresuccessful. Isolates exhibited fast growth rates (up to 3/day at 20 degrees C),though growth was poor in f/2-enriched st<strong>and</strong>ard media. Natural filtered seawater,however, when combined with f/2 media, produced maximal growth rates,suggesting certain growth factors in the bloom water. The organic additive,sodium glycerophosphate, enhanced growth significantly, whereas other glucose<strong>and</strong> succinate did not.Methods: None per se; see text (Research note)QA/QC: N/AContact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,New York 11794 USAAuthor: Cosper, E. M., Dennison, W. C., Carpenter, E. J., Bricelj, V. M., Mitchell, J. G.,Duenstner, S. H., Colflesh, D. <strong>and</strong> Dewey, M.Date: 1987Title: Recurrent <strong>and</strong> persistent brown tide blooms perturb coastal marine ecosystemJournal: Estuaries 10(4):284-290110


Key words: brown tide, bloom, chrysophyte, Long Isl<strong>and</strong>, eelgrass, bay scallop, Rhode Isl<strong>and</strong>,New jersey, microalgaeSummary: Culture experiments with the previously undescribed brown tide microalgaimplicated a role for stimulatory growth factors in bloom seawater, aided bymeterological factors that prompted blooms in Long Isl<strong>and</strong>, Rhode Isl<strong>and</strong> <strong>and</strong> NewJersey waters in the summers <strong>of</strong> 1985 <strong>and</strong> 1986. Bloom concentrations exceeded10 9 cells/liter in Long Isl<strong>and</strong> bays, harming eelgrass beds <strong>and</strong> bay scalloppopulations.Methods: See “Methods <strong>and</strong> Materials.”QA/QC: None per se; see “Methods <strong>and</strong> Materials.”Contact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, StonyBrook, New York 11794 USAAuthor: Cosper, E. M., Carpenter, E. J. <strong>and</strong> Cottrell, M.Date: 1989Title/Ch.: Primary productivity <strong>and</strong> growth dynamics <strong>of</strong> the “brown tide” in Long Isl<strong>and</strong>embaymentsBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 139-158Publisher: Springer-VerlagKey words: brown tide, Long Isl<strong>and</strong>, bloom, Aureococcus anophagefferens, Nannochloris sp.,picoplankton, Pelagococcus subviridisSummary: Throughout the brown tide blooms <strong>of</strong> 1985-88, phytoplankton biomass <strong>and</strong>productivity did not differ from pre-bloom years, but Aureococcusanophagefferens dominated the species composition, apparently controlled bymicr<strong>of</strong>lagellate grazers, yet still <strong>of</strong>ten as dense as 10 8 cells/l. No notable changes ininorganic nutrient levels could explain the blooms, though certain organic nutrientssuch as glycerophosphates <strong>and</strong> chelators were implicated. <strong>Historical</strong> blooms in thesummers <strong>of</strong> “small forms” <strong>and</strong> <strong>of</strong> usually diverse composition have not beenuncommon in Long Isl<strong>and</strong> waters. A. anophagefferens has been detected in nonbloomdensities in northeast coastal waters in general, implying that the LongIsl<strong>and</strong> blooms occurred due to unique <strong>and</strong> persistently favorable conditions for A.anophagefferens, conditions which may have stimulated excystation <strong>of</strong> a possibleresting stage.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794 USAAuthor: Cosper, E. M., Lee, C. <strong>and</strong> Carpenter, E. J.Date: 1990Title/Ch.: Novel “brown tide” blooms in Long Isl<strong>and</strong> embayments: a search for the causes111


Book: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 17-28Publisher: ElsevierKey words: brown tide, bloom, Long Isl<strong>and</strong>, chrysophyte, Aureococcus anophagefferens,Pelagococcus subviridis, trace elements, chelators, organic nutrientsSummary: Once a suite <strong>of</strong> favorable environmental factors initiates a bloom, Aureococcusanophagefferens appears to have a competitive advantage over other possiblycoincident phytoplankton species due to its heterotrophic <strong>and</strong> photoadaptivecapabilities. Factors contributing to blooms may include high salinities resultingfrom drought, rainfall-induced input <strong>of</strong> organic <strong>and</strong>/or micronutrients, reducedgrazing pressure <strong>and</strong> restricted flushing <strong>of</strong> bay waters. A. anophagefferens bearssome resemblance to the open ocean chrysophyte, Pelagococcus subviridis.Methods: N/A (review article)QA/QC: N/AContact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,New York 11794 USAAuthor: Cosper, E. M., Dennison, W., Milligan, A., Carpenter, E. J., Lee, C., Holzapfel, J.<strong>and</strong> Milanese, L.Date: 1989Title/Ch.: An examination <strong>of</strong> the environmental factors important to initiating <strong>and</strong> sustaining“brown tide” bloomsBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 317-340Publisher: Springer-VerlagKey words: brown tide, bloom, Long Isl<strong>and</strong>, flushing rate, Aureococcus anophagefferens,picoplanktonSummary: The authors hypothesize that the brown tide organism, Aureococcusanophagefferens, will likely bloom when long-term anthropogenic <strong>and</strong>/or othersources <strong>of</strong> eutrophication combine with atypical meteorological <strong>and</strong>hydrographical conditions (e.g., high salinity). The competitive superiority <strong>of</strong> A.anophagefferens relative to other phytoplankton does not appear to be the result<strong>of</strong> an allelopathic (toxic) exudate, but is more likely due to miconutrient needs,efficient use <strong>of</strong> low light levels, <strong>and</strong> heterotrophic capabilities, including an unusualability to take up glutamic acid <strong>and</strong> glucose rapidly as energy <strong>and</strong> carbon sources,with a possible reduction in grazing pressure.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794 USA112


Author: Cosper, E. M., Garry, R. T., Milligan, A. J. <strong>and</strong> Doall, M. H.Date: 1993Title/Ch.: Iron, selenium <strong>and</strong> citric acid are critical to the growth <strong>of</strong> the “brown tide”microalga, Aureococcus anophagefferensBook: Toxic Phytoplankton Blooms in the Sea, Proceedings <strong>of</strong> the Fifth InternationalConference on Toxic Marine PhytoplanktonEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 667-673Publisher: ElsevierKey words: iron, selenium, citric acid, brown tide, microalgae, Aureococcus anophagefferens,chelators, trace elements, chrysophyte, bloomSummary: This study evaluated the impact <strong>of</strong> chelators <strong>and</strong> essential trace elements on thegrowth <strong>of</strong> A. anophagefferens, <strong>of</strong> which both iron <strong>and</strong> selenium were found to becritical. Citric acid was the most effective chelator for growth enhancement, <strong>and</strong>,since it has replaced phosphates in some commercial detergents in New York, theauthors suggest that its role in eutrophication should be examined. Using bloom<strong>and</strong> non-bloom water samples filtered through 5 µm mesh <strong>and</strong> manipulated withregard to light <strong>and</strong> Se <strong>and</strong>/or Fe additions, results indicated that Fe alone increasedby several times the growth rates <strong>of</strong> A. anophagefferens in bloom water under highlight; only Fe with Se increased A. anophagefferens growth rates in non-bloomwater, but only at low light levels.Methods: See “Methods.”QA/QC: None per se; see “Methods” <strong>and</strong> “Results.”Contact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794 USAAuthor: Davis, C. C.Date: 1948Title: Gymnodinium brevis sp. nov., a cause <strong>of</strong> discolored water <strong>and</strong> animal mortality inthe Gulf <strong>of</strong> MexicoJournal: Botanical Gazette 109(3):358-360Key words: Gymnodinium brevis, discolored water, fish mortality, Gulf <strong>of</strong> Mexico, Florida,plankton, cell, chromatophoresSummary: The description <strong>and</strong> classification <strong>of</strong> the then-novel red tide din<strong>of</strong>lagellate nowknown as Gymnodinium breve occupies the majority <strong>of</strong> text but is preceded bysome examples <strong>of</strong> mass mortality caused by this organism <strong>of</strong>f the west coast <strong>of</strong>Florida in 1947.Methods: None (taxonomic description/classification)QA/QC: N/AContact: Charles C. DavisSource Inst.: University <strong>of</strong> Miami Marine Laboratory, Coral Gables, Florida, USAAuthor: Dennison, W. C., Marshall, G. J. <strong>and</strong> Wig<strong>and</strong>, C.Date: 1989Title/Ch.: Effect <strong>of</strong> “brown tide” shading on eelgrass (Zostera marina L.) distributions113


Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 675-692Publisher: Springer-VerlagKey words: brown tide, eelgrass, Zostera marina, Aureococcus anophagefferens, Long Isl<strong>and</strong>,distribution, die-backSummary: A lack <strong>of</strong> historical data on eelgrass distributions in Long Isl<strong>and</strong> bays prevented theauthors from making definitive statements about any deleterious effects by browntide on eelgrass populations since the summer blooms <strong>of</strong> 1985-88, yet the browntide did reduce light available to the eelgrass beds in bloom areas. Most bays hadless eelgrass during the 1988 survey than indicated in historic, pre-bloom aerialphotographs.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: William C. DennisonSource Inst.: Horn Point Environmental Laboratories, Center for Environmental <strong>and</strong> EstuarineStudies, University <strong>of</strong> Maryl<strong>and</strong>, P. O. Box 775, Cambridge, MD 21613 USAAuthor: Deyoe, H. R. <strong>and</strong> Suttle, C. A.Date: 1994Title: The inability <strong>of</strong> the Texas “brown tide” alga to use nitrate <strong>and</strong> the role <strong>of</strong> nitrogenin the initiation <strong>of</strong> a persistent bloom <strong>of</strong> this organism.Journal: J. Phycol. 30:800-806Key words: Texas, brown tide, bloom, nitrate, nitrate reductase, nitrogen, plankton,Aureococcus anophagefferens, Laguna Madre, nitrite, ammoniumSummary: The Texas brown tide organism differs slightly but significantly from the NewEngl<strong>and</strong> brown tide alga, Aureococcus anophagefferens, in that the Texas browntide alga apparently lacks the enzyme nitrate reductase, making it unable to useNO - 3 for growth. It can, however, use nitrite <strong>and</strong> ammonium, the latter <strong>of</strong> whichwas in unusually high concentrations after two severe freezes in December 1989<strong>and</strong> January 1990 caused declines in invertebrate biomass <strong>and</strong> mortality in fishpopulations <strong>and</strong> apparently stimulated the Texas brown tide bloom.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Hudson R. DeyoeSource Inst.: CCBNEP, Texas A & M University-Corpus Christi, 6300 Ocean Drive, CorpusChristi, TX 78412 USAAuthor: Draper, C., Gainey, L., Shumway, S. <strong>and</strong> Shapiro, L.Date: 1990Title/Ch.: Effects <strong>of</strong> Aureococcus anophagefferens (“brown tide”) on the lateral cilia <strong>of</strong> 5species <strong>of</strong> bivalve molluscsBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. Anderson114


Pages: 128-131Publisher: ElsevierKey words: Aureococcus anophagefferens, brown tide, cilia, bivalve, mollusc, Mya arenaria,Geukensia demissa, Argopecten irradians, Mercenaria mercenaria, Mytilusedulis, dopamineSummary: Aureococcus anophagefferens caused significant decreases in lateral ciliary activityin isolated gills <strong>of</strong> Mercenaria mercenaria <strong>and</strong> Mytilus edulis as did theneurotransmitter dopamine. Neither A. anophagefferens nor dopamine, however,had any effect on lateral ciliary activity <strong>of</strong> Mya arenaria, Geukensia demissa orArgopecten irradians.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Christy DraperSource Inst.: Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Southern Maine, Portl<strong>and</strong>, ME04103 USAAuthor: Duguay, L. E., Monteleone, D. M. <strong>and</strong> Quaglietta, C.-E.Date: 1989Title/Ch.: Abundance <strong>and</strong> distribution <strong>of</strong> zooplankton <strong>and</strong> ichthyoplankton in Great SouthBay, New York during the brown tide outbreaks <strong>of</strong> 1985 <strong>and</strong> 1986Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 599-623Publisher: Springer-VerlagKey words: zooplankton, ichthyoplankton, Great South Bay, New York, brown tide,chrysophyte, Aureococcus anophagefferens, bivalve larvae, trophodynamics,Acartia sp.Summary: Investigations <strong>of</strong> plankton trophodynamics in Great South Bay, NY wereunderway when brown tide blooms <strong>of</strong> Aureococcus anophagefferens occurred inthe summers <strong>of</strong> 1985 <strong>and</strong> 1986, the former apparently more severe. Summerbivalve larvae were significantly higher in both peak <strong>and</strong> average density in 1986 asopposed to 1985, confirming previous reports <strong>of</strong> mortality <strong>and</strong>/or reduction intissue dry weights in bivalves exposed to brown tide blooms. The authorsspeculate that excess rainfall in May <strong>and</strong> June <strong>of</strong> 1985 initiated the brown tide,with the bloom occurring at a time when phytoplankton grazers were notabundant.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Linda E. DuguaySource Inst.: University <strong>of</strong> Maryl<strong>and</strong>, Center for Environmental <strong>and</strong> Estuarine Studies,Chesapeake Biological Laboratory, P. O. Box 38, Solomons, MD 20688-0038USAAuthor: Durbin, A. G. <strong>and</strong> Durbin, E. G.Date: 1989115


Title/Ch.: Effect <strong>of</strong> the “brown tide” on feeding, size <strong>and</strong> egg laying rate <strong>of</strong> adult femaleAcartica tonsaBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 625-646Publisher: Springer-VerlagKey words: brown tide, Acartia tonsa, picoplankton, Narragansett Bay, chrysophyte,Aureococcus anophagefferens, bloom, copepods, Greenwich Bay,protozooplanktonSummary: Relates the impact <strong>of</strong> a picoalgal bloom in Narragansett Bay beginning in the earlysummer <strong>of</strong> 1985. The then undescribed 2 µm chrysophyte, now known asAureococcus anophagefferens, negatively affected egg laying rate, gut pigments,body size <strong>and</strong> condition factor <strong>of</strong> adult female Acartia tonsa copepods whencomparisons were made between field observations <strong>and</strong> laboratory manipulations.The authors suggest that (1) a regional climatological or hydrographic changeaffected or worked in concert with temperature, light <strong>and</strong> nutrients to favor thebloom, <strong>and</strong>/or (2) predation pressure affected the size composition <strong>of</strong> thephytoplankton community, favoring the picoplankton that are thought to becontrolled by protozooplankton predators.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Ann G. DurbinSource Inst.: University <strong>of</strong> Rhode Isl<strong>and</strong>, Graduate School <strong>of</strong> Oceanography, Narragansett,Rhode Isl<strong>and</strong> 02882 USAAuthor: Dunton, K. H.Date: 1994Title: Seasonal growth <strong>and</strong> biomass <strong>of</strong> the subtropical seagrass Halodule wrightii inrelation to continuous measurements <strong>of</strong> underwater irradianceJournal: Marine BiologyKey words: growth, biomass, seagrass, Halodule wrightii, underwater irradiance, PAR, Texas,chrysophyte, brown tide, bloomSummary: Three different Texas estuaries, including brown tide-stricken Laguna Madre, weresites for studies <strong>of</strong> continuous underwater irradiance measurements <strong>and</strong> concurrentleaf elongation <strong>and</strong> plant biomass for the seagrass Halodule wrightii over one t<strong>of</strong>our years. The brown tide bloom significantly decreased spring leaf elongationrates <strong>and</strong> reduced below-ground biomass by almost 50%.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Ken H. DuntonSource Inst.: The University <strong>of</strong> Texas at Austin, Marine Science Institute, P. O. Box 1267, PortAransas, TX 78373 USAAuthor: Dzurica, S.Date: 1988116


Thesis: Role <strong>of</strong> environmental variables, specifically organic compounds <strong>and</strong>micronutrients, in the growth <strong>of</strong> Aureococcus anophagefferensPages: 122 pp.Source: Marine Environmental Sciences Program, State University <strong>of</strong> New York at StonyBrookKey words: micronutrient, Aureococcus anophagefferens, brown tide, microalga, Long Isl<strong>and</strong>,chrysophyte, organic phosphorus, chelator, trace metal, uptake rateSummary: Noting that st<strong>and</strong>ard culture media (f/2 Enriched Instant Ocean) supported slowgrowth rates in Long Isl<strong>and</strong> brown tide cultures, the author conducted a series <strong>of</strong>experiments in an attempt to determine factors contributing to bloom conditions.Organic rather than inorganic phosphorus compounds greatly enhanced growth asdid adding the chelators nitrilotriacetic acid <strong>and</strong> citric acid. Comparisons <strong>of</strong> uptakerates for C-14 labeled organic compounds between Aureococcus anophagefferens<strong>and</strong> five other species <strong>of</strong> microalgae which co-occur with the brown tide. A.anophagefferens exhibited a faster uptake rate for glutamic acid whenaccompanied by inorganic nitrogen, but not when glutamic acid was the sole Nsource. A. anophagefferens showed the highest uptake rate <strong>of</strong> glucose per unit cellvolume but not with the highest uptake rate constant per cell. For everycompound tested, however, A. anophagefferens demonstrated an advantage inuptake rate per unit cell volume.Methods: See "Methods <strong>and</strong> Materials," pp. 6-16.QA/QC: See "Methods <strong>and</strong> Materials."Contact: Cindy LeeSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794 USAAuthor: Dzurica, S., Lee, C. <strong>and</strong> Cosper E. M.Date: 1989Title/Ch.: Role <strong>of</strong> environmental variables, specifically organic compounds <strong>and</strong>micronutrients, in the growth <strong>of</strong> the chrysophyte Aureococcus anophagefferensBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 229-252Publisher: Springer-VerlagKey words: micronutrients, chrysophyte, Aureococcus anophagefferens, phytoplankton,bloom, brown tide, organic phosphorus, trace metals, chelatorsSummary: With a two-fold purpose, this study sought to tailor culture media to supportoptimal growth <strong>of</strong> Aureococcus anophagefferens in order to identify possiblecommon environmental factors between culture media <strong>and</strong> bloom water <strong>and</strong> toexperimentally evaluate growth <strong>of</strong> A. anophagefferens relative to otherphytoplankton in an attempt to explain long-term dominance. Replacing inorganicphosphate in f/2 EIO medium with the organic phosphorus compoundsglycerophosphate <strong>and</strong> fructose-1,6-diphosphate stimulated growth in A.anophagefferens, as did the chelators citric acid <strong>and</strong> nitrilotriacetic acid. Traceelements accompanied by the chelator EDTA also enhanced growth. In situ117


chelation, therefore, may be an important growth factor. A. anophagefferens has amore rapid uptake rate for glutamic acid <strong>and</strong> glucose than three co-occurringphytoplankton species (diatoms <strong>and</strong> a green alga), perhaps conferring acompetitive advantage.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Susan DzuricaSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, StonyBrook, NY 11794 USAAuthor: Elliott, B. A.Date: 1982Title: Anticyclonic rings in the Gulf <strong>of</strong> MexicoJournal: Journal <strong>of</strong> Physical Oceanography 12:1292-1309Key words: anticyclonic ring, Gulf <strong>of</strong> Mexico, Loop <strong>Current</strong>, heat budget, salt budget,Caribbean subtropical underwater, western anticyclonic cell, circulationSummary: <strong>Historical</strong> data sets provided evidence that three anticyclonic rings separated fromthe Loop <strong>Current</strong> in the eastern Gulf <strong>of</strong> Mexico during the twelve monthsfollowing October <strong>of</strong> 1966. Translating westward at a mean speed <strong>of</strong> 2.1 km/day,the rings' mean radius was 183 km, <strong>and</strong> their estimated life span was one year. Theheat <strong>and</strong> salt provided by these rings to the western Gulf must play important rolesin the heat <strong>and</strong> salt budgets <strong>of</strong> the Gulf. Convective mixing in winter graduallyreduced the higher ring salinity to values typical <strong>of</strong> Gulf <strong>of</strong> Mexico common water.Methods: None (Data interpretation)QA/QC: N/AContact: W. D. Nowlin, Jr.Source Inst.: Department <strong>of</strong> Oceanography, Texas A & M University, College Station, TX77843 USAAuthor: Eng-Wilmot, D. L., McCoy, L. F., Jr. <strong>and</strong> Martin, D. F.Date: 1979Title/Ch.: Isolation <strong>and</strong> synergism <strong>of</strong> a red tide (Gymnodinium breve) cytolytic factor(s) fromcultures <strong>of</strong> Gomphosphaeria aponinaBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 355-360Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: red tide, Gymnodinium breve, cytolytic factor, Gomphosphaeria aponina,din<strong>of</strong>lagellate, blue-green alga, aponinSummary: The blue-green alga Gomphosphaeria aponina yields a cytolytic factor (“aponin”)active towards the unarmored red tide din<strong>of</strong>lagellate Gymnodinium breve.Chlor<strong>of</strong>orm extractions at neutral pH produced maximum yields <strong>of</strong> aponin, purifiedby any one <strong>of</strong> three chromatographic techniques. Separate fractions showedobvious losses <strong>of</strong> total <strong>and</strong> specific activity, <strong>and</strong> a synergistic relationship amongthe separate components is supported by evidence.118


Methods: See “Materials <strong>and</strong> Methods.”QA/QC: See “Materials <strong>and</strong> Methods.”Contact: D. F. MartinSource Inst.: Department <strong>of</strong> Chemistry, University <strong>of</strong> South Florida, Tampa, FL 33620 USAAuthor: Eng-Wilmot, D. L., Henningsen, B. F., Martin, D. F. <strong>and</strong> Moon, R. E.Date: 1979Title/Ch.: Model solvent systems for delivery <strong>of</strong> compounds cytolytic towards GymnodiniumbreveBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 361-366Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: solvent, cytolytic compounds, Gymnodinium breve, aponin, din<strong>of</strong>lagellate, sterol,Gomphosphaeria aponina, liposomes, micelles, red tideSummary: A sterol (C 29 H 49 OR) isolated <strong>and</strong> purified from cultures <strong>of</strong> the cyanobacteriumGomphosphaera aponina is cytolytic towards the toxigenic red tide din<strong>of</strong>lagellateGymnodinium breve. Because the sterol is hydrophobic, it must be delivered toseawater in a non-toxic solvent system. Two are presented as promising fortreatment <strong>of</strong> localized red tide blooms, liposomes <strong>and</strong> micelles, the latter beingsuperior.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: See “Materials <strong>and</strong> Methods.”Contact: D. L. Eng-WilmotSource Inst.: Department <strong>of</strong> Chemistry, University <strong>of</strong> Oklahoma, Norman, Oklahoma 73019USAAuthor: Freeberg, L. R., Marshall, A. <strong>and</strong> Heyl, M.Date: 1979Title/Ch.: Interrelationships <strong>of</strong> Gymnodinium breve (Florida red tide) within thephytoplankton communityBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 139-144Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: Gymnodinium breve, red tide, Florida, phytoplankton, diatom, flagellate,din<strong>of</strong>lagellate, lysis, growth, inhibitionSummary: Medium in which G. breve had been grown was subsequently used as a mediumfor each <strong>of</strong> 28 phytoplankton species in axenic cultures. The medium significantlyinhibited growth in 18 species, though its effect, while species-specific, variedwithin species. The population levels <strong>of</strong> several diatoms <strong>and</strong> din<strong>of</strong>lagellates <strong>and</strong>one flagellate barely increased above inoculum concentrations; two din<strong>of</strong>lagellateinocula suffered lysis. Toxin extracts totally arrested growth in eight <strong>of</strong> twelve119


species (4 diatoms <strong>and</strong> 4 din<strong>of</strong>lagellates). Column chromotography could notseparate the algal inhibition component from the ichthyotoxin.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Larry R. FreebergSource Inst.: Mote Marine Laboratory, 1600 City Isl<strong>and</strong> Park, Sarasota, FloridaAuthor: Gaffney, R. J.Date: 1992Title: <strong>Brown</strong> tide comprehensive assessment <strong>and</strong> management program summaryPages: 40 pp.Publisher: Suffolk County Department <strong>of</strong> Health Services, NYKey words: brown tide, algal bloom, Aureococcus anophagefferens, Peconic Bay, Fl<strong>and</strong>ersBay, Shinnecock Bay, Moriches Bay, Great South BaySummary: This report concludes that the Aureococcus anophagefferens brown tideapparently was not triggered by the conventional macronutrients nitrogen <strong>and</strong>phosphorus, but may have been caused by other factors such as meteorologicalpatterns <strong>and</strong> specific chemicals (chelators, organic nutrients, metals). The reportrecommends, among other things, further research on viral control <strong>of</strong> the browntide <strong>and</strong> the relationship between dimethyl sulfide <strong>and</strong> zooplankton grazers.Methods: N/A (Summary report)QA/QC: N/AContact: Robert J. Gaffney, Suffolk County ExecutiveSource Inst.: Suffolk County Department <strong>of</strong> Health Services, NYAuthor: Gainey, L. F., Jr. <strong>and</strong> Shumway, S. E.Date: 1991Title: The physiological effect <strong>of</strong> Aureococcus anophagefferens (“brown tide”) on thelateral cilia <strong>of</strong> bivalve mollusksJournal: Biol. Bull. 181:298-306Key words: Aureococcus anophagefferens, brown tide, gill cilia, bivalve, mollusk, dopamine,ergometrineSummary: The isolated gills <strong>of</strong> eight bivalve mollusks were exposed to Aureococcusanophagefferens, the brown tide alga, resulting in a significant decrease in theactivity <strong>of</strong> the lateral cilia <strong>of</strong> the gills <strong>of</strong> five <strong>of</strong> the bivalves, whereas those <strong>of</strong> threeother species were unaffected. Only the presence <strong>of</strong> the brown tide cells producedsuch results, <strong>and</strong> the inhibition was similar to that induced by dopamine.Treatment with the dopamine antagonist ergometrine prevented loss <strong>of</strong> ciliaryactivity due to both dopamine <strong>and</strong> a water-soluble compound derived from browntide cells exposed to amylase. This dopamine-like, brown tide inhibitorycompound was likely released upon partial digestion by the isolated gills.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Louis F. GaineySource Inst.: Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Southern Maine, Portl<strong>and</strong>, Maine04103 USA120


Author: Gallager, S. M., Stoecker, D. K. <strong>and</strong> Bricelj, V. M.Date: 1989Title/Ch.: Effects <strong>of</strong> the brown tide alga on growth, feeding physiology <strong>and</strong> locomotorybehavior <strong>of</strong> scallop larvae (Argopecten irradians)Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 511-541Publisher: Springer-VerlagKey words: brown tide, scallop larvae, Argopecten irradians, Aureococcus anophagefferens,growth, mortalitySummary: Using scallop larvae from a local population (Woods Hole), the authorsdemonstrated in laboratory experiments that near-bloom concentrations <strong>of</strong>Aureococcus anophagefferens reduced larval growth <strong>and</strong> survival due toinefficient capture <strong>and</strong> reduced ingestion rates, the former attributed to cell surfacecharacteristics. The presence <strong>of</strong> A. anophagefferens also hindered ingestion, butnot capture, <strong>of</strong> other, nutritious algal species. Ingestion <strong>of</strong> A. anophagefferensresulted in poor nutrition or toxicity in larval scallops, though assimilationefficiency was good <strong>and</strong> equal to that for Isochrysis sp. Scallop veligers exhibitedno swimming behavior indicating recognition <strong>of</strong> the presence <strong>of</strong> A.anophagefferens as they do with common algal prey such as Isochrysis sp.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Scott M. GallagerSource Inst.: Woods Hole Oceanographic Institution, Woods Hole, MA 02543 USAAuthor: Gallegos, S.Date: 1990Dissertation: Evaluation <strong>of</strong> the potential <strong>of</strong> the NOAA-n AVHRR reflective data inoceanographyPages: 189 pp.Source: Department <strong>of</strong> Oceanography, Texas A & M University, College Station, TX,USAKey words: reflective data, advanced very high resolution radiometry, phytoplankton, satelliteobservation, atmospheric effect, albedo, pixel, Coastal Zone Color Scanner,pigment, red tideSummary: The organic <strong>and</strong> physical features <strong>of</strong> the ocean surface as seen synoptically canprovide clues to the interactions <strong>of</strong> those processes governing the ocean. Twosatellite data bases (the NOAA-n Advanced Very High Resolution Radiometer <strong>and</strong>the Coastal Zone Color Scanner data) provide the means to compare informationon sea surface phytoplankton concentrations <strong>and</strong> circulation patterns over a broadarea. In situ data taken during a Florida red tide in 1983 <strong>and</strong> a Texas red tide in1986 serve to evaluate the relative capabilities <strong>of</strong> the two satellite data sources.Methods: See "Materials <strong>and</strong> Methods."QA/QC: N/A (Physical oceanography)121


Contact: Sonia GallegosSource Inst.: Center for Space Research, The University <strong>of</strong> Texas at Austin, Austin, TX,USAAuthor: Gates, J. A. <strong>and</strong> Wilson, W. B.Date: 1960Title: The toxicity <strong>of</strong> Gonyaulax monilata Howell to Mugil cephalusJournal: Limnol. Oceanogr. 5(2):171-174Key words: Gonyaulax monilata, Mugil cephalus, din<strong>of</strong>lagellate, toxin, Galveston, Texas, redtideSummary: “Young” mullet (Mugil cephalus) subjected to in vitro cultures <strong>of</strong> the marine redtide din<strong>of</strong>lagellate Gonyaulax monilata suffered mortality within 4.5 hours in allaliquots except the uninoculated control medium. The authors conclude that atoxic substance produced by Gonyaulax monilata was the causative agent.Methods: See “Methods <strong>and</strong> Materials.”QA/QC: None per se; see “Methods <strong>and</strong> Materials.”Contact: William B. WilsonSource Inst.: U. S. Fish <strong>and</strong> Wildlife Service, Galveston, Texas, USAAuthor: Geesey, M. <strong>and</strong> Tester, P. A.Date: 1993Title/Ch.: Gymnodinium breve: ubiquitous in Gulf <strong>of</strong> Mexico waters?Book: Toxic Phytoplankton Blooms in the Sea, Proceedings <strong>of</strong> the Fifth InternationalConference on Toxic Marine PhytoplanktonEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 251-255Publisher: ElsevierKey words: Gymnodinium breve, Gulf <strong>of</strong> Mexico, Ptychodiscus brevis, red tide, bloomSummary: A common source <strong>of</strong> red tide <strong>of</strong>f the west coast <strong>of</strong> Florida, Gymnodinium breve isfound throughout the Gulf <strong>of</strong> Mexico, but, until this study, its cell density innearshore <strong>and</strong> <strong>of</strong>fshore waters <strong>of</strong> various depths was largely unknown. Overtwelve months (3/90-3/91), NOAA vessels sampled 61 stations throughout theGulf <strong>of</strong> Mexico from depths <strong>of</strong> 0 to >150 m. Density in shallow, well-mixedwaters was constant, but in deeper waters with a thermocline, G. breve was moreabundant near the surface. Concentrations in central Gulf waters remained at


Journal: Science 113:250-251Key words: fish mortality, din<strong>of</strong>lagellate, bloom, Gulf <strong>of</strong> Mexico, Florida, Texas, Gonyaulax,Offats BayouSummary: In this lengthy letter to the editors <strong>of</strong> Science, the author <strong>of</strong>fers some criticism foran article by Connell <strong>and</strong> Cross in a previous volume in which the two stated that arecent case <strong>of</strong> fish mortality in Offats Bayou, a branch <strong>of</strong> Galveston Bay, was likelydue to a red tide <strong>of</strong> Gonyaulax sp. In addition, Gunter added some historical <strong>and</strong>scientific data from what little was known <strong>of</strong> red tides at that time.Methods: None (Letter to editor)QA/QC: N/AContact: N/ASource Inst.: Marine Science Institute, The University <strong>of</strong> Texas, P. O. Box 1267, Port Aransas,TX 78373-1267 USAAuthor: Gunter, G., Smith, F. G. W., <strong>and</strong> Williams, R. H.Date: 1947Title: Mass mortality <strong>of</strong> marine animals on the lower west coast <strong>of</strong> Florida, November1946-January 1947Journal: Science 105:256-257Key words: fish mortality, Florida, Gulf <strong>of</strong> Mexico, discolored water, plankton,Summary: From reports <strong>of</strong> an "odorless but acrid gas" that produced respiratory irritation inan area <strong>of</strong> heavy surf, the discolored water <strong>and</strong> corresponding fish kills thatoccurred <strong>of</strong>f the Gulf Coast <strong>of</strong> South Florida from November <strong>of</strong> 1946 to January<strong>of</strong> 1947 was due to Gymnodinium breve, though this article did not identify theorganism to species.Methods: None (report)QA/QC: N/AContact: University <strong>of</strong> Miami, Coral Gables, Florida, USASource Inst.: SameAuthor: Haddad, K. D. <strong>and</strong> Carder K. L.Date: 1979Title/Ch.: Oceanic intrusion: one possible initiation mechanism <strong>of</strong> red tide blooms on thewest coast <strong>of</strong> FloridaBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 269-274Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: red tide, bloom, Florida, upwelling, Loop <strong>Current</strong>, cyst, excystment, growthSummary: The authors suggest that the Loop <strong>Current</strong>’s intrusion onto the west Florida shelfmay resuspend the resting cysts <strong>of</strong> Gymnodinium breve <strong>and</strong> create conditionsconducive to excystment <strong>and</strong> growth. Such conditions include decreasedtemperature, increased light availability <strong>and</strong> increased nutrients.Methods: N/A (Published hypothesis)QA/QC: N/A123


Contact: Kenneth D. HaddadSource Inst.: Department <strong>of</strong> Marine Science, University <strong>of</strong> South Florida, 830 First StreetSouth, St. Petersburg, FL 33701 USAAuthor: Hallegraeff, G. M.Date: 1993Title: A review <strong>of</strong> harmful algal blooms <strong>and</strong> their apparent global increaseJournal: Phycologia 32(2):79-99Key words: bloom, red tide, aquaculture, eutrophication, climate, transport, din<strong>of</strong>lagellate,cyst, ballast water, shellfishSummary: This review presents evidence for an apparent increase in problem algal blooms inrecent years, though it does not claim to ascribe the trend to an actual increase orto steadily improving means <strong>and</strong> efforts to detect blooms. Increased recreational,commercial <strong>and</strong> aquacultural use <strong>of</strong> coastal waters, however, have yielded moreproblems with negative public health <strong>and</strong> economic impacts from harmful algalblooms. As a result, the author urges the scientific community to respond to theproblem by making efforts to prevent the spread <strong>of</strong> harmful algae from the areaswhere they are indigenous to other novel <strong>and</strong> vulnerable areas throughout theworld <strong>and</strong> to make public <strong>of</strong>ficials aware <strong>of</strong> policies to prevent or ameliorate thenegative impacts <strong>of</strong> blooms. The possible impacts <strong>of</strong> global warming, ozonedepletion, El Nino events, eutrophication <strong>and</strong> ballast water transport must bestudied with respect to their possible stimulation <strong>of</strong> blooms.Methods: None (review)QA/QC: N/AContact: G. M. HallegraeffSource Inst.: Department <strong>of</strong> Plant Science, University <strong>of</strong> Tasmania, GPO Box 252C, Hobart,Tasmania 7001, AustraliaAuthor: Hallegraeff, G. M. <strong>and</strong> Bolch, C. J.Date: 1992Title: Transport <strong>of</strong> diatom <strong>and</strong> din<strong>of</strong>lagellate resting spores in ships' ballast water:implications for plankton biogeography <strong>and</strong> aquacultureJournal: Journal <strong>of</strong> Plankton Research 14(8):1067-1084Key words: diatom, din<strong>of</strong>lagellate, cyst, ballast water, aquaculture, sediment, Gymnodinium,Alex<strong>and</strong>rium, bloom, quarantineSummary: Non-endemic diatoms <strong>and</strong> din<strong>of</strong>lagellates may be introduced to novel regions whencargo ships with cyst-contaminated ballast water <strong>and</strong> associated sedimentsdischarge their ballast water in port. Fifty percent <strong>of</strong> sediment samples taken from343 cargo vessels in 18 Australian ports revealed din<strong>of</strong>lagellate resting spores(cysts); these spores represented at least 53 species, 20 <strong>of</strong> which were successfullygerminated in the laboratory. Cysts <strong>of</strong> toxic din<strong>of</strong>lagellates <strong>of</strong> the generaAlex<strong>and</strong>rium <strong>and</strong> Gymnodinium were found in 16 ships, one <strong>of</strong> which contained anestimated >300 million viable A. tamarense cysts. To counter the threat fromtoxic species, the authors suggest several ways to lessen or prevent the spread <strong>of</strong>cysts from ballast water discharge.Methods: See "Method."124


QA/QC: None per se; see "Method."Contact: G. M. HallegraeffSource Inst.: Department <strong>of</strong> Plant Science, University <strong>of</strong> Tasmania, GPO Box 252C, Hobart,Tasmania 7001, AustraliaAuthor: Harper, D. E., Jr. <strong>and</strong> Guillen, G.Date: 1989Title: Occurrence <strong>of</strong> a din<strong>of</strong>lagellate bloom associated with an influx <strong>of</strong> low salinitywater at Galveston, Texas, <strong>and</strong> coincident mortalities <strong>of</strong> demersal fish <strong>and</strong> benthicinvertebratesJournal: Contributions in Marine Science 31:147-161Key words: din<strong>of</strong>lagellate, bloom, Galveston, Texas, Atlantic threadfin, Polydactylusoctonemus, hypoxia, hydrogen sulfideSummary: Attempting to correlate hydrographic data with din<strong>of</strong>lagellate bloom occurrence tounderst<strong>and</strong> the sequence <strong>of</strong> events leading to the bloom, the authors noted thecorrespondence <strong>of</strong> low salinity water with bloom appearance <strong>of</strong>f Galveston, Texasin early June 1984. The low salinity was likely due to high discharge from theMississippi-Atchafalaya Rivers in preceding months <strong>and</strong> a wind-driven currentdown the Texas coast. Within a week, a die-<strong>of</strong>f <strong>of</strong> the demersal Atlantic threadfin(Polydactylus octonemus) occurred, probably due to hypoxia <strong>and</strong>/or hydrogensulfide production as a result <strong>of</strong> nocturnal din<strong>of</strong>lagellate metabolism <strong>and</strong> anaerobicdecay <strong>of</strong> dead din<strong>of</strong>lagellate cells.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Donald E. Harper, Jr.Source Inst.: Texas A & M Marine Laboratory, 5007 Avenue U, Galveston, TX 77551USAAuthor: Hemmert, W. H.Date: 1975Title/Chap.: The public health implications <strong>of</strong> Gymnodinium breve red tides, a review <strong>of</strong> theliterature <strong>and</strong> recent eventsBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 489-497Publisher:Key words:Summary:The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MApublic health, Gymnodinium breve, red tide, din<strong>of</strong>lagellate, Florida, neurotoxicshellfish poisoning (NSP), respiratory irritation, contact irritation, hemtologicpathology, toxinThe author found the scientific literature depauperate on the public healthproblems caused by G. breve <strong>and</strong> similar organisms. Southwest Florida's major G.breve red tide event <strong>of</strong> 1973-74 revealed three major public health concerns <strong>and</strong> apossible fourth: neurotoxic shellfish poisoning (NSP), respiratory irritation, contactirritation <strong>and</strong> hematologic pathology. Eleven people were apparently stricken withNSP, though none fatally. Clinical reactions to airborne toxins constituted a125


espiratory irritation that disappeared upon departure from the affected area. Alsotemporary was the contact irritation (mild dermatitis, if any, <strong>and</strong>/or mild to severeconjunctivitis) experienced by people exposed to the toxin while in the water;severity <strong>of</strong> conjunctivitis correlates with intensity <strong>of</strong> exposure. Only the potentialfor decreased blood coagulation exists for humans.Methods: None (report)QA/QC: N/AContact: Wynn H. HemmertSource Inst.: Bureau <strong>of</strong> Preventable Diseases, Florida State Division <strong>of</strong> Health, Jacksonville,Florida, USAAuthor: H<strong>of</strong>mann, E. E., <strong>and</strong> Worley, S. J.Date: 1986Title: An investigation <strong>of</strong> the circulation <strong>of</strong> the Gulf <strong>of</strong> MexicoJournal: Journal <strong>of</strong> Geophysical Research 91(C12):14,221-14.236Key words: circulation, Gulf <strong>of</strong> Mexico, hydrography, density, Loop <strong>Current</strong>, mesoscale eddy,Antarctic Intermediate Water, velocity pr<strong>of</strong>ile, gyre, transportSummary: Reanalysis <strong>of</strong> historic hydrographic data on winter circulation in the Gulf <strong>of</strong>Mexico revealed that the large-scale circulation is heavily influenced by ananticyclonic gyre in the upper 500 m. Additional influences come from the Loop<strong>Current</strong> in the eastern gulf <strong>and</strong> a cyclonic eddy in the northwestern gulf, <strong>and</strong> thetransports estimated for these circulation features concur with estimates from otherhydrographic studies in the gulf.Methods: See "Methods."QA/QC: N/A (Physical oceanography)Contact: Eileen F. H<strong>of</strong>mannSource Inst.: Department <strong>of</strong> Oceanography, Texas A & M University, College Station, TX,USAAuthor: Howell, J. F.Date: 1953Title: Gonyaulax monilata, sp. nov., the causative din<strong>of</strong>lagellate <strong>of</strong> a red tide on the eastcoast <strong>of</strong> Florida in August-September, 1951Journal: Transactions <strong>of</strong> the American Microscopical Society 72:153-156Key words: Gonyaulax monilata, din<strong>of</strong>lagellate, red tide, Florida, plankton, Offatts Bayou,Galveston, TexasSummary: Using samples mainly from the Indian <strong>and</strong> Banana Rivers on Florida's east coast,the author described a small din<strong>of</strong>lagellate species known to form chains <strong>of</strong> up to40 cells in length. Some doubt was expressed as to whether it was appropriate toassign the species to the Genus Gonyaulax because the prevalent genericdescription allowed "no variation in the number <strong>of</strong> precingular plates" <strong>and</strong> theorganism exhibited radically different thecal suture patterns when compared to G.catenella. In early September <strong>of</strong> 1952, the author found identical chain-formingdin<strong>of</strong>lagellates in a sample taken from Offatts Bayou, near Galveston, Texas.Methods: None (taxonomic description/classification)QA/QC: N/A126


Contact: John F. HowellSource Inst.: U. S. Fish <strong>and</strong> Wildlife Service, Fort Crockett, Galveston, Texas, USAAuthor: Huntley, M., Sykes, P., Rohan, S. <strong>and</strong> Marin, V.Date: 1986Title: Chemically-mediated rejection <strong>of</strong> din<strong>of</strong>lagellate prey by the copepods Calanuspacificus <strong>and</strong> Paracalanus parvus: mechanism, occurrence <strong>and</strong> significanceJournal: Marine Ecology Progress Series 28:105-120Key words: din<strong>of</strong>lagellate, copepods, Ptychodiscus brevis, Calanus pacificus, Paracalanusparvus, particle rejection, bloom, chemical defense, growth rateSummary: Thirteen species <strong>of</strong> din<strong>of</strong>lagellates were used as possible prey for two species <strong>of</strong>copepods in a large suite <strong>of</strong> experiments. Results <strong>of</strong> one experiment includedPtychodiscus brevis (now called Gymnodinium breve, a source <strong>of</strong> neurotoxicshellfish poisoning) as one <strong>of</strong> several din<strong>of</strong>lagellates consistently rejected by thecopepod Calanus pacificus. Ingestion <strong>of</strong> P. brevis cells produced elevated heartrate <strong>and</strong> loss <strong>of</strong> motor control in the copepod. Not only do noxious din<strong>of</strong>lagellatespecies such as P. brevis gain an interspecific competitive advantage in survivalover edible species, they also seem far more likely to be able to maintain bloomproportions in spite <strong>of</strong> zooplankton grazers when other factors are favorable.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: M. HuntleySource Inst.: Scripps Institution <strong>of</strong> Oceanography, A-002, La Jolla, California 92093 USAAuthor: Jensen, A. C.Date: 1975Title/Ch.: The economic halo <strong>of</strong> a red tideBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 507-516Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: red tide, economics, New Engl<strong>and</strong>, Gonyaulax tamarensis, fishing industry, publichealth, shellfish, halo effect, seafood, seafood industrySummary: An economic halo effect from the 1972 New Engl<strong>and</strong> red tide <strong>of</strong> Gonyaulaxtamarensis (source <strong>of</strong> potentially lethal paralytic shellfish poisoning or PSP)appeared in the form <strong>of</strong> consumer resistance to canned shellfish <strong>and</strong> finfishproducts. Shortly after shellfish that had been harvested in New Engl<strong>and</strong> waterswere removed from the market as a safety measure, consumers in states outside <strong>of</strong>New Engl<strong>and</strong> refused to buy canned seafood regardless <strong>of</strong> where it had beenharvested, hurting the seafood industries <strong>of</strong> states untouched by the red tide. Theauthor concludes with a discussion <strong>of</strong> the necessary informational system neededto counter the rumors <strong>and</strong> misinformation that can cause economic havoc forindustries both suffering from <strong>and</strong> falsely associated with toxic red tide blooms.Methods: None (report)QA/QC: N/A127


Contact: Albert C. JensenSource Inst.: Division <strong>of</strong> Marine <strong>and</strong> Coastal Resources, New York State Department <strong>of</strong>Environmental Conservation, State University <strong>of</strong> New York at Stony Brook,Stony Brook, New York 11794 USAAuthor: Jensen, D. A. <strong>and</strong> Bowman, J.Date: 1975Title: On the occurrence <strong>of</strong> the "red tide" Gonyaulax monilata in the Corpus ChristiInner Harbor (July-September 1975)Pages: 20 pp.Source: Texas Water Quality Board, District 12Key words: red tide, Gonyaulax monilata, Corpus Christi, Texas, bloom, din<strong>of</strong>lagellate, cyst,fish kill, iron, Eutreptia c. f. lanowiiSummary: The maximum recorded density <strong>of</strong> a bloom <strong>of</strong> Gonyaulax (now Alex<strong>and</strong>rium)monilata in the Corpus Christi Inner Harbor turning basin was noted when it wasfirst detected by routine monitoring on 22 July 1975. Thereafter, cellconcentrations <strong>and</strong> the number/length <strong>of</strong> chains decreased until no red tide wasdetected on 17 September. Despite relatively low cell concentrations, a fish killoccurred in the turning basin on 22 August. The most prominent correlationbetween nutrient <strong>and</strong> cell concentration concerned iron, which peaked <strong>and</strong> declinedwith the bloom. A "green tide" occurred immediately after the red tidedisappeared, identified as the euglenoid Eutreptia c. f. lanowii. Benthic cysts aresuggested as the means by which the bloom appeared deep within the inner harbor,though the document <strong>of</strong>fers no evidence <strong>of</strong> cyst presence.Methods: See "Methods <strong>and</strong> Materials."QA/QC: None per se but for modified Winkler dissolved oxygen analysis (USEPA, 1974)Contact: James Bowman, Jr.Source Inst.: Texas Natural Resource Conservation Commission, Field Operations Division,Environmental Assessment Program, 4410 Dillon Lane, Suite 47, Corpus Christi,TX 78415-5326 USAAuthor: Joyce, E. A., Jr., <strong>and</strong> Roberts, B. S.Date: 1975Title/Chap.: Florida Department <strong>of</strong> Natural Resources <strong>Red</strong> <strong>Tide</strong> Research ProgramBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 95-103Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: red tide, Florida, initiation, nutrition, Gymnodinium breve, maintenance, transport,toxin, life cycle, seed bedSummary: About 30 red tides had been reported in Florida waters since 1844, but research oncauses began only in the early 1950's. That research, sponsored by the FloridaDepartment <strong>of</strong> Natural Resources (FDNR) centered on Gymnodinium breve <strong>and</strong>dealt with location <strong>of</strong> bloom initiation, nutrition, hydrologic <strong>and</strong> meteorologicinfluences on bloom maintenance <strong>and</strong> transport, effects on <strong>of</strong>fshore patch reef128


iota <strong>and</strong> inshore shellfish beds, taxonomy <strong>and</strong> ecology <strong>of</strong> associatedphytoplankton <strong>and</strong> bloom prediction. FDNR research at time <strong>of</strong> publicationemphasized the study <strong>of</strong> G. breve life cycles, the search for <strong>of</strong>fshore seed beds,toxin longevity <strong>and</strong> effects on marine life, fisheries repopulation <strong>and</strong> carcasssalvage, <strong>and</strong> evaluation <strong>of</strong> l<strong>and</strong> discharges on bloom growth. The research designemphasized prediction <strong>and</strong> public education <strong>and</strong> reduction <strong>of</strong> negative economicimpact.Methods: None (report)QA/QC: N/AContact: Edwin A. Joyce, Jr.Source Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Science <strong>and</strong>Technology, Tallahassee, Florida, USAAuthor: Keller, A. A. <strong>and</strong> Rice, R. L.Date: 1989Title: Effects <strong>of</strong> nutrient enrichment on natural populations <strong>of</strong> the brown tidephytoplankton Aureococcus anophagefferens (Chrysophyceae)Journal: J. Phycol. 25:636-646Key words: brown tide, phytoplankton, Aureococcus anophagefferens, Chrysophyceae,diatoms, mesocosm, Narragansett Bay, Rhode Isl<strong>and</strong>, picoalgae, dissolvedinorganic nitrogenSummary: Approximately equal populations <strong>of</strong> the brown tide picoalga Aureococcusanophagefferens in twelve 13,000 L mesocosms were exposed to varying levels <strong>of</strong>nutrients Densities in untreated systems were similar to the increased abundances<strong>of</strong> A. anophagefferens in Narragansett Bay (2.6 x 10 9 cells/L max.), the seawatersource for the mesocosms. Nutrient addition tanks also had blooms, but they werebrief in duration <strong>and</strong> did not long remain at densities above the usual level foreukaryotic algae in the bay. Diatom abundance increased in all nutrient-treatedtanks; that plus a significant inverse correlation between dissolved inorganicnitrogen (DIN) levels <strong>and</strong> mean picoalgae abundance revealed that the brown tidecan grow at DIN levels known to limit diatom growth.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Aimee A. KellerSource Inst.: Graduate School <strong>of</strong> Oceanography, University <strong>of</strong> Rhode Isl<strong>and</strong> Bay Campus,South Ferry Road, Narragansett, Rhode Isl<strong>and</strong> 92882-1197 USAAuthor: Keller, M. D., Bellows, W. K. <strong>and</strong> Guillard, R. R. L.Date: 1989Title/Ch.:Book:Editor:Dimethylsulfide production <strong>and</strong> marine phytoplankton: an additional impact <strong>of</strong>unusual bloomsNovel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. Carpenter129


Pages: 101-115Publisher: Springer-VerlagKey words: dimethylsulfide, DMS, phytoplankton, bloom, dimethylsulfoniopropionate, DMSP,acrylic acidSummary: Though Aureococcus anophagefferens is only mentioned twice among thenumerous other phytoplankton species responsible for greater or lesser degrees <strong>of</strong>DMS release, this paper explains the basic mechanism for that release, its generalenvironmental effect, <strong>and</strong> the relative contributions to release by specificphytoplankton <strong>of</strong> twelve classes. A. anophagefferens, as an alga possessingchlorophyll a <strong>and</strong> c, is grouped with other chrysophytes <strong>and</strong> diatoms as DMSproducers <strong>of</strong> species-specific variability, producing less than din<strong>of</strong>lagellates <strong>and</strong>prymnesiophytes, but significantly more than chlorophytes, cryptomonads <strong>and</strong>cyanobacteria.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Maureen D. KellerSource Inst.: Bigelow Laboratory for Ocean Sciences, McKown Point, West BoothbayHarbor, Maine 04575 USAAuthor: Kuenstner, S. H.Date: 1988Thesis: The effects <strong>of</strong> the "brown tide" alga on the feeding physiology <strong>of</strong> Argopectenirradians <strong>and</strong> Mytilus edulisPages: 84 pp.Source: Marine Environmental Sciences Program, SUNY at Stony BrookKey words: brown tide, Argopecten irradians, Mytilus edulis, Aureococcus anophagefferens,bloom, chrysophyte, retention efficiency, clearance rate, absorption efficiency,starvationSummary: This study was prompted by brown tide-induced weight loss in adult bay scallops<strong>and</strong> recruitment failure during the bloom <strong>of</strong> 1985. The author studied the feedingmechanism <strong>of</strong> bay scallops (Argopecten irradians) <strong>and</strong> mussels (Mytilus edulis) todetermine the adverse effects <strong>of</strong> the brown tide alga Aureococcus anophagefferenson both bivalves. Scallops had a lower retention efficiency (36%) than mussels(59%) when both grazed on A. anophagefferens. Juvenile scallops <strong>and</strong> musselshad higher clearance rates for the alga Thalassiosira weissflogii than A.anophagefferens, but both bivalves can digest both algae with 90% efficiencywhen the algae are at low concentrations. Using physiological data to construct anenergy budget for scallops exposed to the brown tide, the results over the shorttermcould not explain the bivalve starvation observed in the field. Chronicexposure, therefore, may have been the key to negative growth during the LongIsl<strong>and</strong> brown tide bloom <strong>of</strong> 1985.Methods: See "Methods <strong>and</strong> Materials," pp. 16-31.QA/QC: See "Methods <strong>and</strong> Materials."Contact: V. Monica BriceljSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, Stony Brook,NY 11794-5000 USA130


Author: Lowe, J. A., Farrow, D. R. G., Pait, A. S., Arenstam, S. J., <strong>and</strong> Lavan, E. F.Date: 1991Title: Fish kills in coastal watersPages: pp. 60-61Publisher: National Oceanic <strong>and</strong> Atmospheric AdministrationKey words: fish, fish-kill, coast, Gulf <strong>of</strong> Mexico, Texas, red tideSummary: Only one table on p. 61 <strong>of</strong> this report categorizes fish kills due to red tide in Texasfrom 1980-89, indicating that less than 100 fish died as a result <strong>of</strong> eight reportedred tide events. [This report is in direct conflict with documented observationalevidence <strong>of</strong> far greater fish mortality ranging from Galveston Isl<strong>and</strong> to south <strong>of</strong> thePadre Isl<strong>and</strong> National Seashore during the period from late August to late October1986. See Trebatoski (1988), among others, listed below.]Methods: None (government report).QA/QC: N/AContact: Strategic Environmental Assessments Division, Office <strong>of</strong> Ocean ResourcesConservation <strong>and</strong> Assessment, National Ocean ServiceSource Inst.: National Oceanic <strong>and</strong> Atmospheric AdministrationAuthor: Lund, E. J.Date: 1936Title/Ch.: Some facts relating to the occurrences <strong>of</strong> dead <strong>and</strong> dying fish on the Texas coastduring June, July, <strong>and</strong> August, 1935Source: Annual Report <strong>of</strong> the Texas Game, Fish <strong>and</strong> Oyster Commission, 1934-35Pages: 47-50Publisher: Texas Game, Fish <strong>and</strong> Oyster CommissionKey words: fish [mortality], Texas, [aerosol irritant], menhaden, mullet, [red tide],[Gymnodinium breve]Summary: This report contains information on a massive fish kill in the summer <strong>of</strong> 1935,ranging from near Port Aransas (St. Joseph Isl<strong>and</strong>) to points from 150 to 200 milessouth. Menhaden <strong>and</strong> mullet made up the largest percentage <strong>of</strong> the estimated 2million pounds <strong>of</strong> dead fish that washed ashore in July <strong>and</strong> August, accompaniedby at least a dozen other fish species. On several instances, observers reportedexperiencing an "irritating 'gas'," a description which is similar to the effects <strong>of</strong>aerosol irritants from Gymnodinium breve toxin. During May <strong>and</strong> June, prior tothe first appearance <strong>of</strong> fish carcasses, rainfall <strong>and</strong> river flood levels were unusuallyhigh, <strong>and</strong> on 28 August, the surface water temperature was recorded at 31 o C. Theauthor raises those physical factors as potentially important influences on the fishkill, but aside from riverine outflow possibly carrying "heavy inshore planktongrowth," he makes no connection between fish mortality <strong>and</strong> phytoplankton.Methods: None (Report)QA/QC: N/AContact: Larry McEachron [For access to this literature]Source Inst.: Texas Parks &Wildlife Department, Coastal Fisheries Division, Rockport, TX78382 USA131


Author: Marvin, K. T.Date: 1965Title/Ch.: Operation <strong>and</strong> maintenance <strong>of</strong> salt-water laboratoriesSource: Fishery Research: Biological Laboratory, Galveston, Fiscal Year 1964, Circular230Pages: 84-86Publisher: United States Department <strong>of</strong> the Interior, Fish <strong>and</strong> Wildlife Service, Bureau <strong>of</strong>Commercial FisheriesKey words: East Lagoon (Galveston Bay), plankton, bloom, chlorophyll, Gonyaulax monilata,din<strong>of</strong>lagellate,Summary: The USFWS experimental facility in Galveston Bay's East Lagoon had sufferedcomplete mortality <strong>of</strong> all oyster stock in previous years due to "poisonous planktonblooms" entering the system via the water intake. This led to installation <strong>of</strong> arecirculating system to temporarily free the system from total dependence on anexterior water source <strong>and</strong> twice-weekly water analyses for chlorophyll <strong>and</strong> otherconstituents. An August 1963 chlorophyll peak was linked to a bloom <strong>of</strong> the redtide din<strong>of</strong>lagellate Gonyaulax (now Alex<strong>and</strong>rium) monilata. The author indicatesthat such blooms seem to be seasonal occurrences in East Lagoon.Methods: None (report)QA/QC: N/AContact: Kenneth T. Marvin, Supervisory ChemistSource Inst.: Bureau <strong>of</strong> Commercial Fisheries Biological Laboratory, Galveston, Texas, USAAuthor: Matsuoka, K., Fukuyo, Y. <strong>and</strong> Anderson, D. M.Date: 1989Title/Ch.: Methods for modern din<strong>of</strong>lagellate cyst studiesBook: <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology. Proceedings <strong>of</strong> theFirst International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14, 1987, inTakamatsu, Kagawa Prefecture, Japan.Editor: T. Okaichi, D. M. Anderson <strong>and</strong> T. NemotoPages: 461-479Publisher: Elsevier Science Publishing Company, Inc., New YorkKey words: din<strong>of</strong>lagellate, cyst, sampling, fixation, preservation, isolation, culture,identification, Gymnodinium breve, Alex<strong>and</strong>rium monilataSummary: This manual includes a key for identification <strong>of</strong> cyst-forming din<strong>of</strong>lagellates basedon cyst shape <strong>and</strong> was the backbone for a din<strong>of</strong>lagellate cyst workshop held at thesymposium. Gymnodinium breve <strong>and</strong> Alex<strong>and</strong>rium monilata are included in thelist <strong>of</strong> all extant cyst-forming din<strong>of</strong>lagellates.Methods: Primarily a methods paper.QA/QC: None per se.Contact: D. M. AndersonSource Inst.: Biology Department, Woods Hole Oceanographic Inst., Woods Hole, MA 02543USAAuthor: Milligan, A. J.132


Date: 1992Thesis: An investigation <strong>of</strong> factors contributing to blooms <strong>of</strong> the “brown tide”Aureococcus anophagefferens (Chrysophyceae) under nutrient saturated (lightlimited) conditionsPages: 84 pp.Source: M. S. Thesis, Department <strong>of</strong> Marine Environmental Science, State University <strong>of</strong>New York at Stony BrookKey words: bloom, brown tide, Aureococcus anophagefferens, Chrysophyceae, nutrientsaturation, light limitation, Long Isl<strong>and</strong>, iron, selenium, nanoplanktonSummary: Following the six-year summer bloom <strong>of</strong> the brown tide chrysophyte Aureococcusanophagefferens in the bays <strong>of</strong> Long Isl<strong>and</strong>, New York, the author tested thehypothesis that A. anophagefferens had a greater ability to photoacclimate <strong>and</strong>maintain high growth rates relative to other phytoplankton. In this attempt, lightutilization efficiency, photoacclimation, carbon metabolism <strong>and</strong> iron/seleniumeffects were determined for A. anophagefferens. The brown tide alga shows goodphotoacclimation, greater carbon fixation (per unit chlorophyll) <strong>and</strong> photosyntheticefficiency in fluctuating rather than constant light <strong>and</strong> enhanced growth withadditions <strong>of</strong> iron but not selenium.Methods: See “Methods” on pp. 12-20.QA/QC: None per se; see “Methods.”Contact: Elizabeth M. CosperSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, StonyBrook, NY 11794 USAAuthor: Moestrup, O. <strong>and</strong> Larsen, J.Date: 1990Title/Ch.: Some comments on the use <strong>of</strong> the generic names Ptychodiscus <strong>and</strong> Alex<strong>and</strong>riumBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 78-81Publisher: Elsevier Science Publishing Co., Inc.Key words: Ptychodiscus brevis, Gymnodinium, Alex<strong>and</strong>rium, genus, toxic algae, plankton,taxonomy, paralytic shellfish poisoning (PSP), Gonyaulax, flagellatesSummary: The authors argue that the red tide din<strong>of</strong>lagellate known as Ptychodiscus brevisshould not be separated from the genus Gymnodinium <strong>and</strong> that the generic nameAlex<strong>and</strong>rium is better suited for PSP-producing species formerly placed within thegenus Gonyaulax. Upon these recommendations, the proper nomenclature for one<strong>of</strong> the two red tide algae that have constituted problem blooms in the “CoastalBend” <strong>of</strong> the Texas Gulf coast supports the name Gymnodinium breve.Methods: N/A (Comments on taxonomy)QA/QC: N/AContact: O. MoestrupSource Inst.: Institut for Sporeplanter, Oster Farimagsgade 2D, DK-1353 Kobenhavn K,Denmark133


Author: Montagna, P. A., Stockwell, D. A. <strong>and</strong> Kalke, R. D.Date: 1993Title: Dwarf surfclam Mulinia lateralis (Say, 1822) populations <strong>and</strong> feeding during theTexas brown tide eventJournal: Journal <strong>of</strong> Shellfish Research 12(2):433-442Key words: dwarf surfclam, Mulinia lateralis, Texas, brown tide, chrysophyte, algae, BaffinBay, Laguna Madre, shellfishSummary: Dwarf surfclam populations perished at the same time that a brown tide bloomoccurred in Baffin Bay <strong>and</strong> Laguna Madre on the Texas Gulf Coast in 1990. Infeeding experiments to determine if the shellfish loss was due to negative feedinginteractions between the clam <strong>and</strong> the brown tide, radioactive tracers <strong>and</strong> anadditional three phytoplankton were used. Grazing rates increased with algalconcentrations


Author: Nixon, S. W., Granger, S. L., Taylor, D. I., Johnson, P. W. <strong>and</strong> Buckley, B. A.Date: 1994Title: Subtidal volume fluxes, nutrient inputs <strong>and</strong> the brown tide--an alternate hypothesisJournal: Estuarine, Coastal <strong>and</strong> Shelf Science 39:303-312Key words: nutrients, nutrient flux, brown tide, Aureococcus anophagefferens, Great SouthBay, mesocosmsSummary: This paper agrees that unusual meteorological <strong>and</strong> hydrological conditionscontributed the the 1985 outbreak <strong>of</strong> brown tide in Long Isl<strong>and</strong> bays, but proposesthat, with a many-fold greater contribution <strong>of</strong> nutrients from the coastal oceanrelative to l<strong>and</strong> drainage, A. anophagefferens achieved bloom proportions becausenutrient input was reduced, not increased. Evidence from field surveys <strong>and</strong>mesocosms suggest that oligotrophic waters favor the growth <strong>of</strong> A.anophagefferens.Methods: N/A (Published hypothesis)QA/QC: N/AContact: S. W. NixonSource Inst.: Graduate School <strong>of</strong> Oceanography, Univeristy <strong>of</strong> Rhode Isl<strong>and</strong>, Narragansett,Rhode Isl<strong>and</strong> 02882-1197 USAAuthor: Nuzzi, R. <strong>and</strong> Waters R. M.Date: 1989Title/Ch.: The spatial <strong>and</strong> temporal distribution <strong>of</strong> “brown tide” in eastern Long Isl<strong>and</strong>Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 117-137Publisher: Springer-VerlagKey words: distribution, brown tide, Long Isl<strong>and</strong>, bloom, phytoplankton, AureococcusanophagefferensSummary: The Suffolk County Department <strong>of</strong> Health Service’s “Bureau <strong>of</strong> MarineResources” reports on its investigations into the distribution <strong>of</strong> the brown tidefrom 1986-1988. Areas affected by brown tide, the Peconic Bay estuary <strong>and</strong> theSouth Shore bays as seen during aerial surveys, remained relatively constant, <strong>and</strong>both unusually large <strong>and</strong> persistent over the three years covered by this study.<strong>Brown</strong> tide abundance seemed to be unlinked to macronutrient concentrations inthe affected areas, with regional meterorological conditions a more likely catalyst,particularly rainfall fluctuations. Micronutrient concentrations in bloom watersmay also be a principal agent <strong>of</strong> bloom formation.Methods: See “Methods.”QA/QC: None per se; see “Methods” <strong>and</strong> “Results.”Contact: Robert NuzziSource Inst.: Suffolk County Department <strong>of</strong> Health Services, Evans K. Griffing County Center,Riverhead, New York 11901 USAAuthor: Olsen, P. S.Date: 1989135


Title/Ch.: Development <strong>and</strong> distribution <strong>of</strong> a brown-water algal bloom in Barnegat Bay, NewJersey with perspective on resources <strong>and</strong> other red tides in the regionBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 189-212Publisher: Springer-VerlagKey words: brown-water, bloom, Barnegat Bay, new Jersey, red tide, phytoplankton,Aureococcus anophagefferens, picoplanktonSummary: The authors conducted a 1987 survey <strong>of</strong> a persistent yellow-brown algal bloom inBarnegat Bay, New Jersey, that had returned in high density every summer since1985, seeking to accurately characterize the responsible picoplankton <strong>and</strong> revealpotential similarities between it <strong>and</strong> the simultaneous brown tide in Long Isl<strong>and</strong><strong>and</strong> the New York Bight region. The 2-3 µm coccoid alga at densities as high asseveral times 10 6 cells/ml could not be identified as a common bloom-producingchlorophyte nor the chrysophyte Aureococcus anophagefferens, among others,<strong>and</strong> greatly reduced algal diversity at peak bloom. High densities seldom coincidedwith chlorophyll maxima, but a trend <strong>of</strong> higher numbers with increasing salinitywas observed. Un<strong>of</strong>ficial reports <strong>of</strong> reduced eelgrass survival <strong>and</strong> sportfishingcatches came from areas affected by the bloom.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Paul S. OlsenSource Inst.: New Jersey State Dept. <strong>of</strong> Environmental Protection, Division <strong>of</strong> WaterResources, Geological Survey, CN 029, Trenton, New Jersey 08625 USAAuthor: Perry, H. M., Stuck, K. C., <strong>and</strong> Howse, H. D.Date: 1979Title: First record <strong>of</strong> a bloom <strong>of</strong> Gonyaulax monilata in coastal waters <strong>of</strong> MississippiJournal: Gulf Research Reports 6(3):313-316Key words: bloom, Gonyaulax [Alex<strong>and</strong>rium] monilata, Mississippi Sound, din<strong>of</strong>lagellate,Gulf <strong>of</strong> Mexico, red tideSummary: Aerial surveys <strong>and</strong> water samples were performed during a widespread bloom <strong>of</strong>the armored din<strong>of</strong>lagellate Gonyaulax [now Alex<strong>and</strong>rium] monilata in MississippiSound as well as Alabama <strong>and</strong> Florida. The bloom persisted in Mississippi Soundfrom about 8 August until the advent <strong>of</strong> Hurricane Frederic on 11 September(1979). During the bloom, surface temperatures ranged from 30.0 to 30.8 o C.Salinities were below normal in Mississippi Sound due to previous heavy rains <strong>and</strong>ranged from 24.0 to 26.0 ppt. The maximum cell count reached 1.65 x 10 7cells/liter on 22 August. Pensacola Bay, Florida, had the maximum cell density forthat state <strong>of</strong> 3.18 x 10 7 cells/liter on 15 August, when water temperature was28.0 o C <strong>and</strong> salinity only 14.0 ppt.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Harriet M. Perry136


Source Inst.: Fisheries Research <strong>and</strong> Development, Gulf Coast Research Laboratory, OceanSprings, MS 39564 USAAuthor: Pierce, R. H., Henry, M. S., Pr<strong>of</strong>fitt, L. S. <strong>and</strong> Hasbrouck, P. A.Date: 1990Title/Ch.: <strong>Red</strong> tide toxin (brevetoxin) enrichment in marine aerosolBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 128-131Publisher: ElsevierKey words: red tide, toxin, brevetoxin, aerosol, din<strong>of</strong>lagellate, Ptychodiscus brevisSummary: Knowing that marine aerosols can carry brevetoxins, the authors measured thelevel <strong>of</strong> toxin enrichment in aerosols created in the laboratory versus originalculture concentrations. A total <strong>of</strong> six toxins were detected in both culture <strong>and</strong>aerosol, enriched by 5 to 50 times in the aerosol relative to the culture.Methods: See “Methods <strong>and</strong> Materials.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Richard H. PierceSource Inst.: Mote Marine Laboratory, 1600 City Isl<strong>and</strong> Park, Sarasota, FL 34236 USAAuthor: Proctor, R. R., Jr.Date: 1965Title/Ch.: Special Report: Biological Indicators in East Lagoon, Galveston Isl<strong>and</strong>Source: Fishery Research: Biological Laboratory, Galveston, Fiscal Year 1964, Circular230Pages: 87-88Publisher: United States Department <strong>of</strong> the Interior, Fish <strong>and</strong> Wildlife Service, Bureau <strong>of</strong>Commercial FisheriesKey words: East Lagoon, [Galveston Bay], oyster, oyster mortality, Gonyaulax monilata, [redtide, din<strong>of</strong>lagellate], bloomSummary: The author reports on the first two years <strong>of</strong> a study by the Bureau <strong>of</strong> CommercialFisheries to determine if water in the East Lagoon was suitable for oyster culture.An August, 1963 bloom <strong>of</strong> Gonyaulax [now Alex<strong>and</strong>rium] monilata compelledthe transfer <strong>of</strong> oysters kept in East Lagoon into the laboratory tank with acorresponding group <strong>of</strong> oysters kept there <strong>and</strong> the conversion <strong>of</strong> the tankcirculation to a closed, recirculating system. Both oyster groups still sufferedmortality thereafter (10 <strong>of</strong> the lagoon group's original 25 oysters; 2 out <strong>of</strong> 25 in thelaboratory group); the author suggests that G. monilata cells were in therecirculated water <strong>and</strong> caused the mortality. The bloom ceased in the lagoon afterabout two weeks, <strong>and</strong> the lagoon oysters were returned to the field. No furthermortality was associated with G. monilata.Methods: None (report)QA/QC: N/AContact: Raphael R. Proctor, Jr., ChemistSource Inst.: Bureau <strong>of</strong> Commercial Fisheries Biological Laboratory, Galveston, Texas, USA137


Author: Proctor, R. R., Jr.Date: 1966Title/Ch.: Special Report: Oyster growth experiment in East LagoonSource: Annual Report <strong>of</strong> the Bureau <strong>of</strong> Commercial Fisheries Biological Laboratory,Galveston, Texas, Fiscal Year 1965, Circular 246Pages: 48-49Publisher: United States Department <strong>of</strong> the Interior, Fish <strong>and</strong> Wildlife Service, Bureau <strong>of</strong>Commercial FisheriesKey words: oyster, oyster mortality, East Lagoon, [Galveston Bay], Gonyaulax monilata,[din<strong>of</strong>lagellate, red tide]Summary: The author states that conditions in a USFWS lagoon <strong>and</strong> the associated lagoonlaboratory are suitable only for short-term oyster experiments due to a commonoyster disease caused by Dermocystidium marinum <strong>and</strong> apparent annual (summer)blooms with subsequent oyster mortality due to the red tide din<strong>of</strong>lagellateGonyaulax [now Alex<strong>and</strong>rium] monilata.Methods: None per se (special report)QA/QC: N/AContact: R. R. Proctor, Jr., ChemistSource Inst.: Bureau <strong>of</strong> Commercial Fisheries Biological Laboratory, Galveston, Texas, USAAuthor: Quick, J. A., Jr., <strong>and</strong> Henderson, G. E.Date: 1975Title/Chap.: Evidences <strong>of</strong> new ichtyointoxicative phenomena in Gymnodinium breve red tidesBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 413-422Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: ichthyotoxin, Gymnodinium breve, red tide, din<strong>of</strong>lagellate, Florida, fishes, distressbehavior, necropsy, neurointoxication, chronic tissue damageSummary: The west coast <strong>of</strong> Florida's peninsula suffered an extended red tide from October1973 through June 1974. Necropsies were performed as quickly as possible on129 fish <strong>of</strong> fifteen genera <strong>and</strong> species that were either severely distressed or freshlydead at time <strong>of</strong> collection. Sixteen pathologies were consistently observed; somefish perished due to neurointoxication, whereas other species succumbed tochronic tissue damage. General symptoms were dehydration, hemolysis <strong>and</strong>disrupted blood clotting mechanisms. Distress behavior was marked <strong>and</strong> couldwell be diagnostic for some species.Methods: See "Methods <strong>and</strong> Materials."QA/QC: None per se; see "Methods <strong>and</strong> Materials."Contact: J. A. Quick, Jr.Source Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research laboratory, St.Petersburg, Florida, USAAuthor: Ray, S. M. <strong>and</strong> Wilson, W. B.138


Date: 1957Title: Effects <strong>of</strong> unialgal <strong>and</strong> bacteria-free cultures <strong>of</strong> Gymnodinium brevis on fish , <strong>and</strong>notes on related studies with bacteriaJournal: USFW Service Fishery Bulletin 123(57):469-496Key words: unialgal culture, Gymnodinium brevis, fish, bacteria, din<strong>of</strong>lagellate, Gulf <strong>of</strong>Mexico, red tide, toxinSummary: With a goal to (1) conduct a laboratory study complementary to field studies <strong>of</strong>marine animal mortality due to Gymnodinium brevis red tide [sp. now referred toas G. breve] <strong>and</strong> (2) increase underst<strong>and</strong>ing <strong>of</strong> why such mass mortalities occur,the authors used unialgal <strong>and</strong> bacteria-free G. brevis cultures to test for toxiceffects on fish. Given that fish mortality in bacteria-free cultures did not differ intoxicity from unialgal cultures, the authors concluded that G. brevis does producea toxic substance directly responsible for mass mortality <strong>of</strong> marine animals duringblooms in the Gulf <strong>of</strong> Mexico. Furthermore, this substance does not lose toxicitywith the death or even the absence <strong>of</strong> the G. brevis cells. The six fish speciestested were differentially sensitive to the toxin.Methods: See methods described within the body <strong>of</strong> the paper with respect to eachexperiment.QA/QC: None per se; see the descriptions <strong>of</strong> methods in the body <strong>of</strong> the paper.Contact: Sammy M. RaySource Inst.: Marine Laboratory, Texas A & M University, Galveston, TXAuthor: Ray, S. M. <strong>and</strong> Aldrich, D. V.Date: 1967Title/Ch.: Ecological interactions <strong>of</strong> toxic din<strong>of</strong>lagellates <strong>and</strong> molluscs in the Gulf <strong>of</strong> MexicoBook: Animal ToxinsEditor: F. Russell <strong>and</strong> P. SaundersPages: 75-83Publisher: Pergamon Press, NYKey words: toxin, din<strong>of</strong>lagellate, mollusc, Gulf <strong>of</strong> Mexico, Gymnodinium breve, Gonyaulaxmonilata, Florida, TexasSummary: The authors investigate several hypotheses as to why red tide din<strong>of</strong>lagellates in theGulf <strong>of</strong> Mexico have not caused comparable harm to its large shellfish industry asother species have to those <strong>of</strong> the Atlantic <strong>and</strong> Pacific coasts. Cultures <strong>of</strong>Gymnodinium breve <strong>and</strong> Gonyaulax monilata [preferably called Alex<strong>and</strong>riummonilata], both Gulf species responsible for fish kills, were introduced into oystercultures. The oysters were then fed to chicks, but poisoning in the chicks onlyoccurred with G. breve. Oysters <strong>and</strong> other shellfish did not filter when exposed toG. monilata in laboratory culture, a behavior apparently identical to wildpopulations exposed to natural G. monilata blooms in situ. Differing toxic effectsbetween the two din<strong>of</strong>lagellates were observed in molluscs, a polychaete <strong>and</strong> fish.G. breve blooms may be too infrequent to severely affect commercial shellfish inthe Gulf, <strong>and</strong> Gulf shellfish avoid any negative effects <strong>of</strong> G. monilata by notfiltering.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”139


Contact: Sammy M. RaySource Inst.: Marine Laboratory, Texas A & M University, Galveston, TXAuthor: Riley, C. M., Holt, S. A., Holt, G. J., Buskey, E. J. <strong>and</strong> Arnold, C. R.Date: 1989Title: Mortality <strong>of</strong> larval red drum (Sciaenops ocellatus) associated with a Ptychodiscusbrevis red tideJournal: Contributions in Marine Science 31:137-146Key words: larvae, red drum, Sciaenops ocellatus, din<strong>of</strong>lagellate, Ptychodiscus brevis, redtide, Texas, recruitmentSummary: Ptychodiscus brevis, preferably known as Gymnodinium breve, is a red tidedin<strong>of</strong>lagellate that can cause paralysis <strong>and</strong> death in laboratory-spawned <strong>and</strong> wildcaughtred drum larvae at all concentrations above 40 cells/ml. A fall outbreak <strong>of</strong>this red tide on the Texas Gulf coast in 1986 achieved average peak densities <strong>of</strong>7000 cells/ml, but concentrations <strong>of</strong> almost 5 x 10 4 cells/ml were detected inisolated patches in bays near Port Aransas throughout October. <strong>Red</strong> tide blooms<strong>of</strong> this sort can reduce red drum recruitment during spawning periods.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Cecelia M. RileySource Inst.: University <strong>of</strong> Texas at Austin, Marine Science Institute, Port Aransas, TX 78373-1267Author: Roberts, B. S.Date: 1979Title/Ch.: Occurrence <strong>of</strong> Gymnodinium breve red tides along the west <strong>and</strong> east coasts <strong>of</strong>Florida during 1976 <strong>and</strong> 1977Book: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 199-202Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: Gymnodinium breve, red tide, Florida, bloomSummary: <strong>Red</strong> tides in 1976 <strong>and</strong> 1977 were analyzed with regard to their most likely means<strong>of</strong> initiation, support <strong>and</strong> maintenance, all three progressing in a predictablefashion. In 1976, for instance, surface concentrations <strong>of</strong> G. breve increasedgradually <strong>of</strong>fshore, were then transported inshore <strong>and</strong>, after winds inducedmovement to the south, gradually dissipated. Movement <strong>and</strong> increasedconcentrations could be attributed to winds <strong>and</strong> currents.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Beverly S. RobertsSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Ave. S. E., St. Petersburg, FL 33701 USAAuthor: Roberts, B. S., Henderson, G. E. <strong>and</strong> Medlyn, R. A.140


Date: 1979Title/Ch.: The effect <strong>of</strong> Gymnodinium breve toxin(s) on selected mollusks <strong>and</strong> crustaceansBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 419-424Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: Gymnodinium breve, toxin, mollusk, crustacean, Fasciolaria hunteria, Melongenacorona, Oliva sayana, Callinectes sapidus, Menippe mercenaria, red tideSummary: The toxin or toxins <strong>of</strong> the red tide din<strong>of</strong>lagellate Gymnodinium breve adverselyaffected three species <strong>of</strong> mollusks (Fasciolaria hunteria, Melongena corona <strong>and</strong>Oliva sayana) when directly exposed to G. breve toxins (55-69% mortality over48 hrs.), but had no affect on two species <strong>of</strong> crustacean (Callinectes sapidus <strong>and</strong>Menippe mercenaria) when all crab species were fed toxic clam meat. Along witha lack <strong>of</strong> any mortality, the crab tissue, upon analysis, showed no accumulation <strong>of</strong>the toxin.Methods: See “Methods <strong>and</strong> Materials.”QA/QC: None per se; see “Methods <strong>and</strong> Materials.”Contact: Beverly S. RobertsSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Ave. S. E., St. Petersburg, FL 33701 USAAuthor: Roszell, L. E., Schulman, L. S. <strong>and</strong> Baden, D. G.Date: 1990Title/Ch.: Toxin pr<strong>of</strong>iles are dependent on growth stages in cultured Ptychodiscus brevisBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 403-406Publisher: ElsevierKey words: toxin, growth, Ptychodiscus brevis, red tide, din<strong>of</strong>lagellateSummary: Characterization from over 3000 liters <strong>of</strong> batch cultures <strong>of</strong> the various toxins, allvery similar, produced by Ptychodiscus brevis (now known as Gymnodiniumbreve) during logarithmic, stationary <strong>and</strong> declining phases <strong>of</strong> population growthreveals that the eight possible toxins appear individually or in combination atdifferent times in the life <strong>of</strong> a population. What the authors do not suggest butwhat seems obvious is extension <strong>of</strong> these findings to characterization <strong>of</strong> the overallgrowth stage <strong>of</strong> blooms in situ.Methods: See “Experimental” <strong>and</strong> “Results.”QA/QC: None per se; see “Experimental.”Contact: Laurie E. RoszellSource Inst.: University <strong>of</strong> Miami Rosenstiel School <strong>of</strong> Marine <strong>and</strong> Atmospheric Science,Division <strong>of</strong> Biology <strong>and</strong> Living Resources, 4600 Rickenbacker Causeway, Miami,FL 33149 USAAuthor: Rounsefell, G. A., <strong>and</strong> Nelson, W. R.141


Date: 1966Title: <strong>Red</strong>-tide research summarized to 1964 including an annotated bibliographyPages: 85 pp.Source: United States Fish <strong>and</strong> Wildlife Service Special Scientific Report, Fisheries No.535, Washington, D. C.Key words: red tide, Florida, bloom, mortality, plankton, din<strong>of</strong>lagellate, toxic, Gymnodiniumbreve, Gulf <strong>of</strong> Mexico, TexasSummary: Using published <strong>and</strong> unpublished data, the authors summarize the status <strong>of</strong> red tideresearch in Florida up to 1964, listing 292 references with some annotations, <strong>and</strong>discuss the influence <strong>of</strong> oceanographic conditions on blooms, seasonal <strong>and</strong> coastaldistribution <strong>of</strong> the Florida red tide <strong>and</strong> progress in various research efforts.Emphasis is upon Gymnodinium breve.Methods: None (research summary)QA/QC: N/AContact: Tom SerotaSource Inst.: U. S. Fish <strong>and</strong> Wildlife Service, Fisheries Resources Office, 6300 Ocean Dr.,Corpus Christi, Texas 78412 USAAuthor: S<strong>and</strong>gren, C. D.Date: 1983Title/Chap.: Survival strategies <strong>of</strong> chrysophycean flagellates: reproduction <strong>and</strong> the formation<strong>of</strong> resistant resting cystsBook: Survival strategies <strong>of</strong> the algaeEditor: G. A. FryxellPages: Pp. 23-48Publisher: Cambridge University Press, CambridgeKey words: chrysophytes, flagellates, survival strategy, reproduction, resting cyst, microalgae,resting stage, statospore, phytoplankton, life historySummary: Many phytoplankton have resistant resting stages (cysts) as a necessary component<strong>of</strong> their life history strategy, enabling them to outlast unfavorable environmentalconditions until recolonization <strong>of</strong> the plankton is possible. These strategies existfor both freshwater <strong>and</strong> marine phytoplankton. Freshwater chrysophytes known asgolden algae are very seasonal species <strong>and</strong> use a siliceous resting cyst known as astatospore. The statospore is the primary subject <strong>of</strong> the review paper, whichcovers its development, surface morphology, <strong>and</strong> responses to factors influencingencystment, both in individual cells <strong>and</strong> populations. The author also refers earlyin the review to several general reference texts in phycology that address restingcysts in marine chrysophytes.Methods: None per se (Review article)QA/QC: N/AContact: Craig D. S<strong>and</strong>grenSource Inst.: Department <strong>of</strong> Biology, University <strong>of</strong> Texas at Arlington, Arlington, TX 76019Author: Seliger, H. H., Tyler, M. A. <strong>and</strong> McKinley, K. R.Date: 1979Title/Ch.: Phytoplankton distributions <strong>and</strong> red tides resulting from frontal circulation patterns142


Book: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 239-248Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: phytoplankton, distribution, red tide, frontal circulation, Gymnodinium breve,Florida, Prorocentrum mariae lebouriae, Chesapeake BaySummary: This paper extrapolates conditions existing for blooms <strong>of</strong> Prorocentrum mariaelebouriae in Chesapeake Bay with the west Florida red tides <strong>of</strong> Gymnodiniumbreve <strong>and</strong> that <strong>of</strong> another species in New Engl<strong>and</strong>. Conditions promoting red tidesinclude (1) downward movement <strong>of</strong> surface concentrations <strong>of</strong> red tidedin<strong>of</strong>lagellates resulting from wind <strong>and</strong>/or tidal influences, (2) transport to depth<strong>and</strong> toward shallow waters, (3) upwelling into the photic zones <strong>of</strong> shallow waters<strong>and</strong> (4) concentration into wind- or tide-driven convergences at the surface,producing dense surface patches. A specific scenario is included for each <strong>of</strong> thethree locales.Methods: N/A (Published hypothesis)QA/QC: N/AContact: Howard H. SeligerSource Inst.: McCollum-Pratt Institute <strong>and</strong> Department <strong>of</strong> Biology, The Johns HopkinsUniversity, Baltimore, MD 21218 USAAuthor: Shanley, E. <strong>and</strong> Vargo, G. A.Date: 1990Title/Ch.: Cellular composition, growth, photosynthesis, <strong>and</strong> respiration rates <strong>of</strong>Gymnodinium breve under varying light levelsBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 831-836Publisher: ElsevierKey words: growth, photosynthesis, respiration, Gymnodinium breve, chlorophyll a,chlorophyll c, carotenoidSummary: Laboratory cultures <strong>of</strong> Wilson’s clone <strong>of</strong> the red tide din<strong>of</strong>lagellate Gymnodiniumbreve were adapted to the full range <strong>of</strong> light available in situ in an attempt tomeasure <strong>and</strong> compare any variations in cellular composition, growth rates,respiration rates <strong>and</strong> rates <strong>of</strong> photosynthesis. Results indicated a proportionalincrease between photosynthesis <strong>and</strong> growth rates over a range <strong>of</strong> irradiance equalto about 2% to 16% <strong>of</strong> surface irradiance. The study’s purpose was to clarify thelow light response for Gymnodinium breve populations in <strong>of</strong>fshore waters atdepths <strong>of</strong> 20-30 m, where light levels can be as low as 2% <strong>of</strong> the mean dailyincident irradiance <strong>and</strong> where bloom initiation is hypothesized. Findings revealthat <strong>of</strong>fshore populations at such depths may indeed serve as seed populations forblooms.Methods: See “Methods.”QA/QC: None per se; see “Methods.”143


Contact: Evanne ShanleySource Inst.: Department <strong>of</strong> Marine Science, University <strong>of</strong> South Florida, 140 Seventh AvenueSouth, St. Petersburg, FL 33701 USAAuthor: Shormann, D. E.Date: 1992Thesis: The effects <strong>of</strong> freshwater inflow <strong>and</strong> hydrography on the distribution <strong>of</strong> brown tidein south Texas baysPages: 111 pp.Source: M. A. Thesis, Department <strong>of</strong> Marine Science, The University <strong>of</strong> Texas at Austin.Key words: hydrography, brown tide, Texas, nitrate, Copano Bay, residence time, phosphate,ammonium, nitrite, silicateSummary: The author created an equation to predict mean bay salinity <strong>and</strong> determine ifCopano Bay mixing processes werre dominated by freshwater influx, tidal mixingor a combination there<strong>of</strong>. A second equation was created to gauge residence timefrom freshwater inflow. Phytoplankton growth (including brown tide) did notappear to be either N- or P-limited. Mean chlorophyll concentrations weresignificantly correlated with with particulate carbon <strong>and</strong> nitrogen <strong>and</strong> thus from insitu phytoplankton production. A growth rate <strong>of</strong> the brown tide was only 0.14day -1 . Copano Bay was eutrophic with respect to phosphorus <strong>and</strong> had a N/P ratioidentical to fertilizer ratios used by local farmers. The mean residence <strong>and</strong>response times <strong>of</strong> Copano Bay were similar to Baffin Bay <strong>and</strong> about one order <strong>of</strong>magnitude greater than Nueces Bay. Such low water circulation could well aidpersistence <strong>of</strong> the brown tide in those bays; notably, Nueces Bay is rarely affectedby any algal bloom.Methods: See “Methods” on pp. 18-25 <strong>and</strong> p. 94.QA/QC: None per se; see “Methods.”Contact: Terry E. WhitledgeSource Inst.: The University <strong>of</strong> Texas Marine Science Institute, P. O. Box 1267, Port Aransas,TX 78373-1267Author: Shumway, S. E.Date: 1990Title: A review <strong>of</strong> the effects <strong>of</strong> algal blooms on shellfish <strong>and</strong> aquacultureJournal: Journal <strong>of</strong> the World Aquaculture Society 21(2):65-104Key words: toxic algae, bloom, shellfish, aquaculture, din<strong>of</strong>lagellates, Gymnodinium breve,Ptychodiscus brevis, Gonyaulax monilata, Aureococcus anophagefferensSummary: This review paper comments on worldwide toxic algal blooms, presents thecommercial management possibilities for coping with such blooms <strong>and</strong> suggestsmeans <strong>of</strong> efficiently <strong>and</strong> successfully using marine resources in spite <strong>of</strong> theinstability caused by blooms. Included are the two red tide din<strong>of</strong>lagellatesresponsible for problem blooms along the Texas Gulf Coast <strong>and</strong> a brown tidechrysophyte similar to the organism that began blooming in the Laguna Madre in1990.Methods: N/A (Review paper)QA/QC: N/A144


Contact: S<strong>and</strong>ra E. ShumwaySource Inst.: Maine Department <strong>of</strong> Marine Resources <strong>and</strong> Bigelow Laboratory for OceanSciences, West Boothbay Harbor, Maine 04575 USAAuthor: Siddall, S. E.Date: 1987Title: Climatology <strong>of</strong> Long Isl<strong>and</strong> related to the "brown tide" phytoplankton blooms <strong>of</strong>1985 <strong>and</strong> 1986Pages: 16 pp.Publisher: Marine Sciences Research Center, SUNY, Stony Brook, NYKey words: climatology, Long Isl<strong>and</strong>, brown tide, phytoplankton, bloom, meteorology,precipitationSummary: Using long-term records to indicate prevailing trends over months or years forLong Isl<strong>and</strong> <strong>and</strong> short-term (daily) records <strong>of</strong> weather events for 1984-86 for fivestations among East End bays, climatological/meterorological data were matchedwith brown tide cell counts in an effort to discover trends <strong>and</strong> suggest causativefactors for the bloom during those two years. No significant anomalies intemperature, precipitation, wind or daily radiation could be associated with browntide initiation or maintenance. Long Isl<strong>and</strong> experienced a moderate to severedrought when the first bloom occurred in 1985, but the report emphasizes that themere existence <strong>of</strong> drought beforeh<strong>and</strong> says nothing about true underlying causesfor the bloom <strong>and</strong> cannot be considered statistically valid. A hypothesis <strong>of</strong> bloomformation <strong>and</strong> persistence can, however, be linked to precipitation <strong>and</strong> resultingterrestrial run<strong>of</strong>f or groundwater effects during an otherwise dry period.Methods: N/A (Report)QA/QC: N/AContact: Scott E. SiddallSource Inst.: Marine Sciences Research Center, State University <strong>of</strong> New York, StonyBrook, NY 11794-5000Author: Sieburth, J. McN., Johnson, P. W. <strong>and</strong> Hargraves, P. E.Date: 1988Title: Ultrastructure <strong>and</strong> ecology <strong>of</strong> Aureococcus anophagefferens Gen. et sp. nov.(Chrysophyceae): the dominant picoplankter during a bloom in Narragansett Bay,Rhode Isl<strong>and</strong>, summer 1985Journal: J. Phycol. 24:416-425Key words: ultrastructure, ecology, Aureococcus anophagefferens, Chrysophyceae,picoplankton, bloom, Narragansett Bay, Rhode Isl<strong>and</strong>, Synechococcus Nägeli,Calycomonas ovalisSummary: The unknown 2 um diameter chrysophyte herein described as the novel speciesAureococcus anophagefferens first came to the authors attention with thecessation <strong>of</strong> feeding in filter feeders exposed to Narragansett Bay seawater <strong>and</strong> adensity <strong>of</strong> 10 9 cell/L. Neither phase contrast nor epifluorescence microscopy coulddiscriminate the chrysophyte from similar sized phytoplankton, though examination<strong>of</strong> thin sections with transmission electron microscopy was successful. About 300attempts to culture the alga with twelve different culture media were unsuccessful.145


Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: John McN. SieburthSource Inst.: Graduate School <strong>of</strong> Oceanography, University <strong>of</strong> Rhode Isl<strong>and</strong> Bay Campus,South Ferry Road, Narragansett, Rhode Isl<strong>and</strong> 02882-1197 USAAuthor: Sieburth, J. McN. <strong>and</strong> Johnson, P. W.Date: 1989Title/Ch.: Picoplankton ultrastructure: a decade <strong>of</strong> preparation for the brown tide alga,Aureococcus anophagefferensBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 1-21Publisher: Springer-VerlagKey words: picoplankton, ultrastructure, brown tide, Aureococcus anophagefferens,transmission electron microscopy, TEMSummary: In an effort to determine the trophic definition <strong>of</strong> pico- <strong>and</strong> nanoplankton, theauthors developed a TEM process to examine thin sections <strong>of</strong> cells in the naturallyoccurring populations <strong>of</strong> uncultured seawater samples. With success depending oncells possessing unique ultrastructural characteristics, the authors, after a decade <strong>of</strong>such work on samples from Narragansett Bay, were in a unique position toexamine samples <strong>of</strong> brown tide upon bloom occurrence there, beginning in 1985.Their timely sampling incriminated Aureococcus anophagefferens as the algaresponsible for cessation <strong>of</strong> filter feeding during the bloom’s peak. Reexamination<strong>of</strong> previous samples several years before the initial brown tide bloomrevealed A. anophagefferens as a likely normal component <strong>of</strong> the picoplankton.Many such picoalgae probably remain unindentified.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: John McN. SieburthSource Inst.: Graduate School <strong>of</strong> OceanographyUniversity <strong>of</strong> Rhode Isl<strong>and</strong>, Narragansett, RI 92882-1197 USAAuthor: Sievers, A. M.Date: 1969Title: Comparative toxicity <strong>of</strong> Gonyaulax monilata <strong>and</strong> Gymnodinium breve to annelids,crustaceans, molluscs <strong>and</strong> a fishJournal: Journal <strong>of</strong> Protozoology 16(3):401-404Key words: toxicity, Gonyaulax monilata, Gymnodinium breve, annelid, crustacean, mollusc,fish, din<strong>of</strong>lagellate, mortality, toxin,Summary: Several kinds <strong>of</strong> marine animals were exposed to concentrations <strong>of</strong> twodin<strong>of</strong>lagellate cultures from undiluted to 90% dilution over 48 hours in order todetermine toxicity <strong>of</strong> the two toxins in terms <strong>of</strong> mortality. Fish were equallysensitive to both Gonyaulax monilata [preferably known as Alex<strong>and</strong>rium146


monilata] <strong>and</strong> Gymnodinium breve toxins, crustaceans were resistant to both, <strong>and</strong>both annelids <strong>and</strong> molluscs were more sensitive to G. monilata toxin.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Anita M. SieversSource Inst.: Marine Laboratory, Texas A&M University, Galveston, TX 77550Author: Smayda, T. J. <strong>and</strong> F<strong>of</strong>on<strong>of</strong>f, P.Date: 1989Title/Ch.: An extraordinary, noxious brown-tide in Narragansett Bay. II. Inimical effects.Book: <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology. Proceedings <strong>of</strong> theFirst International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14, 1987, inTakamatsu, Kagawa Prefecture, Japan.Editor: T. Okaichi, D. M. Anderson <strong>and</strong> T. NemotoPages: 133-136Publisher: Elsevier Science Publishing Company, Inc., New YorkKey words: brown tide, Narragansett Bay, bloom, Aureococcus anophagefferens (appears asA. anorexefferens), zooplankton, Mytilus edulis, anchovy, kelp, benthic larvae,Acartia tonsaSummary: The density <strong>of</strong> an A. anophagefferens bloom was strongly <strong>and</strong> negativelycorrelated (-0.91) with density <strong>of</strong> the predominant copepod Acartia tonsa; A.tonsa adult females fed the brown tide alga suffered lower feeding rates, eggproduction <strong>and</strong> body weight. Compared to 1984 <strong>and</strong> 1986, the 1985 bloomcorresponded with a 60-fold lower mean cladoceran abundance. Natural musselbeds lost 30%-100% <strong>of</strong> their populations, apparently due to starvation,substantiated by cessation <strong>of</strong> filtering in laboratory-raised mussels at brown tidealgal densities exceeded 500 x 10 6 cells/liter. Laminarian kelp died when theeuphotic zone mussels to which they were attached also died <strong>and</strong> sank to deeperwaters. Aureococcus abundance was also negatively correlated with benthic larvalnumbers (r = -0.58). Polychaete <strong>and</strong> bivalve larvae <strong>and</strong> bay anchovy eggs were alllower than normal (1.5x, 3.6x <strong>and</strong> 10x lower, resp.) as was zooplankton <strong>and</strong>benthic grazing.Methods: None (Review paper)QA/QC: N/AContact: Theodore J. SmaydaSource Inst.: Graduate School <strong>of</strong> Oceanography, University <strong>of</strong> Rhode Isl<strong>and</strong>, Kingston, RhodeIsl<strong>and</strong> 02881 USAAuthor: Smayda, T. J. <strong>and</strong> Villareal, T. A.Date: 1989Title/Ch.: An extraordinary, noxious brown-tide in Narragansett Bay. I. The organism <strong>and</strong>its dynamics.Book: <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology. Proceedings <strong>of</strong> theFirst International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14, 1987, inTakamatsu, Kagawa Prefecture, Japan.Editor: T. Okaichi, D. M. Anderson <strong>and</strong> T. Nemoto147


Pages: 129-132Publisher: Elsevier Science Publishing Company, Inc., New YorkKey words: brown tide, Narragansett Bay, bloom, chrysophyte, Aureococcus anophagefferensSummary: The A. anophagefferens bloom during May-September 1985 in Narragensett Bayproduced densities as high as 1.2 x 10 9 cells/liter; mean abundance was stronglycorrelated (r = 0.98) with the observed salinity gradient, though authors suggestthat salinity per se was not a causative factor. Neither was eutrophication, forstrong inverse correlations appeared between mean abundance <strong>and</strong> "ammoniumplus nitrate" <strong>and</strong> phosphate concentrations (-0.76 <strong>and</strong> -0.62, resp.). Diatoms,din<strong>of</strong>lagellates, micr<strong>of</strong>lagellates <strong>and</strong> euglenids bloomed concurrently <strong>and</strong>extensively, with euglenids persisting through November once the brown tidebloom disappeared. Authors postulate that simultaneous brown tide blooms inLong Isl<strong>and</strong> <strong>and</strong> New Jersey indicate a mesoscale scenario associated withclimatologic/hydrographic conditions. Spring increases in quantity or duration <strong>of</strong>light <strong>and</strong> concentrations <strong>of</strong> phagotrophic flagellates may be bloom catalysts.Methods: None specified. See “Materials <strong>and</strong> Methods.”QA/QC: None. See “Materials <strong>and</strong> Methods.”Contact: Theodore J. SmaydaSource Inst.: Graduate School <strong>of</strong> Oceanography, University <strong>of</strong> Rhode Isl<strong>and</strong>, Kingston, RhodeIsl<strong>and</strong> 02881 USAAuthor: Smayda, T. J. <strong>and</strong> Villareal, T. A.Date: 1989Title/Ch.: The 1985 ‘brown-tide’ <strong>and</strong> the open phytoplankton niche in Narragansett Bayduring summerBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 159-187Publisher: Springer-VerlagKey words: brown tide, phytoplankton, Narragansett Bay, chrysophyte, Aureococcusanophagefferens, bloom, nicheSummary: The authors call attention to the lack <strong>of</strong> information on the presence <strong>and</strong>abundance <strong>of</strong> other phytoplankton species before, during <strong>and</strong> after a bloom <strong>and</strong>suggest that such information may be crucial to a complete underst<strong>and</strong>ing <strong>of</strong>bloom dynamics. Smaller-scale but significant blooms <strong>of</strong> several diatom,din<strong>of</strong>lagellate <strong>and</strong> other phyt<strong>of</strong>lagellate species (a total <strong>of</strong> 14 other taxa) cooccurredwith <strong>and</strong>/or followed the major A. anophagefferens bloom at differenttimes <strong>and</strong> for various durations during May-October 1985 in Narragensett Bay.As a result, one may view blooms <strong>of</strong> these other taxa after the A. anophagefferensbloom ceased as exploitations <strong>of</strong> the niche previously dominated by A.anophagefferens. Future blooms should not be considered monospecific unless soproven.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Theodore J. Smayda148


Source Inst.: Graduate School <strong>of</strong> Oceanography, University <strong>of</strong> Rhode Isl<strong>and</strong>, Kingston, RhodeIsl<strong>and</strong> 02881 USAAuthor: Snider, R. (ed.)Date: 1987Title: <strong>Red</strong> tide in Texas: an explanation <strong>of</strong> the phenomenonPages: 4 pp.Publisher: Marine Information Service, Texas A&M Sea Grant College ProgramKey words: red tide, din<strong>of</strong>lagellate, bloom, toxin, Ptychodiscus brevis, Gonyaulax monilata,Gulf <strong>of</strong> Mexico, fish kills, neurotoxic shellfish poisoningSummary: This pamphlet summarizes the pertinent facts about the red tide organisms,Ptychodiscus brevis <strong>and</strong> Gonyaulax monilata [now known as Gymnodinium breve<strong>and</strong> Alex<strong>and</strong>rium monilata, resp.]; causes <strong>of</strong> sudden blooms; locations <strong>and</strong> times<strong>of</strong> bloom occurrence; red tide impacts on marine life, public health <strong>and</strong> theeconomy; myths about red tides; history <strong>and</strong> folklore; <strong>and</strong> control, prevention <strong>and</strong>prediction.Methods: N/A (Public information bulletin)QA/QC: N/AContact: Rhonda SniderSource Inst.: Marine Information Service, Sea Grant College Program, Texas A&MUniversity, College Station, TX 77843-4115Author: Starr, T. J.Date: 1958Title: Notes on a toxin from Gymnodinium breveJournal: Texas Reports in Biology <strong>and</strong> Medicine 16:500-507Key words: toxin, Gymnodinium breve, bloom, algae, culture, mullet, Lebistes reticulatus,Mugel cephalus, exotoxin, endotoxinSummary: This paper describes bioassay procedures for toxin from unialgal cultures <strong>of</strong>Gymnodinium breve as well as some properties <strong>of</strong> the crude toxin. Mullet <strong>and</strong>guppies were used for the toxin bioassay; toxicity increases with any agent thatdestroys G. breve. The organism's fragile nature frustrated attempts toconcentrate the toxin by concentrating intact cells. The author cannot concludethat G. breve toxin is either a true exotoxin (produced or excreted from livingcells) or endotoxin (introduced by cell lysis or autolysis).Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Theodore J. StarrSource Inst.: The University <strong>of</strong> Texas Medical Branch, Department <strong>of</strong> Preventive Medicine <strong>and</strong>Public Health, Virus Research Laboratory, Galveston, Texas, USAAuthor: Stehn, T.Date: 1987Title: Whooping cranes during the 1986-1987 winterPages: 45 pp.Source: Aransas National Wildlife Refuge, U. S. Fish <strong>and</strong> Wildlife Service149


Key words: whooping crane, habitat, red tide, algae, Ptychodiscus brevis, toxin, clamSummary: In an excerpt from the unpublished report, the author reveals that the Ptychodiscusbrevis [now Gymnodinium breve] red tide <strong>of</strong> 1986-87 in South Texas came veryclose to infiltrating whooping crane critical habitat in the Aransas National WildlifeRefuge in the fall <strong>of</strong> 1986. Because clams, swallowed whole, comprise a smallpart <strong>of</strong> a whooping crane's diet in fall <strong>and</strong> early winter, brevetoxin could possiblysicken or kill any cranes consuming contaminated clams. Because scaup <strong>and</strong>cormorants have been known to perish from exposure to brevetoxin in Florida, thepotential threat to the small population <strong>of</strong> endangered whooping cranes calls forfuture monitoring.Methods: None (unpublished report)QA/QC: N/AContact: Tom Stehn, Refuge BiologistSource Inst.: Aransas National Wildlife Refuge, US Fish <strong>and</strong> Wildlife Service, P. O. Box100, Austwell, TX 77950 USAAuthor: Steidinger, K. A.Date: 1964Title: Gymnodinium breve DavisSource: Florida Board <strong>of</strong> Conservation Leaflet Series: Plankton, Vol. 1, No. 1 (2 pp.)Key words: Gymnodinium breve, red tide, cyst, neurotoxin, bloom, din<strong>of</strong>lagellate, Gulf <strong>of</strong>Mexico, salinity, temperatureSummary: This monograph provides a drawing, photomicrographs <strong>and</strong> a description <strong>of</strong> thered tide din<strong>of</strong>lagellate Gymnodinium breve, plus comments on its distribution <strong>and</strong>ecology.Methods: None (description)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K. A.Date: 1975aTitle/Chap.: Basic factors influencing red tidesBook: Proceedings <strong>of</strong> the First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 153-162Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: red tide, bloom, bloom initiation, din<strong>of</strong>lagellate, sexual phase, benthic cyst, seedpopulation, Gymnodinium breve, FloridaSummary: The author sets forth the three aspects common to all toxic red tides: (1) bloominitiation, defined as an increase in population size; (2) support, in terms <strong>of</strong>favorable levels <strong>of</strong> nutrients, growth factors, temperature <strong>and</strong> salinity; <strong>and</strong> (3)bloom maintenance <strong>and</strong> transport by meteorologic <strong>and</strong> hydrologic forces. At time<strong>of</strong> publication, dormant stages in the form <strong>of</strong> benthic cysts had not been150


substantiated for G. breve (though true for many other species) nor had theconcept <strong>of</strong> "seed populations" been proven, but such life cycle work was <strong>of</strong> highpriority. The locality <strong>of</strong> bloom initiation was also considered critical, an examplebeing the general location <strong>of</strong> G. breve blooms, at depths <strong>of</strong> 12-37 m 16-64 km<strong>of</strong>fshore <strong>of</strong> southwest Florida, as a likely place to find dormant cysts. G. breveblooms, at least, were characterized from available data as gradual increases inmotile cells, not sudden increases in cell division rates. Important transporting <strong>and</strong>concentrating mechanisms are listed as winds, currents <strong>and</strong> organism migration.Offshore sampling programs for adequate warning <strong>of</strong> blooms are highlyrecommended.Methods: None (review)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Ave. S. E., St. Petersburg, FL 33701 USAAuthor: Steidinger, K. A.Date: 1975bTitle: Implications <strong>of</strong> din<strong>of</strong>lagellate life cycles on initiation <strong>of</strong> Gymnodinium breve redtidesJournal: Environmental Letters 9(2):129-139Key words: din<strong>of</strong>lagellate, life cycle, initiation, Gymnodinium breve, red tide, Florida, restingcyst, hypnozygote, seed population, sexualitySummary: West Florida coastal waters 18-74 km <strong>of</strong>fshore are the sites <strong>of</strong> red tide bloominitiation, usually in late summer or fall. If G. breve has a sexual cycle involvingalternation <strong>of</strong> generations that includes a resting cyst stage (hypnozygote), thenseed populations or seed "beds" could be located <strong>and</strong> mapped. The author reviewsin detail advances in din<strong>of</strong>lagellate life cycle work.Methods: None (review)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Ave. S. E., St. Petersburg, FL 33701 USAAuthor: Steidinger, K. A.Date: 1979Title/Ch.: Collection, enumeration <strong>and</strong> identification <strong>of</strong> free-living marine din<strong>of</strong>lagellatesBook: Toxic Din<strong>of</strong>lagellate Blooms, Proceedings <strong>of</strong> the Second International Conferenceon Toxic Din<strong>of</strong>lagellate Blooms, Key Biscayne, Florida, Oct. 31-Nov. 5, 1978Editor: D. L. Taylor <strong>and</strong> H. H. SeligerPages: 435-442Publisher: Elsevier/North-Holl<strong>and</strong>, Inc.Key words: din<strong>of</strong>lagellate, Peridinales, theca, armored, unarmored, toxin, bloom, PtychodiscusbrevisSummary: Water samples <strong>of</strong> din<strong>of</strong>lagellates can be collected by numerous mechanical meansdepending on the species <strong>of</strong> interest, the study area <strong>and</strong> the researcher’s purpose.151


Formalin as both fixative <strong>and</strong> preservative is suggested for armored specimens, butunarmored species are difficult to fix <strong>and</strong> preserve. Enumeration methods arevaried for both living <strong>and</strong> preserved cells. Plates <strong>of</strong> armored species may bedistinguished optically, with stains, or by physical or chemical separation; cell size,shape, number <strong>and</strong> organelle placement are among other diagnostic features usedto identify unarmored din<strong>of</strong>lagellates. Cultures may confound identification due toculture-specific anomalous effects. Toxic species are thought to number less than20, though many more produce blooms in estuaries <strong>and</strong> neritic waters. The namePtychodiscus brevis [since renamed as Gymnodinium breve] is proposed for atoxic, bloom-producing species.Methods: N/A (Review paper)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Ave. S. E., St. Petersburg, FL 33701 USAAuthor: Steidinger, K. A.Date: 1983Title/Ch.: A re-evaluation <strong>of</strong> toxic din<strong>of</strong>lagellate biology <strong>and</strong> ecologyBook: Progress in Phycological Research, Vol. 2Editor: Round <strong>and</strong> ChapmanPages: 147-188Publisher: Elsevier Science Publishers B. V.Key words: toxic din<strong>of</strong>lagellate, Ptychodiscus brevis, Gymnodinium breve, Gonyaulaxmonilata, Pyrrhophyta, neurotoxin, hemolytic agent, bloom, systematics, lifehistorySummary: This review summarizes toxic din<strong>of</strong>lagellate research to 1983, commenting onsystematics, life histories, toxins, organismal activity, environmental impacts <strong>and</strong>certain ecological aspects <strong>and</strong> suggesting avenues for future work. The two redtide din<strong>of</strong>lagellates known to afflict the Texas Gulf Coast are included.Methods: N/A (Review paper)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K. A. <strong>and</strong> Haddad, K.Date: 1981Title: Biologic <strong>and</strong> hydrographic aspects <strong>of</strong> red tidesJournal: Bioscience 31(11):814-819Key words: red tide, bloom, din<strong>of</strong>lagellate, life cycle, initiation, endotoxin, Ptychodiscus brevis[now Gymnodinium breve], sexual reproduction, satellite imagerySummary: This article summarizes red tides in general, din<strong>of</strong>lagellate biology, the sequentialdevelopment <strong>of</strong> red tides <strong>and</strong> their occurrence in Florida, providing an excellentintroduction to the state <strong>of</strong> knowledge <strong>and</strong> research on the factors concerning redtides in the eastern Gulf <strong>of</strong> Mexico as <strong>of</strong> 1981.152


Methods: None (review)QA/QC: N/AContact: K. A. SteidingerSource Inst.: Bureau <strong>of</strong> Marine Research, Florida Department <strong>of</strong> Natural Resources, St.Petersburg, Fl 33701 USAAuthor: Steidinger, K. A. <strong>and</strong> Ingle, R. M.Date: 1972Title: Observations on the 1971 summer red tide in Tampa Bay, FloridaJournal: Environmental Letters 3(4):271-278Key words: red tide, Tampa Bay, Florida, din<strong>of</strong>lagellate, Gymnodinium breve, bloom, cyst,estuaries, reefs, bivalvesSummary: Using information from a summer red tide on the west Florida coast that persistedfor 3.5 months, the authors' observations suggest that (1) Gymnodinium breve is aneritic species hindered by low salinity barriers in estuaries, (2) dense cellconcentrations are due more to physical factors than rapid cell division, (3) G.breve temporarily affects inshore <strong>and</strong> nearshore reef fisheries only, (4) commercialshellfish may be safely consumed 1-2 months after the end <strong>of</strong> a red tide event, (5)cyst populations in residence may become G. breve blooms, (6) pollution is not thecatalyst for a Florida red tide, (7) G. breve blooms are probably annualoccurrences <strong>and</strong> (8) monitoring programs can predict major Florida red tides.Methods: None per se; methodological comments appear in "Results <strong>and</strong> Discussion."QA/QC: N/A (Descriptive)Contact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K. A. <strong>and</strong> Joyce, E. A., Jr.Date: 1973Title: Educational Series No. 17: Florida <strong>Red</strong> <strong>Tide</strong>sPages: 26 pp.Publisher: State <strong>of</strong> Florida Department <strong>of</strong> Natural Resources, St. PetersburgKey words: Florida, red tide, Gymnodinium breveSummary: Gymnodinium breve, according to the authors, was very likely responsible fornumerous recorded fish kills in Florida as early as 1844, but was not recognized<strong>and</strong> identified as the causative agent until 1948. Stating that red tides are naturaloccurrences, the authors conclude that there is a lack <strong>of</strong> means to control red tideswith a concomitant warning that such means might harm the environment even ifavailable.Methods: N/A (Review paper)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K. A. <strong>and</strong> Vargo, G. A.153


Date: 1988Title/Ch.: 15. Marine din<strong>of</strong>lagellate blooms: dynamics <strong>and</strong> impactsBook: Algae <strong>and</strong> Human AffairsEditor: C. A. Lembi <strong>and</strong> J. R. Waal<strong>and</strong>Pages: 373-401Publisher: Cambridge University Press, New York, NYKey words: din<strong>of</strong>lagellate, bloom, microalgae, Pyrrhophyta, red tide, paralytic shellfishpoisoning, neurotoxic shellfish poisoning, diarrhetic shellfish poisoning,phytoplanktonSummary: A general review <strong>of</strong> the mechanisms <strong>and</strong> effects <strong>of</strong> marine din<strong>of</strong>lagellate blooms,this paper recommends a multidisciplinary approach to identifying the physicalchemicalforces <strong>of</strong> <strong>and</strong> biological responses to blooms. Hydrographic effects alterdin<strong>of</strong>lagellate behavior <strong>and</strong> location; variable light conditions can beaccommodated by the organism’s ability to modify photosynthetic pigment quality<strong>and</strong> quantity. Din<strong>of</strong>lagellates can migrate <strong>and</strong> store nutrients; certain speciesproduce toxins, <strong>and</strong> some bioluminesce. Reports <strong>of</strong> toxic blooms (red tides) areon the increase, including occurrences in areas previously unaffected; such bloomscan restructure marine communities via selective mortality, threatening publichealth, commercial fishing <strong>and</strong> shellfish industries. Outside <strong>of</strong> negative economicimpacts <strong>and</strong> health concerns, red tides, being natural occurrences, may notnecessarily cause long-term negative ecological impacts.Methods: N/A (Review paper)QA/QC: N/AContact: Karen A. SteidingerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K. A., <strong>and</strong> Williams, J.Date: 1964Title: Gymnodinium breve DavisSource: Florida Board <strong>of</strong> Conservation Leaflet Series: Plankton, Vol. 1, No. 1A(Supplement, 2 pp.)Key words: Gymnodinium breve, din<strong>of</strong>lagellate, red tide, fish mortality, cyst formation,encystment, chromatophores, nucleus, FloridaSummary: The principal author published a monograph in the same leaflet series in early 1964to which this is a supplement providing drawings, photomicrographs <strong>and</strong>documented observations <strong>of</strong> the red tide din<strong>of</strong>lagellate Gymnodinium breve. The63 cells collected <strong>and</strong> observed from a Florida outbreak causing fish mortality inmid-1964 differed from previous published observations, especially with respect tonucleus position, ventral concavity/dorsal convexity, the overhanging apicalprocess <strong>and</strong> the length/breadth <strong>of</strong> the body. Also described was the process <strong>of</strong>encystment.Methods: None (description)QA/QC: N/AContact: Karen A. Steidinger154


Source Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Steidinger, K., Babcock, C., Mahmoudi, B., Tomas, C. <strong>and</strong> Truby, E.Date: 1989Title/Ch.: Conservative taxonomic characters in toxic din<strong>of</strong>lagellate species identificationBook: <strong>Red</strong> tides: biology, environmental science, <strong>and</strong> toxicology. Proceedings <strong>of</strong> theFirst International Symposium on <strong>Red</strong> <strong>Tide</strong>s held November 10-14, 1987, inTakamatsu, Kagawa Prefecture, Japan.Editor: T. Okaichi, D. M. Anderson <strong>and</strong> T. NemotoPages: 285-288Publisher: Elsevier Science Publishing Company, Inc., New YorkKey words: numerical taxonomy, toxic, din<strong>of</strong>lagellate, species identification, Gymnodinium,Gyrodinium, Ptychodiscus, Prorocentrum, optical pattern recognitionSummary: Proper identification is <strong>of</strong> concern to areas experiencing din<strong>of</strong>lagellate blooms,especially those with no historical record <strong>of</strong> nor data on life history, biochemistryor growth <strong>of</strong> the bloom species. All such information is necessary for scientists toestimate negative impacts <strong>and</strong> establish monitoring programs. The taxonomy <strong>of</strong>toxic din<strong>of</strong>lagellates is unfortunately still at the morphospecies level, <strong>and</strong>identification <strong>of</strong> conservative morphological characters/descriptors is still essential.This paper <strong>of</strong>fers some apparently uniform, objective criteria to identify toxicdin<strong>of</strong>lagellate species.Methods: See "Experimental."QA/QC: None per se; see "Experimental."Contact: Karen SteidingerSource Inst.: Bureau <strong>of</strong> Marine Research, Florida Department <strong>of</strong> Natural Resources, St.Petersburg, Fl 33701 USAAuthor: Stockwell, D. A., Buskey, E. J. <strong>and</strong> Whitledge, T. E.Date: 1993Title/Ch.: Studies on conditions conducive to the development <strong>and</strong> maintenance <strong>of</strong> apersistent “brown tide” in Laguna Madre, TexasBook: Toxic Phytoplankton Blooms in the SeaEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 693-698Publisher: Elsevier Science Publishers B. V.Key words: brown tide, Laguna Madre, Texas, Baffin Bay, bloom, chrysophyte, Aureococcusanophagefferens, Pelagoccus subviridis, dimethyl sulfide, microzooplanktonSummary: A 4-5 µm diameter organism similar to the Type III, aberrant group <strong>of</strong>chrysophytes appeared as a bloom from 6/90 <strong>and</strong> persisted for more than 18months thereafter in the hypersaline waters <strong>of</strong> Baffin Bay <strong>and</strong> the Upper LagunaMadre <strong>of</strong> South Texas. Cell densities reached the order <strong>of</strong> 10 9 cells/l withchlorophyll a concentrations as much as 70 µg/l. The organism is capable <strong>of</strong>producing large amounts <strong>of</strong> dimethyl sulfide (DMS) <strong>and</strong>, while at maximum bloomin 7/90, was linked to a great reduction in microzooplankton grazing rates. The155


authors suggest the coincidence <strong>of</strong> regional drought, local hypersalinity <strong>and</strong> theuncoupling <strong>of</strong> water column processes from the benthos led to the bloom.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Dean A. StockwellSource Inst.: Marine Science Institute, The University <strong>of</strong> Texas at Austin, P. O. Box 1267, PortAransas, TX 78373-1267 USAAuthor: Taylor, F. J. R.Date: 1990Title/Ch.: <strong>Red</strong> tides, brown tides <strong>and</strong> other harmful algal blooms: the view into the 1990’sBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 527-533Publisher: ElsevierKey words:Summary:red tide, brown tide, algae, bloom, phytoplankton,A concise overview <strong>of</strong> the proceedings <strong>of</strong> the Fourth International Conference onToxic Marine Phytoplankton, this summary paper identifies what the authorconsidered the major outst<strong>and</strong>ing problems remaining (ca. 1990) in the study <strong>of</strong>toxic marine phytoplankton. Foremost for the purpose <strong>of</strong> this annotation are theelusive goal <strong>of</strong> accurate bloom prediction in most regions, the apparent increase infrequency <strong>of</strong> harmful blooms <strong>and</strong> the possibility <strong>of</strong> artificial dispersal (e.g., in theballast water <strong>of</strong> ships) <strong>of</strong> bloom species from formerly restricted ranges <strong>of</strong>occurrence to distant sites.Methods: N/AQA/QC: N/AContact: F. J. R. TaylorSource Inst.: Departments <strong>of</strong> Oceanography <strong>and</strong> Botany, Univeristy <strong>of</strong> British Columbia,Vancouver, B. C., Canada V6T 1W5Author: Tester, P. A.Date: 1992Title/Chap.: Section VI: Phytoplankton distributionReport: Report on Investigation <strong>of</strong> 1990 Gulf <strong>of</strong> Mexico Bottlenose Dolphin Str<strong>and</strong>ingsEditor: L. J. HansenPages: 44-47; Table 1; Figures 1-4; Appendix VPublisher: National Oceanic <strong>and</strong> Atmospheric Administration, National Marine FisheriesServiceKey words: phytoplankton, Gymnodinium breve, din<strong>of</strong>lagellate, Gulf <strong>of</strong> Mexico, red tide,bloom, brevetoxin, marine mammals, Galveston Bay, Mississippi deltaSummary: Of the 123 phytoplankton samples collected in the primary study area (betweenGalveston Bay <strong>and</strong> the Mississippi delta) in March <strong>of</strong> 1990, 80% contained G.breve cells. Upon detailed examination <strong>of</strong> the 70 samples taken in the upper half<strong>of</strong> the water column, 94% revealed the presence <strong>of</strong> G. breve cells, <strong>and</strong> 65%contained more than 50 cells/liter. Bloom concentrations are considered as at least156


5000 cells/liter, so the cell counts from this study can be categorized as "normalbackground levels," but they were consistently greater than counts <strong>of</strong> samples fromsimilar areas or from the primary study area later in the summer <strong>of</strong> 1990. (At thattime, another toxic din<strong>of</strong>lagellate, Gonyaulax [now Alex<strong>and</strong>rium] monilata wasfound near the Mississippi delta in elevated concentrations.)Methods: See "Methods."QA/QC: None per se; see "Methods."Contact: Patricia A. TesterSource Inst.: Southeast Fisheries Science Center (NOAA/NMFS), Beaufort Laboratory,Beaufort, NC 28516Author: Tester, P. A. <strong>and</strong> Fowler, P. K.Date: 1990Title/Ch.: Brevetoxin contamination <strong>of</strong> Mercenaria mercenaria <strong>and</strong> Crassostrea virginica: amanagement issueBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 499-503Publisher: ElsevierKey words: brevetoxin, Mercenaria mercenaria, Crassostrea virginica, red tide, Ptychodiscusbrevis, bloom, North Carolina, shellfish, clamSummary: The authors examine factors affecting the toxicity <strong>of</strong> Mercenaria mercenaria <strong>and</strong>Crassostrea virginica during <strong>and</strong> after a P. brevis [aka Gymnodinium breve]bloom in the field <strong>and</strong> suggest changes in the American Public Health Associationguidelines concerning what is an acceptable amount <strong>of</strong> neurotoxic shellfishpoisoning (NSP) toxin in shellfish harvested for human consumption. They alsonote that P. brevis requires salinities in excess <strong>of</strong> 24 ppt.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Patricia A. TesterSource Inst.: National Marine Fisheries Service, NOAA, Beaufort, NC 28516 USAAuthor: Tester, P. A., Geesey, M. A. <strong>and</strong> Vukovich, F. M.Date: 1993Title/Ch.: Gymnodinium breve <strong>and</strong> global warming: what are the possibilities?Book: Toxic Phytoplankton Blooms in the Sea, Proceedings <strong>of</strong> the Fifth InternationalConference on Toxic Marine PhytoplanktonEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 67-72Publisher: Elsevier Science Publishers B. V.Key words: Gymnodinium breve, global warming, Gulf Stream, North Carolina, red tide,bloom, South Atlantic Bight, Florida <strong>Current</strong>, Gulf <strong>of</strong> MexicoSummary: An unusual red tide bloom in North Carolina in 1987-88 may have resulted fromtransport <strong>of</strong> a seed population from the west Florida shelf into the South AtlanticBight (SAB) by the Florida <strong>Current</strong>-Gulf Stream System. Continuous transport <strong>of</strong>157


this kind has been implicated by water samples collected by NOAA vessels in thenorthern Gulf <strong>of</strong> Mexico (GOM) <strong>and</strong> the SAB, for G. breve appears to becontinously distributed from the GOM throughout the SAB regardless <strong>of</strong> season.Twelve incursions <strong>of</strong> the warmer Gulf Stream into the shelf waters <strong>of</strong> the SABwere associated with detectable levels <strong>of</strong> G. breve cells during the winter <strong>of</strong> 1990-91. When SAB shelf water temperatures increase in summer, Gulf Streamincursions are more difficult to detect <strong>and</strong> the growth <strong>of</strong> G. breve populations isbetter supported. Variations in the transport <strong>and</strong> distribution <strong>of</strong> G. breve may beuseful for assessing global climate change <strong>and</strong> its effects on the water temperature<strong>and</strong> circulation patterns between the GOM <strong>and</strong> the SAB. The authors add that oneliter was the minimum volume necessary for detection <strong>of</strong> G. breve cells in theSAB, at least ten times the quantity typically examined with the st<strong>and</strong>ard Utermöhlsettling chamber methodology.Methods: See "Introduction."QA/QC: None per se; see "Introduction."Contact: Patricia A. TesterSource Inst.: National Marine Fisheries Service--NOAA, Southeast Fisheries Science Center,Beaufort Laboratory, Beaufort, NC 28516 USAAuthor: Tester, P. A., Stumpf, R. P., Vukovich, F. M., Fowler, P. K., <strong>and</strong> Turner, J. T.Date: 1991Title: An expatriate red tide bloom: transport, distribution, <strong>and</strong> persistenceJournal: Limnol. Oceanogr. 36(5):1053-1061Key words: red tide, bloom, transport, distribution, din<strong>of</strong>lagellate, Gymnodinium breve, NorthCarolina, Gulf <strong>of</strong> Mexico Loop <strong>Current</strong>, Florida <strong>Current</strong>, Gulf StreamSummary: A Gymnodinium breve bloom occurred in North Carolina nearshore waters inNovember <strong>of</strong> 1987, the first to occur north <strong>of</strong> Florida due to G. breve. Theauthors propose that the Gulf Stream Loop <strong>Current</strong> entrained a portion <strong>of</strong> a G.breve bloom in progress <strong>of</strong>f the southwest Florida coast, then transported it, viathe Florida <strong>Current</strong>-Gulf Stream system, around the Florida peninsula <strong>and</strong> northover 800 km <strong>and</strong> 22-54 days to North Carolina. Thirty days after the southwestFlorida bloom, satellite sea-surface temperature images confirmed movement <strong>of</strong> aparcel <strong>of</strong> Gulf Stream water onto the shelf between Cape Hatteras <strong>and</strong> CapeLookout <strong>and</strong> its residence there for >19 days, a likely source <strong>of</strong> the G. breve cells.Once inshore, windspeed <strong>and</strong> direction largely controlled its distribution.Methods: None per se (note).QA/QC: N/AContact: Patricia A. TesterSource Inst.: National Marine Fisheries Service, NOAA, Southeast Fisheries Center, BeaufortLaboratory, Beaufort, NC 28516 USAAuthor: Texas Parks <strong>and</strong> Wildlife DepartmentDate: 1986Title: Commission Agenda Item (Briefing Session): <strong>Red</strong> <strong>Tide</strong>Pages: 3 pp.158


Key words: red tide, bloom, algae, din<strong>of</strong>lagellate, Ptychodiscus brevis, neurotoxin, Gulf <strong>of</strong>Mexico, aerial surveillance, fish mortality, aerosol toxinSummary: This report highlights the impact on marine fishes by the red tide din<strong>of</strong>lagellatePtychodiscus brevis during the extensive bloom that began during Labor Dayweekend <strong>of</strong> 1986. The red tide moved steadily southward from the first fish killnoted <strong>of</strong>fshore <strong>of</strong> Galveston, Texas, <strong>and</strong> by November had reached Mexicanbeaches at Tampico <strong>and</strong> the Bay <strong>of</strong> Campeche. Causative factors were unclear, forthe organism's preference for salinities greater than 28 ppt <strong>and</strong> temperaturesexceeding 60 o F is commonly fulfilled in late summer along Texas shores, yetwithout blooms. Total fish mortality was estimated at more than 7.5 millionindividuals, mainly menhaden <strong>and</strong> mullet, but including pinfish, red drum, spottedseatrout, black drum, southern flounder <strong>and</strong> Atlantic croaker. Oyster harvestingwas prohibited, <strong>and</strong> numerous beaches were closed due to dead fish <strong>and</strong> aerosoltoxin.Methods: N/A (Briefing report)QA/QC: N/AContact: Larry McEachronSource Inst.: Texas Parks & Wildlife Department, Coastal Fisheries Division, 702 NavigationCircle, Rockport, TX 78382Author: Tracey, G. A.Date: 1988Title: Feeding reduction, reproductive failure, <strong>and</strong> mortality in Mytilus edulis during the1985 ‘brown tide’ in Narragansett Bay, Rhode Isl<strong>and</strong>Journal: Mar. Ecol. Prog. Ser. 50:73-81Key words: feeding, reproduction, mortality, Mytilus edulis, brown tide, Narragansett Bay,Rhode Isl<strong>and</strong>, bloom, chrysophyte, Mercenaria mercenaria,Summary: The blue mussel, Mytilus edulis, suffered reduced clearance rates, reproductivefailure <strong>and</strong> high mortality in response to a dense, novel brown tide bloom (10 6cells/ml) in Narragansett Bay, RI in the summer <strong>of</strong> 1985. Using naturalparticulates from Narragansett Bay in the clearance rate experiments with the bluemussel, the authors replicated the experiment <strong>and</strong> saw similar inhibition in the hardclam, Mercenaria mercenaria. Deleterious experimental effects appeared whenparticles exceeded 5.0 x 10 5 particles/ml, whereas bay concentrations ranged from9 x 10 5 to 15 x 10 5 particles/ml. Blue mussel mortality in the bay varied from 30-100% with concomitant reproductive failure, neither <strong>of</strong> which could be attributedto temperature, salinity or dissolved oxygen concentrations.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Gregory A. TraceySource Inst.: Science Applications International Corporation, Marine Services Branch, USEnvironmental Protection Agency, Environmental Research Laboratory,Narragansett, Rhode Isl<strong>and</strong> 92882 USAAuthor: Tracey, G. A., Johnson, P. W., Steele, R. W., Hargraves, P. E. <strong>and</strong> Sieburth, J.McN.159


Date: 1988Title: A shift in photosynthetic picoplankton composition <strong>and</strong> its effect on bivalvemollusc nutrition: the 1985 “brown tide” in Narragansett Bay, Rhode Isl<strong>and</strong>Journal: Journal <strong>of</strong> Shellfish Research 7(4):671-675Key words: photosynthesis, picoplankton, bivalve, mollusc, nutrition, brown tide, NarragansettBay, Synechococcus, bloom, Mytilus edulisSummary: Upon analysis by epifluorescence <strong>and</strong> transmission electron microscopy, watersamples from the brown tide bloom <strong>of</strong> 1985 in Narragansett Bay, Rhode Isl<strong>and</strong>,revealed 10-fold greater concentrations <strong>of</strong> bacteria (10 7 cells/ml) during the peak<strong>of</strong> the bloom. A 1.5-2.0 µm chrysophyte composed 95% <strong>of</strong> the totalphytoplankton by abundance at about 10 6 cells/ml. The photosyntheticcyanobacterium Synechococcus was 10-fold lower in abundance than the usual 5 x105 cells/ml. The brown tide organism reduced feeding in the mussel Mytilusedulis, but clearance rates were optimal for comparable densities <strong>of</strong> a similar-sizedstrain <strong>of</strong> Synechococcus. The mussel’s nutrition <strong>and</strong> growth may thus be affectedby the species composition <strong>of</strong> the picoplankton.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Gregory A. TraceySource Inst.: Science Applications International Corporation, Marine Services Branch, c/o U. S.Environmental Protection Agency, South Ferry Rd., Narragansett, RI 02882USAAuthor: Tracey, G. A., Steele, R. L., Gatzke, J., Phelps, D. K., Nuzzi, R., Waters, M. <strong>and</strong>Anderson, D. M.Date: 1989Title/Ch.: Testing <strong>and</strong> application <strong>of</strong> biomonitoring methods for assessing environmentaleffects <strong>of</strong> noxious algal bloomsBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 557-574Publisher: Springer-VerlagKey words: biomonitoring, bloom, Aureococcus anophagefferens, Peconic Bay, Long Isl<strong>and</strong>,New York, brown tide, mussel, Mytilus edulus, Minutocellus polymorphusSummary: The authors evaluated biomonitoring methods in waters with a history <strong>of</strong> noxiousalgal blooms, in this case the Peconic Bays system <strong>of</strong> Long Isl<strong>and</strong>, New York,where a mild bloom occurred in June-September, 1988. Objectives included (1)the effect on bivalve nutrition by the characteristics <strong>of</strong> algae <strong>and</strong> other suspendedparticulates <strong>and</strong> (2) the effect on bivalve growth <strong>and</strong> physiology by environmentalconditions in different bay locations. <strong>Red</strong>uced feeding <strong>and</strong> slower growth formussels in Peconic Bay were attributed in the main to the deleterious influence <strong>of</strong>Aureococcus anophagefferens, with possible added negative effects due to thenanoplanktonic diatom, Minutocellus polymorphus.Methods: See “Methods.”QA/QC: None per se; see “Methods.”160


Contact: Gregory A. TraceySource Inst.: Science Applications International Corporation, c/o U. S. EnvironmentalProtection Agency, Environmental Research Laboratory-Narragansett,Narragansett, Rhode Isl<strong>and</strong> 02882 USAAuthor: Tracey, G., Steele, R. <strong>and</strong> Wright, L.Date: 1990Title/Ch.: Variable toxicity <strong>of</strong> the brown tide organism, Aureococcus anophagefferens, inrelation to environmental conditions for growthBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 128-131Publisher: ElsevierKey words: toxicity, brown tide, Aureococcus anophagefferens, growth, blue mussel, MytilusedulisSummary: Varying light, temperature <strong>and</strong> nutrient regimes in A. anophagefferens cultures,then checking the feeding response <strong>of</strong> M. edulis on the brown tide organism, theauthors found that Aureococcus toxicity in late-exponential phase growth waslower in low light as opposed to high light, greater at higher versus lowertemperatures <strong>and</strong> increased in high rather than low nutrient concentrations. Thesame kinds <strong>of</strong> conditions may affect the alga’s toxicity in situ.Methods: See “Materials <strong>and</strong> Methods.”QA/QC: None per se; see “Materials <strong>and</strong> Methods.”Contact: Gregory TraceySource Inst.: Science Applications International Corporation, c/o U. S. EnvironmentalProtection Agency, Narragansett Bay, Rhode Isl<strong>and</strong> 02882 USAAuthor: Trainer, V. L. <strong>and</strong> Baden, D. G.Date: 1990Title/Ch.: Enzyme immunoassay <strong>of</strong> brevetoxinsBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 430-435Publisher: ElsevierKey words: enzyme, immunoassay, brevetoxin, Ptychodiscus brevis (aka Gymnodinium breve),red tide, din<strong>of</strong>lagellate, ELISA, enzyme-linked immunosorbent assay, FloridaSummary: At the time <strong>of</strong> the study, the only assays available to detect the presence <strong>of</strong>brevetoxins in seafood were impractical for field use. The authors analyzed theusefulness <strong>of</strong> two different brevetoxin-enzyme conjugates via enzyme-linkedimmunosorbent assays (ELISAs), the results <strong>of</strong> which are visual <strong>and</strong> could possiblybe read in the field. The two toxin-enzyme conjugates can be st<strong>and</strong>ardized to testfor brevetoxin in unknown field samples, though their stability must first beassessed <strong>and</strong> the overall anti-brevetoxin ELISA optimized.Methods: See “Methods.”161


QA/QC: None per se; see “Methods.”Contact: Vera L. TrainerSource Inst.: Department <strong>of</strong> Biochemistry <strong>and</strong> Molecular Biology, University <strong>of</strong> Miami, Miami,FL 33149 USAAuthor: Trebatoski, B.Date: 1988Title: Observations on the 1986-1987 Texas <strong>Red</strong> <strong>Tide</strong> (Ptychodiscus brevis), Report 88-02Pages: 48 pp.Publisher: Texas Water CommissionKey words:Summary:red tide, Ptychodiscus brevis, Texas, din<strong>of</strong>lagellate, Gulf <strong>of</strong> MexicoThis study documents the spread <strong>and</strong> impact <strong>of</strong> a bloom <strong>of</strong> the toxic din<strong>of</strong>lagellatePtychodiscus brevis [now Gymnodinium breve] that affected Gulf <strong>of</strong> Mexicocoastal waters from Galveston to Port Isabel, Texas, <strong>and</strong> on into Mexico fromAugust 1986 through January 1987. Plankton <strong>and</strong> water chemistry samples alongwith boat <strong>and</strong> aerial surveys were the prinicipal means <strong>of</strong> judging the effects <strong>of</strong> thered tide, which produced massive fish mortality, irritating aerosols <strong>and</strong>contamination <strong>of</strong> shellfish beds <strong>and</strong> discouraged tourism <strong>and</strong> seafood consumption.The resulting large-scale bioperturbation actually may have been a positiveinfluence on the environment.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Bob TrebatoskiSource Inst.: Texas Water Commission, P. O. Box 13087, Austin, TX 78711 USAAuthor: Vargo, G. A., Carder, K. L., Gregg, W., Shanley, E., Neil, C., Steidinger, K. A.<strong>and</strong> Haddad, K. D.Date: 1987Title: The potential contribution <strong>of</strong> primary production by red tides to the west Floridashelf ecosystemJournal: Limnol. Oceanogr. 32(3):762-767Key words: primary production, red tide, west Florida shelf, din<strong>of</strong>lagellate, bloom,Ptychodiscus brevis, CZCS imagerySummary: Field <strong>and</strong> laboratory estimations <strong>of</strong> <strong>and</strong> some data for daily <strong>and</strong> monthlyproduction rates <strong>and</strong> theoretical annual carbon input for the red tide din<strong>of</strong>lagellatePtychodiscus brevis [now Gymnodinium breve] revealed that bloom rates were 2-5 times greater than published values or non-bloom rates. Annually, red tideblooms <strong>of</strong> P. brevis could be responsible for a significant proportion <strong>of</strong> annualproduction in the water column <strong>of</strong> the west Florida shelf.Methods: None per se; see text. (Research note)QA/QC: N/AContact: Gabriel A. VargoSource Inst.: Department <strong>of</strong> Marine Science, University <strong>of</strong> South Florida, 140 7th Ave. South,St. Petersburg, FL 33701 USA162


Author: Vargo, G. A. <strong>and</strong> Howard-Shamblott, D.Date: 1990Title/Ch.: Phosphorus dynamics in Ptychodiscus brevis: cell phosphorus, uptake <strong>and</strong> growthrequirementsBook: Toxic Marine Phytoplankton, Proceedings <strong>of</strong> the Fourth International Conferenceon Toxic Marine PhytoplanktonEditor: E. Granéli, B. Sundström, L. Edler <strong>and</strong> D. M. AndersonPages: 324-329Publisher: ElsevierKey words: phosphorus, Ptychodiscus brevis, bloom, West Florida Shelf, red tide,din<strong>of</strong>lagellate, Gulf <strong>of</strong> MexicoSummary: This study concerns nutrient-growth interactions for P. brevis (renamedGymnodinium breve) <strong>and</strong> phosphorus. Blooms <strong>of</strong> this species are initiated 18 to74 km <strong>of</strong>fshore <strong>of</strong> Florida’s west coast in highly oligotrophic waters. Bloompersistence should be related to the ability to efficiently extract, stock <strong>and</strong> usenutrients for growth <strong>and</strong> maintenance, but P. brevis does not possess a highphosphorus storage capacity, though it can apparently store phosphorus in twodifferent cellular storage pools, the significance to growth <strong>of</strong> which has yet to bedetermined. Phosphorus uptake rates for P. brevis, however, are more than anorder <strong>of</strong> magnitude above those <strong>of</strong> other din<strong>of</strong>lagellates. The uptake rates, halfsaturationconstant for growth <strong>and</strong> long generation times <strong>of</strong> P. brevis enable it tobloom in the presence <strong>of</strong> relatively excess phosphorus <strong>and</strong> maintain populations in<strong>of</strong>fshore waters with low phosphorus concentrations.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Gabriel A. VargoSource Inst.: Department <strong>of</strong> Marine Science, University <strong>of</strong> South Florida, 140 Seventh AvenueSouth, St. Petersburg, Florida 33701 USAAuthor: Vieira, M. E. C. <strong>and</strong> Chant, R.Date: 1993Title: On the contribution <strong>of</strong> subtidal volume fluxes to algal blooms in Long Isl<strong>and</strong>estuariesJournal: Estuarine, Coastal <strong>and</strong> Shelf Science 36:15-29Key words: algae, bloom, Long Isl<strong>and</strong>, estuary, subtidal volume flux, picoplankton,Aureococcus anophagefferens, sea level fluctuation, flushing time, wind stressSummary: Tidal data taken from 1980-88 for oceanic locations near Nantucket, MA,Montauk, NY <strong>and</strong> Atlantic City, NJ <strong>and</strong> for four Long Isl<strong>and</strong> estuaries revealedthat sea level fluctuations due to subtidal frequencies were greater than tidalfrequencies <strong>and</strong> were seasonal in their magnitude. When the spring1985 browntide bloom (Aureococcus anophagefferens) occurred along with a drought in LongIsl<strong>and</strong> bays, subtidal variance was at an absolute minimum, subtidal flushing timeswere correspondingly slow <strong>and</strong> freshwater influx was reduced. As a result,inorganic nutrients may have been available for longer than usual, supporting theexpansive bloom <strong>of</strong> A. anophagefferens.Methods: See “Data Base.”163


QA/QC: None per se; see “Data Base” <strong>and</strong> “Analysis <strong>and</strong> Results.”Contact: Mario E. C. VieiraáSource Inst.: Oceanography Department, US Naval Academy, Annapolis, MD 21402-5026USAAuthor: Walker, L. M.Date: 1982Title: Evidence for a sexual cycle in the Florida red tide din<strong>of</strong>lagellate, Ptychodiscusbrevis (= Gymnodinium breve)Journal: Transactions <strong>of</strong> the American Microscopical Society 101(3):287-293Key words: sexual cycle, Florida, red tide, din<strong>of</strong>lagellate, Gymnodinium breve, gametes,planozygotes, cystSummary: Non-clonal stock cultures provided evidence <strong>of</strong> isogamous, haploid gametes <strong>and</strong>diploid planozygotes in the sexual cycle <strong>of</strong> Gymnodinium breve. Though nohypnozygotes were ever positively identified in the cultures, apparenthypnozygotic cysts may have formed in one preliminary experiment <strong>and</strong> were alsorecovered from field samples.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Linda M. WalkerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Avenue South East, St. Petersburg, Florida 33701-5095 USAAuthor: Walker, L. M., <strong>and</strong> Steidinger, K. A.Date: 1979Title: Sexual reproduction in the toxic din<strong>of</strong>lagellate Gonyaulax monilataJournal: Journal <strong>of</strong> Phycology 15:312-315Key words: reproduction, din<strong>of</strong>lagellate, Gonyaulax monilata, sexual cycle, isogametes,planozygote, hypnozygote, archeopyle, red tide, seed bedSummary: The authors observed the induction <strong>of</strong> the sexual cycle <strong>of</strong> Gonyaulax [nowAlex<strong>and</strong>rium] monilata in nonclonal laboratory cultures within one week afterinoculating asexual cells into nitrogen-deficient media. Asexual cells can producetwo armored, motile, isogamous gametes. Two gametes fuse to form a largedouble-flagellated planozygote that encysts as a hypnozygote with a three-layeredcyst wall 1-3 weeks later. The cell that emerges from the cyst divides twice t<strong>of</strong>orm a four-celled chain within three days. The authors report that cysts <strong>of</strong> G.monilata were found in Tampa Bay sediments, were excysted <strong>and</strong> produced chains<strong>of</strong> cells identical to those studied in the nonclonal laboratory cultures.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: Linda M. WalkerSource Inst.: Florida Department <strong>of</strong> Natural Resources, Marine Research Laboratory, 100Eighth Avenue S.E., St. Petersburg, FL 33701Author: Wall, D.Date: 1975164


Title/Chap.: Taxonomy <strong>and</strong> cysts <strong>of</strong> red-tide din<strong>of</strong>lagellatesBook: Proceedings <strong>of</strong> The First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 249-255Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: taxonomy, cyst, red tide, din<strong>of</strong>lagellate, life cycle, theca, Gonyaulax, speciescomplexSummary: Many din<strong>of</strong>lagellate species, both estuarine <strong>and</strong> neritic, produce a cyst at some part<strong>of</strong> their life cycle. Cysts are composed <strong>of</strong> a variety <strong>of</strong> materials, can providetaxonomic information <strong>and</strong> have indicated that the Genus Gonyaulax is geneticallyheterogeneous with seven species complexes. The G. tamarensis complexincludes red-tide species such as Gonyaulax [now Alex<strong>and</strong>rium] monilata <strong>and</strong>possesses a distinctive thecal morphotype common to all species in the complex<strong>and</strong> some others. Taxonomic revision is needed for the Genus Gonyaulax <strong>and</strong>others, <strong>and</strong> cyst cycles should be employed to identify species complexes.Methods: None (review)QA/QC: N/AContact: David WallSource Inst.: Woods Hole Oceanographic Institution, Woods Hole, Massachusetts, USAAuthor: Ward, J. E. <strong>and</strong> Targett, N. M.Date: 1989Title/Ch.: Are metabolites from the brown tide alga, Aureococcus anophagefferens,deleterious to mussel feeding behavior?Book: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 543-556Publisher: Springer-VerlagKey words: metabolites, brown tide, Aureococcus anophagefferens, mussel, bloom, Isochrysisgalbana, Heterosigma akashiwo, Dunaliella tertiolectaSummary: Given the adverse effects <strong>of</strong> A. anophagefferens on bivalves, the lack <strong>of</strong>phytoplankton diversity during the brown tide blooms <strong>and</strong> the high densities <strong>of</strong> thebrown tide alga, the authors hypothesized a metabolite detrimental to bivalvefeeding behavior. They found, however, no evidence <strong>of</strong> a negative effect by A.anophagefferens on mussel filtration rates, particle selection or valve movements,even at densities <strong>of</strong> 10 5 -10 6 cells/ml. Their findings do not rule out toxic effects <strong>of</strong>metabolite exposure for days or weeks, rapidly degrading metabolites or harmfulepicellular constituents.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: J. Evan WardSource Inst.: University <strong>of</strong> Delaware, College <strong>of</strong> Marine Studies, Lewes, DE 19958 USAAuthor: Wardle, W. J., Ray, S. M. <strong>and</strong> Aldrich, A. S.165


Date: 1975Title/Ch.: Mortality <strong>of</strong> marine organisms associated with <strong>of</strong>fshore summer blooms <strong>of</strong> thetoxic din<strong>of</strong>lagellate Gonyaulax monilata Howell at Galveston, TexasBook: Proceedings <strong>of</strong> the First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 257-263Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: mortality, bloom, din<strong>of</strong>lagellate, Gonyaulax monilata, Galveston, Texas, snail,hermit crab, brittle star, red tideSummary: The authors document two red tides <strong>of</strong> the toxic din<strong>of</strong>lagellate Gonyaulaxmonilata [preferred name Alex<strong>and</strong>rium monilata] that occurred in waters nearGalveston, Texas in the summers <strong>of</strong> 1971 <strong>and</strong> 1972, with recorded maximum celldensities <strong>of</strong> 1.2 x 10 6 cells/L <strong>and</strong> 1.88 x 10 6 cells/L, respectively. Twenty-ninespecies <strong>of</strong> cnidarians, annelids, molluscs, crustaceans, echinoderms <strong>and</strong> fishes, allsedentary or slow-moving, were represented among the organisms that perishedfrom the toxin. Monitoring in the summers <strong>of</strong> 1973 <strong>and</strong> 1974 did not detect thepresence <strong>of</strong> G. monilata, perhaps due to lower salinities (


Date: 1993Title/Ch.: The nutrient <strong>and</strong> hydrographic conditions prevailing in Laguna Madre, Texasbefore <strong>and</strong> during a brown tide bloomBook: Toxic Phytoplankton Blooms in the SeaEditor: T. J. Smayda <strong>and</strong> Y. ShimizuPages: 711-716Publisher: Elsevier Science Publishers B. V.Key words: nutrients, Laguna Madre, Texas, brown tide, bloom, pigment, chrysophyte, BaffinBay, dissolved inorganic nitrogen, ammoniumSummary: An unidentified aberrant chrysophyte bloomed widely for more than 18 months inthe hypersaline Laguna Madre. The bloom followed dissolved inorganic nitrogen(DIN) concentrations <strong>of</strong> almost 20 µmol/l, especially in Baffin Bay, withammonium composing 60-95% <strong>of</strong> the DIN. Baffin Bay, subject to annual salinityvariations <strong>of</strong> as much as 65 ppt, also had the greatest salinity <strong>of</strong> the north to southgradient in the Laguna Madre, ranging from 40-60 ppt during the bloom.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Terry E. WhitledgeSource Inst.: Marine Science Institute, The Univerity <strong>of</strong> Texas at Austin, P. O. Box 1267, PortAransas, TX 78373 USAAuthor: Whitledge, T. E. <strong>and</strong> Pulich, W. M., Jr.Date: 1991Title: Report <strong>of</strong> <strong>Brown</strong> <strong>Tide</strong> Symposium <strong>and</strong> Workshop, 15-16 July 1991Pages: 44 pp.Publisher: University <strong>of</strong> Texas Marine Science Institute Technical Report TR/91-002Key words: brown tide, bloom, seagrass, grazing, nutrients, hydrography, light, benthos,zooplankton, ichthyoplanktonSummary: This report consolidates the abstracts <strong>of</strong> all presentations given at thesymposium/workshop <strong>and</strong> summarizes the results in the forms <strong>of</strong> (1)comparisons/contrasts between characteristics <strong>of</strong> geographically isolatedoccurrences <strong>of</strong> brown tide <strong>and</strong> (2) recommendations for future research.Methods: N/AQA/QC: N/AContact: Terry E. WhitledgeSource Inst.: Marine Science Institute, The University <strong>of</strong> Texas, P. O. Box 1267, PortAransas, TX 78373 USAAuthor: Williams, J., <strong>and</strong> Ingle, R. M.Date: 1972Title: Ecological Notes on Gonyaulax monilata (Dinophyceae) Blooms Along the WestCoast <strong>of</strong> FloridaSource: Florida Department <strong>of</strong> Natural Resources Leaflet Series: Phytoplankton, Part 1(Din<strong>of</strong>lagellates), No. 5Key words: Gonyaulax monilata, din<strong>of</strong>lagellate, Florida, fish kill, bloom, Gymnodinium breve167


Summary: The first bloom <strong>of</strong> the toxic din<strong>of</strong>lagellate Gonyaulax monilata [now Alex<strong>and</strong>riummonilata] <strong>of</strong>f the west coast <strong>of</strong> Florida is the subject <strong>of</strong> this note. The organismcaused fish kills <strong>and</strong>, also for the first time anywhere, was detected blooming in<strong>of</strong>fshore waters, not just estuarine or nearshore environments, in the late summer<strong>of</strong> 1966. A Gymnodinium breve bloom followed the G. monilata bloom inOctober.Methods: NoneQA/QC: N/A (Report)Contact: Jean WilliamsSource Inst.: Florida Department <strong>of</strong> Natural Resources, Bureau <strong>of</strong> Marine Research, St.Petersburg, FL 33701 USAAuthor: Wilson, W. B.Date: 1965Title: The suitability <strong>of</strong> sea-water for the survival <strong>and</strong> growth <strong>of</strong> Gymnodinium breveDavis; <strong>and</strong> some effects <strong>of</strong> phosphorus <strong>and</strong> nitrogen on its growthJournal: Florida Board <strong>of</strong> Conservation Pr<strong>of</strong>essional Papers Series No. 7:1-42Key words: Gymnodinium breve, din<strong>of</strong>lagellate, bloom, toxin, growth, phosphorus, nitrogen,chelator, iron, glycerophosphateSummary: The toxic marine din<strong>of</strong>lagellate G. breve was used as a bioassay organism to testthe suitability <strong>of</strong> various seawater samples from the West Florida coast <strong>and</strong>Galveston Isl<strong>and</strong> to support its survival <strong>and</strong> growth. Of the few seawater samplesthat were suitable, experiments determined which substances used for G. breveculture media would promote growth in those samples. Survival <strong>and</strong> growth inmost samples were improved by addition <strong>of</strong> the chelator EDTA or a combination<strong>of</strong> EDTA <strong>and</strong> iron. EDTA, iron <strong>and</strong> an inorganic mixture with nitrogen <strong>and</strong>phosphorus improved growth more than any other additive combination in 1964seawater samples, but inorganic nutrients alone were not generally effective. HighG. breve concentrations, however, were not seen in response to most additives.Higher phosphorus concentrations in artificial media supported relatively greatercell densities, but populations with less than 6.0 µg P/liter did not continue togrow. Glycerophosphate as the sole phosphorus source was acceptable; nitratenitritewas not required, but some form <strong>of</strong> organic nitrogen was essential.Methods: See "Sample Procedure <strong>and</strong> Assay Methods."QA/QC: None per se; see "Sample Procedure <strong>and</strong> Assay Methods."Contact: William B. WilsonSource Inst.: Texas A & M Marine Laboratory, Galveston, Texas, USAAuthor: Wilson, W. B.Date: 1967Title: Forms <strong>of</strong> the din<strong>of</strong>lagellate Gymnodinium breve Davis in culturesJournal: Contributions in Marine Science 12:120-134Key words: din<strong>of</strong>lagellate, Gymnodinium breve, culture, morphology, encystment stage, cyst,cell division, reproduction, bloom, Gulf <strong>of</strong> MexicoSummary: Cultured cells <strong>of</strong> the marine din<strong>of</strong>lagellate Gymnodinium breve are described <strong>and</strong>discussed with respect to the most common encystment stage, general features <strong>of</strong>168


cell division, other possible reproductive stages <strong>and</strong> forms resulting from variedphysical factors <strong>and</strong> reproductive stages. The author includes numerousphotographs <strong>and</strong> some line drawings.Methods: None per se (descriptive paper).QA/QC: N/AContact: William B. WilsonSource Inst.: Marine Laboratory, Texas A & M University, Galveston, Texas, USAAuthor: Wilson, W. B. <strong>and</strong> Ray, S. M.Date: 1956Title: The occurrence <strong>of</strong> Gymnodinium brevis in the western Gulf <strong>of</strong> MexicoJournal: Ecology 87(2):388Key words: Gymnodinium brevis, Gulf <strong>of</strong> Mexico, fish mortality, Port Isabel, Texas, RioGr<strong>and</strong>e RiverSummary: September <strong>of</strong> 1955 saw a widely documented fish kill apparently due to the redtide din<strong>of</strong>lagellate Gymnodinium brevis [now preferably known as Gymnodiniumbreve] that littered the Gulf coast from a point 17 miles north <strong>of</strong> Port Isabel to themouth <strong>of</strong> the Rio Gr<strong>and</strong>e in Texas as well as a 120-mile stretch in the Mexicanstate <strong>of</strong> Tamaulipas. Highest reported densities were found 36 miles south <strong>of</strong> theRio Gr<strong>and</strong>e River mouth at no less than 22,000 cells/ml.Methods: N/A (Report)QA/QC: N/AContact: W. B. WilsonSource Inst.: Gulf Fishery Investigations, U. S. Fish <strong>and</strong> Wildlife Service, Galveston, TX, USAAuthor: Wilson, W. B., Ray, S. M., <strong>and</strong> Aldrich, D. V.Date: 1975Title/Chap.: Gymnodinium breve: population growth <strong>and</strong> development <strong>of</strong> toxicity in culturesBook: Proceedings <strong>of</strong> the First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 127-141Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: Gymnodinium breve, growth, toxicity, culture, din<strong>of</strong>lagellate, toxicity assay,mosquito fish, Gambusia affinisSummary: Eight 12-liter cultures <strong>of</strong> Gymnodinium breve were maintained for five months inthis study <strong>of</strong> population growth <strong>and</strong> toxicity. Each culture received an estimatedinoculation <strong>of</strong> 50,000-60,000 cells per liter, but population levels inexplicablydiminished rapidly to 3000-9000 cells per liter shortly after inoculation. Allpopulations, however, reached at least 5 million cells per liter by five weeks (acalculated population growth rate <strong>of</strong> 0.3/day). Populations varied thereafter,ranging from a half million to 24 million per liter after five months. Cultures wereconsidered replicates, but no cause was apparent for the density variance. Toxicity(determined with the mosquito fish, Gambusia affinis) <strong>and</strong> bacterial populationlevels varied directly with G. breve cell concentration; the ratio <strong>of</strong> large to small169


cells was greater in populations with increasing as opposed to decreasingpopulations.Methods: See "Materials <strong>and</strong> Methods."QA/QC: None per se; see "Materials <strong>and</strong> Methods."Contact: William B. WilsonSource Inst.: Department <strong>of</strong> Marine Science, Texas A & M University, Galveston, Texas, USAAuthor: Wyatt, T.Date: 1975Title/Chap.: The limitations <strong>of</strong> physical models for red tidesBook: Proceedings <strong>of</strong> the First International Conference on Toxic Din<strong>of</strong>lagellateBloomsEditor: V. R. LoCiceroPages: 81-93Publisher: The Massachusetts Science <strong>and</strong> Technology Foundation, Wakefield, MAKey words: physical model, red tide, circulation pattern, vertical stability, advection, diffusion,plankton distibutionSummary: Circulation patterns seen in both shallow water <strong>and</strong> deep water above a pycnocline<strong>and</strong> processes that increase vertical stability in the water column are the two majoroceanographic processes commonly used to address the influence <strong>of</strong> the physicalenvironment on red tide concentrations. The two may not be distinct <strong>and</strong> mayoccur concurrently, but making such a distinction may aid underst<strong>and</strong>ing <strong>of</strong> smallscaleoceanographic features that influence plankton population distribution. Thepaper reviews the two processes (advective <strong>and</strong> diffusive) regarding theirusefulness in underst<strong>and</strong>ing the mechnisms affecting red tides apart from those thatare ecological.Methods: None (review)QA/QC: N/AContact: T. WyattSource Inst.: Ministry <strong>of</strong> Agriculture, Fisheries <strong>and</strong> Food, Fisheries Laboratory, Lowest<strong>of</strong>t, NZAuthor: Yentsch, C. S., Phinney, D. A. <strong>and</strong> Shapiro, L. P.Date: 1989Title/Ch.: Absorption <strong>and</strong> fluorescent characteristics <strong>of</strong> the brown tide chrysophyte: its roleon light reduction in coastal marine environmentsBook: Novel Phytoplankton Blooms: Causes <strong>and</strong> Impacts <strong>of</strong> Recurrent <strong>Brown</strong> <strong>Tide</strong>s <strong>and</strong>Other Unusual Blooms, Coastal <strong>and</strong> Estuarine Studies 35Editor: E. M. Cosper, V. M. Bricelj <strong>and</strong> E. J. CarpenterPages: 77-83Publisher: Springer-VerlagKey words: brown tide, chrysophyte, phytoplankton, bloom, absorption spectra, fluorescencespectra, Skeletonema sp., Aureococcus sp.Summary: This study sought to determine if the brown tide absorption <strong>and</strong> chlorophyllfluorescence spectra differed from typical coastal phytoplankton. Comparisonswere made between the diatom Skeletonema sp. <strong>and</strong> the brown tide organism,Aureococcus sp. The brown tide, unlike the diatom, indeed selectively absorbed170


lue light in the same region as that for theoretically maximal photosynthesis in seagrasses (450 nm), greatly changing the subsurface light quality. Selectiveabsorbance <strong>of</strong> blue light is a typical characteristic <strong>of</strong> oceanic phytoplankton. Usingthese results, the authors speculate that remote sensing could successfully indicatepresence <strong>of</strong> brown tide-like organisms in coastal waters via fluorescence excitationspectra <strong>and</strong>/or colorimetry.Methods: See “Methods.”QA/QC: None per se; see “Methods.”Contact: Charles S. YentschSource Inst.: Bigelow Laboratory for Ocean Sciences, West Boothbay Harbor, Maine 04575USA171


XVI. AcknowledgmentsLIST OF CONTRIBUTORS TO CCBNEP REPORT1. Aransas National Wildlife RefugeTom Stehn, BiologistPO Box 100,Austwell, TX 779502. Cameron County Marine Extension ServiceTony Reisinger, Marine Extension AgentCCMES, 650 East Hwy 77San Benito, TX 785863. Florida Dept. <strong>of</strong> Natural ResourcesKaren SteidingerFlorida Department <strong>of</strong> Natural Resources,Institute <strong>of</strong> Marine Research100 8th Ave. SESt. Petersburg, FL 337014. Louisiana State University / LATEX B programSteve Murray, Program Mgr.Coastal Studies Inst., LSUBaton Rouge, LA 708035. Minerals Management ServiceWalter JohnsonDepartment <strong>of</strong> the Interior, MMS381 Elden St. MS 644Herndon, VA 220706. National Biological ServiceChris OnufSouthern Science CenterCorpus Christi Field StationTexas A&M University-Corpus Christi6300 Ocean Dr.,Corpus Christi, TX 784127. NOAA: National Marine Fisheries ServiceRoger ZimmermanGalveston Laboratory NMFS4700 Ave. UGalveston, TX 77551172


Pat TesterNMFS Southeast Fisheries Science Center., Beaufort Lab.101 Pivers Isl<strong>and</strong> Rd.Beaufort, NC 28516-97228. Newspapers:The <strong>Brown</strong>sville HeraldBox 351, <strong>Brown</strong>sville, TX 78522-0351The Port Isabel/South Padre PressPort Isabel, TX9. Researchers:William Wardle,Department <strong>of</strong> Marine Biology,Texas A & M UniversityP. O. Box 1675Galveston, TX 77553Henry Hildebr<strong>and</strong> (retired.)Quay Dortch,LUMCON8124 Highway 56Chauvin, LA 70344Daniel G. Baden.University <strong>of</strong> Miami Rosenstiel School <strong>of</strong> Marine <strong>and</strong> Atmospheric Science,Division <strong>of</strong> Biology <strong>and</strong> Living ResourcesMiami, FL 3314910. Texas A&M University: College StationDr. Frank J. KellyTexas A&M UniversityGeochemical & Environmental Research Group833 Graham Rd.College Station, TX 7784511. Texas A&M University-Corpus ChristiCenter for Coastal Studies <strong>and</strong> Conrad Blucher InstituteCBI: Nick KrausCCS: Roy LehmanCBI: Chris FawcetteTexas A&M University-Corpus Christi6300 Ocean Dr.Corpus Christi, TX 78412173


12. Texas Department <strong>of</strong> Health & Human ServicesMike Ordner <strong>and</strong> Gary HeidemanTexas Dept. <strong>of</strong> Health, Division <strong>of</strong> Seafood Safety1100 West 49th Street,Austin, TX 7840113. Texas Natural Resource Conservation CommissionJim Bowman4410 Dillon Lane,Corpus Christi, TX 78415Richard Kiesling, Kelly Hutchinson, Patrick RoquesTX NRCC, Water Quality OfficePO Box 13087, Capitol StationAustin, TX 78711-308714. Texas Parks <strong>and</strong> Wildlife Dept.Larry McEachron, Science DirectorTX Pks. &Wldlf, Coastal Fisheries DivisionRockport, TX 78382Todd EngelingTX Pks. & Wldlf. Dept.Marine Development Ctr.4300 Waldron Rd.,Corpus Christi, TX 7841815. Texas Water Development BoardDr. David BrockTWDB, Environmental Section611 S. Congress,Austin, TX 7870416. University <strong>of</strong> Texas Marine Science InstituteKenneth DuntonJoan HoltScott HoltPaul MontagnaDean StockwellTerry Whitledge750 Channel View DrivePort Aransas, Texas 78373174

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!