Collapse - Ronald Feldman Gallery
Transcription
Collapse - Ronald Feldman Gallery
Appendix by Ballengée et al. (B. Ballengée with T.Gardner, J. Rudloe, B. Schiering and P. Warny) 2012 26,162 preserved specimens representing 370+ species or >2.4% of the known 15,419 species of the Gulf of Mexico Date: 5 May 2012 Ronald Feldman Fine Arts, New York, NY Table of contents SPECIES LIST BY TROPHIC LEVEL………………..……………………………..….2-8 SPECIES LIST BY PHYLOGENETIC ORDER…………...………………………….9-15 SELECT PUBLICATIONS Specimen list by trophic level Level 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 Reference Species Col.1.1 Flower coral species Col.1.2 Organ pipe coral Col.1.3 Thin finger coral Col.1.4 Elkhorn coral Col.1.5 Frilly lettuce coral Col.1.6 Scarlet coral species Col.1.7 Boulder coral species Col.1.8 Brain coral species Col.1.9 Cauliflower coral species Col.1.10 Table coral Col.1.11 Branch coral Col.1.12 Brownstem coral Col.1.13 Lace coral species Col.1.14 Branching coral species Col.1.15 Blue ridge coral Col.1.16 Large-cupped fungal coral Col.1.17 Rooster tail conch Col.1.18 West Indian fighting conch Col.1.19 Florida fighting conch Col.1.20 Incongruous ark Col.1.21 Southern quahog Col.1.22 Stout tagelus Col.1.23 Common razor clams Col.1.24 Propeller clam Col.1.25 Soft shell clams Col.1.26 Neapolitan triton Col.1.27 Northern moon snails Col.1.28 Dwarf razor clams Col.1.29 Hawkwing conch Col.1.30 Shark eye sea snail (didyma) Col.1.31 Graceful fig sea snail Col.1.32 Florida crown conch Col.1.33 Lettered olive sea snail Col.1.34 Florida calico scallop Col.1.35 White scallop Col.1.36 Nucleus scallop Col.1.37 Fairyland snail sea snail Col.1.38 Pillow stinking sponge Col.1.39 Florida hermit-crab sponge Col.1.40 Stinking vase sponge Col.1.41 Half-naked pen clam Col.1.42 Atlantic plate limpet Col.1.43 Yellow sea fan species Col.1.44 Green sea fan species Col.1.45 Purple sea fan species Col.1.46 Spiny oyster Col.1.47 Red sea lettuce Col.1.48 Tar balls Scientific Name Type Quantity Origin Mussa spp. Partial specimens 5 Pet-trade Tubipora musica* Partial specimens 15 Pet-trade Porites furcate Partial specimens 8 Pet-trade Acropora palmate Partial specimens 1 Pet-trade Lobophytum crassum Partial specimens 3 Pet-trade Dichocoenia spp. Partial specimens 1 Pet-trade Monstastrea spp. Partial specimens 10 Pet-trade Diplora spp. Partial specimens 1 Pet-trade Pocillopora spp. Partial specimens 3 Pet-trade Acropora hyacinthus Partial specimens 23 Pet-trade Acropora florida Partial specimens 1 Pet-trade Pocillopora verrucosa Partial specimens 22 Pet-trade Family stylasteridae Partial specimens 3 Pet-trade Madracris spp. Partial specimens 9 Pet-trade Heliopora coerula Partial specimens 4 Pet-trade Scolymia lacera Partial specimens 3 Pet-trade Lobatus gallus Shells only 1 Wholesale supplier Strombus pugilis Shells only 4 Wholesale supplier Strombus alatus Shells only 11 Wholesale supplier Anadara brasiliana Whole specimens 24 Markets Mercenaria campechiensis notata Whole specimens 63 Markets Tagelus plebeius Whole specimens 10 Markets Ensis directus Whole specimens 9 Markets Cyrtodaria siliqua Whole specimens 20 Markets Mya arenaria Whole specimens 24 Markets Cymatium parthenopeum Whole specimens 4 Markets Lunatia heros Whole specimens 6 Markets Ensis megistus Whole specimens 93 Markets Strombus raninus Shells only 6 Wholesale supplier Polinices duplicatus Shells only 51 Wholesale supplier Ficus gracilis Shells only 2 Wholesale supplier Melongena corona Shells only 6 Wholesale supplier Oliva sayana Shells only 40 Wholesale supplier Argopecten gibbus Shells only 5 Wholesale supplier Argopecten irradians Shells only 11 Wholesale supplier Argopecten nucleus Shells only 28 Wholesale supplier Achatina achatina Shells only 10 Wholesale supplier Ircinia strobilina Partial specimens 1 Wholesale supplier Pseudospongosorites suberitoides Whole specimens 34 Wholesale supplier Ircinia campana Whole specimens 1 Wholesale supplier Atrina seminude Shells only 3 Ecological surveys Lottia testudinalis Shells only 19 Wholesale supplier Gongonia spp. Whole specimens 1 Wholesale supplier Gongonia spp. Whole specimens 1 Wholesale supplier Gongonia spp. Whole specimens 3 Wholesale supplier Spondylus americanus Partial shells 12 Pet-trade Halymenia floresia Partial specimens 25 Markets N/A N/A 10 Ecological surveys Year 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 1999 1999 1999 1999 1999 1999 2000 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2005 2012 2012 2012 2012 2012 2012 2011 1 Col.1.49 Mixed cone sea snail species Conus spp. Shells only 21 Wholesale supplier 2012 1 1 Col.1.50 Col.1.51 Crassispira spp. Littoraria spp. Shells only Shells only 839 753 Wholesale supplier Wholesale supplier 2012 2012 1 Col.1.52 Haliotis pourtalesii** Shells only 51 Markets 2012 1 1 1 Col.1.53 Col.1.54 Col.1.55 Turrid snail species Yellow sea snails Pourtale's abalone (pearlized) Knobbed triton snail Nerite snail species Eastern auger Cymatium muricinum Nerita spp. Terebra dislocata Shells only Whole specimens Shells only 146 613 17 Markets Markets Wholesale supplier 2012 2012 2012 1 Col.1.56 White donax or Coquina clam Donax variabilis Shells only 556 Wholesale supplier 2012 2 Specimen list by trophic level Level 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 Reference Species Col.1.57 Rockweed Col.1.58 Rough file clam Col.1.59 Southern quahog Crown conch (smooth and Col.1.60 horned forms) Col.1.61 Skate species egg cases Col.1.62 Pleated sea squirt Deep sea hexactinellid Col.1.63 “glass” sponge species Col.1.64 Five-holed keyhole urchin Col.1.65 Black sun coral Col.1.66 Cat's paw coral Col.1.67 Finger coral Col.1.68 Plate coral species Col.1.69 Purple barnacle Col.1.70 Gulf fire coral Col.1.71 Spiral Babylon sea snail Col.1.72 Endive murex sea snail Col.1.73 Millipede conch Col.1.74 Horned starfish Col.1.75 Brown spiny sea star Col.1.76 Inflated sea biscuit Col.1.77 Flat sea biscuit Knobbed whelk egg case Col.1.78 (string) Col.1.79 Knobbed whelk Col.1.80 Channeled whelk Col.1.81 Common or Waved whelk Goose neck barnacles Col.1.82 species and Shipworm species with wood Col.1.83 Giant sea roach Col.1.84 Cannonball jellyfish Atlantic horseshoe crab Col.1.85 (eggs) Col.1.86 Colorful sea rod Col.1.87 Brown rock sea cucumber Col.1.88 Furry sea cucumber Col.1.89 Five-toothed sea cucumber Col.1.90 Moon jellyfish Col.1.91 Black encrusting tunicate Col.1.92 Mangrove tunicate Col.1.93 Sea pork species Primitive deep sea stalked Col.1.94 barnacle species Col.1.95 Portuguese man of war Col.1.96 Mottled sea hare Col.1.97 Atlantic black sea hare Col.1.98 Hauff’s alcyonidium Col.1.99 Lettuce bryozoan Col.1.100 Gulfweed Sea lettuce macro-algae Col.1.101 species Col.1.102 Pincushion urchin species Col.1.103 Spotted linckia starfish Col.1.104 Brown sargassum weed Col.1.105 Paper scallop Col.1.106 Gulf mole crabs Col.1.107 Thin sea lettuce Col.1.108 Soft spaghettiweed Scientific Name Fucus vesiculosus Lima scabra Mecenaria campechiensis Type Quantity Origin Partial specimens 12 Ecological surveys Shells only 1 Wholesale supplier Partial Shells 6 Pet-trade Year 2012 2012 2012 Melongena corona Shells only 90 Wholesale supplier 2012 Raja spp. Styela plicata Whole specimens Whole specimens 81 11 Ecological surveys Markets 2001 1999 Scolymastra spp. Partial specimens 1 Wholesale supplier 2012 Mellita quinquiesperforata Tubastraea micrantha Pocillopora palifera* Acropora humilis* Acropora spp. Conchylepes conchylepes* Millepora alcicornis Babylonia spirata* Murex endiva* Lambis millepeda** Protoreaster nodosus** Echinaster spinulosus Clypeaster rosaceus Clypeaster subdepressus Whole specimens Partial specimens Partial specimens Partial specimens Partial specimens Partial specimens Partial specimens Whole specimens Shells only Shells only Whole specimens Whole specimens Whole specimens Whole specimens 9 3 14 7 2 5 3 84 1 6 27 49 2 24 Wholesale supplier Pet-trade Pet-trade Pet-trade Pet-trade Pet-trade Pet-trade Markets Pet-trade Wholesale supplier Wholesale supplier Wholesale supplier Wholesale supplier Wholesale supplier 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 2012 Busycon carica Eggs 159 Ecological surveys 2001 Busycon carica Busycotypus canaliculatus Buccinum undatum* Whole specimens Whole specimens Whole specimens 6 10 28 Markets Markets Markets 2012 2012 2012 Bankia spp. & Lepas spp. Whole specimens >40 Ecological surveys 2012 Bathynomus giganteous Stomolophus meleagris Whole specimens Whole specimens 1 2 Ecological surveys Ecological surveys 2012 2012 Limulus polyphemus Eggs 241 Ecological surveys 1999 Diodogorgia nodulifera Holothuria glaberrima Astichopus multifudus Actinopyga agassizii Aurelia aurito Botrylloides nigrum Ecteinascidia turbinata Amaroucium spp Whole specimens Whole specimens Whole specimens Whole specimens Partial specimens Colonies Partial colonies Whole specimens 1 2 2 2 3 4 12 1 Wholesale supplier Markets Markets Ecological surveys Markets Markets Markets Ecological surveys 2012 2001 1999 2012 1999 2001 2000 2012 Litoscalpellum spp. Colonies 1 Ecological surveys 2012 Physalia physalis Aplysia brasiliana Aplysia morio Alcyonidium hauffi Thalamoporella gothica Sargassum spp. Whole specimens Whole specimens Whole specimens Partial specimens Partial specimens Partial specimens 1 1 1 1 1 12 Ecological surveys Ecological surveys Ecological surveys Ecological surveys Ecological surveys Ecological surveys 2012 2012 2012 2000 2000 2000 Ulvaria spp. Partial specimens 67 Markets 2012 Lytechinus spp. Linckia multiflora** Sargassum natans Euvola papyracea Emerita brasiliensis Ulva lactuca Liagora farinose Whole specimens Whole specimens Partial specimens Whole specimens Whole specimens Partial specimens Partial specimens 12 3 5 7 116 47 4 Markets Wholesale supplier Markets Markets Ecological surveys Ecological surveys Ecological surveys 2012 2012 2000 2012 2001 2001 2012 3 Specimen list by trophic level Level Reference 1 Col.1.109 1 1 1 1 Col.1.110 Col.1.111 Col.1.112 Col.1.113 1 Col.1.114 1 1 1 1 1 1 1 1 1 1 Col.1.115 Col.1.116 Col.1.117 Col.1.118 Col.1.119 Col.1.120 Col.1.121 Col.1.122 Col.1.123 Col.1.124 1 Col.1.125 1 1 1 1 1 Col.1.126 Col.1.127 Col.1.128 Col.1.129 Col.1.130 1 Col.1.131 1 Col.1.132 1 Col.1.133 1 1 Col.1.134 Col.1.135 1 Col.1.136 1 1 Col.1.137 Col.1.138 1 Col.1.139 1 1 Col.1.140 Col.1.141 1 Col.1.142 1 Col.1.143 1 1 1 1 Col.1.144 Col.1.145 Col.1.146 Col.1.147 1 Col.1.148 1 1 1 1 1 1 Col.1.149 Col.1.150 Col.1.151 Col.1.152 Col.1.153 Col.1.154 1 Col.1.155 1 Col.1.156 Species Florida halemenia red macroalgae Intricate brown macro-algae Green encrusting tunicate Feather duster worm species Line nemerteans worm White shrimp with oil staining with one specimen with assemtrical eyes Common Shore Shrimp Ivory barnacles Northern brown shrimp Northern white shrimp Southern pink shrimp Malaysian prawn Pacific giant clam species Silver lip conch Tulip sea snail Asian shore crab Southern and American horse mussel species Variable coquina clams Bruised nassa Caribbean carpet anemone Curlycue anemone Cassiopea jellyfish Atlantic horseshoe crab (early instar) Basket sea-star Flamescallop or Red file shell clam Brittle Star species Serpent star species Sponge spider crab species & Fire sponges Green reef crab Ridged slipper lobster Deep-water slipper lobster species Yellowline arrow crab West Indian sea urchins Flat-clawed hermit crab inside Northern moon snail shell Acadian hermit crab inside Neapolitan triton snail shell Brown spiny sea star Mud fiddler crab Purple marsh crab Mangrove land crab Mixed southern gulf periwinkle species Caribbean mud fiddler crab Sea nettle jellyfish species Pink vase sponge Boring sponge species Red beard sponge Rope sponge species White shrimp with lesions and one specimen with missing eye "sOil" (marsh sediment mixed with Macondo oil and chemical dispersants) Scientific Name Type Quantity Origin Year Halymenia floridana Partial specimens 3 Ecological surveys 2012 Rosenvingeo intricate Symplegma viride Sabellastarte spp. Lineus spp. Partial specimens Whole specimens Partial specimens Whole specimens 4 54 5 1 Ecological surveys Ecological surveys Pet-trade Markets 2012 2012 2004 2001 Litopenaeus setiferus Whole specimens 2 Anonymous shrimper 2012 Palaemonetes vulgaris Chelonibia patula Farfantepenaeus aztecus Penaeus setiferus Farfantepenaeus duorarum Macrobrachium rosenbergii Tridacna spp.* Strombus lentiginosus* Turbonilla curta Hemigrapsus sanguineus Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Shells only Whole specimens Whole specimens Whole specimens 64 67 1 212 5 1 21 6 5 36 Markets Ecological surveys Markets Markets Ecological surveys Markets Pet-trade Wholesale supplier Wholesale supplier Ecological surveys 2001 2001 2001 2012 2012 2001 2012 2012 2012 2006 Modiolus spp. Whole specimens 56 Ecological surveys 2012 Donax variabilus Nassarius vibex Stichodactyla helianthus Bartholomea annulata Cotylorhiza tuberculata Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 58 31 1 1 1 Ecological surveys Ecological surveys Pet-trade Pet-trade Pet-trade 2012 2012 2012 2012 2012 Limulus polyphemus Whole specimens 1 Pet-trade 2012 Gorgonocephalus eucnemis Whole specimens 1 Pet-trade 2012 Lima scabra Whole specimens 1 Pet-trade 2012 Ophioderma spp. Hemipholis spp. Whole specimens Whole specimens 1 2 Pet-trade Pet-trade 2012 2012 Macrocoeloma spp. & TedaniaWhole ignis specimens 1 Pet-trade 2012 Mithrax sculptus Scyllarides nodifer Whole specimens Whole specimens 1 1 Pet-trade Ecological surveys 2012 2012 Scyllarides spp. Whole specimens 1 Ecological surveys 2012 Stenorhynchus seticornis Tripneustes ventricosus Whole specimens Whole specimens 1 3 Ecological surveys Pet-trade 2012 2012 Pagurus pollicaris inside Lunatia Whole heros specimens 1 Markets 2012 Pagurus acadianus inside Cymatium Whole specimens parthenopeum 1 Markets 2012 Echinaster spinulosus Uca pugnax Afrithelphusa monodosa Ucides cordatus Whole specimens Whole specimens Whole specimens Whole specimens Pet-trade Pet-trade Pet-trade Pet-trade 2012 2012 2005 2004 Littorina spp. Shells only 18,584 Wholesale supplier 2012 Uca rapax Chrysaora spp. Niphates digitalis Cliona spp. Microciona prolifera Aplysina spp. Whole specimens Whole specimens Whole specimens Partial specimens Partial specimens Partial specimens 65 3 1 8 1 2 Markets Ecological surveys Retail store Retail store Retail store Retail store 2012 2012 2012 2012 2012 2012 Whole specimens 2 Anonymous shrimper 2012 Partial specimens 1 Ecological surveys 2012 1 1 1 1 4 Specimen list by trophic level Level 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 Reference Species Col.2.1 Lady crab Col.2.2 Rock crab Col.2.3 Green crab Rugose or Purple swimming Col.2.4 crab Col.2.5 Common mantis shrimp Col.2.6 Hogchoker Col.2.7 Mediterranean sand lance Col.2.8 Rimspine searobin Col.2.9 Atlantic sturgeon (juvenile) Banded coral shrimp or Col.2.10 Cleaner shrimp-like decapod Col.2.11 Atlantic guitarfish Col.2.12 Smooth skate Col.2.13 Clear-nose skate Col.2.14 Lesser electric ray Col.2.15 Red goatfish Col.2.16 Hickory shad Col.2.17 Blueback herring Col.2.18 Balao halfbeak Col.2.19 Southern eagle ray (juvenile) Col.2.20 Underworld windowskate Col.2.21 Masked pufferfish Col.2.22 Regal angelfish Col.2.23 Six barred angel (juvenile) Col.2.24 Sargassum triggerfish Col.2.25 Gray triggerfish Col.2.26 Queen parrotfish Col.2.27 Sharpnose puffer Col.2.28 Sea lamprey Col.2.29 Thornbacked boxfish Col.2.30 Violet goby Col.2.31 White trevally Col.2.32 Pygmy filefish Col.2.33 Long-horned cowfish Col.2.34 Longsnout seahorse Col.2.35 Northern pipefish Col.2.36 Chain pipefish Col.2.37 Lined seahorse Col.2.38 Checkered puffer fish Col.2.39 Striped burrfish Col.2.40 Spotlight parrotfish Col.2.41 Redband parrotfish Col.2.42 Porcupinefish Col.2.43 Blunthead puffer Col.2.44 Opossum pipefish Col.2.45 Gulf pipefish Col.2.46 Sargassum pipefish Col.2.47 Atlantic Moonfish Col.2.48 Lookdown Col.2.49 Blue-striped Angelfish Col.2.50 Pearl wrasse Col.2.51 Yellow tang Col.2.52 Spotfin jawfish species Col.2.53 Striped killifish Col.2.54 Flathead mullet Col.2.55 Naked goby (female & male) Col.2.56 Tidewater silverside Scientific Name Ovalipes ocellatus Cancer irroratus Cancer maenas Type Quantity Origin Whole specimens 12 Markets Whole specimens 3 Markets Whole specimens 10 Markets Year 2005 2012 2005 Callinectes exasperates Whole specimens 1 Markets 2005 Squilla empusa Trinectes maculatus Gymnammodytes cicerelus* Peristedion thompsoni Acipenser oxyrinchus Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 1 2 82 2 1 Markets Markets Markets Ecological surveys Research institute Stenopus hispidus Whole specimens 1 Pet-trade 2012 Ecological surveys Ecological surveys Markets Ecological surveys Markets Markets Markets Markets Ecological surveys Ecological surveys Pet-trade Pet-trade Pet-trade Pet-trade Ecological surveys Pet-trade Pet-trade Markets Pet-trade Markets Markets Pet-trade Wholesale supplier Pet-trade Markets Markets Markets Ecological surveys Pet-trade Markets Markets Pet-trade Pet-trade Markets Ecological surveys Ecological surveys Markets Markets Pet-trade Pet-trade Pet-trade Pet-trade Markets Ecological surveys Ecological surveys Ecological surveys 2012 2012 2005 2012 2012 2012 2012 2012 2012 2012 2005 2005 2005 2005 2012 2005 2005 1999 2005 2012 2012 2005 1980 2005 2012 2012 1999 2012 1990 2012 2012 1990 2005 2000 2011 2002 2005 2012 2003 2004 2003 2004 2000 2011 2011 2011 Rhinobatos lentiginosus Whole specimens Malacoraja senta Whole specimens Raja eglanteria Whole specimens Narcine bancroftii Whole specimens Mullas auratus Whole specimens Alosa mediocris Whole specimens Alosa aestivalis Whole specimens Hemiramphus balao Whole specimens Myliobatis goodei Whole specimens Fenestraja plutonia Whole specimens Arothron diadematus* Whole specimens Pygoplites diacanthus** Whole specimens Pomacanthus sexstriatus** Whole specimens Xanthichthys ringens Whole specimens Balistes capriscus Whole specimens Scarus vetula Whole specimens Bamthigaster rostrata Whole specimens Petromyzon marinus Whole specimens Tetrosomus gibbosus** Whole specimens Gobioides broussoneti Whole specimens Pseudocaranx dentex Whole specimens Monacanthus setifer Whole specimens Lactoria cornuta Whole specimens Hippocampus reidi Whole specimens Synganathus fuscus* Whole specimens Synganathus louisanae Whole specimens Hippocampus erectus Whole specimens Sphoeroides testudineus Whole specimens Chilomycterus schoepfii Whole specimens Sparisoma viride Whole specimens Sparisoma aurofrenatum Whole specimens Diodon hystrix Whole specimens Sphoeroides pachygaster Whole specimens Microphis brachyurus lineatus Whole specimens Syngnathus scovelli Whole specimens Syngnathus pelagicus Whole specimens Selene setapinnis Whole specimens Selene vomer Whole specimens Chaetodontoplus septentrionalis** Whole specimens Anampses cuvier** Whole specimens Zebrasoma flavescens* Whole specimens Opistognathus spp. Whole specimens Fundulus majalis Whole specimens Mugil cephalus Whole specimens Gobiosoma bosci Whole specimens Menidia peninsulae Whole specimens 1 1 1 1 7 1 13 1 1 1 1 1 1 1 1 1 1 2 1 15 8 1 1 1 30 15 2 40 2 1 3 1 1 8 9 31 3 3 1 1 2 1 177 12 2 10 2005 2005 2005/2012 2012 2005 5 Specimen list by trophic level Level 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 Reference Species Col.2.57 Golden topminnow Col.2.58 Texas cichlid Col.2.59 Sailfin molly (female & male) Col.2.60 Rainwater killifish Dwarf livebearer or Least Col.2.61 killifish Col.2.62 Rough silverside Col.2.63 Green sunfish Col.2.64 High hat (juvenile) Col.2.65 Foureye butterfly fish Col.2.66 Dwarf seahorse Col.2.67 Pancake batfish Col.2.68 Wahoo (larvae) Col.2.69 Gulf killifish Col.2.70 Striped Mullet Col.2.71 Inland silverside Col.2.72 Atlantic silverside Col.2.73 Spanish sardine Col.2.74 Dusky pipefish Col.2.75 Gulf menhaden Col.2.76 Fantail mullet Col.2.77 Mummichog Col.2.78 Blackwing flyingfish Col.2.79 American gizzard shad Col.2.80 Butterfish Col.2.81 Harvestfish Col.2.82 Blackcheek tonguefish Col.2.83 Blue reef chromis damselfish Col.2.84 Creole wrasse Col.2.85 Sharksucker Col.2.86 Atlantic thread herring Col.2.87 Whitecheek surgeonfish Col.2.88 Cunner Col.2.89 Mountain mullet Col.2.90 Giant or Red hermit crab Col.2.91 Rosy lobsterette Red swamp crayfish or Col.2.92 Louisiana crawfish Common Australian yabby Col.2.93 crayfish Longear sunfish or Creek Col.2.94 perch Col.2.95 Yellowtail coris wrasse Col.2.96 Cownose ray Caribbean spiny lobster Col.2.97 (larvae) Col.2.98 Coral mithrax crab Col.2.99 Whitetail damselfish Col.2.100 Lined porcelain crab Bigclaw pistol or snapping Col.2.101 shrimp Col.2.102 Mosquitofish Armored catfish or Col.2.103 Plecostomus species Col.2.104 Bay anchovy Col.2.105 Birdmouth Wrasse Col.3.1 Freshwater drum Col.3.2 Permit Col.3.3 Florida pomano Col.3.4 Atlantic needlefish Col.3.5 Red lionfish Scientific Name Fundulus chrysotus Herichthys cyanoguttatus Poecilia latipinna Lucania parva Type Quantity Origin Whole specimens 1 Ecological surveys Whole specimens 1 Ecological surveys Whole specimens 2 Ecological surveys Whole specimens 5 Ecological surveys Year 2011 2011 2011 2011 Heterandria formosa Whole specimens 6 Ecological surveys 2011 Membras martinica Lepomis cyanellus Pareques acuminatus Chaetodon capistratus Hippocampus zosterae Halieutichthys aculeatus Acanthocybium solandri Fundulus grandis Mugil cephalus Menidia beryllina Menidia menidia Sardinella aurita Synganathus floridae Brevoortia patronus Mugil trichodon Fundalulus heteroclitus* Hirundichthys rondeletii Dorosoma cepedianum Peprilus triacanthus Peprilus paru Symphurus plagiusa Chromis cyanea Clepticus parrae Echeneis naucrates Opisthonema oglinum Acanthurus nigricans* Tautoglabrus adspersus Agonostomus moniticola Petrochirus diogenes Nephropsis rosea Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 4 3 1 2 2 2 61 8 4 4 816 7 3 15 5 147 3 3 14 7 5 2 1 1 1 1 5 29 1 1 Ecological surveys Ecological surveys Pet-trade Pet-trade Wholesale supplier Ecological surveys Markets Ecological surveys Markets Ecological surveys Markets Markets Ecological surveys Markets Markets Markets Markets Markets Markets Markets Markets Pet-trade Pet-trade Ecological surveys Markets Pet-trade Markets Markets Ecological surveys Ecological surveys 2011 2011 2012 2012 2012 2012 2000 2002 2012 2011 2002 2012 2001 2012 2012 2004 2004 2012 2012 2012 2012 2012 2012 2012 2012 2003 2005/2012 2004 2012 2012 Procambarus clarkia Whole specimens 1 Markets 2012 Cherax destructor** Whole specimens 1 Markets 2002 Lepomis megalotis Whole specimens 4 Markets 1999 Coris gaimard** Rhinoptera bonasus Whole specimens Partial specimens 1 1 Pet-trade Ecological surveys 2003 1996 Panulirus argus Whole specimens 1 Ecological surveys 1999 Mithrax coryphe Dascyllus aruanus* Petrolisthes galathinus Whole specimens Whole specimens Whole specimens 5 1 1 Ecological surveys Pet-trade Pet-trade 1999 2003 2012 Alpheus heterochaelis Whole specimens 1 Ecological surveys 2012 Gambusia affinis Whole specimens 70 Ecological surveys 2012 Hypostomus ssp. Whole specimens 1 Pet-trade 2012 Anchoa mitchilli Gomphosus caeruleus** Aplodinotus grunniens Trachinotus falcatus Trachinotus carolinus Strongylura marina Pterois volitans Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 20 1 1 1 2 1 1 Markets Pet-trade Ecological surveys Markets Markets Markets Pet-trade 2000 2004 2011 2012 2012 2005 2012 6 Specimen list by trophic level Level 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 Reference Species Col.3.6 Atlantic blue-clawed crab Col.3.7 Three-toed amphiuma Col.3.8 White perch Royal dorade or Col.3.9 Mediterranean gilt-head seabream Col.3.10 Rainbow smelt Col.3.11 Ocellated frogfish Col.3.12 Winter flounder Col.3.13 Summer flounder Col.3.14 Gulf flounder Col.3.15 Southern flounder Col.3.16 Windowpane flounder Col.3.17 Spotted hake Col.3.18 Spotted pig grunt Col.3.19 White grunt Col.3.20 Sailor’s choice grunt Col.3.21 Puddingwife Col.3.22 Bigeye Col.3.23 Short bigeye Col.3.24 Striped searobin Col.3.25 Southern kingfish Col.3.26 Northern kingfish Col.3.27 Atlantic croaker 3 Col.3.28 Deep-sea lantern fish species Diaphus or Lampanyctus spp. Whole specimens 3 3 3 3 3 3 3 3 3 3 3 3 Col.3.29 Col.3.30 Col.3.31 Col.3.32 Col.3.33 Col.3.34 Col.3.35 Col.3.36 Col.3.37 Col.3.38 Col.3.39 Col.3.40 3 Col.3.41 3 3 Col.3.42 Col.3.43 3 Col.3.44 3 3 3 3 3 3 3 Col.3.45 Col.3.46 Col.3.47 Col.3.48 Col.3.49 Col.3.50 Col.3.51 3 Col.3.52 3 3 3 3 4 4 4 Col.3.53 Col.3.54 Col.3.55 Col.3.56 Col.4.1 Col.4.2 Col.4.3 Pinfish Sea bream Yellowfin morjarra Atlantic saury Atlantic chub mackerel Mackerel scad Bigeye scad Fat sleeper Tautog or Blackfish Chinese softshell turtle American eel Shrimp eel Asian swamp or Synbranchidae eel species Common Atlantic octopus Atlantic Oval Squid Gulf deep-water octopus species Caribbean reef octopus Brownstripe octopus Caribbean reef scorpion fish Chestnut Moray eel Spanish grunt Northern sennet Deepreef scorpionfish Bonnethead shark or shovelhead Blackbelly rosefish Butter hamlet Clown leaflip soapfish Longfin squid St Pierre’s fish or John Dory Guachanche barracuda Yellow-tail snapper 3 Scientific Name Callinectes sapidus Amphiuma tridactylum Morone Americana* Type Quantity Origin Whole specimens 9 Markets Whole specimens 1 Pet-trade Whole specimens 2 Markets Year 2005 2000 2000 Sparus aurata* Whole specimens 4 Markets 2012 Osmerus mordax* Whole specimens Fowlerichthys ocellatus Whole specimens Pseudopleuronectes americanus* Whole specimens Paralichthys dentatus Whole specimens Paralichthys albigutta Whole specimens Paralichthys lethostigma Whole specimens Scophthalmus aquosus Whole specimens Urophycis regius Whole specimens Orthopristis chrysoptera Whole specimens Haemulon plumierrii Whole specimens Haemulon parra Whole specimens Halichoeres poeyi Whole specimens Priacanthus arentus Whole specimens Pristigenys alta Whole specimens Prionotus evolans Whole specimens Mentricirrhus americanus Whole specimens Mentricirrhus saxitilis Whole specimens Micropogonias undulates Whole specimens 5 1 2 1 1 1 1 3 1 1 1 1 1 1 1 4 1 1 Markets Ecological surveys Markets Markets Markets Markets Markets Markets Ecological surveys Ecological surveys Ecological surveys Markets Markets Markets Markets Markets Ecological surveys Markets 2012 2012 2005 2012 2002 2003 2012 2012 2012 2012 2012 2000 1999 2006 2012 2012 2004 2012 3 Markets 2012 Lagodon rhomboids Archosargus rhomboidalis Gerres cinereus Scomberesox saurus Scomber colias Decapterus macarellus Selar crumenophthalmus Dormitator maculatus Tautoga onitis Pelodiscus sinensis** Anguilla rostrata Ophichthus gomesii Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 2 1 1 4 2 3 3 1 1 1 3 1 Ecological surveys Ecological surveys Ecological surveys Markets Markets Markets Markets Markets Markets Markets Markets Ecological surveys 1999 1999 1999 2012 2012 2012 2012 2012 2012 2012 2012 2012 Monopterus spp. Whole specimens 3 Markets 2002/2012 Octopus vulgaris Sepioteuthis sepioidea Whole specimens Whole specimens 1 1 Markets Markets 2012 2000 Benthoctopus spp. Whole specimens 1 Ecological surveys 2012 Octopus briareus Whole specimens Octopus burryi Whole specimens Scorpaenodes caribbaeus Whole specimens Enchelycore carychroa Whole specimens Haemulon macrostomum Whole specimens Sphyraena boreali Whole specimens Scorpaenodes tredecimspinosus Whole specimens 1 2 1 1 1 1 12 Ecological surveys Markets Pet-trade Pet-trade Ecological surveys Ecological surveys Pet-trade 2012 2000 2012 2012 2012 1999 2012 Sphyrna tiburo Whole specimens 1 Ecological surveys 2012 Helicolenus dactylopterus Hypoplectrus unicolor Pogonoperca punctate** Loligo pealei Zeus faber Sphyraena guachancho Ocyurus chrysurus Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens 3 1 1 1 1 4 1 Ecological surveys Pet-trade Pet-trade Markets Markets Markets Markets 2012 2004 2004 2012 2012 2000 2012 7 Specimen list by trophic level Level 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 5 5 5 5 5 5 5 5 5 5 5 5 5 6 6 6 6 6 Reference Species Col.4.4 Cottonwick Col.4.5 Conger eel Col.4.6 Gulf toadfish Col.4.7 Gray snapper Col.4.8 Red snapper Col.4.9 Freckled stargazer Col.4.10 Vermilion snapper Col.4.11 Barmundi or Asian seabass Col.4.12 Oyster toadfish Col.4.13 Bearded brotula Coney Grouper or Leopard Col.4.14 Hind Col.4.15 Spinycheek scorpionfish Col.4.16 Inshore lizardfish Deep-sea dagger tooth or Col.4.17 Pharaoh fish Col.4.18 Pilot fish Col.4.19 Black sea bass Chain catshark or Chain Col.4.20 dogfish Col.4.21 Caribbean ocellated moray Col.4.22 Hourglass moray eel Col.4.23 Bluefish Col.4.24 Great northern tilefish Col.4.25 Wrenchman Col.4.26 Deep-sea squid species Col.4.27 Mutton hamlet Col.4.28 Graysby Col.4.29 Red hind grouper Col.4.30 Redfish species Deep-water needlefish or Col.5.1 Houndfish Col.5.2 Blue runner Col.5.3 Crevalle jack Col.5.4 Blackfin goosefish Atlantic cutlassfish or Col.5.5 Ribbonfish Col.5.6 Burmese python Col.5.7 Greater amber jack Dusky smooth-hound or Col.5.8 Smooth dogfish shark Col.5.9 Atlantic Spanish mackerel Col.5.10 Yellowfin grouper Col.5.11 Yellowmouth grouper Col.5.12 Red grouper Col.5.13 Atlantic striped bass Col.6.1 Blacktip shark (juvenile) Col.6.2 Mako shark species Col.6.3 Atlantic wreckfish Col.6.4 Wahoo Mahi-mahi or Common Col.6.5 dolphinfish Scientific Name Haemulor melanurum Conger oceanicus Opsanus beta Lutjanus griseus Lutjanus campechanus Gnathagnus egreglus Rhomboplites aurorubens Lates calcarifer** Opsanus tau Brotula barbata Type Quantity Origin Whole specimens 3 Markets Whole specimens 2 Markets Whole specimens 1 Ecological surveys Whole specimens 2 Markets Whole specimens 2 Markets Whole specimens 1 Ecological surveys Whole specimens 3 Markets Whole specimens 1 Markets Whole specimens 1 Markets Whole specimens 1 Ecological surveys Year 2000 2000 2012 2000 2000 2012 2012 2012 2012 2012 Cephalopholis fulva Whole specimens 1 Markets 2012 Neomerinthe hemingway Synodus foetens Whole specimens Whole specimens 3 1 Ecological surveys Markets 2012 2000 Anotopterus pharao Whole specimens 3 Ecological surveys 2012 Naucrates doctor Centropristis striata Whole specimens Whole specimens 3 2 Markets Markets 2012 2012 Scyliorhinus rotifer Whole specimens 1 Ecological surveys 2012 Gymnothorax ocellatus Whole specimens Muraena clepsydra* Whole specimens Pomatomus saltatrix Whole specimens Lopholatilus chamaeleonticepsWhole specimens Pristipomoides aquilonaris Whole specimens llex spp. Whole specimens Alphestes afer Whole specimens Cephalopholis cruentata Whole specimens Epinephelus guttatus Whole specimens Sebastes spp. Whole specimens 1 1 6 1 1 1 1 1 1 4 Pet-trade Pet-trade Markets Markets Markets Ecological surveys Markets Markets Markets Markets 2012 2004 1999 2012 2000 2012 2000 2000 2012 2012 Tylosurus crocodilus Whole specimens 1 Ecological surveys 2012 Caranx crysos Caranx hippos Lophius gastrophysus Whole specimens Whole specimens Whole specimens 3 2 1 Markets Markets Markets 2012 2012 2012 Trichiurus lepturus Whole specimens 3 Markets 2012 Python molurus bivittatus Seriola dumerili Whole specimens Whole specimens 1 1 Pet-trade Markets 2012 2012 Mustelus canis Whole specimens 1 Markets 2000 Scomberomorus maculatus Mycteroperca venenosa Mycteroperca interstitialis Epinephelus morio Morone saxatilis Carcharhinus limbatus Isurus spp. Polyprion americanus Acanthocybium solandri Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Whole specimens Partial specimens Partial specimens Partial specimens 1 1 1 1 2 1 1 1 1 Markets Markets Markets Markets Markets Ecological surveys Ecological surveys Markets Markets Coryphaena hippurus Partial specimens 1 Markets TOTAL 2012 2012 2000 2012 2000 2012 Historic 2012 2012 2012 26,162 * non-native species ** non-native species anecdotally reported but not scientifically confirmed or species with a high-probability of introduction from the pet-trade or seafood industry 8 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Ulvaceae Liagoraceae Halymeniaceae Sargassaceae Scytosiphonaceae Rossellidae Aplysinidae Clionaidae Irciniidae Microcionidae Niphatidae Suberitidae Acroporidae Aiptasiidae Alcyoniidae Anthothelidae Dendrophylliidae Faviidae Gorgoniidae Helioporidae Meandrinidae Mussidae Pocilloporidae Poritidae Stichodactylidae Tubiporidae Milleporidae Physaliidae Stylasteridae Cepheidae Pelagiidae Stomolophidae Ulmaridae Lineidae Aplysiidae Babyloniidae Buccinidae Conidae Cypraeidae Ficidae Haliotidae Reference Col.1.101 Col.1.107 Col.1.108 Col.1.109 Col.1.100 Col.1.104 Col.1.110 Col.1.63 Col.1.154 Col.1.152 Col.1.38 Col.1.40 Col.1.153 Col.1.151 Col.1.39 Col.1.4 Col.1.10 Col.1.11 Col.1.67 Col.1.68 Col.1.129 Col.1.5 Col.1.86 Col.1.65 Col.1.7 Col.1.8 Col.1.43 Col.1.44 Col.1.45 Col.1.15 Col.1.6 Col.1.1 Col.1.16 Col.1.9 Col.1.12 Col.1.66 Col.1.14 Col.1.3 Col.1.128 Col.1.2 Col.1.70 Col.1.95 Col.1.13 Col.1.130 Col.1.150 Col.1.84 Col.1.90 Col.1.113 Col.1.96 Col.1.97 Col.1.71 Col.1.78 Col.1.79 Col.1.80 Col.1.81 Col.1.49 Col.1.47 Col.1.31 Col.1.52 Species Sea lettuce macro-algae species Thin sea lettuce Soft spaghettiweed Florida halemenia red macro-algae Gulfweed Brown sargassum weed Intricate brown macro-algae Deep sea hexactinellid “glass” sponge species Rope sponge species Boring sponge species Pillow stinking sponge Stinking vase sponge Red beard sponge Pink vase sponge Florida hermit-crab sponge Elkhorn coral Table coral Branch coral Finger coral Plate coral species Curlycue anemone Frilly lettuce coral Colorful sea rod Black sun coral Boulder coral species Brain coral species Yellow sea fan species Green sea fan species Purple sea fan species Blue ridge coral Scarlet coral species Flower coral species Large-cupped fungal coral Cauliflower coral species Brownstem coral Cat's paw coral Branching coral species Thin finger coral Caribbean carpet anemone Organ pipe coral Gulf fire coral Portuguese man of war Lace coral species Cassiopea jellyfish Sea nettle jellyfish species Cannonball jellyfish Moon jellyfish Line nemerteans worm Mottled sea hare Atlantic black sea hare Spiral Babylon sea snail Knobbed whelk egg case (string) Knobbed whelk Channeled whelk Common or Waved whelk Mixed cone sea snail species Red sea lettuce Graceful fig sea snail Pourtale's abalone (pearlized) Scientific Name Ulvaria spp. Ulva lactuca Liagora farinose Halymenia floridana Sargassum spp. Sargassum natans Rosenvingeo intricate Scolymastra spp. Aplysina spp. Cliona spp. Ircinia strobilina Ircinia campana Microciona prolifera Niphates digitalis Pseudospongosorites suberitoides Acropora palmate Acropora hyacinthus Acropora florida Acropora humilis* Acropora spp. Bartholomea annulata Lobophytum crassum Diodogorgia nodulifera Tubastraea micrantha Monstastrea spp. Diplora spp. Gongonia spp. Gongonia spp. Gongonia spp. Heliopora coerula Dichocoenia spp. Mussa spp. Scolymia lacera Pocillopora spp. Pocillopora verrucosa Pocillopora palifera* Madracris spp. Porites furcate Stichodactyla helianthus Tubipora musica* Millepora alcicornis Physalia physalis Family stylasteridae Cotylorhiza tuberculata Chrysaora spp. Stomolophus meleagris Aurelia aurito Lineus spp. Aplysia brasiliana Aplysia morio Babylonia spirata* Busycon carica Busycon carica Busycotypus canaliculatus Buccinum undatum* Conus spp. Halymenia floresia Ficus gracilis Haliotis pourtalesii** 9 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Littorinidae Reference Col.1.148 Col.1.51 Lottiidae Col.1.42 Melongenidae Col.1.32 Col.1.60 Muricidae Col.1.72 Nassariidae Col.1.127 Naticidae Col.1.27 Col.1.30 Neritidae Col.1.54 Olividae Col.1.33 Pseudomelatomidae Col.1.50 Pyramidellidae Col.1.123 Ranellidae Col.1.26 Col.1.53 Strombidae Col.1.17 Col.1.18 Col.1.19 Col.1.29 Col.1.73 Col.1.122 Terebridae Col.1.55 Achatinidae Col.1.37 Arcidae Col.1.20 Cardiidae Col.1.57 Col.1.121 Donacidae Col.1.56 Col.1.126 Hiatellidae Col.1.24 Limidae Col.1.58 Col.1.133 Myidae Col.1.25 Mytilidae Col.1.125 Pectinidae Col.1.34 Col.1.35 Col.1.36 Col.1.105 Pharidae Col.1.23 Col.1.28 Pinnidae Col.1.41 Solecurtidae Col.1.22 Spondylidae Col.1.46 Tellinidae Col.1.48 Veneridae Col.1.21 Col.1.59 Loliginidae Col.3.43 Octopodidae Col.3.44 Col.3.42 Col.3.45 Col.3.46 Ommastrephidae Col.4.26 Sabellidae Col.1.112 Limulidae Col.1.131 Col.1.85 Penaeidae Col.1.117 Col.1.118 Col.1.119 Stenopodidae Col.2.10 Species Scientific Name Mixed southern gulf periwinkle species Littorina spp. Yellow sea snails Littoraria spp. Atlantic plate limpet Lottia testudinalis Florida crown conch Melongena corona Crown conch (smooth and horned forms) Melongena corona Endive murex sea snail Murex endiva* Bruised nassa Nassarius vibex Northern moon snails Lunatia heros Shark eye sea snail (didyma) Polinices duplicatus Nerite snail species Nerita spp. Lettered olive sea snail Oliva sayana Turrid snail species Crassispira spp. Tulip sea snail Turbonilla curta Neapolitan triton Cymatium parthenopeum Knobbed triton snail Cymatium muricinum Rooster tail conch Lobatus gallus West Indian fighting conch Strombus pugilis Florida fighting conch Strombus alatus Hawkwing conch Strombus raninus Millipede conch Lambis millepeda** Silver lip conch Strombus lentiginosus* Eastern auger Terebra dislocata Fairyland snail sea snail Achatina achatina Incongruous ark Anadara brasiliana Rockweed Fucus vesiculosus Pacific giant clam species Tridacna spp.* White donax or Coquina clam Donax variabilis Variable coquina clams Donax variabilus Propeller clam Cyrtodaria siliqua Rough file clam Lima scabra Flamescallop or Red file shell clam Lima scabra Soft shell clams Mya arenaria Southern and American horse mussel species Modiolus spp. Florida calico scallop Argopecten gibbus White scallop Argopecten irradians Nucleus scallop Argopecten nucleus Paper scallop Euvola papyracea Common razor clams Ensis directus Dwarf razor clams Ensis megistus Half-naked pen clam Atrina seminude Stout tagelus Tagelus plebeius Spiny oyster Spondylus americanus Tar balls N/A Southern quahog Mercenaria campechiensis notata Southern quahog Mecenaria campechiensis Atlantic Oval Squid Sepioteuthis sepioidea Gulf deep-water octopus species Benthoctopus spp. Common Atlantic octopus Octopus vulgaris Caribbean reef octopus Octopus briareus Brownstripe octopus Octopus burryi Deep-sea squid species llex spp. Feather duster worm species Sabellastarte spp. Atlantic horseshoe crab (early instar) Limulus polyphemus Atlantic horseshoe crab (eggs) Limulus polyphemus Northern brown shrimp Farfantepenaeus aztecus Northern white shrimp Penaeus setiferus Southern pink shrimp Farfantepenaeus duorarum Banded coral shrimp or Cleaner shrimp-like decapod Stenopus hispidus 10 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Palaemonidae Reference Col.1.120 Col.1.115 Alpheidae Col.2.101 Nephropidae Col.2.91 Cambaridae Col.2.92 Parastacidae Col.2.93 Palinuridae Col.2.97 Scyllaridae Col.1.138 Col.1.139 Diogenidae Col.2.90 Paguridae Col.1.142 Col.1.143 Porcellanidae Col.2.100 Hippidae Col.1.106 Majidae Col.2.98 Col.1.136 Col.1.137 Inachidae Col.1.140 Cancridae Col.2.2 Col.2.3 Portunidae Col.3.6 Col.2.4 Col.2.1 Varunidae Col.1.124 Potamonautidae Col.1.146 Ocypodidae Col.1.149 Col.1.145 Ucididae Col.1.147 Cirolanidae Col.1.83 Squillidae Col.2.5 Chelonibiidae Col.1.116 Lepadidae Col.1.82 Scalpellidae Col.1.94 Balanidae Col.1.69 Alcyonidiidae Col.1.98 Thalamoporellidae Col.1.99 Echinasteridae Col.1.75 Col.1.144 Oreasteridae Col.1.74 Ophiactidae Col.1.135 Ophidiasteridae Col.1.103 Ophiodermatidae Col.1.134 Gorgonocephalidae Col.1.132 Clypeasteridae Col.1.76 Col.1.77 Mellitidae Col.1.64 Toxopneustidae Col.1.102 Col.1.141 Holothuriidae Col.1.87 Col.1.89 Stichopodidae Col.1.88 Perophoridae Col.1.92 Polyclinidae Col.1.93 Styelidae Col.1.91 Col.1.62 Col.1.111 Petromyzontidae Col.2.28 Lamnidae Col.6.2 Scyliorhinidae Col.4.20 Species Scientific Name Malaysian prawn Macrobrachium rosenbergii Common Shore Shrimp Palaemonetes vulgaris Bigclaw pistol or snapping shrimp Alpheus heterochaelis Rosy lobsterette Nephropsis rosea Red swamp crayfish or Louisiana crawfish Procambarus clarkia Common Australian yabby crayfish Cherax destructor** Caribbean spiny lobster (larvae) Panulirus argus Ridged slipper lobster Scyllarides nodifer Deep-water slipper lobster species Scyllarides spp. Giant or Red hermit crab Petrochirus diogenes Flat-clawed hermit crab inside Northern moon snailPagurus shell pollicaris inside Lunatia heros Acadian hermit crab inside Neapolitan triton snail shell Pagurus acadianus inside Cymatium parthenopeum Lined porcelain crab Petrolisthes galathinus Gulf mole crabs Emerita brasiliensis Coral mithrax crab Mithrax coryphe Sponge spider crab species & Fire sponges Macrocoeloma spp. & Tedania ignis Green reef crab Mithrax sculptus Yellowline arrow crab Stenorhynchus seticornis Rock crab Cancer irroratus Green crab Cancer maenas Atlantic blue-clawed crab Callinectes sapidus Rugose or Purple swimming crab Callinectes exasperates Lady crab Ovalipes ocellatus Asian shore crab Hemigrapsus sanguineus Purple marsh crab Afrithelphusa monodosa Caribbean mud fiddler crab Uca rapax Mud fiddler crab Uca pugnax Mangrove land crab Ucides cordatus Giant sea roach Bathynomus giganteous Common mantis shrimp Squilla empusa Ivory barnacles Chelonibia patula Goose neck barnacles species and Shipworm species Bankia with spp. wood & Lepas spp. Primitive deep sea stalked barnacle species Litoscalpellum spp. Purple barnacle Conchylepes conchylepes* Hauff’s alcyonidium Alcyonidium hauffi Lettuce bryozoan Thalamoporella gothica Brown spiny sea star Echinaster spinulosus Brown spiny sea star Echinaster spinulosus Horned starfish Protoreaster nodosus** Serpent star species Hemipholis spp. Spotted linckia starfish Linckia multiflora** Brittle Star species Ophioderma spp. Basket sea-star Gorgonocephalus eucnemis Inflated sea biscuit Clypeaster rosaceus Flat sea biscuit Clypeaster subdepressus Five-holed keyhole urchin Mellita quinquiesperforata Pincushion urchin species Lytechinus spp. West Indian sea urchins Tripneustes ventricosus Brown rock sea cucumber Holothuria glaberrima Five-toothed sea cucumber Actinopyga agassizii Furry sea cucumber Astichopus multifudus Mangrove tunicate Ecteinascidia turbinata Sea pork species Amaroucium spp Black encrusting tunicate Botrylloides nigrum Pleated sea squirt Styela plicata Green encrusting tunicate Symplegma viride Sea lamprey Petromyzon marinus Mako shark species Isurus spp. Chain catshark or Chain dogfish Scyliorhinus rotifer 11 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Triakidae Carcharhinidae Sphyrnidae Narcinidae Rhinobatidae Rajidae Myliobatidae Acipenseridae Anguillidae Muraenidae Ophichthidae Congridae Engraulidae Clupeidae Loricariidae Osmeridae Synodontidae Anotopteridae Myctophidae Phycidae Ophidiidae Batrachoididae Lophiidae Antennariidae Ogcocephalidae Mugilidae Atherinopsidae Exocoetidae Hemiramphidae Belonidae Scomberesocidae Fundulidae Poeciliidae Reference Col.5.8 Col.6.1 Col.3.52 Col.2.14 Col.2.11 Col.1.61 Col.2.20 Col.2.12 Col.2.13 Col.2.96 Col.2.19 Col.2.9 Col.3.39 Col.4.21 Col.4.22 Col.3.48 Col.3.40 Col.4.5 Col.2.104 Col.2.73 Col.2.75 Col.2.79 Col.2.86 Col.2.16 Col.2.17 Col.2.103 Col.3.10 Col.4.16 Col.4.17 Col.3.56 Col.3.28 Col.3.17 Col.4.13 Col.4.6 Col.4.12 Col.5.4 Col.3.11 Col.2.67 Col.2.70 Col.2.76 Col.2.89 Col.2.54 Col.2.71 Col.2.72 Col.2.56 Col.2.62 Col.2.78 Col.2.18 Col.5.1 Col.3.4 Col.3.32 Col.2.77 Col.2.69 Col.2.53 Col.2.57 Col.2.60 Col.2.102 Col.2.59 Col.2.61 Species Dusky smooth-hound or Smooth dogfish shark Blacktip shark (juvenile) Bonnethead shark or shovelhead Lesser electric ray Atlantic guitarfish Skate species egg cases Underworld windowskate Smooth skate Clear-nose skate Cownose ray Southern eagle ray (juvenile) Atlantic sturgeon (juvenile) American eel Caribbean ocellated moray Hourglass moray eel Chestnut Moray eel Shrimp eel Conger eel Bay anchovy Spanish sardine Gulf menhaden American gizzard shad Atlantic thread herring Hickory shad Blueback herring Armored catfish or Plecostomus species Rainbow smelt Inshore lizardfish Deep-sea dagger tooth or Pharaoh fish Long-finned lantern fish species Deep-sea lantern fish species Spotted hake Bearded brotula Gulf toadfish Oyster toadfish Blackfin goosefish Ocellated frogfish Pancake batfish Striped Mullet Fantail mullet Mountain mullet Flathead mullet Inland silverside Atlantic silverside Tidewater silverside Rough silverside Blackwing flyingfish Balao halfbeak Deep-water needlefish or Houndfish Atlantic needlefish Atlantic saury Mummichog Gulf killifish Striped killifish Golden topminnow Rainwater killifish Mosquitofish Sailfin molly (female & male) Dwarf livebearer or Least killifish Scientific Name Mustelus canis Carcharhinus limbatus Sphyrna tiburo Narcine bancroftii Rhinobatos lentiginosus Raja spp. Fenestraja plutonia Malacoraja senta Raja eglanteria Rhinoptera bonasus Myliobatis goodei Acipenser oxyrinchus Anguilla rostrata Gymnothorax ocellatus Muraena clepsydra* Enchelycore carychroa Ophichthus gomesii Conger oceanicus Anchoa mitchilli Sardinella aurita Brevoortia patronus Dorosoma cepedianum Opisthonema oglinum Alosa mediocris Alosa aestivalis Hypostomus ssp. Osmerus mordax* Synodus foetens Anotopterus pharao Diaphus or Lampanyctus spp. Diaphus or Lampanyctus spp. Urophycis regius Brotula barbata Opsanus beta Opsanus tau Lophius gastrophysus Fowlerichthys ocellatus Halieutichthys aculeatus Mugil cephalus Mugil trichodon Agonostomus moniticola Mugil cephalus Menidia beryllina Menidia menidia Menidia peninsulae Membras martinica Hirundichthys rondeletii Hemiramphus balao Tylosurus crocodilus Strongylura marina Scomberesox saurus Fundalulus heteroclitus* Fundulus grandis Fundulus majalis Fundulus chrysotus Lucania parva Gambusia affinis Poecilia latipinna Heterandria formosa 12 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Zeidae Syngnathidae Synbranchidae Scorpaenidae Sebastidae Triglidae Peristediidae Latidae Moronidae Polyprionidae Serranidae Opistognathidae Centrarchidae Priacanthidae Malacanthidae Pomatomidae Coryphaenidae Echeneidae Carangidae Lutjanidae Gerreidae Reference Col.4.1 Col.2.74 Col.2.66 Col.2.44 Col.2.45 Col.2.46 Col.2.34 Col.2.35 Col.2.36 Col.2.37 Col.3.41 Col.4.15 Col.3.47 Col.3.51 Col.3.5 Col.4.30 Col.3.53 Col.3.24 Col.2.8 Col.4.11 Col.5.13 Col.3.8 Col.6.3 Col.5.10 Col.5.11 Col.5.12 Col.4.29 Col.4.27 Col.4.28 Col.4.14 Col.4.19 Col.3.54 Col.3.55 Col.2.52 Col.2.94 Col.2.63 Col.3.22 Col.3.23 Col.4.24 Col.4.23 Col.6.5 Col.2.85 Col.5.7 Col.5.2 Col.5.3 Col.4.18 Col.3.34 Col.3.35 Col.3.2 Col.3.3 Col.2.47 Col.2.48 Col.2.31 Col.4.25 Col.4.7 Col.4.8 Col.4.10 Col.4.3 Col.3.31 Species St Pierre’s fish or John Dory Dusky pipefish Dwarf seahorse Opossum pipefish Gulf pipefish Sargassum pipefish Longsnout seahorse Northern pipefish Chain pipefish Lined seahorse Asian swamp or Synbranchidae eel species Spinycheek scorpionfish Caribbean reef scorpion fish Deepreef scorpionfish Red lionfish Redfish species Blackbelly rosefish Striped searobin Rimspine searobin Barmundi or Asian seabass Atlantic striped bass White perch Atlantic wreckfish Yellowfin grouper Yellowmouth grouper Red grouper Red hind grouper Mutton hamlet Graysby Coney Grouper or Leopard Hind Black sea bass Butter hamlet Clown leaflip soapfish Spotfin jawfish species Longear sunfish or Creek perch Green sunfish Bigeye Short bigeye Great northern tilefish Bluefish Mahi-mahi or Common dolphinfish Sharksucker Greater amber jack Blue runner Crevalle jack Pilot fish Mackerel scad Bigeye scad Permit Florida pomano Atlantic Moonfish Lookdown White trevally Wrenchman Gray snapper Red snapper Vermilion snapper Yellow-tail snapper Yellowfin morjarra Scientific Name Zeus faber Synganathus floridae Hippocampus zosterae Microphis brachyurus lineatus Syngnathus scovelli Syngnathus pelagicus Hippocampus reidi Synganathus fuscus* Synganathus louisanae Hippocampus erectus Monopterus spp. Neomerinthe hemingway Scorpaenodes caribbaeus Scorpaenodes tredecimspinosus Pterois volitans Sebastes spp. Helicolenus dactylopterus Prionotus evolans Peristedion thompsoni Lates calcarifer** Morone saxatilis Morone Americana* Polyprion americanus Mycteroperca venenosa Mycteroperca interstitialis Epinephelus morio Epinephelus guttatus Alphestes afer Cephalopholis cruentata Cephalopholis fulva Centropristis striata Hypoplectrus unicolor Pogonoperca punctate** Opistognathus spp. Lepomis megalotis Lepomis cyanellus Priacanthus arentus Pristigenys alta Lopholatilus chamaeleonticeps Pomatomus saltatrix Coryphaena hippurus Echeneis naucrates Seriola dumerili Caranx crysos Caranx hippos Naucrates doctor Decapterus macarellus Selar crumenophthalmus Trachinotus falcatus Trachinotus carolinus Selene setapinnis Selene vomer Pseudocaranx dentex Pristipomoides aquilonaris Lutjanus griseus Lutjanus campechanus Rhomboplites aurorubens Ocyurus chrysurus Gerres cinereus 13 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Haemulidae Sparidae Sciaenidae Mullidae Chaetodontidae Pomacanthidae Cichlidae Labridae Scaridae Ammodytidae Uranoscopidae Eleotridae Gobiidae Acanthuridae Sphyraenidae Trichiuridae Scombridae Stromateidae Scophthalmidae Paralichthyidae Pleuronectidae Achiridae Cynoglossidae Balistidae Monacanthidae Ostraciidae Reference Col.3.49 Col.4.4 Col.3.18 Col.3.19 Col.3.20 Col.3.9 Col.3.29 Col.3.30 Col.3.25 Col.3.26 Col.3.27 Col.3.1 Col.2.64 Col.2.15 Col.2.65 Col.2.49 Col.2.22 Col.2.23 Col.2.99 Col.2.83 Col.2.58 Col.3.21 Col.3.37 Col.2.105 Col.2.95 Col.2.84 Col.2.50 Col.2.32 Col.2.40 Col.2.41 Col.2.26 Col.2.7 Col.4.9 Col.3.36 Col.2.55 Col.2.30 Col.2.87 Col.2.51 Col.3.50 Col.4.2 Col.5.5 Col.5.9 Col.6.4 Col.3.33 Col.2.68 Col.2.80 Col.2.81 Col.3.16 Col.3.13 Col.3.14 Col.3.15 Col.3.12 Col.2.6 Col.2.82 Col.2.24 Col.2.25 Col.2.32 Col.2.29 Col.2.33 Species Scientific Name Spanish grunt Haemulon macrostomum Cottonwick Haemulor melanurum Spotted pig grunt Orthopristis chrysoptera White grunt Haemulon plumierrii Sailor’s choice grunt Haemulon parra Royal dorade or Mediterranean gilt-head seabream Sparus aurata* Pinfish Lagodon rhomboids Sea bream Archosargus rhomboidalis Southern kingfish Mentricirrhus americanus Northern kingfish Mentricirrhus saxitilis Atlantic croaker Micropogonias undulates Freshwater drum Aplodinotus grunniens High hat (juvenile) Pareques acuminatus Red goatfish Mullas auratus Foureye butterfly fish Chaetodon capistratus Blue-striped Angelfish Chaetodontoplus septentrionalis** Regal angelfish Pygoplites diacanthus** Six barred angel (juvenile) Pomacanthus sexstriatus** Whitetail damselfish Dascyllus aruanus* Blue reef chromis damselfish Chromis cyanea Texas cichlid Herichthys cyanoguttatus Puddingwife Halichoeres poeyi Tautog or Blackfish Tautoga onitis Birdmouth Wrasse Gomphosus caeruleus** Yellowtail coris wrasse Coris gaimard** Creole wrasse Clepticus parrae Pearl wrasse Anampses cuvier** Pygmy filefish Monacanthus setifer Spotlight parrotfish Sparisoma viride Redband parrotfish Sparisoma aurofrenatum Queen parrotfish Scarus vetula Mediterranean sand lance Gymnammodytes cicerelus* Freckled stargazer Gnathagnus egreglus Fat sleeper Dormitator maculatus Naked goby (female & male) Gobiosoma bosci Violet goby Gobioides broussoneti Whitecheek surgeonfish Acanthurus nigricans* Yellow tang Zebrasoma flavescens* Northern sennet Sphyraena boreali Guachanche barracuda Sphyraena guachancho Atlantic cutlassfish or Ribbonfish Trichiurus lepturus Atlantic Spanish mackerel Scomberomorus maculatus Wahoo Acanthocybium solandri Atlantic chub mackerel Scomber colias Wahoo (larvae) Acanthocybium solandri Butterfish Peprilus triacanthus Harvestfish Peprilus paru Windowpane flounder Scophthalmus aquosus Summer flounder Paralichthys dentatus Gulf flounder Paralichthys albigutta Southern flounder Paralichthys lethostigma Winter flounder Pseudopleuronectes americanus* Hogchoker Trinectes maculatus Blackcheek tonguefish Symphurus plagiusa Sargassum triggerfish Xanthichthys ringens Gray triggerfish Balistes capriscus Pygmy filefish Monacanthus setifer Thornbacked boxfish Tetrosomus gibbosus** Long-horned cowfish Lactoria cornuta 14 Specimen list in phylogenetic order prepared by Bryan Weatherwax Ichthyology Department, New York State Museum, Albany, NY Family Tetraodontidae Diodontidae Amphiumidae Trionychidae Pythonidae Reference Col.2.38 Col.2.43 Col.2.21 Col.2.27 Col.2.39 Col.2.42 Col.3.7 Col.3.38 Col.5.6 Species Checkered puffer fish Blunthead puffer Masked pufferfish Sharpnose puffer Striped burrfish Porcupinefish Three-toed amphiuma Chinese softshell turtle Burmese python Scientific Name Sphoeroides testudineus Sphoeroides pachygaster Arothron diadematus* Bamthigaster rostrata Chilomycterus schoepfii Diodon hystrix Amphiuma tridactylum Pelodiscus sinensis** Python molurus bivittatus * non-native species ** non-native species anecdotally reported but not scientifically confirmed or species with a high-probability of introduction from the pet-trade or seafood industry 15 AQUATIC MICROBIAL ECOLOGY Aquat Microb Ecol Vol. 63: 101–109, 2011 doi: 10.3354/ame01482 Published online March 31 OPEN ACCESS FEATURE ARTICLE Effects of COREXIT® EC9500A on bacteria from a beach oiled by the Deepwater Horizon spill Leila J. Hamdan1,*, Preston A. Fulmer 2 1 Marine Biogeochemistry Section, Code 6114, U.S. Naval Research Laboratory, Overlook Ave. SW, Washington, DC 20375, USA 2 Bioenergy and Biofabrication Section, Code 6115, U.S. Naval Research Laboratory, Overlook Ave. SW, Washington, DC 20375, USA ABSTRACT: Hydrocarbon-degrading bacteria are important for controlling the fate of natural and anthropogenic hydrocarbons in the marine environment. In the wake of the Deepwater Horizon spill in the Gulf of Mexico, microbial communities will be important for the natural attenuation of the effects of the spill. The chemical dispersant COREXIT® EC9500A was widely deployed during the response to the Deepwater Horizon incident. Although toxicity tests confirm that COREXIT® EC9500A does not pose a significant threat to invertebrate and adult fish populations, there is limited information on its effect on microbial communities. We determined the composition of the microbial community in oil that had been freshly deposited on a beach in Louisiana, USA, as a result of the Deepwater Horizon spill. The metabolic activity and viability in cultures obtained from oil samples were determined in the absence and presence of COREXIT® EC9500A at concentrations ranging from 0.001 to 100 mg ml–1. In length heterogeneity PCR (LH-PCR) fingerprints of oil samples, the most abundant isolates were those of Vibrio, followed by hydrocarbon-degrading isolates affiliated with Acinetobacter and Marinobacter. We observed significant reductions in production and viability of Acinetobacter and Marinobacter in the presence of the dispersant compared to controls. Of the organisms examined, Marinobacter appears to be the most sensitive to the dispersant, with nearly 100% reduction in viability and production as a result of exposure to concentrations of the dispersant likely to be encountered during the response to the spill (1 to 10 mg ml–1). Significantly, at the same concentration of dispersant, the non-hydrocarbon-degrading Vibrio isolates proliferated. These data suggest that hydrocarbon-degrading bacteria are inhibited by chemical dispersants, and that the use of dispersants has the potential to diminish the capacity of the environment to bioremediate spills. The mobile offshore drilling unit ‘Deepwater Horizon’ experienced an explosion on April 20, 2010, and sank 2 d later (Coastal Response Research Center 2010). As a result of the explosion, and failure of a blowout preventer on the sea floor, crude oil began leaking from a broken riser pipe into the Gulf of Mexico at an estimated rate of 19 000 to 70 000 barrels (~3000 to 11 000 m3) of oil d–1 (Labson et al. 2010). This resulted in the largest oil spill in the coastal waters of the USA. *Email: [email protected] © Inter-Research 2011 · www.int-res.com Oil from the Deepwater Horizon spill washing ashore on Elmer's Island, Louisiana, USA, in May 2010. Inset: viability assay for the hydrocarbon-degrading bacterium Acinetobacter venetianus before (left) and after (right) exposure to COREXIT, revealing inhibition by the chemical dispersant. Photo: Warren Wood; inset: Leila J. Hamdan KEY WORDS: Deepwater Horizon · Gulf of Mexico · Dispersant · Hydrocarbon degraders · Vibrio · COREXIT · Toxicity Resale or republication not permitted without written consent of the publisher INTRODUCTION Report Documentation Page Form Approved OMB No. 0704-0188 Public reporting burden for the collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources, gathering and maintaining the data needed, and completing and reviewing the collection of information. Send comments regarding this burden estimate or any other aspect of this collection of information, including suggestions for reducing this burden, to Washington Headquarters Services, Directorate for Information Operations and Reports, 1215 Jefferson Davis Highway, Suite 1204, Arlington VA 22202-4302. Respondents should be aware that notwithstanding any other provision of law, no person shall be subject to a penalty for failing to comply with a collection of information if it does not display a currently valid OMB control number. 1. REPORT DATE MAR 2011 3. DATES COVERED 2. REPORT TYPE 00-00-2011 to 00-00-2011 4. TITLE AND SUBTITLE 5a. CONTRACT NUMBER Effects of COREXIT EC9500A on bacteria from a beach oiled by the Deepwater Horizon spill 5b. GRANT NUMBER 5c. PROGRAM ELEMENT NUMBER 6. AUTHOR(S) 5d. PROJECT NUMBER 5e. TASK NUMBER 5f. WORK UNIT NUMBER 7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES) U.S. Naval Research Laboratory,Marine Biogeochemistry Section, Code 6114,Washington,DC,20375 9. SPONSORING/MONITORING AGENCY NAME(S) AND ADDRESS(ES) 8. PERFORMING ORGANIZATION REPORT NUMBER 10. SPONSOR/MONITOR’S ACRONYM(S) 11. SPONSOR/MONITOR’S REPORT NUMBER(S) 12. DISTRIBUTION/AVAILABILITY STATEMENT Approved for public release; distribution unlimited 13. SUPPLEMENTARY NOTES 14. ABSTRACT Hydrocarbon-degrading bacteria are important for controlling the fate of natural and anthropogenic hydrocarbons in the marine environment. In the wake of the Deepwater Horizon spill in the Gulf of Mexico, microbial communities will be important for the natural attenuation of the effects of the spill. The chemical dispersant COREXIT? EC9500A was widely deployed during the response to the Deepwater Horizon incident. Although toxicity tests confirm that COREXIT? EC9500A does not pose a significant threat to invertebrate and adult fish populations there is limited information on its effect on microbial communities. We determined the composition of the microbial community in oil that had been freshly deposited on a beach in Louisiana, USA, as a result of the Deepwater Horizon spill. The metabolic activity and viability in cultures obtained from oil samples were determined in the absence and presence of COREXIT? EC9500A at concentrations ranging from 0.001 to 100 mg ml?1. In length heterogeneity PCR (LH-PCR) fingerprints of oil samples, the most abundant isolates were those of Vibrio, followed by hydrocarbon-degrading isolates affiliated with Acine - to bac ter and Marinobacter. We observed significant reductions in production and viability of Acineto - bacter and Marinobacter in the presence of the dis - persant compared to controls. Of the organisms examined, Marinobacter appears to be the most sensitive to the dispersant, with nearly 100% reduction in viability and production as a result of exposure to concentrations of the dispersant likely to be encountered during the response to the spill (1 to 10 mg ml?1). Significantly, at the same concen tration of dispersant the non-hydrocarbon-degrading Vibrio isolates proliferated. These data suggest that hydro - carbon-degrading bacteria are inhibited by chemical dispersants, and that the use of dispersants has the potential to diminish the capacity of the environment to bioremediate spills. 15. SUBJECT TERMS 16. SECURITY CLASSIFICATION OF: a. REPORT b. ABSTRACT c. THIS PAGE unclassified unclassified unclassified 17. LIMITATION OF ABSTRACT 18. NUMBER OF PAGES Same as Report (SAR) 9 19a. NAME OF RESPONSIBLE PERSON Standard Form 298 (Rev. 8-98) Prescribed by ANSI Std Z39-18 102 Aquat Microb Ecol 63: 101–109, 2011 Natural seepage of hydrocarbons is an important source of carbon for benthic environments in the Gulf of Mexico and is a structuring factor for benthic and near-bottom microbial communities (National Research Council Committee on Oil in the Sea 2003). In marine environments, bacteria are the predominant degraders of hydrocarbons (Leahy & Colwell 1990) and thus are of great importance in controlling the fate of natural and anthropogenic seepage. In such environments, microbial communities contain members predisposed to hydrocarbon metabolism (Aharon & Fu 2000, Lanoil et al. 2001). Microbial assemblages associated with hydrocarbon seeps are well characterized in Gulf of Mexico sediments (Hollaway et al. 1980, Aharon & Fu 2000, Lanoil et al. 2001, Mills et al. 2003, Joye et al. 2004, Reed et al. 2006, Lloyd et al. 2010, Orcutt et al. 2010). The impact of natural seepage on water-column communities has been relatively less studied, although microorganisms with the ability to degrade hydrocarbons are ubiquitously distributed in waters along continental shelves (Venkateswaran et al. 1991). Because large-scale impacts of hydrocarbons on Gulf of Mexico beaches are rare, little is known about the microbial communities which live in proximity to beached oil. In the USA, the Environmental Protection Agency (EPA) maintains a list of chemicals and spill-mitigating devices which may be deployed during an oil spill in coastal waters of the USA; this is a part of the National Contingency Plan (NCP) (US EPA 2010b). NCPapproved dispersants are applied to break up masses of oil and reduce the formation of surface slicks. This is done to reduce the incidence of oil coating on populations of birds, mammals and invertebrates. In the wake of the Deepwater Horizon spill, dispersants were deployed widely in surface and sub-surface waters (Coastal Response Research Center 2010). As of early June 2010, more than 990 000 gallons (~3700 m3) of dispersant were used in the response. The most commonly used dispersant was COREXIT® EC9500A (Nalco). There is limited knowledge regarding the communities that metabolize oil in the water column and on coastal beaches. Even less is known about the impact of dispersants in general — and specifically COREXIT® EC9500A — on hydrocarbon-degrading microbial communities in the Gulf of Mexico. Dispersants have been suggested as a means to improve microbial biodegradation of oil contamination in the water column by forming small oil droplets with high surface-to-volume ratios which increase their lability to microplankton (Brakstad 2008). However, this same process is likely to increase the concentration and lability of spill-related toxic compounds in the water column, which may, in turn, affect the microbial hydrocarbon-degrading community (Zahed et al. 2010). Furthermore, the dispersants themselves may impact hydrocarbon-degrading microorganisms. Thus, the goal of this study was to understand the impact of COREXIT® EC9500A (henceforth referred to as ‘COREXIT’) on bacterial viability and metabolic function in oiled beach samples from the area affected by the Deepwater Horizon spill. These samples may be ideal candidates for such analysis because microbial communities have been acclimatized to conditions of high contamination by hydrocarbons, environmental weathering and possibly in situ exposure to dispersants. Future studies of the impact of dispersants on microbial communities should also include analysis of ‘pristine’ environmental communities. Studies of both exposed and pristine communities will be of importance in understanding the long-term affects of the spill and the use of chemical dispersants in aquatic environments. Previous studies have documented clear evidence of the deleterious effect of the dispersant on wildlife and microbial communities (Lindstrom & Braddock 2002, Couillard et al. 2005, Jung et al. 2009). Due to the unknown effects of the large-scale use of dispersant on microbial communities in affected areas, more study is warranted on the effects of the widespread application of dispersant. MATERIALS AND METHODS Sample collection and handling. Beached oil samples were collected on May 22, 2010, from the south end of the Elmer’s Island Wildlife Refuge (EIWR), Louisiana, USA, on the seaward facing shore. Personal accounts from EIWR staff indicate that accumulations of oil reached the beach mid-day on the previous day. During sampling, small brown pea-sized floating droplets of oil were observed in the water within 0.5 m of the high-water mark, and pools of thick oil were deposited at the high-water mark along the beach. Oil on the beach appeared in several different forms, related to how long it had been deposited (EIWR staff personal account). Oil that had been on the beach longer than 24 h had soaked into the sand. Fresher oil rested on the surface of the sand. Freshly deposited oil was collected in cellulose acetate butyrate core liners, capped at both ends, and later sub-sampled into sterile 50 ml polyethylene tubes. Samples were maintained at ambient temperature during transport from the field (24 h) and subsequently held in the dark on ice during shipping to the laboratory (48 h). The total time between sample collection and analysis was ~72 h. Because of these hold times, alterations in the structure of the microbial community may have occurred; this may have resulted in the enhancement of some phylotypes, masking of the appearance of phylotypes accounting for less than 1% Hamdan & Fulmer: Effects of COREXIT® EC9500A on bacteria of the population in LH-PCR and culture analysis, or loss of cultivability of others. Bacterial abundance. A modification of the method of Hobbie et al. (1977), described in Hamdan & Jonas (2006), was used to determine bacterial abundance. Oil and oil-water (liquid surrounding the oil) was diluted 1:100 with buffered sterile seawater, stained with 0.1%, acridine orange, collected on black polycarbonate filters of pore size 0.2 µm (Osmonics) and observed at 1000× magnification. Cultivation conditions. Marine agar (MA) (Difco) was used as a complete medium for bacterial growth. Bushnell-Haas (BH) agar (Difco) supplemented with 1% w/v hexadecane (Fisher Scientific) was used as a selective medium to isolate hydrocarbon degraders. All media were prepared according to the manufacturer’s specifications. Culturable bacteria were assessed using standard microbial culture techniques. Approximately 10 µl of oil or oil-water was plated on MA and BH using a quadrant streak method. Plates were incubated at 30°C. Colonies were isolated based on morphology and re-plated to produce pure cultures. Isolation of genomic DNA and 16S rDNA sequencing. To identify members of the microbial community, individual colonies from MA and BH plates were added to 5 ml of marine broth (MB) and incubated for 18 h at 30°C. Genomic DNA was isolated by alkaline lysis. For this, 2 ml of MB culture was centrifuged at 4000 × g. The medium was removed and cells were resuspended in 1 ml sterile phosphate-buffered saline (PBS). Cells were pelleted again at 4000 × g and resuspended in 50 mM Tris-Cl (pH 8.0), 10 mM EDTA, and 1 mg l–1 (1 µg ml–1) RNase A. Cells were then lysed in a buffer containing 200 mM NaOH and 1% sodium dodecyl sulfate (SDS), and the solution was neutralized with 3.0 M potassium acetate. Lysates were centrifuged at 15 000 × g to remove cell debris, and supernatants were collected. DNA was precipitated with isopropyl alcohol and centrifuged at 15 000 × g, washed with ethanol, and dried at room temperature prior to resuspension in 10 mM Tris-Cl (pH 8.5). DNA (1 µg) was used as a template for a PCR as follows: 1 U Failsafe Enzyme mix (Epicentre Biotechnologies), 2× Failsafe Premix E, 2 mM 27F 16s rDNA universal bacterial primer (5’-AGA GTT TGA TCC TGG CTC AG-3’), and 2 mM 1492R 16s rDNA universal bacterial primer (5’ACG GCT AGC TTG TTA CGA CTT-3’). Reactions were run as follows: 94°C for 5 min; 20 cycles of 94°C for 30 s, 50°C for 30 s, and 72°C for 90 s; 72°C for 7 min. PCR product was visualized on a 1% agarose gel containing ethidium bromide. Bands of ~1500 bp size were excised and purified using a gel purification kit (Qiagen). Excised PCR products were sequenced by Genewiz and checked for homology to known sequences in GenBank using the BLASTn algorithm. 103 Length heterogeneity polymerase chain reaction (LH-PCR). Genomic DNA was extracted from ~500 mg of oil and oil-water mixtures using the Bio 101 FastDNA® SPIN kit for soil. DNA was quantitated on a 1% agarose gel with ethidium bromide and diluted with DEPCtreated water such that ~10 ng of DNA was used as template for LH-PCR. Environmental samples (oil and oilwater) and isolates were analyzed by LH-PCR. Hamdan et al. (2008) provides a detailed description of LH-PCR. Briefly, amplification of variable regions V1 and V2 of the small subunit rRNA gene was performed using the primers 6-FAM-27F (5’-6-FAM-AGA GTT TGA TCM TGG CTC AG-3’) and 355R (5’-GCTGCC TCC CGT AGG AGT-3’). Controls accompanied reactions to determine PCR efficiency and calibrate peak intensity. PCR mixtures consisted of 1 × Gold buffer, 2.5 mM MgCl2, 0.2 mM dNTPs, 0.5 U AmpliTaq Gold, 0.5 µM primers, 0.01% BSA, and diethylpyrocarbonate (DEPC)-treated water. PCR was performed on a GeneAmp System 9700 (Applied Biosystems) with the following program: 20 to 30 cycles of 95°C (30 s), 48°C (30 s), 72°C (2 min plus 5 s per cycle) and final extension at 72°C (30 min). PCR product was diluted, mixed with ILS-600 (Promega) and HiDi formamide and loaded on an ABI 3130xl Analyzer (Applied Biosystems). LH-PCR was performed on the isolates, and the resulting amplicons were virtually aligned with LHPCR amplicons from the environmental samples so that the latter could be attributed to one or more isolates. In this manner, the contribution of isolates to the environmental community was determined. LH-PCR was also performed using universal archaeal primers. No product was obtained with archaeal primers 1HKF and 589R (see Litchfield et al. 2005 for details); thus, we conclude that most of the environmental community belongs to the bacterial domain. Bacterial viability. The effect of COREXIT (Nalco, batch # SLOE 1184) on bacterial growth was determined using a Bac Light™ (Invitrogen) Live/Dead Bacterial Viability assay. COREXIT is a mixture of light petroleum distillates (10 to 30%), propylene glycol (1 to 5%) butanedioic acid, 2-sulfo-1, 4-bis(2-ethylhexyl) ester, sodium salt (10 to 30%), in addition to unspecified amounts of propanol and sorbitan (www.nalco.com, www.epa.gov). Assays were conducted according to the manufacturer’s instructions, including standardization curves for each strain. To determine the toxicity of COREXIT for isolates, 10-fold serial dilutions of COREXIT, ranging from 1:10 to 1:100 000, in MB were added to wells of a 96-well microtitre plate. Approximately 106 colony-forming units (cfu) from a mid-log phase culture of each isolate (in triplicate) were added to the wells and incubated at 30°C for 18 h. Cells were pelleted at 3000 × g and resuspended in sterile 0.85% NaCl, followed by staining using the Bac Light™ kit. Bacterial 104 Aquat Microb Ecol 63: 101–109, 2011 viability was determined using a FLx800 (BioTek) microplate reader. To determine the effect of COREXIT on hydrocarbon utilization by isolates with known hydrocarbondegrading capabilities, we used Bushnell-Haas (BH) agar supplemented with 1% w/v hexadecane. COREXIT was added to the medium–hexadecane mix at ratios of 1:10, 1:25 and 1:50 COREXIT:hexadecane. Cultures were assayed as above. Heterotrophic bacterial production. Production was measured as leucine incorporation (Smith & Azam 1992). Briefly, 0.50 ml aliquots of oil, oil-water and MB cultures were added to microcentrifuge tubes (3 experimental and 1 abiotic control) that contained [3H-4, 5]–1 L-leucine (154 mCi mmol ) and incubated for 1 h at 25°C. Incubations were terminated by the addition of 100% trichloroacetic acid (TCA). Samples were centrifuged to pellet cells and washed with 5% TCA and ethanol to remove unincorporated radiolabel. Radioactivity was determined on a Beckman-Coulter LS6500 liquid scintillation counter. MB cultures of isolates containing no COREXIT and containing 1 of 2 dilutions of COREXIT (final conc. ~1 and 10 mg ml–1 in MB) were used. These were held for 48 h and sampled at t = 0, 6, 12, 24 and 48 h. RESULTS Cell abundance in the oil-water fraction of the sample was 2.9 × 1011 cells l–1. Measurements of cell abundance on the oil sample were not possible even when dispersed, as the sample matrix clogged filters, making accurate measurements impossible. Heterotrophic secondary production in the oil-water was 2.7 × 109 cells l–1 d–1, indicating a high standing stock of cells in the stationary phase of growth. Eight isolates were identified from the environmental samples (oil and oil-water) (Table 1). Most isolates were members of the class Gammaproteobacteria. With the exception of the isolates related to Exiguobacterium arabatum and Acinetobacter venetianus, all matched most closely (> 96%) with environmental clones obtained from marine or coastal waters. Among the isolates were 3 known hydrocarbon degraders (Table 1). A total of 15 LH-PCR amplicons were observed in the samples (Table 1). The Simpson’s Index (D) was used to estimate bacterial diversity. Diversity was relatively low in the oil and oil-water samples (0.15 and 0.20, respectively), and 4 amplicons accounted for the majority of the LH-PCR peak area. The 4 peaks that accounted for the majority of peak abundance corresponded with 3 isolates and 1 unidentified amplicon (Table 1). Vibrio sp. was highly abundant in both samples and, alone, accounted for ≥31% of the peak area in both samples. Experiments to assess the acute toxicity of COREXIT for environmental isolates are summarized in Fig. 1. The addition of COREXIT resulted in varying levels of reduction in cell viability, and an increase in cell numbers in some cases. A reduction in live cells was seen for all isolates at all concentrations tested with the Table 1. Summary of isolates obtained from oil and oil-water samples and LH-PCR (length heterogeneity PCR) analysis of samples and isolates Isolate Phylogenetic group Nearest Alignment Amplicon Hydrocarbon Cumulative peak relative (%) length degrader abundance (%) (GenBank (bp) Oil-water Oil Accession no.) Acinetobacter venetianus Exiguobacterium arabatum Marinobacter hydrocarbonoclasticus Pseudidiomarina sp. Pseudoalteromonas sp. Pseudomonas pseudoalcaligenes Shewanella algae Vibrio sp. Unmatched amplicon Unmatched amplicon Unmatched amplicon Unmatched amplicon Unmatched amplicon Unmatched amplicon Unmatched amplicon Gammaproteobacteria Bacillales Gammaproteobacteria AM909651 FM203124 GQ901055 99 99 100 338 359 341 Yes No Yes 4 <1 12 4 0 12 Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria EF212001 GQ245921 EU440977 96 99 99 357 344 336 No No Yes <1 17 <1 <1 13 <1 Gammaproteobacteria Gammaproteobacteria DQ386137 EU834012 98 99 346 353/354 302 310 312 320 322 327 329 No Yes Unknown Unknown Unknown Unknown Unknown Unknown Unknown 5 31 0 5 9 1 10 2 1 10 32 2 3 16 0 3 2 3 Hamdan & Fulmer: Effects of COREXIT® EC9500A on bacteria 105 Fig. 1. Toxicity of COREXIT® EC9500A to isolated strains. Dispersant dilutions are 1:10 to 1:100 000 w/v. Values in parentheses are the concentration of dispersant in each treatment. Percentages of live cells in dispersant-amended cultures are relative to the controls. Error bars are ±1 SD exception of Vibrio sp. Dilutions of COREXIT of 1:10 and 1:100 w/v in Marine broth resulted in near total cell death for all isolates (≥99%). However, at the 1:1000 dilution, corresponding to the addition of 0.964 mg ml–1 COREXIT, there was an increase in live cells of Vibrio sp. compared to the control, and a decrease in live cells (40 to 90%) in all other cultures (relative to controls). To address the effects of COREXIT on hydrocarbon degraders, dilutions of COREXIT were added to minimal medium supplemented with hexadecane as a carbon source. COREXIT was added at the concentrations suggested by the Environmental Protection Agency (EPA), i.e. ratios of 1:10, 1:25 and 1:50 COREXIT:hexadecane (US EPA 1995); the results are summarized in Fig. 2. Such ratios of COREXIT to hydrocarbons may have been encountered during surface and deepwater application in the response to the spill, but dispersant concentrations may have been significantly lower within the water column and on affected beaches. Thus, we suggest that these concentrations represent the maximum encountered in the environment. As before, the addition of COREXIT was toxic to all hydrocarbon degraders in a dose-dependent manner. A second experiment was conducted to determine the impact of COREXIT on heterotrophic secondary production. The results of this experiment are summarized in Fig. 3. At t = 0 h in all 3 cultures, there were statistically significant differences between treatments and controls. In the Acinetobacter venetianus and Marinobacter hydrocarbonoclasticus cultures, addition of the dispersant moderately stimulated production compared to the controls. In the Vibrio sp. culture, a Fig. 2. Toxicity of COREXIT® EC9500A to hydrocarbon degraders when present at concentrations suggested by the Environmental Protection Agency (EPA). Dispersant dilutions are 1:10, 1:25 and 1:50 ratios of COREXIT to hexadecane. Values in parentheses are the concentration of dispersant in each treatment. Error bars are ±1 SD 106 Aquat Microb Ecol 63: 101–109, 2011 8x109 Control + COREXIT® EC9500A (1:1000) + COREXIT® EC9500A (1:100) A 6x109 4x109 2x109 Heterotrophic secondary production (cells l–1 d–1) 0 0 8x1010 10 20 30 40 50 DISCUSSION B 6x1010 4x1010 2x1010 0 0 10 20 30 40 50 10 20 30 40 50 C 3x1010 2x1010 1x1010 0 0 hydrocarbonoclasticus remained similar in all treatments until t = 12 h. After t = 12 h, a biofilm appeared slightly above the medium in treatments containing COREXIT, and production declined significantly, while in the control it remained consistent for the duration of the experiment. In Vibrio sp., production in all treatments declined during the first 12 h of the experiment. After t = 12 h, a concentration-dependent decline in production was observed in the dispersant treatments while the control rebounded. A biofilm was also observed in COREXIT-amended Vibrio sp. tubes after t = 12 h. Time (h) Fig. 3. Time course experiment tracking heterotrophic bacterial production in 3 cultures: (A) Acinetobacter venetianus, (B) Marinobacter hydrocarbonoclasticus, and (C) Vibrio sp. No sample for t = 48 h for A. venetianus + dispersant at 1:100 was available due to evaporative loss. Error bars are ±1 SD moderate decline in production was observed at t = 0 h. At t = 12 h COREXIT clearly inhibited A. venetianus secondary production compared to the controls; such inhibition remained significant in the highest dilution of COREXIT until t = 24 h. Interestingly, in the 1:1000 dispersant dilution, production of A. venetianus rebounded to the level of the control at t = 24 h and exceeded the control at t = 48 h. Production by M. LH-PCR revealed a relatively low-diversity population. This is not surprising given the nature of the samples and that others document low diversity in oil-rich environments (Orcutt et al. 2010). Experiments to determine microbial responses to oil contamination using pristine marine sediments demonstrate rapid increases in the abundance of hydrocarbon degraders (Brakstad 2008). Thus, we hypothesized that the environmental samples would be enriched in microorganisms capable of hydrocarbon degradation. Over 80 bacterial genera have been confirmed to degrade hydrocarbons (Head et al. 2006, Brakstad 2008). The most commonly observed hydrocarbon degraders in aquatic environments belong to the Gammaproteobacteria and include the genera Pseudomonas, Acinetobacter, Marinobacter and Alcanivorax (Atlas 1981, Venkateswaran et al. 1991, Head et al. 2006). Fewer reports indicate that Vibrio spp. are involved in hydrocarbon degradation (Atlas 1981, Venkateswaran et al. 1991). Virtual alignment of LH-PCR amplicons with isolate amplicons reveals that a total of 17% of peak abundance was associated with 3 known hydrocarbon degraders (Table 1). The most abundant of these was Marinobacter hydrocarbonoclasticus. The M. hydrocarbonoclasticus isolate exhibited 100% homology with an isolate obtained from coral tissue, infected with Black Band Disease, obtained proximate to the Gulf of Mexico (Richardson et al. 2009). M. hydrocarbonoclasticus can degrade a variety of aliphatic and aromatic hydrocarbons and produce a nondialyzable bioemulsifier when grown on hydrocarbons (Gauthier et al. 1992). One of the main features that distinguish the genus Marinobacter from closely related genera is the ability of these bacteria to tolerate high levels of salt and to grow at temperatures up to 45°C. These factors suggest that the location in which the sample was found — and high ambient temperatures on the beach at the time of collection — may have selected for this isolate. Hamdan & Fulmer: Effects of COREXIT® EC9500A on bacteria The second most abundant hydrocarbon degrader in the environmental fingerprint was related 99% to Acinetobacter venetianus. Other studies have demonstrated that A. venetianus proliferates in oil-degrading consortia and is capable of metabolizing complex marine hydrocarbon mixtures (Vaneechoutte et al. 2009). Some strains of A. venetianus produce bioemulsifiers to aid in hydrocarbon metabolism. Pseudomonas pseudoalcaligenes, a confirmed naphthalene degrader (Garcia-Valdes et al. 1988), was detected in the LHPCR profile, although its amplicon accounted for less than 1% of peak abundance. Vibrio sp. accounted for the majority of LH-PCR peak abundance. The Vibrio isolate was 99% related to 4 Vibrio phylotypes: Vibrio natriegens, alginolyticus, fluvialis and the pathogenic vulnificus. All GenBank entries which matched the Vibrio sp. isolate are capable of rapid growth, with generation times of less than 10 min (Aiyar et al. 2002). Each is capable of forming biofilms, which may explain the biofilm observed in the production tubes. V. fluvialis and V. natriegens have been identified in hydrocarbon-degrading communities in experimental seawater mesocosms (Venkateswaran et al. 1991); however, their role as hydrocarbon degraders is not well recognized. In experimental studies, V. natriegens was capable of metabolizing insoluble surfactants at air–water interfaces (Salter et al. 2009). This ability to metabolize surfactants may explain why the Vibrio sp. isolate showed greater viability in the presence of COREXIT compared to other phylotypes in this study. At some concentrations, all isolates were susceptible to impairment as a result of exposure to COREXIT. The addition of dispersant, at all concentrations, resulted in a significant reduction in the live:dead cell ratio in all isolates tested — with the exception of Vibrio sp., which, of the organisms examined, appear to be the most tolerant to COREXIT (Figs. 1 & 2). Because of the importance of microorganisms to the natural attenuation of hydrocarbons, toxicity tests were focused on hydrocarbon-degrading isolates. COREXIT at concentrations <1 mg ml–1 was capable of killing at least 60% of cells in cultures of the 2 most abundant hydrocarbon-degrading isolates observed in the LH-PCR profile (Fig. 2). This is a significant finding given that the 1:50 dilution, ~0.2 mg ml–1, is within the concentration range for COREXIT® EC9500A classified as ‘practically non-toxic’ to the standard toxicity test organism Menidia beryllina (US EPA 2010a). The result for M. beryllina was included in a recent Environmental Protection Agency (EPA) report identifying the environmental hazards associated with the use of National Contingency Plan (NCP)-approved dispersants during the Deepwater Horizon response. The results of the current study demonstrate that, at concentrations which do not present a significant hazard 107 to some adult test organisms, dispersants may be highly toxic to communities directly involved in natural hydrocarbon bioremediation. In the second live:dead experiment, in which hexadecane was added to growth media (Fig. 2), we confirm that cell death results from the application of the dispersant, not from limitation of hydrocarbons. Interestingly, in the presence of hexadecane, Pseudomonas pseudoalcaligenes exhibited the greatest resistance to COREXIT toxicity, suggesting that the presence of hydrocarbons may reduce the impact of COREXIT on some populations. In microcosm studies conducted on Alaskan North Slope crude oil treated with COREXIT® EC9500A, the dispersant did increase the surface area of crude oil droplets and enhanced microbial droplet colonization (Lindstrom & Braddock 2002). However, the dispersants resulted in a negligible increase in biodegradation of oil when compared to non-dispersed oil. COREXIT can be a highly labile substrate for microbial growth and metabolism (Zahed et al. 2010), and increases in carbon mineralization in oil samples may be attributed to mineralization of the chemical dispersant alone (Lindstrom & Braddock 2002, Brakstad 2008). This may explain the results observed for Acinetobacter venetianus where, after t = 24 h, production in the highest dilution of COREXIT rebounded to levels that exceeded the control (Fig. 3). However, in the case of Marinobacter hydrocarbonoclasticus, no such recovery occurred. In this respect, COREXIT may impart positive effects on hydrocarbon-degrading cultures capable of withstanding its initial toxicity. Because some Vibrio spp. are capable of metabolizing dispersant (Salter et al. 2009), it is reasonable to hypothesize that the increase in viability in cultures of Vibrio sp. in the live/dead staining analysis may be due to metabolism of COREXIT. However, even within this group, a reduction in heterotrophic secondary production was observed in the presence of COREXIT (Fig. 3), although the impact on Vibrio sp. was less significant than that on M. hydrocarbonoclasticus. The high turnover rates observed for numerous Vibrio spp. may likewise explain the reduced toxicity of COREXIT over other isolates. CONCLUSIONS Hydrocarbon degradation in the marine environment is dependent on the ability of microorganisms to utilize hydrocarbons for growth and metabolism. The results of the current study demonstrate that microbial populations are susceptible to toxicity from the use of COREXIT® EC9500A when applied at prescribed concentrations. While the short-term goals of dispersants Aquat Microb Ecol 63: 101–109, 2011 108 for counting bacteria by fluorescence microscopy. Appl may be achieved, this study provides evidence that Environ Microbiol 33:1225–1228 COREXIT® EC9500A differentially impacts hydrocarHollaway SL, Faw GM, Sizemore RK (1980) The bacterial ➤ bon-degrading microorganisms. Although toxicity community composition of an active oil field in the northtesting with bacteria is only a minor component of the western Gulf of Mexico. Mar Pollut Bull 11:153–156 development of new dispersants, these experiments ➤ Joye SB, Boetius A, Orcutt BN, Montoya JP, Schulz HN, Erickson MJ, Lugo SK (2004) The anaerobic oxidation of demonstrate the importance of understanding the methane and sulfate reduction in sediments from Gulf of impact of dispersants on microbial communities, as Mexico cold seeps. Chem Geol 205:219–238 there is potential to diminish the capacity of the ➤ Jung JH, Yim UH, Han GM, Shim WJ (2009) Biochemical environment to mitigate spills. changes in rockfish, Sebastes schlegeli, exposed to disAcknowledgements. We thank W. Wood (NRL) for acquiring the samples used in this study, P. Gillevet and M. Sikaroodi (George Mason University) for assistance with LH-PCR analysis, B. Ringeisen (NRL) for helpful discussions and support of this work, R. Jonas (George Mason University) for comments on the draft manuscript and L. Tender (NRL) for providing the dispersant used in these experiments. We acknowledge the Naval Research Laboratory base program for support of this project. ➤ LITERATURE CITED ➤ Aharon ➤ ➤ ➤ ➤ ➤ ➤ ➤ ➤ P, Fu B (2000) Microbial sulfate reduction rates and sulfur and oxygen isotope fractionations at oil and gas seeps in deepwater Gulf of Mexico. Geochim Cosmochim Acta 64:233–246 Aiyar SE, Gaal T, Gourse RL (2002) rRNA promoter activity in the fast-growing bacterium Vibrio natriegens. J Bacteriol 184:1349–1358 Atlas RM (1981) Microbial degradation of petroleum hydrocarbons — an environmental perspective. Microbiol Rev 45:180–209 Brakstad OG (2008) Natural and stimulated biodegradation of petroleum in cold marine environments. In: Margesin R, Schinner F, Marx JC, Gerday C (eds) Psychrophiles: from biodiversity to biotechnology. Springer Verlag, Berlin, p 389–407 Coastal Response Research Center (2010) Deepwater Horizon dispersant use. Meeting Report. University of New Hampshire, Durham, NH Couillard CM, Lee K, Legare B, King TL (2005) Effect of dispersant on the composition of the water-accommodated fraction of crude oil and its toxicity to larval marine fish. Environ Toxicol Chem 24:1496–1504 Garcia-Valdes E, Cozar E, Rotger R, Lalucat J, Ursing J (1988) New naphthalene-degrading marine Pseudomonas strains. Appl Environ Microbiol 54:2478–2485 Gauthier MJ, Lafay B, Christen R, Fernandez L, Acquaviva M, Bonin P, Bertrand JC (1992) Marinobacter hydrocarbonoclasticus gen. nov., sp. nov., a new, extremely halotolerant, hydrocarbon-degrading marine bacterium. Int J Syst Bacteriol 42:568–576 Hamdan LJ, Jonas RB (2006) Seasonal and interannual dynamics of free-living bacterioplankton and microbially labile organic carbon along the salinity gradient of the Potomac River. Estuar Coast 29:40–53 Hamdan LJ, Gillevet PM, Sikaroodi M, Pohlman JW, Plummer RE, Coffin RB (2008) Geomicrobial characterization of gas hydrate-bearing sediments along the mid-Chilean margin. FEMS Microbiol Ecol 65:15–30 Head IM, Jones DM, Roling WFM (2006) Marine microorganisms make a meal of oil. Nat Rev Microbiol 4:173–182 Hobbie JE, Daley RJ, Jasper S (1977) Use of Nuclepore filters ➤ ➤ ä ➤ persed crude oil. Comp Biochem Physiol Part C: Toxicol Pharmacol 150:218–223 Labson VF, Clark RN, Swayze GA, Hoefen TM and others (2010) Estimated lower bound for leak rates from the Deepwater Horizon spill—Interim report to the Flow Rate Technical Group from the Mass Balance Team. U.S. Geological Survey Open-File Report 2010–1132 Lanoil BD, Sassen R, La Duc MT, Sweet ST, Nealson KH (2001) Bacteria and archaea physically associated with Gulf of Mexico gas hydrates. Appl Environ Microbiol 67: 5143–5153 Leahy JG, Colwell RR (1990) Microbial degradation of hydrocarbons in the environment. Microbiol Mol Biol Rev 54: 305–315 Lindstrom JE, Braddock JF (2002) Biodegradation of petroleum hydrocarbons at low temperature in the presence of the dispersant Corexit 9500. Mar Pollut Bull 44: 739–747 Litchfield CD, Sikaroodi M, Gillevet PM (2005) The microbial diversity of a solar saltern on San Francisco Bay. In: Gunde-Cimerman N, Oren A, Plemenita$ A (eds) Adaptation to life at high salt concentrations in Archaea, Bacteria, and Eukarya. Springer, Dordrecht, p 59–69 Lloyd KG, Albert DB, Biddle JF, Chanton JP, Pizarro O, Teske A (2010) Spatial structure and activity of sedimentary microbial communities underlying a Beggiatoa spp. mat in a Gulf of Mexico hydrocarbon seep. PLoS ONE 5(1):e8738 Mills HJ, Hodges C, Wilson K, MacDonald IR, Sobecky PA (2003) Microbial diversity in sediments associated with surface-breaching gas hydrate mounds in the Gulf of Mexico. FEMS Microbiol Ecol 46:39–52 National Research Council Committee on Oil in the Sea (2003) Oil in the sea III: inputs, fates and effects. National Academies Press, Washington, DC Orcutt BN, Joye SB, Kleindienst S, Knittel K and others (2010) Impact of natural oil and higher hydrocarbons on microbial diversity, distribution, and activity in Gulf of Mexico cold-seep sediments. Deep-Sea Res II 57:2008–2021 Reed AJ, Lutz R, Vetriani C (2006) Vertical distribution and diversity of bacteria and archaea in sulfide and methanerich cold seep sediments located at the base of the Florida Escarpment. Extremophiles 10:199–211 Richardson LL, Miller AW, Broderick E, Kaczmarsky L, Gantar M, Stanić D, Sekar R (2009) Sulfide, microcystin, and the etiology of black band disease. Dis Aquat Org 87: 79–90 Salter I, Zubkov MV, Warwick PE, Burkill PH (2009) Marine bacterioplankton can increase evaporation and gas transfer by metabolizing insoluble surfactants from the air–seawater interface. FEMS Microbiol Lett 294:225–231 Smith DC, Azam F (1992) A simple, economical method for measuring bacterial protein synthesis rates in seawater using 3H-leucine. Mar Microb Food Webs 6:107–114 US EPA (U.S. Environmental Protection Agency) (1995) COREXIT® EC9500A NCP Product schedule. http://www. epa.gov/osweroe1/content/ncp/products/corex950.htm Hamdan & Fulmer: Effects of COREXIT® EC9500A on bacteria ➤ US EPA (2010a) Comparative toxicity of eight oil dispersant products on two Gulf of Mexico aquatic test species. http:// www.epa.gov/bpspill/reports/phase2dispersant-toxtest.pdf US EPA (2010b) National Contingency Plan product schedule. http://www.epa.gov/emergencies/content/ncp/product_schedule.htm Vaneechoutte M, Nemec A, Musilek M, van der Reijden TJK and others (2009) Description of Acinetobacter venetianus ex Di Cello et al. 1997 sp. nov. Int J Syst Editorial responsibility: Antje Boetius, Bremen, Germany 109 Evol Microbiol 59:1376–1381 ➤ Venkateswaran K, Iwabuchi T, Matsui Y, Toki H, Hamada E, ➤ Tanaka H (1991) Distribution and biodegradation potential of oil-degrading bacteria in North Eastern Japanese coastal waters. FEMS Microbiol Lett 86:113–121 Zahed MA, Aziz H, Isa M, Mohajeri L (2010) Effect of initial oil concentration and dispersant on crude oil biodegradation in contaminated seawater. Bull Environ Contam Toxicol 84:438–442 Submitted: August 25, 2010; Accepted: December 9, 2010 Proofs received from author(s): February 18, 2011 Home Search Collections Journals About Contact us My IOPscience Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill This article has been downloaded from IOPscience. Please scroll down to see the full text article. 2010 Environ. Res. Lett. 5 045301 (http://iopscience.iop.org/1748-9326/5/4/045301) View the table of contents for this issue, or go to the journal homepage for more Download details: IP Address: 193.191.134.1 The article was downloaded on 19/04/2011 at 10:24 Please note that terms and conditions apply. IOP PUBLISHING ENVIRONMENTAL RESEARCH LETTERS Environ. Res. Lett. 5 (2010) 045301 (6pp) doi:10.1088/1748-9326/5/4/045301 Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill William M Graham1,2,3 , Robert H Condon1, Ruth H Carmichael1,2, Isabella D’Ambra1,2, Heather K Patterson1,2, Laura J Linn1 and Frank J Hernandez Jr1,2 1 2 Dauphin Island Sea Lab, Dauphin Island, AL 36528, USA Department of Marine Sciences, University of South Alabama, Mobile, AL 36688, USA E-mail: [email protected] Received 27 September 2010 Accepted for publication 1 November 2010 Published 8 November 2010 Online at stacks.iop.org/ERL/5/045301 Abstract The Deepwater Horizon oil spill was unprecedented in total loading of petroleum hydrocarbons accidentally released to a marine ecosystem. Controversial application of chemical dispersants presumably accelerated microbial consumption of oil components, especially in warm Gulf of Mexico surface waters. We employed δ 13 C as a tracer of oil-derived carbon to resolve two periods of isotopic carbon depletion in two plankton size classes. Carbon depletion was coincident with the arrival of surface oil slicks in the far northern Gulf, and demonstrated that subsurface oil carbon was incorporated into the plankton food web. Keywords: zooplankton, petroleum hydrocarbon, stable isotope, Gulf of Mexico S Online supplementary data available from stacks.iop.org/ERL/5/045301/mmedia present quantitative data collected as a rapid-response effort to track carbon isotopic signal in two size classes representing a pathway into the bulk zooplankton community of the northern Gulf of Mexico. The present study reflects only the initial steps of a larger and continuing laboratory experimental and field effort aimed at understanding how oil affects pelagic communities of the northern Gulf and vice versa effects of the biological community on the fate of the oil. 1. Introduction Following the sinking of Deepwater Horizon (DWH) on 22 April 2010, an estimated 780 000 m3 of Sweet Louisiana Crude (SLC) and 205 000 mT of methane [1] were released into the northern Gulf of Mexico over an 85 d period. General agreement exists that ∼25% was directly recovered or burned at sea, leaving ∼75% to be degraded naturally or with the aid of chemical dispersants [2]. Recent publications document the scope of deep subsea oil and methane along the northern Gulf slope [1, 3, 4], but scant evidence exists for the presence of subsea oil in warm (>25 ◦ C), shallow shelf waters. A large pool of isotopically depleted carbon from dispersed oil and methane is presumably available for biological consumption via prokaryotic consumers [5]. Isotopic depletion extending into marine zooplankton grazers, a pathway mediated by the microbial food web [6], is a good proxy for food web modification by the spill. Here we 2. Methodology We employed δ 13 C as a tracer of oil-derived carbon incorporation into the lower marine food web across the middle and inner continental shelf. During June–August 2010, we followed two plankton size classes: the nominally 1 μm– 0.2 mm ‘small suspended particulate’ and the >0.2–2 mm ‘mesozooplankton’ fractions, with the former considered likely food for the latter. The study region, >100 km north of the DWH well head, had three defined northward pulses of surface 3 Author to whom any correspondence should be addressed. 1748-9326/10/045301+06$30.00 1 © 2010 IOP Publishing Ltd Printed in the UK Environ. Res. Lett. 5 (2010) 045301 W M Graham et al Table 1. The location and depth of surface and deep sampling stations in the Gulf of Mexico and reference sites within Mobile Bay, AL, including mean (±SD) salinity at each depth. n.d. = no data, dash = not deep enough to collect separate Deep sample. Symbols next to station names are for reference to figure 1. Surface Station, symbol Latitude Gulf sites T35, 29.7989 T20, 30.0902 T10, 30.1609 Buoy M (BM), 30.1306 Mobile Bay reference sites Cedar Point Reef, + 30.3256 Sand Reef, − 30.2772 • Longitude Deep Depth (m) Mean salinity (psu) Depth (m) Mean salinity −88.2083 −88.2116 −88.1229 −88.1097 1 1 1 1 27 ± 3 25 ± 3 24 ± 3 23 ± 4 33 18 8 15 36 ± 0 33 ± 4 32 ± 3 n.d. −88.1327 −88.1052 1 1 12 ± 6 18 ± 6 — — — — oil: two prior to the well’s 15 July shut-in, and one in August (figures 1(A), (B), movie S1 available at stacks.iop.org/ERL/ 5/045301/mmedia). Samples were collected in surface and bottom waters of the Gulf of Mexico at four shelf and two inner Mobile Bay reference sites (figure 1(A), table 1). 2.2. Stable isotope analysis Bulk carbon (C) stable isotope ratios (δ 13 C, ) were measured by continuous flow isotope ratio mass spectrometry at the University of Utah Stable Isotope Facility (USA) and on a Picarro cavity ringdown spectrometer coupled to a Costech elemental analyzer at Dauphin Island Sea Lab (DISL). Bulk carbon isotopic composition in organisms reflects both shortterm energy stores (i.e., lipids) and relatively longer turnover in tissues [7]. Since lipids are isotopically depleted and do not necessarily reflect time-integrated diet of organisms, variation in lipid content may introduce bias into stable isotope analyses. Established mathematical normalization techniques allow correction of δ 13 C values in lipid-rich samples, but preserve sample integrity for other analyses [7]. Here, bulk δ 13 C values in mesozooplankton were lipid-corrected according to [7], after comparison to C:N in mesozooplankton samples (figure 2(A). Comparison of δ 13 C values to the C:N and relative C content (mg l−1 ) in suspended particulates indicated no correction was needed for the smaller fraction (figures 2(A) and (B)). C and N content were obtained during stable isotope analysis. 2.1. Plankton and suspended particle collection Suspended particulates (1 μm–0.2 mm) were collected using 1.7 l vertical Niskin bottles deployed at target depths of one meter above the bottom and one meter below the surface at stations T10, T20, and T35 at two-week intervals from 2 June to 15 August, 2010. At Buoy M, and the two bay reference sites, Cedar Point Reef and Sand Reef, only water from one meter below the surface was collected owing to shallowness of the water column; these were collected at 1–2 week intervals. At the Sand Reef reference station, samples were collected using a 1 l horizontal sampler. Air-filled balloons, released at depth from bottle ends, were used to avoid surface oil contamination during sampling. Water was vacuum filtered (5 psi) onto pre-combusted GF/F filters, and dried to a constant weight at 60 ◦ C. Mesozooplankton were collected at the same stations and with the same timing as above with opening–closing plankton nets (3.5 m long, 0.25 m diameter), using 333 μm for surface and bottom and 202 μm for oblique samples at Gulf sites. A 202 μm mesh (1 m long, 0.5 m diameter) ring plankton net was used at BM and reference sites (mesozooplankton were not collected at Sand Reef). Samples were rinsed with ultrapure water, dried at 60 ◦ C, and homogenized by mortar and pestle. Additional historical pre-spill mesozooplankton samples from Buoy M and suspended particulate samples from T35, T20 and Buoy M were collected in May–August of both 2008 and 2009 (all data within each station were pooled as ‘prespill’. Collections were similar to those described for 2010. These samples were historically analyzed only for carbon (C) and nitrogen (N) content; no pre-spill stable isotope data exist for continental shelf stations. All mesozooplankton samples are being processed for community assemblage changes with respect to the spill; however, assemblage analysis is beyond the scope of this study. That said, a cursory review of the samples for presence of contaminating oil droplets revealed the samples were clear of both oil and resuspended sediments. In addition, the zooplankton samples were dominated by organisms typical of spring–summer assemblages such as calanoid copepods. 2.3. Source crude oil samples We analyzed δ 13 C of oil in both weathered and fresh condition. Weathered surface oil was collected in the nearshore waters off Dauphin Island, Alabama, on 11 June 2010, and stored in the dark at 5 ◦ C. Prior to carbon stable isotope analysis, the weathered oil was further dried at 60 ◦ C for 48 h to remove residual water. This sample was analyzed for δ 13 C at the University of Utah Stable Isotope Facility. Fresh Source Oil B (SOB) supplied by BP was collected on 22–23 May 2010, from the riser insertion tube on board the drill ship Enterprise. Accompanying documentation for two samples (SOB-20100716-067 and SOB-20100716-130) reported Nalco EC9323A defoamer was injected topsides, and subsea injections included methanol with 10 000 ppm VX9831 oxygen scavenger/catalysts solution. Fresh oil was stored at 5 ◦ C until analysis by saturating a small piece of precombusted GF/F filter with oil and analyzed using a Picarro cavity ringdown spectrometer coupled to a Costech elemental analyzer at DISL. 2 Environ. Res. Lett. 5 (2010) 045301 W M Graham et al Figure 1. (A) Sample sites, symbols relating to (D) and (E); box defines area used to calculate oil % coverage in (B) with example of peak oil coverage 28 June 2010. A full animation of oil movement around these sites can be found in the supplementary data (movie S1 available at stacks.iop.org/ERL/5/045301/mmedia). (B) Timing of three shoreward pulses of oil, I–III (cf movie S1 available at stacks.iop.org/ERL/5/ 045301/mmedia). (C) Daily averaged river discharge into Mobile Bay. (D) δ 13 C values for mesozooplankton fraction (0.2–2 mm). (E) δ 13 C values for suspended particulate fraction (1 μm–0.2 mm). Both (D) and (E) referenced against δ 13 C values from Mobile Bay and weathered and fresh SLC oil. Error bars show standard deviation. 3 Environ. Res. Lett. 5 (2010) 045301 W M Graham et al Figure 2. Analysis of C stable isotope ratios. (A) Bulk δ 13 C in mesozooplankton (0.2–2 mm) and suspended particulates (1 μm–0.2 mm) compared to C:N in surface, deep, and oblique (mesozooplankton only) samples. (B) δ 13 C compared to carbon content in suspended particulates. (C) Corrected δ 13 C in mesozooplankton and δ 13 C in smaller suspended particulates compared to salinity at surface and deep sampling locations. (D) Bulk δ 13 C in suspended particulates compared to chlorophyll a (chl a ) in water samples from stations T10, T20, and T35. 2.4. Chlorophyll a Claiborne Dam on the Alabama River (# 02429500). There is an estimated 5–10 d flow-dependent lag between readings at gauging stations and reference sites in Mobile Bay [9]. Daily mean discharge is reported in figure 1(C) (not including lags as it did not change the pattern). Salinity was measured at Gulf stations T10–T35 using a Sea Bird SBE 25 conductivity– temperature–depth (CTD) probe and at Buoy M and the two Mobile Bay reference sites using a handheld YSI 85 salinity probe. Whole water samples collected at Stations T10, T20 and T35 (1 m below surface and 1 m above bottom) were stored on ice in the dark and filtered in the laboratory onto GF/F filters within 2 h of collection. Extracted chlorophyll a was fluorometrically determined with a Turner Designs fluorometer [8]. 2.5. Oil proximity data Surface layer slick distribution was defined from Geographical Information System (GIS) data (ftp://satepsanone.nesdis.noaa. gov/OMS/disasters/DeepwaterHorizon/). Per cent coverage of oil and distance from the nearest slick to station T20 were measured by mapping oil layers in ESRI ArcMap v9.3 GIS software. 2.7. Statistical analyses Analysis of variance (ANOVA) was used for comparison of δ 13 C values, C:N, and chl a among sample types. If ANOVAs were significant, post hoc pairwise comparison of means using Tukey’s test of variability were performed. All linear correlations were tested using the Z-test. These analyses were performed in StatView 5.0.1. An additional three-way ANOVA with location, sample depth and condition (pre- or post-spill years) was performed on C:N values for suspended particulates and mesozooplankton using Minitab 15. All data were normally distributed and did not require transformation, and also conformed to the assumption of homogeneity of variance. 2.6. Freshwater discharge To determine potential effects of freshwater discharge on δ 13 C, riverine discharge into Mobile Bay was estimated by adding daily discharge rates at two US Geological Survey gauging stations (http://waterdata.usgs.gov/usa/nwis/sw), the Coffeeville Dam on the Tombigbee River (#02469765) and the 4 Environ. Res. Lett. 5 (2010) 045301 W M Graham et al 3. Results and discussion suspended particles, and some primary consumers in Mobile Bay and elsewhere [13, 14]. δ 13 C values in mesozooplankton were not related to salinity (figure 2(C)), and δ 13 C in suspended particulates showed a weak positive correlation only when surface and bottom samples were considered together (r = 0.37, P = 0.03). This finding is consistent with the relatively low discharge to Mobile Bay during most of the sampling period (figure 1(C)). The similar patterns of δ 13 C depletion in both mesozooplankton and small suspended fractions, despite decreasing discharge to Mobile Bay and little or no relationship to salinity during the period of greatest oil proximity and coverage in the region, supports an oil-derived C source mediating this shift as opposed to a freshwater-derived source from Mobile Bay. Since phytoplankton were a component of the mixed small suspended particulate fraction (1 μm–0.2 mm), isotopic depletion of C in this fraction and subsequently in the mesozooplankton could result from dominance by isotopically depleted phytoplankton. Given the range of δ 13 C values typical in marine phytoplankton (−20 to −24; [15, 16]) and the lack of evidence for a significant freshwater influence during the period of depletion, phytoplankton alone are unlikely to account for the observed depletion (figures 1(D), (E)). The concentration of chlorophyll pigments extracted from water samples collected at stations T10 through T35 varied relatively little during the study period and did not indicate a bloom of shelf phytoplankton (1.50 ± 1.02 μg l−1 ; figure 2(D)). Chl a concentration also was not related to δ 13 C of the suspended particulate fraction ( Freg1,24 = 2.33, P = 0.14). Surface and bottom samples were not different and were combined for further analysis (two-way ANOVA with station and depth as variables: F3,26 = 1.30, P = 0.29) (figure 2), indicting that a change in phytoplankton abundance made no measurable contribution to the isotopic depletion of C shown in figure 1. To elucidate whether the observed δ 13 C depletion was due to the contaminating presence of oil in the water column or to the assimilation and incorporation of oil-derived C by resident biota, we compared C:N for mesozooplankton and suspended particulate fractions during pre- (2008–2009) and post-oil spill (2010) years between May and August. The expectation was that presence of SLC oil on or inside the animals would yield anomalously high C:N values. For the suspended particulate fraction, there was no difference in C:N (by weight) between any of the sampling stations within and across pre- and postspill years (three-way ANOVA with station, pre- and postspill years, and sample depth as variables: F9,89 = 1.46, P = 0.18) (figure 3). Similarly, there was no significant difference in mesozooplankton C:N at station BM in pre- and post-oil spill years (ANOVA: F1,17 = 0.69, P = 0.42), and while we do not have pre-spill C:N data for stations T35, T20 and T10, post-spill C:N data from these sites were the same as those at station BM (ANOVA: F1,68 = 2.30, P = 0.09) (figure 3). Combined, these results suggest that the depleted C isotope values were not driven by direct oil contamination in the samples (e.g., oil micro-droplets collected on the filter). That similar results were found for both mesozooplankton and suspended particulate fractions suggests oil-derived C was transferred through the food web. δ 13 C depletion occurred in each size fraction at middle and inner shelf stations coincident with two sequential northward pulses of surface oil slicks from DWH (figures 1(D), (E)). Relative to early June, an isotopic shift of −1 to −4 (toward weathered and fresh oil, −27.23 ± 0.03 and −27.34 ± 0.34, respectively) occurred during the peak of areal coverage of oil over the sites (figures 1(B), (D), (E)). Recovery from this depletion to the pre-spill baseline was 2– 4 wks. A third pulse of residual oil occurred in late July, and depleted δ 13 C was observed in mid-August at the furthest offshore stations. Depletion and recovery cycles on the order of a few weeks are consistent with published warm water petroleum hydrocarbon decay timescales [10]. The apparent oil-related δ 13 C depletion occurred in both fractions and throughout the water column. The pattern was consistent despite the differences in both δ 13 C and C:N between the two size fractions. δ 13 C and C:N values differed between mesozooplankton and suspended particulates, with mesozooplankton having heavier δ 13 C (−20.44 ± 1.37 compared to −23.19 ± 1.26) and lower C:N (4.8 ± 0.6 compared to 6.9 ± 0.9) than suspended particles (ANOVA: δ 13 C:F4,84 = 23.29, P < 0.001; C:N:F4,84 = 43.04, P < 0.001; figure 2(A). Comparisons among surface, bottom, and oblique samples did not differ for either size fraction (Tukey’s post hoc test: P > 0.05 for all comparisons). Bulk δ 13 C values were correlated with C:N in surface (r = −0.64), bottom (r = −0.61), and oblique (r = −0.63) mesozooplankton samples ( P < 0.01) for all comparisons; figure 2(A), consistent with a mesozooplankton fraction (largely composed of animals) and requiring correction for lipid-related depletion of δ 13 C [7]. In contrast, δ 13 C in suspended particulates was not correlated with C:N (figure 2(A)), and was weakly correlated with the relative C content in the sample only when surface and bottom fractions were considered together (r = −0.44, P < 0.01; figure 2(B)). These findings suggest a small particle fraction of mixed composition, including algal and detrital matter that did not demand lipid correction despite a higher C:N ratio than mesozooplankton [7, 11]. The mean correction applied to bulk δ 13 C in mesozooplankton samples was 1.49 ± 1.73. The relative shift in sample values can be seen by comparing panels A and C in figure 2. 3.1. Discounting masking effects or sample contamination In comparison to reference sites inside Mobile Bay, offshore depletion of δ 13 C was not related to timing of freshwater discharge from the Bay, phytoplankton blooms, or direct contamination of samples with external oil. Corrected δ 13 C in mesozooplankton and bulk δ 13 C values in suspended particulates were compared to salinity at each station to detect potential freshwater influence (figure 2(C)). The hydrology of Mobile Bay is dominated by freshwater inputs, which lead to salinity stratification [12] and may convey isotopically light suspended particles and biota from the upper reaches of the Bay to the Gulf [13]. δ 13 C in reference oil samples was similar to δ 13 C typically found in freshwater-derived vegetation, 5 Environ. Res. Lett. 5 (2010) 045301 W M Graham et al the Richard C Shelby Center for Ecosystem-Based Fisheries Management and the Northern Gulf Institute’s rapid-response funding from BP’s Gulf Research Initiative. We thank J Burchfield, M Miller, K Weis, J Hermann, B Dzwonkowski, K Park, K Robinson, L McCallister and the crew of the R/V E O Wilson for laboratory, field and logistical support. References [1] Adcroft A, Hallberg R, Dunne J P, Samuels B L, Galt J A, Barker C H and Payton D 2010 Geophys. Res. Lett. 37 L18605 [2] Kerr R A 2010 A lot of oil on the loose, not so much to be found Science 329 734–5 [3] Camilli R, Reddy C M, Yoerger D R, Van Mooy B A S, Jakuba M V, Kinsey J C, McIntyre C P, Sylva S P and Maloney J V 2010 Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon Science 330 201–4 [4] Hazen T C et al 2010 Deep-sea oil plume enriches indigenous oil-degrading bacteria Science 330 204–8 [5] Spies R B and DesMarais D J 1983 Natural isotope study of trophic enrichment of marine benthic communities by petroleum seepage Mar. Biol. 73 67–71 [6] Sherr E B and Sherr B F 1991 Planktonic microbes: tiny cells at the base of the ocean’s food webs Trends Ecol. Evol. 6 50–3 [7] Post D M, Layman C A, Arrington D A, Takimoto G, Quattrochi J and Montana C G 2007 Getting to the fat of the matter: models, methods and assumptions for dealing with lipids in stable isotope analyses Oecologia 152 179–89 [8] Welschmeyer N A 1994 Fluorometric analysis of chlorophyll a in the presence of chlorophyll b and phaeopigments Limnol. Oceanogr. 39 1985–92 [9] Schroeder W W 1979 Dispersion and Impact of Mobile River System Waters in Mobile Bay, Alabama (Water Resources Research Institute vol 37) (Auburn, AL: Auburn University) p 48 [10] Atlas R M 1981 Microbial degradation of petroleum hydrocarbons: an environmental perspective Microbiol. Rev. 45 180–209 [11] Søreide J E, Tamelander T, Hop H, Hobson K A and Johansen I 2007 Sample preparation effects on stable C and N isotope values: a comparison of methods in Arctic marine food web studies Mar. Ecol.-Prog. Ser. 328 17–28 [12] Schroeder W W, Dinnel S P and Wiseman W J Jr 1990 Salinity stratification in a river-dominated estuary Estuar. Coasts 13 145–54 [13] Goecker M E, Valentine J F, Sklenar S A and Chaplin G I 2009 Influence from hydrological modification on energy and nutrient transference in a deltaic food web Estuar. Coasts 32 173–87 [14] Michener R H and Schell D M 1994 Stable isotope ratios as tracers in marine aquatic food webs Stable Isotopes in Ecology and Environmental Science ed K Lajtha and R H Michener (Oxford: Blackwell Scientific) pp 138–58 [15] Moncreiff C A and Sullivan M J 2001 Trophic importance of epiphytic algae in subtropical seagrass beds: evidence from multiple stable isotope analyses Mar. Ecol. Prog. Ser. 215 93–106 [16] Fry B 2006 Stable Isotope Ecology (New York: Springer) p 308 [17] Azam F, Fenchel T, Field J G, Gray J S, Meyer-Reil L A and Thingstad F 1983 The ecological role of water-column microbes in the sea Mar. Ecol. Prog. Ser. 10 257–63 Figure 3. Particulate organic carbon to particulate nitrogen ratios (C:N) of (A) mesozooplankton (0.2–2 mm) and (B) smaller suspended particulates (1 μm–0.2 mm) collected at stations T35, T20, T10 and BM (cf figure 1 and table 1). There was no significant difference between surface and bottom C:N within each station, thus data were pooled for all pre- and post-spill analyses (ANOVA: F1,97 = 0.04, P = 0.85). ND indicates data were not available. 4. Conclusions Carbon isotopic depletion in mesozooplankton and suspended particulate samples throughout the water column (figures 1(D) and (E)) indicates trophic transfer of oil carbon into the planktonic food web. A similar response found in benthic communities around natural seeps [5] suggests that carbon isotopic shifts in the plankton fractions are likely due to the duration and magnitude of depleted carbon released into the system. These data provide strong evidence that labile fractions of the oil extended throughout the shallow water column during northward slick transport and that this carbon was processed relatively quickly at least two trophic levels beyond prokaryotic hydrocarbon consumers given our understanding of microbial-zooplankton trophic linkages [6, 17]. Further, this study provides a launching point for follow up experimental laboratory and field exercises aimed at understanding the fate and transport of petroleum hydrocarbons in marine planktonic ecosystems under the influence of natural or human-mediated chemical dispersion. Acknowledgments This work was supported by a grant from the National Science Foundation through the RAPID program funding for oil spill research (OCE-1043413). Additional funding and resources were from the Alabama Department of Conservation and Natural Resources, Marine Resources Division, and the National Oceanic and Atmospheric Administration through 6 GEOPHYSICAL RESEARCH LETTERS, VOL. 39, L01605, doi:10.1029/2011GL049505, 2012 Macondo-1 well oil-derived polycyclic aromatic hydrocarbons in mesozooplankton from the northern Gulf of Mexico Siddhartha Mitra,1 David G. Kimmel,2,3 Jessica Snyder,2 Kimberly Scalise,1 Benjamin D. McGlaughon,2 Michael R. Roman,4 Ginger L. Jahn,4 James J. Pierson,4 Stephen B. Brandt,5 Joseph P. Montoya,6 Robert J. Rosenbauer,7 Thomas D. Lorenson,7 Florence L. Wong,7 and Pamela L. Campbell8 Received 7 October 2011; revised 7 December 2011; accepted 11 December 2011; published 14 January 2012. [1] Mesozooplankton (>200 mm) collected in August and September of 2010 from the northern Gulf of Mexico show evidence of exposure to polycyclic aromatic hydrocarbons (PAHs). Multivariate statistical analysis revealed that distributions of PAHs extracted from mesozooplankton were related to the oil released from the ruptured British Petroleum Macondo-1 (M-1) well associated with the R/V Deepwater Horizon blowout. Mesozooplankton contained 0.03–97.9 ng g 1 of total PAHs and ratios of fluoranthene to fluoranthene + pyrene less than 0.44, indicating a liquid fossil fuel source. The distribution of PAHs isolated from mesozooplankton extracted in this study shows that the 2010 Deepwater Horizon spill may have contributed to contamination in the northern Gulf of Mexico ecosystem. Citation: Mitra, S., et al. (2012), Macondo-1 well oil-derived polycyclic aromatic hydrocarbons in mesozooplankton from the northern Gulf of Mexico, Geophys. Res. Lett., 39, L01605, doi:10.1029/2011GL049505. 1. Introduction [2] An estimated 4.93 million barrels (1 barrel = 42 US gallons) of crude oil were released into the Northern Gulf of Mexico (nGOM) from the British Petroleum (BP) Macondo-1 (M-1) site (Federal Interagency Solutions Group, Oil budget calculator—Deepwater Horizon, 2010, http://www.restorethegulf.gov/sites/default/files/documents/ pdf/OilBudgetCalc_Full_HQ-Print_111110.pdf), the location of the R/V Deepwater Horizon (DWH) blowout. The total extent of the surface oil slick, derived from wind, ocean currents, aerial photography, and satellite imagery, was estimated to be 180,000 km (J. Amos, BP spill was greater disaster than public knew, 2010, http://mcbi.org/news/PRNorse-Amos-2010.pdf). Chemical evidence of subsurface oil from the leak was found as far away as 30 km south of the M-1 well [Diercks et al., 2010]. Oil which is a complex mixture of hydrocarbons and other chemicals, contains numerous PAHs [Connell, 1997; National Research Council, 2002]. These PAHs may be used as chemical fingerprints of specific types of oil released into natural environments [Blumer, 1976]. For example, PAH distributions, or relative abundances of low and high molecular weight PAHs, have been used to fingerprint oil and determine the provenance of various oil spills in the environment [Bennett et al., 2000; Stout et al., 2001; Christensen et al., 2004]. Despite extensive efforts at completely delineating the extent of the DWH oil spill, there have been no studies published to date addressing whether or not the spill introduced oil-derived toxic compounds into the pelagic food web of the nGOM. [3] Mesozooplankton are useful sentinel organisms for oil-derived pollution [Carls et al., 2006] that serve as food for early life stages of fish and shrimp. Furthermore, they act as conduits for the movement of oil-derived contamination and other persistent organic pollutants through the marine food web [Clayton et al., 1977; Borgå et al., 2004; Sobek et al., 2006; Hallanger et al., 2011a, 2011c]. The objective of this study was to extract and analyze PAHs in mesozooplankton collected in the nGOM after the 2010 DWH spill. The null hypothesis of this study was that the relative PAH distributions in mesozooplankton collected throughout the Gulf of Mexico would not resemble the relative PAH distributions in oil collected from the M-1 well. 2. Methods Department of Geological Sciences, East Carolina University, Greenville, North Carolina, USA. 2 Department of Biology, East Carolina University, Greenville, North Carolina, USA. 3 Institute for Coastal Sciences and Policy, East Carolina University, Greenville, North Carolina, USA. 4 Horn Point Laboratory, Center for Environmental Science, University of Maryland, Cambridge, Maryland, USA. 5 Oregon Sea Grant, Oregon State University, Corvallis, Oregon, USA. 6 School of Biology, Georgia Institute of Technology, Atlanta, Georgia, USA. 7 Pacific Coastal Marine Science Center, U.S. Geological Survey, Menlo Park, California, USA. 8 Water Resources, National Research Program, U.S. Geological Survey, Menlo Park, California, USA. 2.1. Mesozooplankton Sampling [4] The mesozooplankton samples in this study were collected from within 20 km from the M-1 site and at distal stations grouped around 300 km south of the M-1 site (Figure 1). Mesozooplankton were sampled with a Multiple Open and Closing Net Environmental Sampling System (MOCNESS) [Wiebe et al., 1976]. The MOCNESS had a 1 m2 opening and a mesh size of 200 mm. Five to ten minute tows were conducted at the surface, and forty to sixty minute tows were conducted between 500 m and the surface in August and September of 2010 (auxiliary material, Table S1).1 All mesozooplankton (>200 mm) were collected from the cod end of the nets and immediately frozen in glass vials. Copyright 2012 by the American Geophysical Union. 0094-8276/12/2011GL049505 1 Auxiliary materials are available in the HTML. doi:10.1029/ 2011GL049505. 1 L01605 1 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON L01605 Figure 1. Map of sampling area. Oil extent as of May 2010. Numbered symbols represent stations from which samples were analyzed for PAHs in this study. This procedure generally resulted in two sets of samples from each station: one group of mesozooplankton samples from the surface water and one group integrated across the top 500 m water depth. 2.2. Oil, Surface Slick, and Seep Samples [5] A sample of M-1 subsurface oil was provided by B & B Laboratory, College Station, Texas. The well oil was obtained by BP from the riser insertion tube aboard the drillship Discoverer Enterprise on May 21, 2010, and was absent of any defoamer or dispersant. All samples were collected, processed, and shipped under standard chain-ofcustody protocols according to methods listed in the USGS National Field Manual for the Collection of Water-Quality Data (NFM) (http://pubs.water.usgs.gov/twri9A/) as well as other USGS standard operation procedures [Wilde et al., 2010]. Surface slick samples were obtained manually in pre-cleaned I-Chem jars on May 8. The only available natural seep samples from the Gulf of Mexico were collected in 2002. These had been stored frozen at 20 °C at the USGS Menlo Park. Santa Barbara Channel samples, which consisted of sludge, seep oil, and produced oil, were collected in 2001, 2005, and 2008 and frozen until they were extracted for PAHs in this study. 2.3. Polycyclic Aromatic Hydrocarbon (PAH) Extraction [6] Less than a gram of each oil or seep sample was weighed out in pre-cleaned (450 °C for 4h) aluminum foil tins and transferred to vials to be extracted. Following determination of the sample weight, two mL of deuterated PAH surrogate standard (in acetone) were added to each oil, zooplankton, and seep sample. Then, zooplankton samples were macerated in a Sentry tissue macerator after which PAHs were extracted from them using an Accelerated Solvent Extractor (ASE) or sonication using hexane: acetone (1:1, v:v). All extracts were concentrated by rotaryevaporation followed by a N2 stream. Extracts were purified via silica gel and sodium sulfate chromatography on an open column. The aromatic fraction was eluted with 75 mL of 80:20 hexane:methylene chloride. Zooplankton PAH concentrations are reported on a wet weight basis. Additional details of sample collection and extractions, including extraction recoveries, may be found in the auxiliary material. 2.4. Data Analysis [7] Data analyses were conducted using R statistical software version 2.10.1 (©R Foundation for Statistical Computing, 2009). The proportion of each of the 24 PAHs was calculated in each mesozooplankton extract by dividing individual PAH concentrations by the total concentration. We created a data matrix of columns containing proportional concentrations of each PAH compound and rows representing individual samples. Relationships between the PAH distributions in mesozooplankton and oil samples were characterized using a two-stage multivariate analysis. First, we performed an agglomerative, hierarchical cluster analysis 2 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON of a matrix of distances calculated using the dist.prop function in the R library ade4 [Dray and Dufour, 2007], we calculated the distances between each row based on the suite of PAH compounds. The cluster analysis was conducted using the hclust function using Ward’s linkage. To identify particular groupings of data, we used the resultant cluster analysis fit and cut the tree into 4 distinct groups using the cutree function. Using the resulting 4 groupings from this analysis, we calculated the average ( SD) proportion of each PAH within each group. A non-metric multidimensional scaling (NMDS) analysis was performed on the matrix of calculated distances using the isoMDS function in the R library MASS [Venables and Ripley, 1999]. 3. Results and Discussion [8] There were no systematic trends in PAH distributions in nGOM mesozooplankton as a function of radial distance from the M-1 well or with depth in the water column from which the organisms were collected. The lack of any difference in PAH distributions in mesozooplankton collected from surface or deeper (0–500 m) waters is not surprising given that many species of mesozooplankton exhibit diel vertical migration patterns [Haney, 1988]. In contrast to oil spills occurring at the sea surface, petroleum hydrocarbons originating from the M-1 well were subjected to several unique environmental processes that may account for the unique PAH distributions within each cluster group. First, petroleum was released from the Macondo well at 1.5-km depth. This resulted in partitioning of hydrocarbons into the aqueous phase in the absence of atmospheric evaporation [Reddy et al., 2011]. Furthermore, the dispersants applied to this spill may have enhanced aqueous dissolution of oil droplets, affecting overall water column concentrations of PAHs in a manner in which there was no systematic geographical pattern with distance from the M-1 well. [9] Cluster analysis suggests that PAH distributions in all samples analyzed could be divided into four distinct groups (Figure 2, top). Group A corresponded to mesozooplankton that were dissimilar to any oil samples analyzed in this study. The PAHs that were detected in natural seeps in the Gulf of Mexico (Mississippi Canyon, Figure 1, Station 64) clustered in Group B along with PAH distributions in produced oil, crude oil and tar balls from several locations in the Santa Barbara Channel. The distributions of PAHs in mesozooplankton from the other stations led to their emplacement in Groups C and D. Distributions of PAHs in Groups C and D were similar to PAH distributions in from the DWH surface slick and from the broken M-1 riser pipe, respectively. [10] We used a non-parametric multidimensional scaling analysis on the distance matrix calculated for the cluster L01605 analysis [Cox and Cox, 1994] to help visualize relationships in PAH distributions between mesozooplankton, DWH oil, and seep samples (Figure 2, bottom). For a majority of the mesozooplankton samples, PAHs detected in the nGOM mesozooplankton scaled closely to either Cluster C (PAHs in the DWH surface slicks) or Cluster D (PAHs in oil from the M-1 riser pipe). The PAHs in the nGOM seep and SBC samples clustered together, but did not scale closely to DWH oil-derived PAH distributions (Cluster B). Our results indicate that oil derived from the DWH incident was associated with mesozooplankton collected as far as 180 km from the M-1 well. Zooplankton in Cluster A do not have a PAH signature associated with DWH riser oil or DWH-derived surface slicks. This observation suggests that the oil release from the DWH incident, although spatially-extensive, may have been patchy or that these zooplankton were exposed to oil, but no longer have a relative PAH distribution that matches that of the DWH oil. [11] The PAH distributions detected in some nGOM mesozooplankton in this study match the PAH distributions extracted in oil released from the broken riser pipe collected from the M-1 well and from surface oil slicks originating from the DWH incident (Figure 3). A two-sample KolmogorovSmirnov goodness-of-fit (KS-GOF) test comparing the mean PAH distributions revealed no differences in PAH signatures between Cluster B and C ( p = 0.68), Cluster B and D ( p = 0.45), and Cluster C and D ( p = 0.26). Mesozooplankton PAH signatures from Cluster A (Figure 3a), although relatively abundant in naphthalene, did not possess the ratios of fluorene:phenanthrene resembling those in the DWH surface slicks or oil from the broken riser pipe (Figures 3b and 3c). Comparisons of these PAH distributions showed significant differences when compared with the KS-GOF, Cluster A and B ( p = 0.03), Cluster A and C ( p = 0.004), and Cluster A and D were not statistically different at a = 0.05, but did have a low p-value of 0.07. The PAH distributions from the SBC and the nGOM seep sample had much higher proportions of 1-methyl and 2-methylnapthalene (Figure 3b) than the M-1 samples or the nGOM surface slicks (Figures 3c and 3d). The dissimilarity in these particular PAHs between leaking oil from the M-1 well and the natural GOM seep samples can be explained by chemical heterogeneity resulting from varying oil source rock, age, and migration history that likely occurs between the seep location and the M-1 site [Aharon et al., 1997; Hood et al., 2002]. [12] We are not aware of any existing studies examining PAH distributions in zooplankton collected from the water column of the nGOM, therefore we have no means of comparing PAH body burdens in nGOM zooplankton collected prior to the spill. Distributions of PAHs in fluvial suspended sediments such as those exported from the Mississippi River generally contain elevated relative Figure 2. (top) Agglomerative, hierarchical cluster analysis of PAH distributions in oil from the broken riser pipe at the M-1 well (DWH.D), Deepwater Horizon surface slick oil (DWH.S), mesozooplankton collected from surface tows (ZP.S) and tows from 0–500 m water depth in nGOM (ZP.D). A natural seep sample from the nGOM (GMS) and oil samples derived from the Santa Barbara channel (SBC). The four letters in both plots represent the 4 groups identified in the cluster analysis (see text for explanations). (bottom) Non-metric multi-dimensional scaling (NMDS) biplot of PAH distributions in oil from the broken riser pipe at the M-1 well (3, oval), Deepwater Horizon surface slick oil (1,2, circles), nGOM seep samples (64, squares) and oil samples derived from a variety of locations in the Santa Barbara channel (56–63, hexagons). Other numbers are representative of mesozooplankton samples (see Figure 1 for collection location). The stress of the NMDS fit was 0.14. The principal variables structuring the data are the PAH signatures within each group. 3 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON Figure 2 4 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON L01605 L01605 Figure 3. Mean ratio of individual PAHs to total PAHs (R) within groups identified using the cluster analysis. (a) Mesozooplankton samples only, (b) Gulf of Mexico seeps and Santa Barbara Channel, (c) mesozooplankton samples and surface slick oil, and (d) mesozooplankton samples and oil from the leaking riser pipe. abundances of higher molecular weight PAHs [Mitra et al., 2002] than found in mesozooplankton analyzed in this study. Thermodynamic modeling of PAH distributions on particles has demonstrated that PAHs are strongly-sorbed onto Mississippi River and Gulf of Mexico sediments, which contain a non-trivial fraction of combustion-derived black carbon [Mitra and Bianchi, 2003]. Furthermore, ratios of fluoranthene:pyrene in these mesozooplankton samples are less than 0.5 (Table 1), as is expected in PAH distributions originating from a petrogenic source [Yunker et al., 2002]. Taken together, this suggests that PAHs from suspended sediments did not contribute to the PAH signal detected in these mesozooplankton samples and that the PAH distributions in these samples are petrogenic. [13] There are numerous factors that may affect PAH body burdens in zooplankton isolated for this study. Unlike the elevated levels of PAHs accumulated in mesozooplankton collected in Port Valdez, Alaska [Carls et al., 2006], the mesozooplankton PAH concentrations in our study (Table 1) are similar to that found in zooplankton collected near other oil spills in temperate and tropical environments globally [Davenport, 1982; Guzman del Proo et al., 1986]. Several processes (e.g. exchange with water through passive partitioning, ingestion of contaminated food, and production of Table 1. Table of PAH Concentrations in Each Hierarchal Cluster Groupa Cluster Sample Type N Mean Standard Deviation A B mesozooplankton Santa Barbara Channel seep samples nGOM seeps mesozooplankton surface slicks mesozooplankton M-1 well 19 8 4 7 2 26 1 17.4 1.23E + 06 1.19E + 04 29.0 5.45E + 05 13.8 3.61E + 06 28.9 6.91E + 05 1.15E + 04 26.1 C D a Concentrations are given in ng g 1 19.7 Minimum Maximum flu/flu + pyr 1.64 4.96E + 1.60E + 11.1 5.40E + 2.04 3.61E + 120.0 2.61E + 06 2.70E + 04 85.7 5.439 + 05 97.9 3.61E + 06 0.19 0.23 0.25 0.14 0.29 0.11 0.17 0.22 0.41 0.44 0.29 0.44 05 03 05 06 wet weight. N signifies number of samples extracted from within that group. 5 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON fecal pellets and eggs), occurring simultaneously, may affect the final body burden of PAHs derived from oil spills [Sobek et al., 2006; Berrojalbiz et al., 2011]. Moreover, all of these processes may vary as a function of ambient temperature and seasonality [Hallanger et al., 2011b]. [14] Mesozooplankton species in microcosm experiments have been shown to excrete PAHs on time scales of days [Berrojalbiz et al., 2009]. However, we detected the presence of PAHs in mesozooplankton collected in August and early September 2010, well after the M-1 well was capped on 15 July, 2010. This was surprising given that mesozooplankton population turnover times may be quite rapid in the warmer waters of the Gulf of Mexico (e.g. Acartia tonsa has a generation time of 7 days at 25 °C [Heinle, 1966]). We offer several possible explanations for the persistence of a low but DWH-derived PAH signal so long after capping of the well. First, the mesozooplankton samples in this study consisted of several species homogenized together; thus, cross-species metabolism of PAHs may vary resulting in relatively lower body burdens than seen in many single species laboratory studies. Second, PAHs from DWH oil may have remained in the system at significant levels long after the well was capped. Lastly, mesozooplankton may have been accumulating PAHs in their bodies and passing them across generations via eggs, which are relatively lipid rich compared to individual mesozooplankton. Determining which of these explanations is responsible for the ng g 1 observed levels of PAHs in nGOM mesozooplankton is beyond the scope of this study. However, our study demonstrates that there was a signature distribution of DWH-derived PAHs in zooplankton. [15] As of August 2010, the U.S. National Incident Command Center estimated that 26% of the residual oil could be found either on or below the surface as light sheen and weathered tar balls, washed ashore or collected from the shore, or buried in sand and sediments (Federal Interagency Solutions Group, Oil budget calculator, 2010). Although a subsurface oil plume was identified [Camilli et al., 2010] and its composition recently elucidated [Reddy et al., 2012], the ultimate fate of the oil and its presence in the ecosystem has yet to be comprehensively determined. The presence of this PAH signature in nGOM mesozooplankton samples in the patterns noted in our study suggests that the spatial and temporal extent of the 2010 spill in the nGOM was extensive and patchy. A recent study reported a depleted d 13C isotopic signature in coastal mesozooplankton collected north of the M-1 well showing that carbon in oil collected at depth was incorporated up to two trophic levels above prokaryotic hydrocarbon consumers and into the planktonic food web [Graham et al., 2010]. That study, combined with ours, suggests that the potential for movement of DWH-derived carbon and PAHs to higher trophic levels, existed after the well at M-1 had been capped. [16] Acknowledgments. The authors thank the National Science Foundation RAPID grants OCE-1043249, OCE-1047736, OCE-1057461, and the captain and crew of the R/V Oceanus. [17] The Editor thanks two anonymous reviewers for their assistance in evaluating this paper. References Aharon, P., H. P. Schwarcz, and H. H. Roberts (1997), Radiometric dating of submarine hydrocarbon seeps in the Gulf of Mexico, Geol. Soc. Am. Bull., L01605 109(5), 568–579, doi:10.1130/0016-7606(1997)109<0568:RDOSHS>2.3. CO;2. Bennett, A., T. S. Bianchi, and J. C. Means (2000), The effects of PAH contamination and grazing on the abundance and composition of microphytobenthos in salt marsh sediments (Pass Fourchon, LA, USA): II: The use of plant pigments as biomarkers, Estuarine Coastal Shelf Sci., 50(3), 425–439, doi:10.1006/ecss.1999.0572. Berrojalbiz, N., S. Lacorte, A. Calbet, E. Saiz, C. Barata, and J. Dachs (2009), Accumulation and cycling of polycyclic aromatic hydrocarbons in zooplankton, Environ. Sci. Technol., 43(7), 2295–2301, doi:10.1021/ es8018226. Berrojalbiz, N., J. Dachs, M. Ojeda, M. C. Valle, J. Castro-Jimenez, J. Wollgast, M. Ghiani, G. Hanke, and J. M. Zaldivar (2011), Biogeochemical and physical controls on concentrations of polycyclic aromatic hydrocarbons in water and plankton of the Mediterranean and Black Seas, Global Biogeochem. Cycles, 25, GB4003, doi:10.1029/ 2010GB003775. Blumer, M. (1976), Polycyclic aromatic compounds in nature, Sci. Am., 234, 34–45, doi:10.1038/scientificamerican0376-34. Borgå, K., A. T. Fisk, P. F. Hoekstra, and D. C. G. Muir (2004), Biological and chemical factors of importance in the bioaccumulation and trophic transfer of persistent organochlorine contaminants in Arctic marine food webs, Environ. Toxicol. Chem., 23(10), 2367–2385, doi:10.1897/03-518. Camilli, R., C. M. Reddy, D. R. Yoerger, B. A. S. Van Mooy, M. V. Jakuba, J. C. Kinsey, C. P. McIntyre, S. P. Sylva, and J. V. Maloney (2010), Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon, Science, 330, 201–204, doi:10.1126/science.1195223. Carls, M. G., J. W. Short, and J. Payne (2006), Accumulation of polycyclic aromatic hydrocarbons by Neocalanus copepods in Port Valdez, Alaska, Mar. Pollut. Bull., 52, 1480–1489, doi:10.1016/j.marpolbul.2006.05.008. Christensen, J. H., A. B. Hansen, G. Tomasi, J. Mortensen, and O. Andersen (2004), Integrated methodology for forensic oil spill identification, Environ. Sci. Technol., 38(10), 2912–2918, doi:10.1021/es035261y. Clayton, J. R., Jr., S. P. Pavlou, and N. F. Breitner (1977), Polychlorinated biphenyls in coastal marine zooplankton: Bioaccumulation by equilibrium partitioning, Environ. Sci. Technol., 11(7), 676–682, doi:10.1021/ es60130a008. Connell, D. (1997), Basic Concepts of Environmental Chemistry, Lewis, Boca Raton, FL. Cox, T., and M. Cox (1994), Multidimensional Scaling, Chapman and Hall, London. Davenport, J. (1982), Oil and planktonic ecosystems, Philos. Trans. R. Soc. B, 297, 369–384, doi:10.1098/rstb.1982.0048. Diercks, A.-R., et al. (2010), Characterization of subsurface polycyclic aromatic hydrocarbons at the Deepwater Horizon site, Geophys. Res. Lett., 37, L20602, doi:10.1029/2010GL045046. Dray, S., and A. Dufour (2007), The ade4 package: Implementing the duality diagram for ecologists, J. Stat. Software, 22, 1–20. Graham, W. M., R. H. Condon, R. H. Carmichael, I. D’Ambra, H. K. Patterson, L. J. Linn, and F. J. Hernandez (2010), Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill, Environ. Res. Lett., 5, 045301, doi:10.1088/1748-9326/5/4/045301. Guzman del Proo, S. S., E. A. Chavez, F. M. L. Alatriste, S. de la Campa, L. G. De la Cruz, R. Duadarrama, A. Guerra, S. Mille, and D. Torruco (1986), The impact of the Ixtox-1 oil spill on zooplankton, J. Plankton Res., 8, 557–581, doi:10.1093/plankt/8.3.557. Hallanger, I. G., A. Ruus, N. A. Warner, D. Herzke, A. Evenset, M. Schoyen, G. W. Gabrielsen, and K. Borga (2011a), Differences between Arctic and Atlantic fjord systems on bioaccumulation of persistent organic pollutants in zooplankton from Svalbard, Sci. Total Environ., 409(14), 2783–2795, doi:10.1016/j.scitotenv.2011.03.015. Hallanger, I. G., N. A. Warner, A. Ruus, A. Evenset, G. Christensen, D. Herzke, G. W. Gabrielsen, and K. Borga (2011b), Seasonality in contaminant accumulation in Arctic marine pelagic food webs using trophic magnification factor as a measure of bioaccumulation, Environ. Toxicol. Chem., 30(5), 1026–1035, doi:10.1002/etc.488. Hallanger, I. G., A. Ruus, D. Herzke, N. A. Warner, A. Evenset, E. S. Heimstad, G. W. Gabrielsen, and K. Borga (2011c), Influence of season, location, and feeding strategy on bioaccumulation of halogenated organic contaminants in Arctic marine zooplankton, Environ. Toxicol. Chem., 30(1), 77–87, doi:10.1002/etc.362. Haney, J. F. (1988), Diel patterns of zooplankton behavior, Bull. Mar. Sci., 43(3), 583–603. Heinle, D. R. (1966), Production of a calanoid copepod, Acartia tonsa, in the Patuxent River estuary, Chesapeake Sci., 7, 59–74, doi:10.2307/ 1351126. Hood, K. C., L. M. Wenger, O. P. Gross, and S. C. Harrison (2002), Hydrocarbon systems analysis of the northern Gulf of Mexico: Delineation of hydrocarbon migration pathways using seeps and seismic imaging, in Surface Exploration Case Histories: Applications of Geochemistry, 6 of 7 L01605 MITRA ET AL.: DWH PAHS IN GOM MESOZOOPLANKTON Magnetics, and Remote Sensing, edited by D. Schumacher and L. A. LeSchack, AAPG Stud. Geol., 48, 25–40. Mitra, S., and T. S. Bianchi (2003), A preliminary assessment of polycyclic aromatic hydrocarbon distribution in the lower Mississippi River and Gulf of Mexico, Mar. Chem., 82, 273–288, doi:10.1016/S03044203(03)00074-4. Mitra, S., T. S. Bianchi, B. A. McKee, and M. Sutula (2002), Black carbon from the Mississippi River: Quantities, sources, and potential implications for the global carbon cycle, Environ. Sci. Technol., 36(11), 2296–2302, doi:10.1021/es015834b. National Research Council (2002), Oil, in The Sea: Inputs, Fates, and Effects, pp. 89–117, Natl. Acad. of Sci., Washington, D. C. Reddy, C. M., et al. (2012), Composition and fate of gas and oil released to the water column during the Deepwater Horizon oil spill, Proc. Natl. Acad. Sci. U. S. A., doi:10.1073/pnas.1101242108, in press. Sobek, A., G. Cornelissen, P. Tiselius, and Ö. Gustafsson (2006), Passive partitioning of polychlorinated biphenyls between seawater and zooplankton, a study comparing observed field distributions to equilibrium sorption experiments, Environ. Sci. Technol., 40(21), 6703–6708, doi:10.1021/ es061248c. Stout, S. A., A. D. Uhler, and K. J. McCarthy (2001), A strategy and methodology for defensibly correlating spilled oil to source candidates, Environ. Forensics, 2(1), 87–98, doi:10.1006/enfo.2001.0027. Venables, W. N., and B. D. Ripley (1999), Modern Applied Statistics With S-PLUS, 3rd ed., Springer, New York. Wiebe, P. H., K. H. Burt, S. H. Boyd, and A. W. Morton (1976), Multiple opening-closing net and environmental sensing system for sampling zooplankton, J. Mar. Res., 34(3), 313–326. L01605 Wilde, F. D., S. C. Skrobialowski, and J. S. Hart (2010), Sampling protocol for post-landfall Deepwater Horizon oil release, Gulf of Mexico 2010: Addendum to standard USGS methods for the collection of water, benthic invertebrates, and microorganisms, U.S. Geol. Soc. Open File Rep., O2010-1191, U.S. Geological Survey, Reston, VA. Yunker, M. B., S. M. Backus, E. Graf Pannatier, D. S. Jeffries, and R. W. Macdonald (2002), Sources and significance of alkane and PAH hydrocarbons in Canadian arctic rivers, Estuarine Coastal Shelf Sci., 55(1), 1–31, doi:10.1006/ecss.2001.0880. S. Mitra and K. Scalise, Department of Geological Sciences, East Carolina University, Greenville, NC 27858, USA. ([email protected]) D. G. Kimmel, B. D. McGlaughon, and J. Snyder, Department of Biology, East Carolina University, Greenville, NC 27858, USA. G. L. Jahn, J. J. Pierson, and M. R. Roman, Horn Point Laboratory, Center for Environmental Science, University of Maryland, Cambridge, MD 21613, USA. S. B. Brandt, Oregon Sea Grant, Oregon State University, Corvallis, OR 97331–2131, USA. J. P. Montoya, School of Biology, Georgia Institute of Technology, Atlanta, GA 30332, USA. T. D. Lorenson, R. J. Rosenbauer, and F. L. Wong, Pacific Coastal Marine Science Center, U.S. Geological Survey, Menlo Park, CA 94025, USA. P. L. Campbell, Water Resources, National Research Program, U.S. Geological Survey, Menlo Park, CA 94025, USA. 7 of 7 Helen K. Whitea,1, Pen-Yuan Hsingb, Walter Choc, Timothy M. Shankc, Erik E. Cordesd, Andrea M. Quattrinid, Robert K. Nelsone, Richard Camillif, Amanda W. J. Demopoulosg, Christopher R. Germanh, James M. Brooksi, Harry H. Robertsj, William Sheddk, Christopher M. Reddye, and Charles R. Fisherb a Department of Chemistry, Haverford College, Haverford, PA 19041; bDepartment of Biology, Pennsylvania State University, University Park, PA 16802; Biology Department, eDepartment of Marine Chemistry and Geochemistry, fApplied Ocean Physics and Engineering, and hDepartment of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, MA 02543; dBiology Department, Temple University, Philadelphia, PA 19122; gSoutheast Ecological Science Center, US Geological Survey, Gainesville, FL 32653; iTDI-Brooks International Inc., College Station, TX 77845; jDepartment of Oceanography and Coastal Sciences, Coastal Studies Institute, Louisiana State University, Baton Rouge, LA 70803; and kBureau of Ocean Energy Management, US Department of the Interior, New Orleans, LA 70115 c SPECIAL FEATURE SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE Impact of the Deepwater Horizon oil spill on a deep-water coral community in the Gulf of Mexico To assess the potential impact of the Deepwater Horizon oil spill on offshore ecosystems, 11 sites hosting deep-water coral communities were examined 3 to 4 mo after the well was capped. Healthy coral communities were observed at all sites >20 km from the Macondo well, including seven sites previously visited in September 2009, where the corals and communities appeared unchanged. However, at one site 11 km southwest of the Macondo well, coral colonies presented widespread signs of stress, including varying degrees of tissue loss, sclerite enlargement, excess mucous production, bleached commensal ophiuroids, and covering by brown flocculent material (floc). On the basis of these criteria the level of impact to individual colonies was ranked from 0 (least impact) to 4 (greatest impact). Of the 43 corals imaged at that site, 46% exhibited evidence of impact on more than half of the colony, whereas nearly a quarter of all of the corals showed impact to >90% of the colony. Additionally, 53% of these corals’ ophiuroid associates displayed abnormal color and/or attachment posture. Analysis of hopanoid petroleum biomarkers isolated from the floc provides strong evidence that this material contained oil from the Macondo well. The presence of recently damaged and deceased corals beneath the path of a previously documented plume emanating from the Macondo well provides compelling evidence that the oil impacted deep-water ecosystems. Our findings underscore the unprecedented nature of the spill in terms of its magnitude, release at depth, and impact to deep-water ecosystems. hopane | sterane | Paramuricea | sediment B etween October 15 and November 1, 2010, approximately 6 months after the Deepwater Horizon blowout and 3 months after the Macondo well was capped, nine sites hosting deep-water coral communities were examined with the remotely operated vehicle (ROV) Jason II. This effort was part of an ongoing study funded by the Bureau of Ocean Energy Management (BOEM) and the National Oceanic and Atmospheric Administration’s Ocean Exploration and Research program. These sites, located between 93.60 °W and 87.31 °W and between −27.42 °N and −29.16 °N (Fig. S1), were >20 km from the Macondo well, ranged in depth from 290 to 2600 m, and hosted coral communities including scleractinian, gorgonian, and antipatharian corals. At these sites, no visible evidence of impact to the corals and associated communities was observed (Fig. 1). However, on November 2, 2010, the ROV Jason II investigated an area in lease blocks Mississippi Canyon (MC) 294 and 338, 11 km to the SW of the site of the Deepwater Horizon spill. This area was explored because 3D seismic reflectivity data (Fig. S1) suggested there was a strong likelihood of hard grounds, and hence likely coral substrate present. Its location (28.40N, 88.29W, 1,370 m) also placed it in the path of a 100-m-thick deep-water plume of neutrally buoyant water enriched with petroleum hydrocarbons from the Macondo well that was documented at 1,100 m in June 2010 www.pnas.org/cgi/doi/10.1073/pnas.1118029109 (1, 2). Numerous coral colonies were discovered at this location and many were partially or completely covered in a brown, flocculent material (hereafter referred to as floc). They showed signs of recent and ongoing tissue damage (Fig. 2) not observed elsewhere at this time (Fig. 1) or in the previous 10 y of baseline studies in the Gulf of Mexico (GoM) (3–5). Between December 8 and 14, 2010 additional surveys were performed with the deep submergence vehicle (DSV) Alvin at MC 294 and a newly discovered site 22 km to the ESE of the Macondo well in MC 388 (1,850 m depth). Visible signs of recent impact or stress were not evident in the corals imaged at MC 388. To determine whether the cause of the overall decrease in coral health at MC 294 was related to the Deepwater Horizon oil spill, the floc covering the corals and nearby sediment was examined for the presence of petroleum hydrocarbons originating from the Macondo well. Determining the source of petroleum hydrocarbons in these samples posed a significant challenge. The complexity of the petrogeochemical signatures in the GoM environment is considerable (6). Specific crude oils can be differentiated from their source rock groups using biomarkers (molecular fossils), which are highly resistant to abiotic and biotic processes and have been invaluable tools for characterizing and fingerprinting crude oils (7). For example, sterane biomarkers are derived primarily from marine phytoplankton and vary depending on geologic age. Hopanes, which are another class of biomarkers, can be used individually or in concert with sterane distributions to provide even greater certainty in characterizing oils (7). The use of biomarkers by the petroleum industry and subsequently in environmental forensics has, however, been performed in much different environments than the Deepwater Horizon spill, where oil and gas at 105 °C were released at pressure into 5 °C seawater at ∼1,400 m depth (2). We used traditional 1D gas chromatography (GC) and comprehensive two-dimensional gas chromatography (GC×GC, as in refs. 8, 9 and 10) to analyze floc and sediment samples from MC 294. These samples were compared Author contributions: H.K.W., T.M.S., E.E.C., A.W.J.D., C.R.G., and C.R.F. designed research; H.K.W., P.-Y.H., W.C., T.M.S., E.E.C., A.M.Q., R.K.N., R.C., A.W.J.D., C.R.G., C.M.R., and C.R.F. performed research; J.M.B., H.H.R., and W.S. contributed new reagents/analytic tools; H.K.W., P.-Y.H., W.C., T.M.S., E.E.C., A.M.Q., R.K.N., C.M.R., and C.R.F. analyzed data; and H.K.W. and C.R.F. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Data deposition: The octocoral and ophiuroid sequences reported in this paper have been deposited in the GenBank database (accession nos. JQ241244–52, JQ411462–9 and JQ771615–JQ771617) and all images have been submitted to the US National Oceanographic Data Center (accession no. 0084636). 1 To whom correspondene should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1118029109/-/DCSupplemental. PNAS Early Edition | 1 of 6 ENVIRONMENTAL SCIENCES Edited by Paul G. Falkowski, Rutgers, State University of New Jersey, New Brunswick, NJ, and approved February 28, 2012 (received for review November 1, 2011) A 0 50 100 Kilometers E B D C Fig. 1. Healthy deep-water coral communities observed in November 2010 from various sites >20 km from the Macondo well (shown as a star on map). (A) Paramuricea sp. type E and Asteroschema sp. at 360 m depth in Garden Banks (GB) 299; (B) Paramuricea sp. type E and Asteroschema sp. at 440 m depth in Mississippi Canyon (MC) 751; (C) Paramuricea sp. type A and Eumunida picta at 530 m depth in Green Canyon (GC) 354; (D) Paramuricea sp. type E and Asteroschema sp., along with a brisingid basket star at 360 m depth in Viosca Knoll (VK) 906; (E) P. biscaya and A. clavigerum at 2,300 m depth in Desoto Canyon (DC) 673. Phylogenetic species identifications of corals and associated ophiuroids are given in Fig. S2. with oil collected from directly above the broken riser pipe at the Macondo well (2) and field samples from surface water and salt marshes in areas oiled by the Deepwater Horizon spill. Here we report on the analyses of the visible impact to the gorgonian corals and coral associates at MC 294 based on in situ video imagery, shipboard microscopic analyses, and petroleum biomarker analysis of the floc adherent to the coral. In addition, we compare the petroleum hydrocarbon content and biomarkers with the surrounding surface and subsurface sediments and compare the condition of the corals and associates between November and December 2010 visits. Results and Discussion Gorgonian and other corals present at MC 294 are predominantly found in a central area 10 × 12 m in extent, composed of two adjacent carbonate slabs. Scattered boulders surround this region over an area of 50 × 50 m, and some of the isolated boulders host one or two additional coral colonies. The majority of the colonial corals were Paramuricea biscaya, with one or two colonies of Swiftia pallida, Paragorgia regalis, Acanthogorgia aspera, and Clavularia rudis (Fig. S2). The majority of these colonies exhibited signs of stress response, including excessive mucous production and retracted polyps, which have been observed in corals experimentally exposed to crude oil (11). Impact to the corals was quantified from close-up images (<1 m away) for 43 of the 58 coral colonies identified in the central area (Fig. S3) (not all of the corals could be approached for close-up imaging with ROV Jason II or DSV Alvin without disturbing other colonies). The level of impact to individual colonies was ranked from 0 (least impact) to 4 (greatest impact) according to the percentage of the colony exhibiting one or more of the following visual 2 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1118029109 indications of stress: bare skeleton above the basal region, loose tissue or heavy mucous hanging from the skeleton, and/or coverage with brown flocculent material (Fig. 3). Eighty-six percent of the coral colonies imaged in the central area exhibited signs of impact. Forty-six percent exhibited impact to at least 50% of the colony (impact level 3 or 4), and 23% of the colonies sustained impact to more than 90% of the colony (impact level 4). Between the November and December 2010 research cruises, changes in condition were assessed for all corals or portions of colonies for which high-resolution imagery was available from similar perspectives. Although differences in camera placement November 2010 December 2010 Fig. 2. Impacted corals at MC 294. Brown flocculent material and tissue loss is observed on the larger coral, A10, in November and December 2010. Although there is no evidence of recovery on A10, note that the tips of some branches that were living in November were still living in December (arrows). Coral A14 in the red box was the only colony showing apparent signs of recovery from coverage by the floc between visits. White et al. 10 8 6 23% 23% 3 4 14% 9% 4 2 0 0 1 2 Impact ranking Fig. 3. Impact assessment for coral colonies (n = 43) where high-quality images could be obtained from at least the November or December 2010 cruise. The levels of impact are ranked according to the proportion of a coral exhibiting obvious tissue damage, bare skeleton above basal region, or covered by brown flocculation: rank 0 (0–1%), rank 1 (<10%), rank 2 (10– 50%), rank 3 (50–90%), rank 4 (>90%). Numbers above bars are percentages of corals in a rating relative to all assessed colonies. on the two underwater vehicles, lighting, and quality of images limited the size of this data set to 18 colonies, neither progression of the visible damage nor clear evidence of recovery or growth was apparent in the majority of corals. Possible recovery was noted for one colony (A14, highlighted by box in Fig. 2). The relatively light covering of floc over more than 50% (impact level 3) of this colony in November was ranked as less than 10% impacted (impact level 1) by the time it was revisited in December, when extended polyps were visible in areas that had been partially covered with floc in November. Sampling of a P. biscaya coral (E3) in December enabled microscopic analysis to be made after removal of the floc. Varying degrees of tissue loss and sclerite enlargement were observed (Fig. S4). The skeleton was bare and entirely devoid of tissue at the base and along the main axis of the colony. At increasing distances from the basal point of attachment, less extensive tissue loss resulted in the exposure of the calcite skeletal elements that are normally embedded in the tissue layers and coenenchyme. These sclerites were still in their normal form of a polyp but appeared enlarged. The localized alteration of growth form, including excessive secretion of gorgonin and sclerite production to form granuloma-like structures, has previously been observed in gorgonians as an acute stress response (12, 13). Near the tips of some branches, which were not covered by the floc in situ, a few polyps on this coral appeared normal. Coral associates at MC 294 included 13 actinarian anemones and 78 Asteroschema clavigerum (a symbiotic ophiuroid). Of the 52 individual corals examined for coral associates, 25% hosted none, 2% hosted actinarian anemones, and 73% hosted A. clavigerum, with 70% of the ophiuroids present on P. biscaya, 18% on the single individual of P. regalis, and 12% on A. aspera. A. clavigerum is typically tan to red in color (Fig. 1); however, at this site only 47% were tan to red, whereas 44% had distinctly white arms (Fig. 2), and 9% (all hosted by P. biscaya), were bleached almost entirely white. In November, 27% of the ophiuroids displayed behaviors other than their normal attached posture of arms tightly coiled around their coral host (Fig. 1). Between visits, 13% of the ophiuroids transitioned from tightly to loosely coiled (i.e., Fig. 2). Two ophiuroids (Fig. S5) transitioned from tightly coiled to a posture with splayed out arms, a previously undocumented behavior in this species. The floc samples collected (>72 μm in size) were removed from the surface of the corals in situ and when filtered were found to contain dead coral polyp fragments, detached sclerites, and small brown droplets (Fig. S6). Solvent extracts of all of the floc examined were dominated by C16 and C18 saturated and unsaturated fatty acids and cholesterol, which are dominant lipids in biological tissue. Petroleum residues were also present and quantified via 1D gas chromatography coupled to a flame White et al. PNAS Early Edition | 3 of 6 SPECIAL FEATURE ionization detector (GC-FID; Table 1). An unresolved complex mixture (UCM) with n-alkane carbon range of C15–C42 indicates the presence of weathered petroleum (e.g., ref. 8; Table 1). Slight variations in UCM carbon range and distributions of n-alkanes among samples showed no consistent relationship to the pure Macondo Well oil (described in ref. 2; Table 1). Rather, it is evident that the n-alkanes in the samples represent input from a mixture of sources such as plants, biofilms, and differentially weathered subsurface hydrocarbons, including some that may have come from natural seeps. Acoustic mapping cruises performed from late May to August 2010 mapped several natural gas seeps in near proximity to both the Macondo well and the sample sites presented here, which could provide additional sources of subsurface hydrocarbons (14). Polycyclic aromatic hydrocarbon (PAH) distributions from coral E3 and sediment sample 4664 0–2 cm show good correspondence to Macondo well oil, with similar relative abundances of naphthalene, phenanthrene, and their alkylated derivatives as well as dibenzothiophenes, benzo[a]anthracene, and chrysene. The remaining coral samples are inconclusive owing to the small quantity of sample available for analysis, as well as the fact that these samples have been extensively weathered, as evidenced by the dominance of biodegradation-resistant chrysene in all extracts. Petroleum systems in the GoM do not display significant differences in the presence or absence of specific biomarkers; instead, differences in the relative amounts of biomarkers present have previously allowed sources to be determined (15, 16). Analysis of biomarkers such as hopanes is critical because these compounds are more resistant to biodegradation and water washing than n-alkanes and PAHs and provide insight into petroleum source determination (as in ref. 17). At the Macondo well, oil sampled from above the broken riser pipe (2) contains abundant hopanoids, diasteranes, and steranes (Fig. S7). Hopanoid biomarker ratios have been calculated for comparison with coral and sediment samples, as well a reference surface water (S1) and two reference coastal water (M1 and M2) samples (shown in Fig. 1 and described in ref. 18). These reference samples represent Macondo well oil that has undergone vertical transport from the seabed to the ocean surface (∼1,400 m) and subsequent lateral dispersion over ranges of 1–175 km, respectively (Table 1). Comparison of the hopanoid portion of the GC×GC chromatographic plane for the Macondo well oil to the S1 and M1 samples indicates a high degree of similarity (Fig. S8 A–C). This similarity is also seen in the floc from coral B8 (Fig. S8D) and in the surface sediment sample taken in the immediate vicinity of the corals (core 4664 0–2 cm; Fig. S8E). Slight but significant differences in hopanoid biomarker ratios are observed, by contrast, both in comparable core-top sediments collected away from the impacted corals at the MC 294 site (core 4662 0–2 cm; Table 1 and Fig. S8F) and at greater depths (2–5 cm and 5–10 cm; Table 1) in the core 4664 sediments. Further, the concentrations of oil present in the uppermost sediments of core 4664 (0–2 cm) are much higher (9.25 mg/g; Table 1) than the deeper sediments (2–5 cm and 5–10 cm) in the same core, which range in concentration from 0.02 to 0.03 mg/g (Table 1). They are also higher than the oil concentrations observed in surface sediments (0–2 cm) collected away from the impacted corals at the MC 294 site (3.46 mg/g; Table 1), where a bimodal n-alkane distribution indicative of inputs from mixed sources is observed. Significant variations in sediment oil concentrations have been previously documented in the GoM, particularly in areas of known natural oil seepage such as Green Canyon, where oil concentrations may be as high as 39.0 mg/g (19). The oil concentration and biomarker data from sediments collected away from the impacted corals and sediments at depth at MC294, are, however most consistent with long-term background inputs of oil derived from petroleum sources that are quite distinct to that present in the most superficial (hence, recent) core-top sediments and floc SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE Number of corals 12 ENVIRONMENTAL SCIENCES 30% 14 Table 1. Oil content, alkane distribution, and hopanoid biomarker ratios of brown flocculent material and sediment samples compared with Macondo Well oil Sample Source oil Macondo well‡ Surface water samples{ 1 km from well (S1) 175 km from well (M1) 175 km from well (M2) Flocculent material samples MC 294 coral (B8) MC 294 coral (F6) MC 294 coral (A5) MC 294 coral (E3) Sediment samples MC 294 4662 0–2 cm MC 294 4664 0–2 cm MC 294 4664 2–5 cm MC 294 4664 5–10 cm Oil content* (mg/g extract) Oil content* (mg/g dry weight sediment) UCM n-alkane carbon range C17-n-alkane/ pristane C18-n-alkane/ phytane Carbon Preference Index (CPI)† Ts/(Ts+Tm) C29-Ts/NH NA§ NA NA 1.71 2.33 0.86 0.59 0.49 NA NA NA NA NA NA C12–C40 C16–C40 C16–C40 2.21 1.89 1.89 2.66 2.72 2.72 0.81 0.83 0.83 0.59 0.59 0.58 0.50 0.50 0.51 310 8.0 74 73 NDjj ND ND ND C15–C40 C17–C34 C16–C41 C15–C42 2.42 2.41 2.43 1.34 2.69 2.21 1.87 2.28 1.15 1.09 0.38 1.22 0.60 0.58 0.59 0.58 0.43 0.48 0.45 0.48 630 570 68 120 3.46 9.25 0.03 0.02 C15–C37 C17–C42 C16–C42 C11–C42 0.42 0.44 0.56 0.80 0.47 0.29 0.81 1.21 1.12 0.99 1.31 1.37 0.57 0.59 0.52 0.54 0.46 0.50 0.38 0.42 Abbreviations for biomarkers: C29-Ts, 18α(H),21β(H)-30-norneohopane; NH, 17α(H),21β(H)-30-norhopane; Tm, 17α(H)-22,29,30-trinorhopane; Ts, 18α(H)22,29,30-trinorneohopane. *Oil content was calculated by integration of the UCM observed via GC-FID. † CPI = Σ(odd numbered alkane abundances from n-C23 to n-C35)/Σ(even numbered alkane abundances from n-C22 to n-C34). ‡ Described in ref. 2. § Not applicable to sample as pure oil was collected. { Described in ref. 18. jj Not determined due to collection protocol of flocculent onto filters, which did not allow for dry weights of the flocculent material to be taken post collection and before extraction. samples collected from site MC 294. Similarly, a comparison of the sterane portion of the GC×GC chromatographic plane for the Macondo well oil and floc from the coral samples also shows significant differences, particularly in the relative distributions of DiaC29βα-20S, C27αββ-20R, and C27αββ-20S in steranes (e.g., B8, Fig. S9B). Although preferential loss of steranes and diasteranes relative to hopanoids would not be expected from traditional biodegradation sequences, this trend has been observed previously for oil that undergoes severe weathering in energetic and aerobic conditions (20). This could result from either biodegradation or chemical and physical processes arising from the precipitation of the wax component of oil at the low temperatures present in the deep GoM (21, 22). Wax formation may have resulted from turbulent mixing of the well’s hot source-jet fluids with the surrounding cold seawater, fractionating constituent hydrocarbons according to their molecular characteristics (1, 2). Nevertheless, the data from the hopanoids, which have a greater fidelity, confirm the presence of Macondo well oil in the floc and surrounding surface sediment samples. The constant, albeit relatively low level, input of hydrocarbons from natural seepage in the GoM may also complicate these biomarker ratios (14). Conclusions Observations of recently damaged corals and the presence of Macondo well oil on corals indicates impact at a depth of 1,370 m, 11 km from the site of the blowout. This finding provides insight into the extent of the impact of the spill, which is significantly complicated by physical mixing processes (23) and fractionation of the oil constituents (24). Because deep-water corals are sessile and release mucous that may trap material from the water column, these corals may provide a more sensitive indicator of the impact from petroleum hydrocarbons than marine sediment cores and may record impacts from water masses passing through a community, even if no deposition to the sediment occurs. Deep-water colonial corals exhibit extreme longevity as sessile adults (hundreds to thousands of years; 25–27) and typi4 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1118029109 cally inhabit areas exposed to a moderate current regime (28). The presence of a deep-water coral community dominated by recently impacted, visibly unhealthy, and recently dead individuals (as evidenced by skeletons free of encrusting organisms), together with ophiuroid symbionts with unhealthy color and atypical posture, provides evidence of a recent waterborne impact. Although the spatial and temporal proximity of this impact to the Deepwater Horizon oil spill might be coincidental, the normal longevity of deep-water corals and the lack of visual evidence of impact to deep-water corals elsewhere in the GoM suggest that this may not be the case. Importantly, even though there are multiple inputs of oil to the GoM, the use of hopanoid biomarker compositions and ratios in the floc collected from the surface of corals allows us to establish a connection to the oil spill even though other biomarkers for characterizing oil in these environments (e.g., PAHs and sterane biomarker ratios) are affected by severe weathering (20) and, hence, are not robust under the conditions of this spill. The data suggest the Deepwater Horizon oil spill impacted a community of deep-water corals near the Macondo well. The numerous apparently healthy deep-water coral communities in other parts of the GoM may indicate that the localized impact in MC 294 found to date, is not part of a much larger, acute, GoMwide event. However, life in deep-water coral ecosystems is known to operate at a slow pace. Consequently it is too early to fully evaluate the footprint and long-term effects of acute and subacute exposure to potential waterborne contaminants resulting from the Deepwater Horizon oil spill. Materials and Methods Discovery. Areas for exploration were chosen according to examination of 3D seismic data in the BOEM database. Areas of high reflectivity and bathymetric relief were targeted for visual examination, and during the ROV dive, onboard sonar was used to find exposed carbonates that might host corals. Image Analyses. A down-looking mosaic (as in ref. 29) was constructed from 379 partially overlapping images, taken 3 m above the seafloor using White et al. Coral and Invertebrate Associate Collection. Corals and their associate ophiuroids were collected by the ROV Jason II and DSV Alvin in October and December 2010, respectively. The manipulator claws were modified with a cutting blade to aid in the collection of host corals and attached ophiuroid associates. Individuals were collected into temperature-insulated bioboxes on the sea floor and processed immediately after recovery of the vehicles. Approximately 2 to 3 cm of a coral branch or arms of individual ophiuroids were subsampled and frozen at −80 °C or in 70% ethanol for shore-based morphological and genetic analyses. Voucher specimens were either preserved in 95% ethanol or dried. Microscopic Examination. Tissue necrosis and the presence of bare skeleton were documented on a Leica S6D microscope with an attached Nikon D300 camera. Octocoral Identification. Octocorals were identified to the lowest possible taxon using molecular barcodes and morphological characters (following refs. 30–32). DNA was extracted from frozen or preserved (95% ethanol) specimens using the Qiagen DNeasy kit. The 5′ end of the mitochondrial msh gene and the COI+igr region were PCR amplified (33). Sequences were edited, combined with related sequences from GenBank, and aligned by ClustalW, resulting in a 1,430-bp region. Bayesian phylogenetic inference was conducted using the GTR+G+I model (MrBayes v3; number of generations = 2,000,000; sample frequency = 100; burnin = 5,000). Because the COI+igr region has not been previously amplified for many octocorals, these regions were coded as missing data for the appropriate GenBank specimens. Sediment and Floc Collection. Floc was collected at depth through 1.5-m-long precleaned tubing into a 4-L carboy, where it was collected onto two 15-cm diameter precombusted glass fiber filters (GF/C, >1 μm) mounted between two layers of Nytex (63 μm). The majority of the floc did not sorb to the filters and remained suspended in the seawater collected in the carboy. Once onboard ship, this material was immediately filtered onto 47-mm-diameter precombusted GF/F filters (0.7 μm), which were then placed in combusted foil and frozen at –20 °C before further analysis. Push cores (6.35 cm diameter) were used to collect sediment samples. Immediately after reaching the ship’s deck, the cores were sectioned as 0–2 cm, 2–5 cm, and 5–10 cm intervals into combusted glass jars and then frozen at −20 °C until analysis. Macondo well reference oil was collected directly above the well on June 21, 2010 (as in ref. 2). This sample is the reference Macondo well oil with which samples are compared throughout the study. Other samples (described in ref. 18) were also obtained on May 31, 2010 from a saltmarsh ≈175 km from the spill near Cocodrie Louisiana (29.29 °N; −90.49 °W) as a 2-cm-diameter droplet of surface oil (M1) and from a scraping of Spartina alterniflora saltmarsh grass taken within meters of the droplet (M2). Another sample (S1) was collected on June 20, 2010 with the R/V Endeavor from a 1-cm-thick layer of oil floating on the surface water at 28.74 °N, 88.38 °W (18). All samples were placed in combusted glass jars and frozen until further analysis. Oil Analysis. Samples were solvent extracted and purified with fully activated silica gel. Extracts were analyzed for hydrocarbons via GC-FID, gas chromatography–mass spectrometry (GC-MS), and comprehensive GC×GC. For quantification and identification, GC×GC was coupled to an FID (GC×GCFID), and identities of biomarkers were confirmed by coupling with MS (GC×GC-MS). SI Materials and Methods provides a complete discussion of these analyses. Ophiuroid Identification. Ophiuroids were identified to the lowest possible taxon using morphological and genetic data. Morphological examination of type and voucher specimens was conducted at the Smithsonian Institution (Washington, DC), and DNA from voucher specimens was obtained for ACKNOWLEDGMENTS. We thank the crew and captains of the R/V Atlantis and R/V Ron Brown; the pilots and crew of DSV Alvin and ROV Jason II; and J. Abbassi, C. Carmichael, O. Chegwidden, D. Cowart, C. Doughty, T. Enderlein, P. Etnoyer, J. Frometa, K. Halanych, K. Reuter, M. Rittinghouse, A. Sen, C. Sheline, K. Stamler and J. Thoma for their contributions to this work. This work was supported by Bureau of Ocean Energy Management Contract 1435-01-05-CT-39187 (to TDI-Brooks), the National Oceanic and Atmospheric Administration Office of Ocean Exploration, the Census of Marine Life (ChEss Project), USGS-Terrestrial, Water, and Marine Environments Program through the BOEM/Outer Continental Shelf, Northeast Gulf of Mexico Deep Offshore Reef Ecology, Lophelia II Study and National Science Foundation RAPID Grants OCE-1045131 (to H.K.W.), OCE-1045083 and OCE-1064041 (to C.R.F.), OCE-1043976 (to C.M.R.), OCE-1045025 (to R.C.), OCE-1045329 (to T.M.S.), OCE-1044289 (to C.R.G.), and OCE-1045079 (to E.E.C.). 1. Camilli R, et al. (2010) Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon. Science 330:201–204. 2. Reddy CM, et al. (2011) Composition and fate of gas and oil released to the water column during the Deepwater Horizon oil spill. Proc Natl Acad Sci USA, 10.1073/ pnas.1101242108. 3. Schroeder WW, et al. (2005) Cold-Water Corals and Ecosystems, eds Freiwald A, Roberts JM (Springer, Heidelberg), pp 297–307. 4. Cordes EE, et al. (2008) Coral communities of the deep Gulf of Mexico. Deep Sea Res Part I Oceanogr Res Pap 55:777–787. 5. Lessard-Pilon SA, Podowski EL, Cordes EE, Fisher CR (2010) Megafauna community composition associated with Lophelia pertusa colonies in the Gulf of Mexico. Deep Sea Res Part II Top Stud Oceanogr 57:1882–1890. 6. Cole GA, et al. (2001) Constraining source and charge risk in deepwater areas. World Oil 222:69–77. 7. Peters KE, Walters CC, Moldowan JM (2005) The Biomarker Guide: Biomarkers and Isotopes in Petroleum Exploration and Earth History (Cambridge Univ Press, Cambridge, UK), Vol 2. 8. Frysinger GS, Gaines RB, Xu L, Reddy CM (2003) Resolving the unresolved complex mixture in petroleum-contaminated sediments. Environ Sci Technol 37:1653–1662. 9. Reddy CM, et al. (2002) The West Falmouth oil spill after thirty years: The persistence of petroleum hydrocarbons in marsh sediments. Environ Sci Technol 36:4754–4760. 10. Nelson RK, et al. (2006) Tracking the weathering of an oil spill with comprehensive two-dimensional gas chromatography. Environ Forensics 7:33–44. 11. Ducklow HW, Mitchell R (1979) Bacterial populations and adaptations in the mucus layers on living corals. Limnol Oceanogr 24:715–725. 12. Goldberg WM, Makemson JC, Colley SB (1984) Entocladia endozoica sp nov, a pathogenic Chlorophyte: Structure, life history, physiology, and effects on its coral host. Biol Bull 166:368–383. 13. Petes LE, Harvell CD, Peters EC, Webb MAH, Mullen KM (2003) Pathogens compromise reproduction and induce melanization in Caribbean sea fans. Mar Ecol Prog Ser 264: 167–171. 14. Weber TC, De Rovertis A, Greenaway SF, Smith S, Mayer L, Rice G (2011) Estimating oil concentration and flow rate with calibrated vessel-mounted acoustic echo sounders. Proc Natl Acad Sci USA, 10.1073/pnas.1108771108. 15. Sassen R, Sweet ST, DeFreitas DA, Morelos JA, Milkov AV (2001) Gas hydrate and crude oil from the Mississippi Fan Foldbelt, downdip Gulf of Mexico Salt Basin: Significance to petroleum system. Org Geochem 32:999–1008. 16. Warburton GA, Zumberge JE (1983) Determination of petroleum sterane distribution by mass spectrometry with selective metastable ion monitoring. Anal Chem 55: 123–126. 17. Cometa PA, Rafalskaa JK, Brooks JM (1993) Sterane and triterpane patterns as diagnostic tools in the mapping of oils, condensates and source rocks of the Gulf of Mexico region. Org Geochem 20:1265–1296. 18. Carmichael CA, et al. (2012) Floating oil-covered debris from Deepwater Horizon: Identification and application. Environ Res Lett 7:015301. 19. Sassen R, et al. (1994) Organic geochemistry of sediments from chemosythetic communities, Gulf of Mexico slope. Geo-Mar Lett 14:110–119. White et al. PNAS Early Edition | 5 of 6 SPECIAL FEATURE SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE genetic comparison. DNA was extracted from frozen or preserved (95% ethanol) ophiuroids living on MC 294 corals using the Qiagen DNeasy kit. A fragment of the mitochondrial 16S rRNA (16S) gene was amplified with the universal primers 16SarL and 16SbrH, sequenced, and aligned using SEQUENCHER 4.8 (Gene Codes Corporation) and previously described methods (34). Phylogenetic inference (and subsequent species identification) was conducted using parsimony (PAUP* 4.0b10), neighboring-joining distance-based and Bayesian approaches associated with Geneious (version 5). ENVIRONMENTAL SCIENCES a Nikon E995 camera in pressure housing mounted on the ROV Jason II. Individual coral colonies were labeled (Fig. S3). Close-up images of individual corals from a side-looking perspective were obtained from frame grabs using a dedicated NDSF/AIVL Adimec 2000 HDTV digital video camera mounted on the ROV Jason II vehicle frame and the starboard manipulator of DSV Alvin. Close-up imagery was used for assessment of impact to all corals that could be approached by ROV Jason II or DSV Alvin without damaging other corals. Bare skeleton above the coral’s basal region, obviously damaged tissue (strands of mucous or loose tissue hanging from the skeleton), and areas covered by floc were scored as impacted. Levels of impact were broadly binned into five categories according to the percentage of the imaged portion of the colony showing impact: rank 0, 0–1% of the colony impacted; rank 1, 1–10% of the colony impacted; rank 2, 10–50% of the colony impacted; rank 3, 50–90% of the colony impacted; or rank 4, >90% of the colony impacted. In three cases in which the ranking category changed between the November and December visits, reexamination of images and the original video did not substantiate significant changes in the corals, and the higher-quality images obtained from DSV Alvin in December were used for the overall rankings (Fig. 3). All images can be accessed through the US National Oceanographic Data Center (accession no. 0084636). 20. Wang Z, Fingas M, Owens EH, Sigouin L, Brown CE (2001) Long-term fate and persistence of the spilled metula oil in a marine salt marsh environment degradation of petroleum biomarkers. J Chromatogr A 926:275–290. 21. Strom-Kristiansen T, Lewis A, Daling PS, Hokstad JN, Singsaas I (1997) Weathering and dispersion of naphthenic, asphaltenic, and waxy crude oils. Int Oil Spill Conf631–636. 22. Daling PS, Faksness LG, Hansen AB, Stout SA (2002) Improved and standardized methodology for oil spill fingerprinting. Environ Forensics 3:263–278. 23. Valentine DL, et al. (2012) Dynamic autoinoculation and the microbial ecology of a deep water hydrocarbon irruption. Proc Natl Acad Sci USA, 10.1073/pnas.1108820109. 24. Ryerson TB, et al. (2012) Chemical data quantify Deepwater Horizon hydrocarbon flow rate and environmental distribution. Proc Natl Acad Sci USA, 10.1073/ pnas.1110564109. 25. Andrews AH, et al. (2002) Age, growth and radiometric age validation of a deep-sea, habitat-forming gorgonian (Primnoa resedaeformis) from the Gulf of Alaska. Hydrobiologia 471:101–110. 26. Roark EB, Guilderson TP, Dunbar RB, Fallon SJ, Mucciarone DA (2009) Extreme longevity in proteinaceous deep-sea corals. Proc Natl Acad Sci USA 106:5204–5208. 6 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1118029109 27. Prouty NG, Roark EB, Buster NA, Ross SW (2011) Growth-rate and age distribution of deep-sea black corals in the Gulf of Mexico. Mar Ecol Prog Ser 423:101–115. 28. Roberts JM, Wheeler AJ, Freiwald A (2006) Reefs of the deep: The biology and geology of cold-water coral ecosystems. Science 312:543–547. 29. Pizarro O, Singh H (2003) Toward large-area mosaicing for underwater scientific applications. IEEE J Oceanic Eng 28:651–672. 30. Madsen FJ (1970) Remarks on Swiftia rosea (Grieg) and related species (Coelenterata, Gorgonaria). Steenstrupia (Cph) 1:1–10. 31. Grasshoff M (1977) Die Gorgonarien des ostlichen Nordatlantik und des Mitteleeres III Die Familie Paramuriceidae (Cnidaria, Anthozoa). Meteor Forsch-Ergebnisse 27:5–75. 32. Sanchez J (2005) Systematics of the bubblegum corals (Cnidaria: Octocorallia: Paragorgiidae) with description of new species from New Zealand and the Eastern Pacific. Zootaxa 1014:1–72. 33. McFadden CS, et al. (2011) Limitations of mitochondrial gene barcoding in Octocorallia. Mol Ecol Resour 11:19–31. 34. Cho W, Shank TM (2010) Incongruent patterns of genetic connectivity among four ophiuroid species on North Atlantic Seamounts. Mar Ecol (Berl) 31:121–143. White et al. Environmental Toxicology and Chemistry, Vol. 24, No. 5, pp. 1219–1227, 2005 q 2005 SETAC Printed in the USA 0730-7268/05 $12.00 1 .00 COMPARATIVE TOXICITY OF TWO OIL DISPERSANTS, SUPERDISPERSANT-25 AND COREXIT 9527, TO A RANGE OF COASTAL SPECIES ALAN SCARLETT,† TAMARA S. GALLOWAY,*‡ MARTIN CANTY,‡ EMMA L. SMITH,† JOHANNA NILSSON,‡ and STEVEN J. ROWLAND† †School of Earth, Ocean, and Environmental Sciences, and ‡School of Biological Sciences, University of Plymouth, Drake Circus, Plymouth, PL4 8AA United Kingdom ( Received 2 July 2004; Accepted 5 November 2004) Abstract—The acute toxicity of the oil dispersant Corexit 9527 reported in the literature is highly variable. No peer-reviewed data exist for Superdispersant-25 (SD-25). This study compares the toxicity of the two dispersants to a range of marine species representing different phyla occupying a wide range of niches: The marine sediment-dwelling amphipod Corophium volutator (Pallas), the common mussel Mytilus edulis (L.), the symbiotic snakelocks anemone Anemonia viridis (Forskål), and the seagrass Zostera marina (L.). Organisms were exposed to static dispersant concentrations for 48-h and median lethal concentration (LC50), median effect concentration (EC50), and lowest-observable-effect concentration (LOEC) values obtained. The sublethal effects of 48-h exposures and the ability of species to recover for up to 72 h after exposure were quantified relative to the 48-h endpoints. Results indicated that the anemone lethality test was the most sensitive with LOECs of 20 ppm followed by mussel feeding rate, seagrass photosynthetic index and amphipod lethality, with mussel lethality being the least sensitive with LOECs of 250 ppm for both dispersants. The results were consistent with current theory that dispersants act physically and irreversibly on the respiratory organs and reversibly, depending on exposure time, on the nervous system. Superdispersant-25 was found overall to be less toxic than Corexit 9527 and its sublethal effects more likely to be reversible following short-term exposure. Keywords—Dispersants Anemonia viridis Corophium volutator Mytilus edulis Zostera marina es of Warren Spring Laboratory Specification LR 448(OP) and has been approved as a type-2, as well as a type-3, dispersant under test qualification CSR 4600/8902798 (Oil Slick Dispersants Ltd. product profile, cited May 5, 2004; http:// www.croftpark.co.uk/osd-products.html). Superdispersant-25 has been tested by the Center for Environment, Fisheries, and Aquaculture Science and has been found to be of low toxicity to Crangon crangon (brown shrimp) for use at sea and on beaches, and Patella vulgata (common limpet) for use on rocky shores; it is licensed under the Ministry of Agriculture, Fisheries, and Food, Food and Environment Protection Act 53/ 98 (Oil Slick Dispersants Ltd. product profile, cited May 5, 2004; www.croftpark.co.uk/osd-products.html). Corexit 9527 has been tested extensively in the laboratory and used on oil spills since 1978 [4]. A considerable number of toxicity reports exist concerning a wide variety of species, reviewed by George-Ares and Clark [3]. Thus Corexit 9527 provides a useful comparative toxicant for the study of SD-25. The use of dispersants within enclosed bodies of water may pose a threat to a diverse range of species. This study compares the toxicity of SD-25 with that of Corexit 9527 to the marine sediment-dwelling amphipod Corophium volutator (Pallas), the blue mussel Mytilus edulis (L), the symbiotic snakelocks anemone Anemonia viridis (Forskål), and the seagrass Zostera marina (L.). The mudshrimp C. volutator is distributed widely around the coasts of western Europe and northeast America, and is significant in structuring and sustaining the ecology of near-shore sediment communities [5,6]. Corophium volutator is now used commonly as a European acute toxicity test organism [7–12]. Amphipods occupying a similar niche exist in other regions, e.g., Ampelisca abdita (Mills) also are used for toxicity testing. Blue mussels mainly occur on exposed rocky INTRODUCTION Faced with the prospect of an oil spill coming ashore or passing over reefs, decisions have to be made swiftly as to how best to deal with the situation. One option is to use chemical dispersants to break up the slick into a large number of small droplets. Once broken up, the slick poses less of a physical risk to seabirds or marine mammals but may transfer oil into the water column and possibly to the benthos. Within estuaries, inlets, enclosed bays, or shallow water reefs, the concentration of the dispersants alone may be sufficient to cause toxic effects. In the United Kingdom, dispersants cannot be used in water less than 20 m deep or within one nautical mile of such [1] without the permission of the Department for Environment Food and Rural Affairs; similar rules relating to sensitive habitats such as coral reefs and mangroves exist in tropical regions [2]. Hence, the option to use dispersants within estuaries, inlets, and shallow water does exist and it is in such circumstances that difficult decisions on how best to protect the environment and commercial operations have to be made. The handling of large volumes of dispersant under difficult conditions may result in accidental release of potentially toxic chemicals into the sea. Research into the toxicity of dispersants has been reported widely [3,4] and companies continue to improve the efficiency of the chemicals and reduce their toxicity. In the United Kingdom, the oil dispersant Superdispersant-25 (SD-25) is now the Maritime and Coastguard Agency’s main stockpiled chemical for spraying onto oil slicks at sea. No data exist within peer-reviewed literature for SD-25. However, SD-25 in association with oil meets all the relevant claus* To whom correspondence may be addressed ([email protected]). 1219 1220 Environ. Toxicol. Chem. 24, 2005 shores and are distributed widely from the Arctic to the Mediterranean with related species distributed worldwide. Mussels have been used for long-term monitoring projects such as the Global Mussel Watch [13], field surveys [14], and Scope for Growth studies [15], allowing a large body of knowledge to accumulate regarding their acute and sublethal responses to stressors. Snakelocks anemones occur on intertidal rocky shores and are associated closely with seagrass beds [16]. As symbionts possessing zooxanthellae, anemones have been used as surrogate organisms for the study of coral organisms (e.g., [17,18]). Eelgrass, Z. marina, is a marine angiosperm with a worldwide distribution and is protected strictly under the Berne Convention [19]. North Atlantic populations have suffered great losses during the last century [20] and, as a consequence, Z. marina is now deemed to be scarce. Eelgrass plants grow well under laboratory conditions and have proved to be a robust test species if chlorophyll fast fluorescence transient is used to measure plant health [21]. Taken together, these species represent a broad range of organisms that might be affected by dispersant use. Some degree of extrapolation to other similar species can be justified from the toxicity data obtained. The aim of this study was to compare the acute and sublethal toxicities of the two dispersants, Superdispersant-25 and Corexit 9527, following 48-h static exposures. The extent of recovery of each species was documented for up to 72 h with regard to changes in the no-observable-effect concentration (NOEC) and LOEC values. Behavioral responses of test animals also were recorded. MATERIALS AND METHODS Collection and maintenance of organisms Corophium volutator and sediment were collected from an intertidal area of the Avon estuary near Aveton Gifford, South Devon, United Kingdom (ordinance survey grid reference: SX 683 467). Amphipods were sieved from the upper 5 cm of sediment and transported back to the laboratory within 1 h, where they were placed in 5-L culture tanks lined with fieldcollected, sieved (,300-mm) sediment. The tanks were filled with filtered seawater 25 6 1 practical salinity units (psu), which was aerated and maintained at 15 6 18C with a 12:12h light:dark cycle. The animals were fed weekly with five drops standard aquarium invertebrate food (Waterlife Invert Food, Waterlife Research Industries, Longford, UK; Liquifry Marine, Interpret Ltd., Dorking, UK; Roti-Rich, Florida Aqua Farms, Dade City, FL, USA; and dried algae) and the water was replaced 24 h after feeding. Amphipods were maintained under the above conditions for one to two weeks after removal from the field to acclimate them to experimental conditions. Mussels were collected from Port Quin, on the North Cornwall coast, United Kingdom (ordinance survey grid reference: SW 972 905). Organisms of 30-mm 6 5 mm length were removed carefully from the rock by cutting the byssal threads and transported back to the laboratory within 2 h. Any epifauna and epiphytes were removed from the mussels, which then were placed in 20-L tanks filled with filtered 34 6 1 psu seawater. Other conditions were as above. Anemones were collected from Jennycliff Bay, Plymouth Sound on the South Devon coast, United Kingdom (ordinance survey grid reference: SX 491 523) and transported back to the laboratory within 1 h. Maintenance conditions were same as for the mussels. Eelgrass was collected from the Yealm estuary, South Devon, United Kingdom (ordinance survey grid reference: SX 530 A. Scarlett et al. 476) and transported to the laboratory within 1 h. The outer leaves were removed so that the three youngest leaves remained. These leaves were wiped with paper to remove epiphytes and shortened to a maximum length of 300 mm above the sheath. Roots were standardized to three rhizome segments. The plants were placed in 20-L tanks filled with filtered 34 6 1 psu seawater without any sediment or additional nutrient. Other conditions were as above. Preparation of test solutions Corexit 9527 was supplied gratis from the U.S. Minerals Management Service and SD-25 was obtained from the U.K. Maritime and Coastguard Agency. Nominal exposure concentrations were prepared by direct syringe injection or pipetting of dispersant into seawater adjusted to the required salinity for the toxicity tests as appropriate (25 6 1 psu for Corophium, 34 6 1 psu for other spp.) and vortex mixing for 3 min at high speed. Rationale and general test conditions Toxicity tests of dispersants have been performed using short-term, static, continuous flow-through and spiked declining flow-through exposures. Exposure conditions have a considerable influence on the reported toxicity values with spiked declining exposures giving much higher LC50 values, approximately 3 to 23 times greater, than continuous exposures [3]. Here we used static exposures of 48-h duration noting that the length of exposure does not necessarily produce an incipient LC50 for any of the species tested during the 48-h exposure and longer exposure, therefore, may result in lower LC50 values. As well as acute toxicity, the ability of organisms to recover in clean seawater for up to 72 h was assessed relative to the 48-h endpoints. The species-specific sublethal endpoints for each test are described in the following paragraphs. Test vessels (2-L Pyrex beakers) were maintained at 15 6 18C with a 12:12 h light:dark cycle. Beakers were sealed loosely with Parafilmt M (Pechiney Plastic Packaging, Menasha, WI, USA) and aerated via a glass Pasteur pipette. Dissolved oxygen, pH, temperature, and salinity were measured after 0, 24, and 48 h in one replicate from each treatment. Dissolved oxygen was measured in all replicates at the beginning and end of the experiment. Three replicate vessels per treatment were used in all tests except the mussel feeding-rate bioassay in which nine replicate vessels per treatment were used. Test results were analyzed statistically using one-way analysis of variance after checking for normality and homogeneity of variances; 48-h LC50 values were derived using the trimmed Spearman-Kärber method. The EC50 values were obtained from regression of log-transformed data. Corophium volutator In order to obtain a concentration-response relationship, nominal exposure concentrations of 0, 50, 125, 175, 213, 250, 375, and 500 ppm Corexit 9527 and SD-25 were prepared as above. Sieved sediment (,300 mm), approximately 160 ml, was placed in the test vessels to a depth of 15 mm and 1 L of test solution added and allowed to settle for 2 h before 20 amphipods (size range 3–7 mm) were introduced via a plastic Pasteur pipette to each test vessel, a total of 60 amphipods per treatment. Beakers were inspected for amphipod activity and behavior after 18, 24, and 42 h. At the end of the exposure period of 48 h, the sediment was sieved (300 mm) and the number of live, dead, and missing individuals recorded. Miss- Environ. Toxicol. Chem. 24, 2005 Comparative toxicity of Superdispersant-25 and Corexit 9527 ing individuals were assumed to have died. Surviving amphipods were placed in clean sediment and seawater for a 24h recovery period to assess their ability to burrow. Mytilus edulis Nominal exposure concentrations of 0, 80, 130, 200, 250, and 320 ppm Corexit 9527 and SD-25 were prepared as above. Three mussels, length 30 mm 6 5 mm, were placed in the test vessels with 1 L of test solution, a total of nine mussels per treatment. At the end of the exposure period, the mussels were categorized as: Closed, open but able to close when stimulated, or open but failed to close when stimulated. Mussels in the former two categories were placed in clean seawater for a further 72 h. Open mussels that failed to respond to stimuli were considered dead and were excluded from the recovery trial. Following the recovery period of 72 h, mussels were reassessed. An additional test was performed with a single toxicant treatment of 50 ppm for both dispersants plus controls using only one mussel per test vessel but with nine replicates per treatment. At the end of the 48-h exposure period, the feeding rate of the mussels was assessed based on methodology described by Donkin et al. [22]. Mussels were placed individually in 400-ml glass beakers containing 350 ml of clean seawater. After a 10-min acclimation period with slow vortex mixing, 500 ml of Isochrysis sp. algal solution was added to give approximately 2 3 104 cells ml21. A 20-ml water sample was removed immediately from all the beakers upon the addition of the algae and retained in vials for algae enumeration. Further samples were taken after 15 and 30 min. Algal particles were analyzed using a Beckman Z2 Coulter particle count and size analyzer (Beckman Coulter, Wycombe, UK), which was set to count particles between 3 to 10.43 mm. From the loss of algal particles during the 30-min period, the feeding rates of the mussels were determined. Anemonia viridis Scoping tests had indicated that anemones were highly sensitive to both dispersants; therefore, a lower concentration range of 0, 10, 20, 40, 80, and 160 ppm was used followed by a 24-h recovery period. One anemone per test vessel was exposed to 1 L of test solution, a total of three anemones per treatment. At the end of the 48-h exposure period, the anemones were classified as having extended or retracted tentacles. The tentacles were stimulated gently with a glass rod and their response recorded. Anemones that were insensitive to stimuli were classed as moribund because they would be unable to feed or defend themselves from predation and, therefore, were unlikely to survive in their natural environment. The anemones then were placed in clean seawater for a 24-h recovery period and reassessed. A confirmatory test was performed using three anemones per test vessel at 20 and 30 ppm, a total of nine anemones per treatment. The recovery period was extended to 72 h to test if further recovery was possible. The effect of the dispersants on the density of zooxanthallae within the tentacles of A. viridis also was assessed using anemones exposed to the 0 to 160 ppm concentration range following 48-h exposure and 24-h recovery. Zooxanthallae were separated from the host material after homogenization of weighed tentacles in a Wheaton glass homogenizer in 1 ml of buffer solution (50 nM HEPES, 1 nM ethylenediaminetetraacetate, pH 7.4). Whole homogenate was centrifuged for 3 min at 735 G (Eppendorf 5415C, Eppendorf UK, Cambridge, UK). The supernatant was discarded and the insoluble plant material retained and resus- 1221 pended in buffer. Centrifugation was repeated until samples were free of host tissue. The resulting pellet was resuspended in buffer to give a cell density of about 1 3 106 cells ml21. Zooxanthallae numbers were determined by replicate counts (n 5 5) and converted to zooxanthallae per wet weight of tentacle. Zostera marina Because mortality is not a practical parameter for a plant toxicity test, the photosynthetic efficiency of eelgrass was examined using the chlorophyll fast fluorescence JIP transient (from 50 ms to 1 s; J is a characteristic step at 2 ms, I is an intermediate step at 30 ms, and P is maximum fluorescence) measurements [23,24]. Three plants per exposure vessel were exposed to a concentration range of 0, 80, 130, 200, 320, and 500 ppm dispersant, a total of nine plants per treatment, followed by a 24-h recovery period in clean seawater. Plants were dark-adapted for 5 min using a clip system before illumination with 660 nm light from a light-emitting diode source built into the fluorometer sensor. Fluorescence measurements at F0 (50 ms), F300ms Fj (2 ms), and Fm (tFmax) were recorded using a Handy PEA plant efficiency analyzer (Hansatech Instruments, Kings Lynn, Norfolk, UK) after 24 and 48 h. The Biolyzerq software (Ronald Maldonado-Rodriguez, Bioenergetics Laboratory, University of Geneva, Geneva, Switzerland, 2002) was used to load the full fluorescence transients and to calculate the JIP parameters according to the equations of the JIP-test [23,24]. The effect of the dispersants on the plant’s photosynthetic apparatus was assessed by statistical comparison of the main parameters: Yield of primary photochemistry expressed in ratio of variable to maximal fluorescence (Fv/ Fm), performance index (PI), and the area between the fluorescence curve and the maximal fluorescence intensity. Ten measurements per exposure vessel were taken at random points on the leaves avoiding any necrotic lesions (minimum of 10 mm distance). The plants then were placed in clean seawater and further measurements recorded after 24 h. RESULTS Physical conditions Dissolved oxygen was above 70% saturation, pH in the range 7.8 to 8.2, salinity 25 6 1 psu (Corophium test) or 34 6 1 psu (other spp. tests), and temperature 15 6 18C during the exposure period in all test vessels. Corophium volutator No activity was observed after 18 h at nominal dispersant concentrations at or below 125 ppm, or within the controls. At dispersant concentrations of 175 ppm and above, individuals were showing signs of stress: Either crawling on the sediment surface or swimming erratically. The activity caused greater turbidity within treatments compared to the controls. After 24 h, several moribund individuals were visible in exposure treatments of 375 and 500 ppm for both Corexit 9527 and Superdispersant-25. After 42 h of exposure, moribund individuals also were visible at and above 175 ppm for both dispersants, although this clearly was greater within the Corexit 9527 treatments. By the end of the experiment, little activity was observed, resulting in a reduction in turbidity, especially at the higher concentrations. Mortality was less than 1% within the controls and 100% at the highest nominal concentration of 500 ppm for both Corexit 9527 and SD-25 (Fig. 1). Although there was a sharp 1222 Environ. Toxicol. Chem. 24, 2005 A. Scarlett et al. Table 2. Responses of the mussel Mytilus edulis after 48-h dispersant exposures (48 h) and a 72-h recovery period (Rec.) Mussels from all concentrations were pooled (Superdispersant-25 [SD-25] n 5 45 and Corexit 9527 n 5 45; Control n 5 9 and classified as closed (closed), open and responsive to stimuli (active), or open and unresponsive to stimuli (dead) Closed (%) Control SD-25 Corexit 9527 a Fig. 1. Mortality (mean 6 standard error) of the mudshrimp Corophium volutator following a 48-h static exposure to the oil dispersants Corexit 9527 and Superdispersant-25 (each treatment contained three replicates of 20 amphipods). Active (%) Dead (%) 48 h Rec. 48 h Rec. 48 h Rec.a 0 44 67 0 58 51 100 36 11 100 0 0 0 20 22 0 42 49 Dead mussels from the 48-h exposure were excluded from the recovery vessels but are recorded in the recovery column. close promptly when touched) were able to close very slowly and then remain closed. Therefore, only those that absolutely failed to close when touched were deemed to have died (Fig. 2a). At the lowest concentration of 80 ppm, the bivalves exposed to Corexit 9527 experienced about 20% mortality but variation was high and significant mortality ( p , 0.05) only occurred at 250 ppm for Corexit 9527–and SD-25–exposed organisms (Table 1). High variation within treatments occurred at all treatment levels and reliable LC50 values could not be calculated. Following 72 h in clean seawater, the number of dead mussels had increased, but NOEC and LOEC values were unchanged (Fig. 2b). Sublethal effects were assessed at dispersant concentrations of 50 ppm for 48 h using a feeding rate bioassay (n 5 9). Two mussels within the Corexit 9527 exposure died and were omitted from the bioassay. Feeding rates were reduced significantly ( p , 0.05) for both dispersant exposures with SD-25 rates reduced to 9.8% of controls and Corexit 9527–exposed mussels only 2.6% (Fig. 3). Corexitexposed mussels had feeding rates significantly ( p , 0.05) less than that of SD-25–exposed organisms. fall in survival of C. volutator exposed to 175 ppm Corexit 9527, a more gradual decline was observed within the SD-25 treatments. The LC50 (48-h) values of 159 (95% confidence limits 145–173 ppm) and 260 ppm (95% confidence limits 240–282 ppm) were calculated for Corexit 9527 and SD-25, respectively (Table 1). The NOEC was 125 ppm for both dispersants. Mortality of Corexit 9527 exposed amphipods was significantly ( p , 0.05) greater than those exposed to SD-25. All of the surviving individuals exposed to SD-25 concentrations up to 175 ppm were able to swim normally and were able to rebury in clean sediment, although swimming activity still was observed 3 h following introduction into the clean seawater; all survived the recovery period. Survivors from Corexit 9527 also were able to rebury and survive the recovery period with exposures up to 125 ppm. However, above 175 ppm, 100% of amphipods failed to recover from Corexit 9527. Anemonia viridis Mytilus edulis The anemones were highly sensitive to both dispersants with 55% of SD-25–exposed and 100% of Corexit 9527–exposed organisms insensitive to stimuli at 20 ppm (Fig. 4a). At 40 ppm for both dispersants, all organisms had retracted tentacles that failed to respond to stimuli and were assessed to be moribund. Above 80 ppm, anemone tissue was starting to decompose. The concentration range between no effect and Control mussels appeared active and healthy (i.e., open and responsive to stimuli) throughout the test period but all dispersant-exposed organisms were classified as either closed or dead at the end of the 72-h recovery period (Table 2). Assessment of mortality after 48 h was problematic because some mussels that initially appeared dead (i.e., open and failing to Table 1. Comparison of toxicity estimates for Superdispersant-25 and Corexit 9527 with four marine species Superdispersant-25 concn. (ppm) Corexit 9527 concn. (ppm) Species Anemonia viridis Corophium volutator Mytilus edulis M. edulis Zostera marina Test No response to stimuli Mortality Mortality Feeding rate JIP-testf PIg NOECa LOECb 10 20 125 200 ,50 ,80 175 250 50 80 LC50 or EC50c 15d 159 — NA 55 NOEC 5 no-observed-effect-concentration. LOEC 5 lowest-observed-effect-concentration. c LC50 5 median lethal concentration; EC50 5 median effect concentration. d Interpolated from data only, not derived from model. e NA 5 not applicable. f JIP-test 5 measurements acquired from the fast fluorescence transient. g PI 5 performance index. a b Confidence limits NAe 145–173 — 28–150 NOEC LOEC LC50 or EC50 10 20 20a 125 200 ,50 ,80 175 250 50 80 260 — NA 386 Confidence limits NA 240–282 — 339–439 Environ. Toxicol. Chem. 24, 2005 Comparative toxicity of Superdispersant-25 and Corexit 9527 1223 Fig. 2. Mortality (mean 6 standard error) of the mussel Mytilus edulis following a 48-h static exposure (a) and a further 72-h recovery period (b) to the oil dispersants Corexit 9527 and Superdispersant-25 (each treatment contained three replicates of three mussels). 100% mortality was too small to calculate LC50 values using the Spearman-Kärber method. However, interpolation of the data indicated EC50 values of about 20 ppm for SD-25 and about 15 ppm for Corexit 9527. The NOEC and LOEC (48h) values were 10 ppm and 20 ppm, respectively for both dispersants (Table 1). Following recovery in clean seawater, Fig. 4. Percentage of moribund Anemonia viridis (mean 6 standard error), following a 48-h static exposure (a) and a further 24-h recovery period (b) to the oil dispersants Corexit 9527 and Superdispersant25 (the 20- and 30-ppm treatments contained three replicates of three anemones; all other treatments contained three replicates of one anemone). all of the anemones previously exposed to 30 ppm SD-25 and below were able to respond to stimuli, as did all organisms previously exposed to Corexit 9527 at 20 ppm (Fig. 4b); however, 89% of the anemones previously exposed to 30 ppm Corexit 9527 failed to respond to stimuli (Fig. 4b). No significant differences were observed in zooxanthallae densities. Zostera marina Fig. 3. Feeding rates (mean 6 standard error) of the mussel Mytilus edulis following a 48-h static exposure of 50 ppm to the oil dispersants Corexit 9527 and Superdispersant-25 (each treatment contained nine replicates of one mussel). All of the main parameters, Fv/Fm, PI, and area, were reduced at the lowest exposure of 80 ppm after 24 and 48 h for both dispersants (Fig. 5); the NOEC for both dispersants was ,80 ppm. The PI was the most-sensitive parameter giving 48h EC50 values of 386 ppm and 55 ppm for SD-25 and Corexit 9527, respectively. At the lowest exposure concentration there was no significant difference between the dispersants; however, at 130 ppm and above, Corexit was significantly ( p , 0.05) more toxic for all main parameters. Leaves of the Corexit-exposed plants turned brown at 200 ppm and the outer leaves started to become detached, leaving only the more protected inner leaf from which to take measurements. As well as the main parameter values obtained from the PEA, the JIPtest allows calculation of various bio-physical expressions, such as specific fluctuations and yields, and phenomenological fluctuations. To visualize the effect of the dispersants on the 1224 Environ. Toxicol. Chem. 24, 2005 A. Scarlett et al. Fig. 5. Photosynthetic parameters (Fv/Fm), performance index (PI), and area (mean 6 standard error), measured in the leaves of Zostera marina after 24- and 48-h exposure, and 24-h recovery in clean seawater, to the oil dispersants Corexit 9527 and Superdispersant-25 (each treatment contained three replicates of three plants from which 10 measurements were taken). eelgrass leaves, pipeline models have been calculated and drawn on the basis of experimental signals from the 200-ppm concentrations and compared to control leaves (Fig. 6). The SD-25–exposed plants were able to recover slightly after 24 h in clean seawater, e.g., mean PI values rose from 0.88 (standard error [SE] 0.01) after 48 h at 80 ppm exposure to 1.07 (SE 0.04) following recovery. Corexit-exposed plants failed to recover, e.g., mean PI values were 0.84 (SE 0.01) after the 48-h exposure to 80 ppm and fell to 0.76 (SE 0.07) at the end of the 24-h recovery period (Fig. 4). DISCUSSION AND CONCLUSION This is, to our knowledge, the first report of toxicity data for SD-25 that reveals its acute toxicity to a range of marine Comparative toxicity of Superdispersant-25 and Corexit 9527 Fig. 6. Comparison of membrane models of the Photosystem II (PSII) apparatus in control plants (top), and 48-h exposure to 200 ppm Superdispersant-25 (middle) and Corexit 9527 (bottom). The models represent the specific activities expressed as fluctuation per reaction center (RC). The relative magnitude of each activity or fluctuation is shown by the width of the corresponding arrow. Absorption flux (ABS) is proportional to the concentration of antenna chlorophyll and the average antenna size is given as ABS/RC. This expresses the total absorption flux of PSII antennae chlorophyll divided by the number of active, in the sense of primary quinone (QA)-reducing, reaction centers. The absorption and trapping by PSII units with a heat sink center (non-QA–reducing) is indicated as the hatched parts of the arrows ABS/RC and energy flux for trapping (TRo/RC). The antenna belonging to the PSII units with heat sink centers is drawn in black and the antenna that belongs to the active centers is drawn in white. The degree of stress, thus, is indicated by the reduction in ABS/RC and TRo/RC leading to a greater dissipation of energy in the form of heat (Dlo/RC) and a decrease in the useful energy available to the plant (ETo/RC). species to be significantly lower ( p , 0.05), in the majority of tests, or equivalent (mussel mortality and anemone tests) to the widely used dispersant Corexit 9527. Dispersants contain surfactants, which may be nonionic or anionic, dissolved, or suspended in solvents. Corexit 9527 is a mixture of both nonionic (48%) and anionic (35%) surfactants in an aqueous solvent containing ethylene glycol monobutyl ether (17%); the surfactants include ethoxylated sorbitan mono- and trioleates (nonionic) and sodium dioctyl sulfoccinate (anionic) [25], but little is known about the constituents of SD-25 except that it is a blend of glycol and glycol ether solvents, combined with nonionic and anionic surfactants [26]. Dispersants are thought to act physically and irreversibly on the respiratory organs and reversibly, depending on exposure time, on the nervous system Environ. Toxicol. Chem. 24, 2005 1225 [27]. Surfactants can bind to and disrupt cellular phospholipid bilayers altering the transmembrane sodium gradient [28]. A review of Corexit 9527 toxicity data by George-Ares and Clark [3] covering 28 reports and 37 aquatic species, found no apparent trend in the reported sensitivity between taxa. Toxicity (24–96–h LC50 or EC50) ranged from 1.6 to .1,000 ppm. Amphipod LC50 values varied between 3 ppm for Allorchestes compressa [29] and .175 ppm for Boekosimus sp. [30], but no data for C. volutator were reported. Mollusc (adults) LC50 values were in the range 33.8 ppm for the sand snail Polinices conicus [29] to 2,500 ppm for the scallop Argopecten irradians [31]; however, no data for Mytilus edulis were reported. The class Anthozoa (anemones and true corals) were not represented in the reviewed papers and the only seagrass was Thalassia testudinum with a 96-h LC50 of 200 ppm [32]. In general, the reviewed data show that Corexit 9527 is of low acute toxicity to most species, although embryo and larval stages are more sensitive. The results from this study are consistent with the literature data but suggest an order of sensitivity, based on NOEC and LOEC values, of: Anthozoa . macrophyte . crustacean . mollusc for both Corexit 9527 and SD-25, although the order of sensitivity partly is due to the differences in test methodology and endpoints. Both dispersants caused a sharp increase in mortality over a narrow concentration range; this was most pronounced for Corexit 9527. Therefore, it is of more use to refer to NOEC or LOEC than LC50 values. The NOEC values for A. viridis were much lower than that observed for the other species tested (Table 1) and the sensitivity of the anemones was similar to that of the embryo or larval stages of molluscs and fish [33,34]. Of the anemones with extended tentacles that were unresponsive to stimuli following 48-h dispersant exposure of 20 ppm and, therefore, classed as moribund (Fig. 4a), all were able to recover when placed in clean seawater (Fig. 4b). All of the SD-25–exposed anemones also were able to recover from the 30-ppm exposure, but this was true of only 11% of Corexitexposed anemones. The ability of the organisms to recover implies that the dispersants act reversibly on the neural receptors at low concentrations but cause irreversible membrane damage at higher concentrations, consistent with reported surfactant toxicity modes of action [27,28]. No effect on zooxanthallae density was found and the cells appeared undamaged even when their host tissue was damaged severely. It is likely that the host tissue protected the plant cells and that, during short-term exposure, the algae do not migrate from their hosts as happens with coral bleaching episodes. Anemones have been used as a surrogate organism for symbiotic coral assemblages [17,18], although it is not known if the skeletal cup of reef-building corals would afford the same protection for the zooxanthallae as the anemone. The M. edulis mortality data were very variable, but Corexit 9527 clearly was more toxic at low concentrations (Fig. 2a). The total percentage of SD-25–exposed mussels that were closed at the end of the 48-h exposure period increased from 44 to 58%, but Corexit-exposed closed mussels decreased from 67 to 51% with a corresponding rise in the percentage dead (Table 2). A very pronounced effect was observed in the feeding rates of mussels exposed to 50 ppm of both dispersants, with Corexit-exposed organisms’ filtration rates reduced so much that they were not significantly greater than system blanks (Fig. 3). Although the highly reduced filtration rates could be a behavioral response to the dispersants, the mussels’ failure to open within the 72-h recovery period following the 1226 Environ. Toxicol. Chem. 24, 2005 48-h concentration-response test (Table 2) suggests that the mussels may have been suffering physiological damage as they would be expected to recover from nonspecific narcosis. Although there was no sublethal component to the C. volutator test, it was observed that many individuals failed to burrow quickly when exposed to $175 ppm dispersants. It was impossible to tell if it was the swimming amphipods that died later, although these likely had a greater exposure than those that burrowed rapidly into the sediment. Briggs et al. [8] used turbidity caused by stressed Corophium as a measure of toxicity and found a good correlation between turbidity during the first 24 h of exposure and mortality after 10 d. The failure to burrow and increased swimming activity observed at moderate contamination levels, are most likely an escape behavioral response rather than a physiological response because they occur with a range of toxicants [8,35]. The dispersant concentrations in this study that caused the greatest turbidity were within the confidence limits of the calculated LC50 values (Table 1). Under test conditions where escape was impossible, the behavioral response not to burrow probably proved to be detrimental, but in the wild it may enable the organisms to escape from localized areas of high concentration. Both dispersants disrupted the Photosystem II (PSII) apparatus, known to be a primary casualty during stress conditions [36], within the leaves of Z. marina and this was found to be largely irreversible, especially for Corexit 9527 (Fig. 5). The PI is a standard composite of several JIP-test parameters and has been shown to be a highly sensitive measure that is correlated strongly with other measurements of plant health [24]. Indeed, at concentrations $200 ppm Corexit 9527, it was evident by visual inspection that the leaves were damaged severely and the outer leaves were becoming detached. Leaves of Z. marina possess a thin cuticle [37] that may afford a degree of protection from the dispersants, although it is clear that 24 h was sufficient for photosynthesis to be affected. Algae and spores with less protection likely are more sensitive, as found by Singer et al. [34], exposing zoospores of giant kelp, Macrocystis pyrifera, to Corexit 9527. Extrapolating the results from this and other studies suggests that the most sensitive organisms to dispersants are those with the least protective tissue or shell. As well as the spore, embryo, and larval stages, it is likely that unprotected organisms such as nudibranchs and seaslugs would be vulnerable and, therefore, extra care may need to be taken where rare or endangered species are present. Under current dispersant-use guidelines that exist in the United Kingdom and elsewhere [1,2], dispersant concentrations are unlikely to be sufficiently high to cause harm to aquatic organisms within open waters. Measured concentrations at sea are rare but Bocard et al. [38] reported a maximum concentration of 13 ppm. However, if dispersants are used or spilled accidentally within estuaries, enclosed bays, or shallow coral reefs, it may be possible for concentrations to become elevated long enough to cause mortality of sensitive species and stress to more tolerant animals and plants. Any harm to aquatic species as a result of dispersant use must be measured against the potential harm that may arise from the toxic or physical effects of oil. The difficult decision of when, where, and under what conditions to use dispersants will remain as long as oil is transported via the sea. Spill response agencies are better able to make such decisions when provided with relevant ecotoxicological data. To our knowledge, these are the first reported results concerning the toxicity of Superdispersant-25; it has shown that this dispersant gen- A. Scarlett et al. erally is of lower toxicity than Corexit 9527 to a range of coastal species, although equivalent in terms of NOEC and LOEC values; also, at maximum concentrations likely to be found at sea, any sublethal effects upon organisms are more likely to be reversible. Acknowledgement—This study was financed by the Maritime and Coastguard Agency, the Department for Environment Food and Rural Affairs, the Department of Trade and Industry, and Minerals Management Service Grant RP 480. The Photosynthesis Membrane model was derived using Biolyzert software provided by R. MaldonadoRodriguez, Bioenergetics Laboratory, University of Geneva, Geneva, Switzerland. REFERENCES 1. Maritime and Coastguard Agency. 2002. Contingency Planning for Marine Pollution Preparedness and Response: Guidelines for Ports, 2002. Southampton, UK. 2. Etkin DS. 1998. Factors in the dispersant use decision-making process: Historical overview and look to the future. Proceedings of the Twenty-First Arctic and Marine Oil Spill Program Technical Seminar. Environment Canada, Ottawa, ON, pp. 281–304. 3. George-Ares A, Clark JR. 2000. Acute toxicity of two Corexit dispersants. Chemosphere 40:897–906. 4. U.S. National Research Council. 1989. Using Oil Dispersants on the Sea. National Academy, Washington, DC. 5. Limia JM, Raffaelli D. 1997. The effects of burrowing by the amphipod Corophium volutator on the ecology of intertidal sediments. J Mar Biol Assoc UK 128:147–156. 6. Raffaelli D, Milne H. 1997. An experimental investigation of the effects of shorebird and flatfish predation on estuarine invertebrates. Estuar Coast Shelf Sci 24:1–13. 7. Bat L, Raffaelli D. 1998. Sediment toxicity testing: A bioassay approach using the amphipod Corophium volutator and the polychaete Arenicola marina. J Exp Mar Biol Ecol 226:217–239. 8. Briggs AD, Greenwood N, Grant A. 2002. Can turbidity caused by Corophium volutator (Pallas) activity be used to assess sediment toxicity rapidly? Mar Environ Res 55:181–192. 9. Roddie BD, Thain JE. 2001. Biological effects of contaminants: Corophium sp. sediment bioassay and toxicity test. Techniques in Marine Environmental Sciences 28. International Council for the Exploration of the Sea, Copenhagen, Denmark. 10. Center for Environment, Fisheries, and Aquaculture Science. 2001. Monitoring and surveillance of nonradioactive contaminants in the aquatic environment and activities regulating the disposal of wastes at sea, 1998. Aquatic Environment Monitoring Report 53. Center for Environment, Fisheries, and Aquaculture Science, Lowestoft, UK. 11. Conradi M, Depledge MH. 1998. Population responses of the marine amphipod Corophium volutator (Pallas, 1766) to copper. Aquat Toxicol 44:34–45. 12. Ciarelli S, Vonck WAPMA, van Stralen NM. 1997. Reproducibility of spiked-sediment bioassays using the marine benthic amphipod, Corophium volutator. Mar Environ Res 43:329–343. 13. Goldberg ED. 1975. The Mussel Watch—a first step in global marine monitoring. Mar Pollut Bull 6:111–113. 14. Galloway TS, Sanger RC, Smith KL, Fillmann G, Readman JW, Ford TE, Depledge MH. 2002. Rapid assessment of marine pollution using multiple biomarkers and chemical immunoassays. Environ Sci Technol 36:2219–2226. 15. Widdows J, Donkin P, Brinsley MD, Evans SV, Salkeld PN, Franklin A, Law RJ, Waldock MJ. 1995. Scope for growth and contaminant levels in North Sea mussels Mytilus edulis. Mar Ecol Prog Ser 127:131–148. 16. Connor DW, Brazier DP, Hill TO, Northen KO. 1997. Marine Nature Conservation Review: Marine Biotope Classification for Britain and Ireland, Vol 1—Littoral Biotopes. Version 97.06, Report 229. Joint Nature Conservation Committee, Peterborough, UK. 17. Morrall CE, Galloway TS, Trapido-Rosenthal HG, Depledge MH. 2000. Characterization of nitric oxide synthase activity in the tropical sea anemone Aiptasia pallida. Comp Biochem Physiol B 125:483–491. 18. Whitehead LF, Douglas AE. 2003. Metabolite comparisons and Comparative toxicity of Superdispersant-25 and Corexit 9527 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. the identity of nutrients translocated from symbiotic algae to an animal host. J Exp Biol 206:3149–3157. Davison DM, Hughes DJ. 1998. An Overview of Dynamics and Sensitivity Characteristics for Conservation Management of Marine SACs, Vol I—Zostera Biotopes. UK Marine SACs Project. Scottish Association for Marine Science, Argyll, UK. Vergeer LHT, Aarts TL, deGroot JD. 1995. The wasting disease and the effect of abiotic factors (light-intensity, temperature, salinity) and infection with Labyrinthula zosterae on the phenolic content of Zostera marina shoots. Aquat Bot 52:35–44. Scarlett A, Donkin P, Fileman TW, Evans SV, Donkin ME. 1999. Risk posed by the antifouling agent Irgarol 1051 to the seagrass Zostera marina. Aquat Toxicol 45:159–170. Donkin P, Widdows J, Evans SV, Worrall CM, Carr M. 1989. Quantitative structure activity relationships for the effect of hydrophobic organic chemicals on rate of feeding by mussels (Mytilus edulis). Aquat Toxicol 14:277–294. Strasser RJ, Srivastava A, Govindjee B. 1995. Polyphasic chlorophyll a fluorescence transient in plants and cyanobacteri. Photochem Photobiol 61:32–34. Hermans C, Smeyers M, Rodriguez RM, Eyletters M, Strasser RJ, Delhaye JP. 2003. Quality assessment of urban trees: A comparative study of physiological characterization, airborne imaging, and on-site fluorescence monitoring by the OJIP-test. J Plant Physiol 160:81–90. Canevari GP. 1971. Oil slick dispersants and methods. Patent 3,793,218. U.S. Patent Office, Washington, DC. As cited in Scelfo GM, Tjeedema RS. 1991. A simple method for determination of Corexit 9527t in natural waters. Mar Environ Res, 31:69–78. Ayles Fernie International. 2003. Superdispersant-25 Safety Data Sheet. Kent, UK. Wells PG. 1984. The toxicity of oil spill dispersants to marine organisms: A current perspective. In Allen TE, ed, Oil Spill Chemical Dispersants: Research, Experience, and Recommendations. STP 840. American Society for Testing and Materials, Philadelphia, PA, pp 177–202. Borseth JF, Aunaas T, Denstad JP, Nordtug T, Olsen AJ, Schmid R, Skjaervo G, Zachariassen KE. 1995. Transmembrane sodium energy gradient and calcium content in the adductor muscle of Mytilus edulis L. in relation to the toxicity of oil and organic chemicals. Aquat Toxicol 31:263–276. Environ. Toxicol. Chem. 24, 2005 1227 29. Gulec I, Leonard B, Holdway DA. 1994. Oil and dispersed oil toxicity to amphipods and snails. Spill Science & Technology Bulletin 4:1–6. 30. Foy MG. 1982. Acute lethal toxicity of Prudhoe Bay crude oil and Corexit 9527 to Arctic marine fish and invertebrates. Technology Development Report EPS 4-EC-82-3. Environment Canada, Ottawa, ON. 31. Ordsie CJ, Garofalo GC. 1981. Lethal and sublethal effects of short-term acute doses of Kuwait crude oil and a dispersant Corexit 9527 on bay scallops, Argopecten irradians, and two predators at different temperatures. Mar Environ Res 5:195–210. 32. Baca BJ, Getter CD. 1984. The toxicity of oil and chemically dispersed oil to the seagrass Thalassia testudinum. In Allen TE, ed, Oil Spill Chemical Dispersants: Research, Experience, and Recommendations. STP 840. American Society for Testing and Materials, Philadelphia, PA, pp 314–323. 33. Singer MM, Smalheer DL, Tjeerdema RS, Martin M. 1990. Toxicity of an oil dispersant to the early life stages of four Californian marine species. Environ Toxicol Chem 9:1389–1397. 34. Singer MM, Smalheer DL, Tjeerdema RS, Martin M. 1991. Effects of spiked exposure to an oil dispersant on the early life stages of four marine species. Environ Toxicol Chem 10:1367– 1374. 35. Kravitz MJ, Lamberson JO, Ferraro SP, Swartz RC, Boese BL, Specht DT. 1999. Avoidance response of the estuarine amphipod Eohaustorius estuarius to polycyclic aromatic hydrocarbon–contaminated, field-collected sediments. Environ Toxicol Chem 18: 1232–1235. 36. Morales F, Belkhodja R, Abadia A, Abadia J. 2000. Photosystem II efficiency and mechanisms of energy dissipation in iron-deficient, field-grown pear trees (Pyrus communis L.). Photosynth Res 63:9–21. 37. Larkum AWD, Roberts G, Kuo J, Strother S. 1989. Gaseous movement in seagrasses. In Larkum AWD, McComb AJ, Shepherd SA, eds, The Biology of Seagrasses with Special References to the Australian Region. Elsevier, Amsterdam, The Netherlands, pp 686–722. 38. Bocard C, Castaing G, Gatellier C. 1984. Chemical oil dispersion in trials at sea and in laboratory tests: The key role of dilution processes. In Allen TE, ed, Oil Spill Chemical Dispersants: Research, Experience, and Recommendations. STP 840. Philadelphia, PA, pp 125–142. Departments of aBiological Sciences, cEnvironmental Sciences, and eOceanography and Coastal Sciences, and dCoastal Studies Institute, Louisiana State University, Baton Rouge, LA 70803; bDepartment of Chemistry and Biochemistry, Texas State University, San Marcos, TX 78666; and fDepartment of Biological Sciences, Clemson University, Clemson, SC 29634 Edited by Paul G. Falkowski, Rutgers, The State University of New Jersey, New Brunswick, NJ, and approved September 1, 2011 (received for review June 13, 2011) The biological consequences of the Deepwater Horizon oil spill are unknown, especially for resident organisms. Here, we report results from a field study tracking the effects of contaminating oil across space and time in resident killifish during the first 4 mo of the spill event. Remote sensing and analytical chemistry identified exposures, which were linked to effects in fish characterized by genome expression and associated gill immunohistochemistry, despite very low concentrations of hydrocarbons remaining in water and tissues. Divergence in genome expression coincides with contaminating oil and is consistent with genome responses that are predictive of exposure to hydrocarbon-like chemicals and indicative of physiological and reproductive impairment. Oil-contaminated waters are also associated with aberrant protein expression in gill tissues of larval and adult fish. These data suggest that heavily weathered crude oil from the spill imparts significant biological impacts in sensitive Louisiana marshes, some of which remain for over 2 mo following initial exposures. ecological genomics toxicogenomics | ecotoxicology | microarray | RNA-seq | F ollowing the Deepwater Horizon (DWH) drilling disaster on April 20, 2011, in the Gulf of Mexico, acute oiling and the resulting mortality of marine wildlife were evident. In contrast, the sublethal effects, critically important for predicting longterm population-level impacts of oil pollution (1), have not been well described following the DWH disaster. Here, we report the results of a 4-mo field study monitoring the biological effects of oil exposure on fish resident in Gulf of Mexico coastal marsh habitats. Gulf killifish (Fundulus grandis) were used as our model species because they are among the most abundant vertebrate animals in Gulf of Mexico-exposed marshes (2–4). Furthermore, the Atlantic-distributed sister species to F. grandis (Fundulus heteroclitus) has a narrow home range and high site fidelity, especially during the summer (5, 6), and, among fishes, it is relatively sensitive to the toxic effects of organic pollutants (7). Although home range and toxicology studies are lacking for F. grandis, we infer that F. grandis is also relatively sensitive to pollutants and exhibits high site fidelity, such that the biology of this species is likely affected primarily by the local environment, given the recent shared ancestry of F. grandis with F. heteroclitus (8) and similar physiology, life history, and habitat (9–13). We sampled from populations resident in Gulf of Mexico-exposed marshes before oil landfall (May 1–9, 2010), during the peak of oil landfall (June 28–30, 2010), and after much of the surface oil was no longer apparent 2 mo later (August 30–September 1, 2010) at six field sites from Barataria Bay, Louisiana, east to Mobile Bay, Alabama (Fig. 1 and Dataset S1). Results and Discussion Remote sensing and analytical chemistry were used to characterize exposure to DWH oil, where remote sensing data are www.pnas.org/cgi/doi/10.1073/pnas.1109545108 spatially and temporally comprehensive but of low resolution and chemistry data are of high resolution but patchy in space and time. Ocean surface oil was remotely detected through the analysis of images from synthetic aperture radar (SAR) (14). Proximity of the nearest oil slick to each field site (e.g., Fig. S1) was measured for each day that SAR data were available, from May 11 through August 13, 2010, to approximate the location, timing, and duration of coastal oiling (Fig. 1C). Although surface oil came close to many of our field sites in mid-June, only the Grande Terre (GT) site was directly oiled (Fig. 1 B and C). Although the GT site had been clearly contaminated with crude oil for several weeks before our sampling (Fig. 1C and Fig. S2) and retained much oil in sediments (Dataset S2), only trace concentrations of oil components were detected in subsurface water samples collected from the GT site on June 28, 2010, and tissues did not carry abnormally high burdens of oil constituents at any site or time point (Dataset S2). Despite a low chemical signal for oil in the water column and tissues at the time of sampling, we detected significant biological effects associated with the GT site postoil. We sampled multiple tissues from adult Gulf killifish (average weight of 3.5 g) from each of six field sites for each of three time points [only the first two time points for the Mobile Bay (MB) site] spanning the first 4 mo of the spill event (Fig. 1C). We compared biological responses across time (before, at the peak, and after oiling) and across space (oiled sites and sites not oiled) and integrated responses at the molecular level using genome expression profiling with complimentary protein expression and tissue morphology. Genome expression profiles, using microarrays and RNAseq, were characterized for livers because the organ is internal and integrates xenobiotic effects from multiple routes of entry (gill, intestine, and skin), and because liver is the primary tissue for metabolism of toxic oil constituents. Tissue morphology and expression of CYP1A protein, a common biomarker for exposure to select polycyclic aromatic hydrocarbons (PAHs), was characterized for gills, the organ that provides the greatest surface area in direct contact with the surrounding aquatic environment. In addition, we exposed developing embryos to field-collected water samples to document bioavailability and bioactivity of oil contaminants for this sensitive early life stage. Author contributions: A.W. and F.G. designed research; A.W., B.D., C.B., T.I.G., S.M., C.P., V.R., J.L.R., N.W., R.B.W. and F.G. performed research; C.D.R. contributed new reagents/ analytic tools; A.W., B.D., C.B., T.I.G., S.M., C.P., V.R., N.W., R.B.W. and F.G. analyzed data; and A.W., B.D., C.B., and F.G. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Data deposition: Microarray data have been deposited to ArrayExpress (accession no. E-MTAB-663). 1 To whom correspondence should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1109545108/-/DCSupplemental. PNAS Early Edition | 1 of 5 SPECIAL FEATURE SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE Andrew Whiteheada,1, Benjamin Dubanskya, Charlotte Bodiniera, Tzintzuni I. Garciab, Scott Milesc, Chet Pilleyd, Vandana Raghunathane, Jennifer L. Roacha, Nan Walkere, Ronald B. Walterb, Charles D. Ricef, and Fernando Galveza ENVIRONMENTAL SCIENCES Genomic and physiological footprint of the Deepwater Horizon oil spill on resident marsh fishes Fig. 1. Location of field study sites and incidence of oil contamination. (A) Location of field sampling sites, which include Grand Terre (GT), Bay St. Louis (BSL), Belle Fontaine Point (BFP), Bayou La Batre (BLB), Mobile Bay (MB), and Fort Morgan (FMA). Color coding is consistent with other figures. The red star indicates the DWH spill site. (B) Photograph (by A.W.) of the GT field site on June 28, 2010, showing contaminating oil and minnow traps in the marsh. (C) Proximity of nearest surface oil to each field site was determined by SAR, where rows are field sites and columns are days. Light gray represents no data, and black represents the nearest surface oil at a distance of >4 km; the increasing intensity of red indicates closer proximity of oil. Three field sampling trips are highlighted (blue boxes). BSL; BFP; FMA. The oiling of the GT site at the end of June 2010 is associated with a clear functional genomic footprint. Of the 3,296 genes included in our analysis, expression of 1,600 and 1,257 genes varied among field sites and throughout the time course, respectively (P < 0.01) (Dataset S3). For the 646 genes that varied in expression only among sites (no significant time effect or siteby-time interaction), site variation followed a pattern of population isolation by distance, which is consistent with neutral evolutionary divergence (Fig. 2A) and population genetic expectations (15). Most importantly, 1,500 genes indicated a pattern of site-dependent time course expression (significant interaction, false discovery rate <0.01), where the trajectory of genome expression through time was divergent at the GT site compared with all other sites (Fig. 2 B and C), particularly at the second time point, which coincides with oil contamination (Fig. 1C). Previous studies have identified genes that are transcriptionally responsive to planar polychlorinated biphenyl (PCB) exposures in killifish (16). Planar PCBs, dioxins, and PAHs (the primary toxic constituents in crude oil) are all mechanistically related insofar as they exert biological effects, in whole or in part, through aryl-hydrocarbon receptor (AHR) signaling pathways; indeed, morpholino knockdown of the AHR is protective of the toxic effects of PAHs and PCBs in killifish (17), and exposures to PCBs and PAHs induce common genome expression responses in flounder (18). Of the genes that were transcriptionally responsive to PCB exposures (16), 380 were included in the current analysis. Expression of this subset of genes is predictive of transcriptional divergence in fish from the GT site coincident with oil contamination compared with other field sites (Fig. S3), especially for the top 10% of PCB-responsive genes (Fig. 2D). Transcriptional activation of these planar PCB-responsive genes in developing killifish embryos is predictive of induction of developmental abnormalities, decreased hatching success, and decreased embryonic and larval survival (16, 19). This set of genes includes members of the canonical battery of genes that are transcriptionally induced by ligand-activated AHR signaling, 2 of 5 | www.pnas.org/cgi/doi/10.1073/pnas.1109545108 such as cytochrome P450s, cytochrome B5, and UDP-glucuronosyltransferase (Fig. 2F, set 1), for which increased transcription is particularly diagnostic of exposure to select hydrocarbons (20). Indeed, many genes that are transcriptionally induced or repressed by AHR activators (dioxins, PCBs, and PAHs) show induction or repression at the GT site coincident with crude oil contamination (Fig. 2F, set 1). An independent measure of genome expression, RNAseq, also indicates AHR activation in GT fish from June 28, 2010, compared with reference RNA (e.g., upregulation of cytochrome P450s, UDP-glucuronosyltransferase (UGT), and AHR itself; Fig. 2E). In parallel, up-regulation of CYP1A protein was detected in gills from GT fish sampled postoil and in early life-stage fish following controlled exposures to GT waters (Figs. 3 and 4). These data appear to be diagnostic of exposure to the toxic constituents in contaminating oil (PAHs) at a sufficient concentration and duration to induce biological responses in resident fish. Sustained activation of the CYP1A gene (Figs. 2F and 3) was predictive of persistent exposure to sublethal concentrations of crude oil components and negative population-level impacts in fish, sea otters, and harlequin ducks following the Exxon Valdez oil spill (reviewed in 1), although PAH toxicity may be mediated through AHR-independent pathways as well (21). Transcriptional responses in other sets of coexpressed genes offer insights into the potential biological consequences of contaminating oil exposure at the GT site. Several gene ontology (GO) categories were enriched in the subset of genes that showed GT-specific expression divergence coincident with siteand time-specific oil contamination (Dataset S4). GO enrichment indicates activation of the ubiquitin-proteasome system (Fig. 2F, set 2), which, among diverse functions, is important for cellular responses to stress, cell cycle regulation, regulation of DNA repair, apoptosis, and immune responses (22). The AHR protein itself plays a role as a unique ligand-dependent E3 ubiquitin ligase that targets sex steroid (estrogen and androgen) receptor proteins for proteasomal destruction, thereby impairing Whitehead et al. SPECIAL FEATURE SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE normal cellular responses to sex hormones in reproductive tissues, and this response can be activated by planar PAHs (23). Significant down-regulation of transcripts for egg envelope proteins zona pellucida (ZP3 and ZP4) and choriogenin (ChgHm and ChgH) that we detect at the GT site coincident with oil exposure (Fig. 2F, set 1) may be linked to this AHR-dependent proteolytic pathway because their transcription is estrogen-dependent (24, 25) and is down-regulated by exposure to PAHs in fish (25–27). In corroboration, RNAseq detects dramatically down-regulated ZP, ChgH, and vitellogenin transcripts in GT fish (Fig. 2E). Although the transcriptional response that we detect is in male fish, these proteins are synthesized in male livers (reviewed in 25, 27) and down-regulation is consistent with antiestrogenic effects from exposure to PAHs (28). Possible impacts on reproduction merit attention because water only from the GT site induced CYP1A protein in the gills of developing killifish (Fig. 3) at low concentrations of total aromatics and alkanes (Dataset S2) and more than 2 mo after initial oiling, indicating persistent bioavailability of PAHs. Marsh contamination with DWH oil coincided with the spawning season for many marsh animals, including killifish (29), and reproductive effects are predictive of long-term population-level impacts from oil spills (1). Controlled exposures of developing killifish to water collected from GT on June 28 and August 30, 2010, induced CYP1A protein expression in larval gills relative to fish exposed to GT water preoil and exposed to Bayou La Batre (BLB) site water that was not oiled (Fig. 3). This response is consistent with the location and timing of oil contamination, and it indicates that the remaining oil constituents dissolved at very low concentrations at GT after landfall (Dataset S2) were bioavailable and bioactive to developing fish. Although exposures to PAHs stereotypically induce cardiovascular system abnormalities in developing fish at relatively high concentrations (e.g., 21), none were observed in these animals. However, even very low-concentration exposures during development, insufficient to induce cardiovascular abnormalities Whitehead et al. in embryos, can impair cardiac performance in adulthood (30). The adult fish sampled in situ from the oil-contaminated GT site showed divergent regulation of several genes involved in blood vessel morphogenesis and heme metabolism coincident with oil contamination (Fig. 2F, set 3). Multigeneration field studies are necessary to confirm cardiovascular effects from DWH oil contamination of marshes that coincided with spawning. Fig. 3. CYP1A protein expression (dark red staining) in larval killifish gills (24 d postfertilization) exposed to waters collected from GT (oiled) and BLB (not oiled) during development. (Magnification 40×, scale bars = 10 μm.) CYP1A expression is elevated in the lamellae of larvae exposed during development to waters collected from GT postoil (trips 2 and 3) compared with background levels of CYP1A expression in larvae exposed to GT water preoil (trip 1), compared with CYP1A in fish exposed to waters collected from BLB (which was not directly oiled), and compared with CYP1A in fish reared in laboratory control water. Nuclei were stained using hematoxylin (blue). Analytical chemistry of exposure waters is reported in Dataset S2. PNAS Early Edition | 3 of 5 ENVIRONMENTAL SCIENCES Fig. 2. Genome expression between field sites and across time. Field sites include Grand Terre (GT), Bay St. Louis (BSL), Belle Fontaine Point (BFP), Bayou La Batre (BLB), Mobile Bay (MB), and Fort Morgan (FMA). GT was the only site to be directly oiled, which occurred between the first and second sampling times (Fig. 1 and Dataset S2). (A) For genes that vary only among sites (no expression change with time or interaction), pairwise site-specific transcriptome divergence along principal component (PC) 1, as a function of pairwise geographical distance, shows a pattern consistent with isolation by distance. (B) Trajectory of genome expression responses through time for each of six field sites from the preoil sample time (dot at base of arrow) through the peak-oil sample time (middle dot), to the latest postevent sample time (dot at head of arrow) following PC analysis of genes showing statistically significant main effects (site and time) and interaction terms. (C) Divergence along PC1 is isolated, where bars for each site from left to right represent sampling times from the earliest to the latest. (D) Expression divergence along PC1 for the subset of genes that is dose-responsive to PCB exposure (top 10% of PCB-responsive genes). (E) RNAseq data showing genes up- and down-regulated (x axis positive and negative, respectively) in fish from GT sample time 2 (coincident with oil) compared with reference RNA, where select genes are identified. (Inset) Genes are dramatically down-regulated at GT (detailed RNAseq data are presented in Dataset S5). (F) Expression levels for specific genes (rows) and treatments (columns), where cell color indicates up-regulation (yellow) or down-regulation (blue) scaled according to site-specific expression level at the preoil sample time, for genes with divergent expression at the GT site. Genes are grouped into functional categories, and scale bars indicate N-fold up- or down-regulation. Coastal salt marsh habitats are dynamic and stressful, where changes in environmental parameters, such as temperature, hypoxia, and salinity, can continuously challenge resident wildlife. Regulation of ion transport in fish is particularly important for facilitating homeostasis in response to the salinity fluctuations that are common in estuaries. We found altered regulation of multiple ion transport genes in fish from the GT site coincident with oil contamination (Fig. 2F, set 4). For example, Vtype proton ATPases are up-regulated and Na+,K+-ATPase subunits and tight-junction proteins are down-regulated, coincident with oiling at the GT site, in the absence of substantial changes in environmental salinity (Dataset S2). Other genes important for osmotic regulation in killifish (31) are also divergently down-regulated at the GT site, including type II iodothyronine deiodinase (DIO2), transcription factor jun-B (JUNB), and arginase 2 (ARG2). In corroboration, RNAseq data show down-regulation of DIO2, JUNB, and ARG2 in GT fish compared with reference fish (Fig. 2E). Although the physiological consequences of oil exposures are typically studied in isolation, it is reasonable to predict that exposure to oil may compromise the ability of resident organisms to adjust physiologically to natural stressors. Induction of CYP1A protein expression is a hallmark of AHR signaling pathway activation, making it a sensitive biomarker of exposure to select planar PAHs and other hydrocarbons (20). Although the liver is the key organ for CYP1A-mediated metabolism of these substrates, gill tissues represent the most proximate site of exposure to PAHs. As a result of direct contact with the environment and the nature of the gill as a transport epithelium, the gill may be a more sensitive indicator of exposure to contaminants than the liver (32). CYP1A protein was markedly elevated in GT fish postoil compared with GT fish preoil and compared with fish from other field sites that were not directly oiled (Fig. 4). CYP1A induction was localized predominantly to pillar cells of the gill lamellae and within undifferentiated cells underlying the interlamellar region, which may have contributed to the filamental and lamellar hyperplasia observed during trips 2 and 3, as well as the gross proliferation of the interlamellar region observed during trip 2 in GT fish (Fig. 4). These effects imply a decrease in the effective surface area of the gill, a tissue that supports critical physiological functions, such as ion homeostasis, respiratory gas exchange, systemic acidbase regulation, and nitrogenous waste excretion (33). Currently, the degree to which oil-induced effects may interact with commonly encountered challenges, such as fluctuations in hypoxia and salinity, to compromise physiological resilience is unclear. By integrating remote sensing and in situ chemical measures of exposure, and linking these with integrated measures of biological effect (genome expression and tissue morphology), we provide evidence that links biological impacts with exposure to contaminating oil from the DWH spill within coastal marsh habitats. Although body burdens of toxins are not high, consistent with reports indicating that seafood from the Gulf of Mexico is safe for consumption (34), this does not mean that negative biological impacts are absent. Our data reveal biologically relevant sublethal exposures causing alterations in genome expression and tissue morphology suggestive of physiological impairment persisting for over 2 mo after initial exposures. Sublethal effects were predictive of deleterious population-level impacts that persisted over long periods of time in aquatic species following the Exxon Valdez spill (1) and must be a focus of long-term research in the Gulf of Mexico, especially because high concentrations of hydrocarbons in sediments (Dataset S2) may provide a persistent source of exposures to organisms resident in Louisiana marshes. Methods The locations (latitude and longitude) of our field sampling sites and dates for sampling at each site are summarized in Dataset S1. Gulf killifish (F. grandis) 4 of 5 | www.pnas.org/cgi/doi/10.1073/pnas.1109545108 Fig. 4. CYP1A protein expression in adult killifish gills (dark red staining) sampled in situ from all sampling times (columns) and locations (rows). Locations include Grand Terre (GT), Bay St. Louis (BSL), Belle Fontaine Point (BFP), Bayou La Batre (BLB), Mobile Bay (MB), and Fort Morgan (FMA). (Magnification 40×, scale bars = 10 μm.) The MB site was only sampled on trips 1 and 2, and gills from trip 1 at the BLB site were not available for processing. Fish gills from the GT site during trips 2 and 3 showed high CYP1A expression and an elevated incidence of hyperplasia of the lamellae and interlamellar space on the gill filaments coincident with oil contamination. CYP1A protein was elevated at the GT site postoil (trips 2 and 3) compared with GT preoil (trip 1) as well as with other field sites, none of which were directly oiled. Nuclei were stained using hematoxylin (blue). Exact site locations and sampling dates are reported in Dataset S1. were caught by minnow trap, and tissues were excised immediately. Liver was preserved in RNAlater (Ambion, Inc.) for genome expression (microarray and RNAseq) analysis. Gill tissues were fixed in situ in buffered zinc-based formalin Z-Fix (Anatech LTD). Succinct methods follow, and more detailed methods are available online. Satellite imagery (SAR) was analyzed to provide estimation of the timing, location, and duration of coastal oil contamination. The calculated distance from each field sampling site to the nearest oil slick was calculated from the “straight-line” distance from the global positioning system position of the station (Dataset S1) to that of the observed oil across any and all intervening geographical barriers (e.g., Fig. S1). Whitehead et al. 21. Incardona JP, et al. (2005) Aryl hydrocarbon receptor-independent toxicity of weathered crude oil during fish development. Environ Health Perspect 113:1755–1762. 22. Glickman MH, Ciechanover A (2002) The ubiquitin-proteasome proteolytic pathway: Destruction for the sake of construction. Physiol Rev 82:373–428. 23. Ohtake F, et al. (2007) Dioxin receptor is a ligand-dependent E3 ubiquitin ligase. Nature 446:562–566. 24. Modig C, et al. (2006) Molecular characterization and expression pattern of zona pellucida proteins in gilthead seabream (Sparus aurata). Biol Reprod 75:717–725. 25. Yu RMK, Wong MML, Kong RYC, Wu RSS, Cheng SH (2006) Induction of hepatic choriogenin mRNA expression in male marine medaka: A highly sensitive biomarker for environmental estrogens. Aquat Toxicol 77:348–358. 26. Holth TF, et al. (2008) Differential gene expression and biomarkers in zebrafish (Danio rerio) following exposure to produced water components. Aquat Toxicol 90:277–291. 27. Sanchez BC, Carter B, Hammers HR, Sepúlveda MS (2011) Transcriptional response of hepatic largemouth bass (Micropterus salmoides) mRNA upon exposure to environmental contaminants. J Appl Toxicol 31:108–116. 28. Thomas P (1990) Teleost model for studying the effects of chemicals on female reproductive endocrine function. J Exp Zool Suppl 4(Suppl 4):126–128. 29. Greeley MS, Macgregor R (1983) Annual and semilunar reproductive-cycles of the Gulf killifish, Fundulus grandis, on the Alabama Gulf Coast. Copeia (3):711–718. 30. Hicken CE, et al. (2011) Sublethal exposure to crude oil during embryonic development alters cardiac morphology and reduces aerobic capacity in adult fish. Proc Natl Acad Sci USA 108:7086–7090. 31. Whitehead A, Roach JL, Zhang S, Galvez F (2011) Genomic mechanisms of evolved physiological plasticity in killifish distributed along an environmental salinity gradient. Proc Natl Acad Sci USA 108:6193–6198. 32. Levine SL, Oris JT (1999) CYP1A expression in liver and gill of rainbow trout following waterborne exposure: Implications for biomarker determination. Aquat Toxicol 46:279–287. 33. Evans DH, Piermarini PM, Choe KP (2005) The multifunctional fish gill: Dominant site of gas exchange, osmoregulation, acid-base regulation, and excretion of nitrogenous waste. Physiol Rev 85:97–177. 34. State of Louisiana Department of Health and Hospitals (2011) Louisiana Seafood Safety Surveillance Report (Louisiana Department of Health and Hospitals, Baton Rouge, LA). 35. Storey JD, Tibshirani R (2003) Statistical significance for genomewide studies. Proc Natl Acad Sci USA 100:9440–9445. 36. Saeed AI, et al. (2006) TM4 microarray software suite. Methods Enzymol 411:134–193. 37. Huang W, Sherman BT, Lempicki RA (2009) Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc 4:44–57. 38. Langmead B, Trapnell C, Pop M, Salzberg SL (2009) Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10:R25. 39. Anders S, Huber W (2010) Differential expression analysis for sequence count data. Genome Biol 11:R106. 40. Rice CD, Schlenk D, Ainsworth J, Goksoyr A (1998) Cross-reactivity of monoclonal antibodies against peptide 277-294 of rainbow trout CYP1A1 with hepatic CYP1A among fish. Mar Environ Res 46:87–91. Whitehead et al. PNAS Early Edition | 5 of 5 SPECIAL FEATURE 1. Peterson CH, et al. (2003) Long-term ecosystem response to the Exxon Valdez oil spill. Science 302:2082–2086. 2. Rozas LP, Reed DJ (1993) Nekton use of marsh-durface habitats in Louisiana (USA) deltaic salt marshes undergoing submergence. Mar Ecol Prog Ser 96:147–157. 3. Rozas LP, Zimmerman RJ (2000) Small-scale patterns of nekton use among marsh and adjacent shallow nonvegetated areas of the Galveston Bay Estuary, Texas (USA). Mar Ecol Prog Ser 193:217–239. 4. Subrahmanyam CB, Coultas CL (1980) Studies on the animal communities in 2 North Florida salt marshes. 3. Seasonal fluctuations of fish and macroinvertebrates. Bull Mar Sci 30:790–818. 5. Lotrich VA (1975) Summer home range and movements of Fundulus heteroclitus (Pisces: Cyprinodontidae) in a tidal creek. Ecology 56:191–198. 6. Teo SLH, Able KW (2003) Habitat use and movement of the mummichog (Fundulus heteroclitus) in a restored salt marsh. Estuaries 26:720–730. 7. Van Veld PA, Nacci DE (2008) Toxicity resistance. The Toxicology of Fishes, eds Di Giulio RT, Hinton DE (Taylor and Francis, Boca Raton, FL), pp 597–641. 8. Whitehead A (2010) The evolutionary radiation of diverse osmotolerant physiologies in killifish (Fundulus sp.). Evolution 64:2070–2085. 9. Able KW, Hata D (1984) Reproductive behavior in the Fundulus heteroclitus-F. grandis complex. Copeia (4):820–825. 10. Kneib RT (1997) The role of tidal marshes in the ecology of estuarine nekton. Oceanography and Marine Biology: An Annual Review 35:163–220. 11. Nordlie FG (2006) Physicochemical environments and tolerances of cyprinodontoid fishes found in estuaries and salt marshes of eastern North America. Reviews in Fish Biology and Fisheries 16:51–106. 12. Rozas LP, Lasalle MW (1990) A comparison of the diets of Gulf killifish, Fundulus grandis Baird and Girard, entering and leaving a Mississippi brackish marsh. Estuaries 13:332–336. 13. Weisberg SB, Lotrich VA (1982) The importance of an infrequently flooded intertidal marsh surface as an energy source for the mummichog Fundulus heteroclitus: An experimental approach. Mar Biol 66:307–310. 14. Brekke C, Solberg AHS (2005) Oil spill detection by satellite remote sensing. Remote Sensing of Environment 95:1–13. 15. Williams DA, Brown SD, Crawford DL (2008) Contemporary and historical influences on the genetic structure of the estuarine-dependent Gulf killifish Fundulus grandis. Mar Ecol Prog Ser 373:111–121. 16. Whitehead A, Pilcher W, Champlin D, Nacci D (2011) Common mechanism underlies repeated evolution of extreme pollution tolerance. Proc R Soc B, 10.1098/rspb.2011.0847. 17. Clark BW, Matson CW, Jung D, Di Giulio RT (2010) AHR2 mediates cardiac teratogenesis of polycyclic aromatic hydrocarbons and PCB-126 in Atlantic killifish (Fundulus heteroclitus). Aquat Toxicol 99:232–240. 18. Williams TD, et al. (2008) Transcriptomic responses of European flounder (Platichthys flesus) to model toxicants. Aquat Toxicol 90:83–91. 19. Whitehead A, Triant DA, Champlin D, Nacci D (2010) Comparative transcriptomics implicates mechanisms of evolved pollution tolerance in a killifish population. Mol Ecol 19:5186–5203. 20. Varanasi U (1989) Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment (CRC, Boca Raton, FL), p 341. ACKNOWLEDGMENTS. K. Carman helped facilitate early field studies. The authors thank R. Brennan, D. Roberts, E. McCulloch, Y. Meng, A. Rivera, C. Elkins, H. Graber, R. Turner, D. Crawford, and M. Oleksiak, for technical assistance. Funding was from the National Science Foundation (Grants DEB1048206 and DEB-1120512 to A.W., Grant EF-0723771 to A.W. and F.G., and Grant DEB-1048241 to R.B.W.), the National Institutes of Health (R15ES016905-01 to C.D.R.), and the Gulf of Mexico Research Initiative (A.W., F.G., and N.W.). SCIENCE APPLICATIONS IN THE DEEPWATER HORIZON OIL SPILL SPECIAL FEATURE each differentially expressed target using a negative binomial method with P values adjusted by the Benjamini–Hochberg procedure. Gill tissues were sampled from all field sites for morphological analysis and immunohistochemical analysis of CYP1A protein expression. Gill tissues from the first and second gill arches were sectioned along the longitudinal axis at a thickness of 4 μm and probed with mAb C10-7 against fish CYP1A (40). Sections were counterprobed using the Vectastain ABC immunoperoxidase system (Vector Laboratories), utilizing the ImmPACT Nova RED peroxidase substrate kit (Vector Laboratories) to visualize the CYP1A protein in red. Tissue sections were counterstained with Vector Hematoxylin QS (Vector Laboratories). F. grandis embryos obtained from parents not exposed to oil (collected from Cocodrie, LA) were exposed to water samples from the GT and BLB sites collected subsurface on the dates indicated in Dataset S1. Following fertilization, 20 embryos were randomly transferred in triplicate to one of the six field-collected waters (2 field sites × 3 time points) at 3 h postfertilization. Embryos were also exposed to a laboratory control consisting of artificial 17 parts per thousand (ppt) water. Larvae were sampled at 24 d postfertilization and fixed in Z-Fix solution. Sectioning and staining were as described in the previous section. ENVIRONMENTAL SCIENCES Analytical chemistry of water, tissue, and sediment samples was performed to offer detailed characterization of exposure to contaminating oil (data reported in Dataset S2). Sample dates and locations are summarized in Dataset S1. All sample extracts were analyzed using GC interfaced to an MS detector system. Spectral data were processed by Chemstation Software (Agilent Technologies), and analyte concentrations were calculated based on the internal standard method. Genome expression across sites and time was characterized using custom oligonucleotide microarrays. Genome expression was measured in liver tissues from five replicate individual male fish per site-time treatment (5 biological replicates) hybridized in a loop design, including a dye swap. Data were lowess-normalized and then mixed model-normalized using linear mixed models to account for fixed (dye) effects and random (array) effects. Normalized data were then analyzed using mixed model ANOVA, with “site” [Grand Terre (GT), Bay St. Louis (BSL), Belle Fontaine Point (BFP), Bayou La Batre (BLB), Mobile Bay (MB), and Fort Morgan (FMA)] and “sampling time” (sampling trips 1, 2, and 3) (Dataset S1) as main effects, including an interaction (site-by-time) term. The false discovery rate was estimated using Q-value (35). Principal components analysis was performed using MeV (36). GO enrichment was tested using DAVID (37). For RNAseq, transcript abundance was compared between liver mRNA from three replicate fish (RNA was not pooled) from the GT site from June 28, 2010, and mRNA from two control samples. All RNA samples were sequenced on the Illumina Gene Analyzer platform (Expression Analysis, Inc.). Following quality control filtering, quantitative transcript abundance analysis was performed by mapping sequence reads to target sequences (6,810 unique F. heteroclitus target EST sequences, Dataset S5) using the Bowtie short read alignment software (38). A custom Perl script determined the number of fragments mapped to each target sequence. The Bioconductor package DESeq (version 2.8) (39) was used to determine the statistical significance of BP Sees a Return to Grandeur as Gulf Fishermen Reel From Disaster | Bridge The Gulf Project 4/30/12 10:38 PM BP Sees a Return to Grandeur as Gulf Fishermen Reel From Disaster Fri, 2012-04-27 13:22 The second memorial of the nation’s worst oil catastrophe has come and gone, forever linked to Earth Day and seared into the psyches of millions of Gulf residents and fishermen. In recent weeks, the media has unleashed a torrent of stories about the devastating impacts (http://www.southernstudies.org/2012/04/troubled-waters-gulf-communities-still-reeling-two-years-into-bp-disaster.html) of the nation’s worst oil spill disaster; deaths, disease and deformities in the fisheries (http://www.tampabay.com/news/environment/water/oil-from-deepwater-horizon-spill-stillcausing-damage-in-gulf-2-years/1225134) ; a two-year record-setting die off (http://www.pbs.org/newshour/rundown/2012/04/baby-dolphin-die-offscontinue-in-the-gulf.html ) in dolphin populations; medical emergencies and family health crises in coastal communities (http://www.huffingtonpost.com/2012/04/20/gulf-oil-spill-anniversary-children_n_1438959.html) ; and ongoing Congressional wrangling over tens of millions of dollars in fines needed to save and rebuild (http://www.nola.com/opinions/index.ssf/2012/03/resolve_to_pass_the_restore_ac.html) the rapidly disappearing Gulf coast. But it won’t be long before these stories fade from the consciousness of a nation once riveted to the volcanic well spewing out Louisiana crude a mile below the sea. Instead we will see more stories like this one BP published in the Alabama Press-Register last week: After Two Years, The Grandeur of the Gulf is Returning. (http://blog.al.com/press-registercommentary/2012/04/bp_exec_after_two_years_grande.html) "These days, we don’t see oily sheens and miles of orange containment boom; we see sparkling water and clean sand, dotted with deck chairs and beach towels. On the horizon, we don’t see an armada of ships skimming oil; we see fishing vessels at work gathering the day’s catch. And, in the skies and on the ground, we don’t see planes and large cleanup crews; we see birds and other wildlife at play. "But one thing is clear: Many of the dire predictions for the Gulf, made in the days and weeks after the accident, have not turned out to be true. Indeed, after two years of hard work alongside local, state and federal officials, the scientific community and the people of the region, substantial progress has been made. And the grandeur of the Gulf is steadily returning." You can expect the media and the airwaves to be clogged with happy talk about the Gulf in the months ahead. We all wish it were true, but the facts—and perceptions of those toiling in the fisheries—just don't support it. After reading BP’s latest polemic, veteran Alaska marine toxicologist and author Riki Ott (http://www.rikiott.com/) remembered Exxon's tactics after the Valdez disaster in an email this week; "Reminds me of when the state of Alaska officials under gag order NOT to speak about impacts complained about Exxon's conclusion that just b/c wildlife are "cavorting" on beaches, everything is "hunky-dory" in Prince William Sound. Exxon flew in 3 or 4 British "scientists" (biostitutes) to fly-over Prince William Sound and visit beaches in spring 1990. Exxon sponsored them to go to our state capital and tell all 60 legislators about the "remarkable recovery of Prince William Sound.'" BP is sponsoring a $500 million scientific initiative (http://www.gulfresearchinitiative.org/) to investigate the impacts of the Gulf oil catastrophe, but it's way too early to be waving the victory flag. Fishermen in the Gulf of Mexico still are reeling after the Deepwater Horizon blew up, and "grandeur" isn't exactly the term they're using down there. Here's how Louisiana's Plaquemines Parish charter fisherman Ryan Lambert described it in the Times Picayunne (http://www.nola.com/news/gulf-oilspill/index.ssf/2012/04/fishing_guides_say_their_busin.html) this week: "'The oil may have stopped flowing, but the spill still goes on down here every day,' said Lambert, owner and operator of Cajun Fishing Adventures, a sprawling lodge and charter business in Buras. 'My fishing business is still down 50 to 60 percent, we're still finding oil and tar balls on the beach and in the marsh, people still think the fish are polluted (http://www.nola.com/news/gulf-oilspill/index.ssf/2012/04/2_years_after_gulf_oil_spill_l_1.html) , and now we can't find speckled trout in nearly the numbers we had before the spill. So don't tell me the disaster is over. Maybe for BP it is. Maybe for the oil business people it is. But for me and other charter businesses, it's never stopped.'" Shrimp with no eyes and tumors caught in Barataria Bay, April, 2012. http://bridgethegulfproject.org/node/633 Photo by Xuan Chen Page 1 of 4 BP Sees a Return to Grandeur as Gulf Fishermen Reel From Disaster | Bridge The Gulf Project 4/30/12 10:38 PM So which is it, gasping for air or returning to grandeur? My bet is most people outside the Gulf will believe the latter as BP continues its PR assault on the national airwaves. Although thousands of fishermen are struggling, their stories are easily overpowered by the massive economic and political forces aligned against them. This Sen. David Vitter (R-LA) missive to his constituents pretty much sums up where the Powers That Be stand in the Gulf: "The good news is that I don’t think anyone would have predicted that the Gulf would have rebounded to where it is today. That goes for our tourism industry, which is thriving, and of course our Gulf seafood, which is as safe and delicious as ever…” Yes, that's the same seafood that NRDC's Miriam Rotkin-Ellman and Gina Solomon reported on in Environmental Health Perspectives (http://ehp03.niehs.nih.gov/article/info%3Adoi%2F10.1289%2Fehp.1104539R) last year; the very same seafood that can expose vulnerable populations like children and pregnant women to up to 10,000 times the allowable levels of cancer-causing polycyclic aromatic hydrocarbons (PAHs). You can read Miriam's excellent Gulf health update here (http://switchboard.nrdc.org/blogs/mrotkinellman/bp_oil_disaster_two_years_late.html) . Of course, it's no surprise that Sen. Vitter ignores this. The oil and gas industry ranks numero uno on his list of top campaign contributors (http://www.opensecrets.org/politicians/summary.php?cid=n00009659#ind) . Most of his colleagues down there float in the same boat. But that's not the boat Gulf fishermen are working in, the ones pulling up tumor and oil-encrusted shrimp (http://www.aljazeera.com/indepth/features/2012/04/201241682318260912.html) . It's not the reality residents who still witness record numbers of dolphins (http://www.pbs.org/newshour/rundown/2012/04/baby-dolphin-die-offs-continue-in-the-gulf.html) washing ashore on beaches littered with tar balls after every storm. As the carnage continues, BP and other oil industry behemoths are busier than ever drilling in the Gulf, dragging their mammoth platforms into deeper water where they can jam cement-reinforced pipelines into even more remote high-pressure subsea oil and gas deposits. Oil slicks are routinely reported (http://onwingsofcare.org/protection-a-preservation/gulf-of-mexico-oil-spill-2010/gulf-2012/244-shell-macondomars-ursa-green-canyon-gulf-oil-wings-care.html) in drilling areas offshore. Can the disaster happen again? You bet. Lured by the high prices and increased global demand, more and more rigs are at work, while Congress still has passed no new drilling safety laws. NRDC's David Pettit reports (http://switchboard.nrdc.org/blogs/dpettit/what_if_another_bp_deepwater_h.html) that while some new federal drilling regulations are in place, we may be setting ourselves up for a "repeat performance." For instance, in an Orlando Sentinel article, (http://articles.orlandosentinel.com/2012-04-17/news/os-bp-gulf-spill-anniversary-20120414_1_deepwater-horizon-rig-gulf-ofmexico-oil-bob-dudley) Pettit questioned why the industry doesn't redesign the critical blowout preventers that failed in the first place. "You can have 10 of them, and if they are all subject to being jammed by a pipe doing something that we know actually happened [once before], then you're not much safer." Requiring twice as many safety mechanisms that failed once before doesn’t exactly inspire confidence. But that’s life in our oil-addicted, rush-to-drill world. Our country’s energy policy is stuck on dig, extricate and burn. The petrochemical industry will become more dangerous as we rush to extract harder to find deposits of crude in places like the harsh environs of the Arctic, where Shell Oil plans (http://switchboard.nrdc.org/blogs/rkistner/arctic_oil_drilling_threatens.html) to start drilling this summer. Have we learned nothing from the historic BP blowout? But there is some good news; disasters can change attitudes, even in the oil-dominated Gulf of Mexico. The BP debacle blew a hole through Gulf residents' confidence in the oil industry and changed perspectives on the importance of environmental protections, according to a recent University of New Hampshire study reported in Science Daily (http://www.sciencedaily.com/releases/2012/04/120412105227.htm) : "If disasters teach any lessons, then experience with the Gulf oil spill might be expected to alter opinions about the need for environmental protection. About one-fourth of our respondents said that as a result of the spill, their views on other environmental issues such as global warming or protecting wildlife had changed," said Lawrence Hamilton, professor of sociology at the University of New Hampshire. This proportion rose to 35 percent among those most affected economically by the spill. People reporting changed views also expressed greater concern about sea level rise due to climate change, more support for a moratorium on deepwater drilling, and were more likely to favor alternative energy rather than increased oil exploration." Shrimp with tumors bought in New Orleans grocery store. Photo by Mac MacKenzie. Changes in attitudes can lead to positive change in the oil patch. But there's also another way to get Big Oil to change its behavior: jail http://bridgethegulfproject.org/node/633 Page 2 of 4 BP Sees a Return to Grandeur as Gulf Fishermen Reel From Disaster | Bridge The Gulf Project 4/30/12 10:38 PM time. Oil industry experts say fines and penalties do little to reduce reckless decisions and risky policies that threaten the lives and livelihoods of fishermen and workers. Pro Publica reporter Abrahm Lustgarten (http://www.propublica.org/site/author/Abrahm_Lustgarten/) laid it all out succinctly in this New York Times op-ed (http://www.nytimes.com/2012/04/20/opinion/a-stain-that-wont-wash-away.html? _r=1&partner=rss&emc=rss) : "What the gulf spill has taught us is that no matter how bad the disaster (and the environmental impact), the potential consequences have never been large enough to dissuade BP from placing profits ahead of prudence. That might change if a real person was forced to take responsibility — or if the government brought down one of the biggest hammers (http://www.propublica.org/article/epa-officials-weighing-sanctions-against-bps-us-operations) [4] in its arsenal and banned the company from future federal oil leases and permits altogether. Fines just don't matter." Although low-level prosecutions (http://www.usatoday.com/money/industries/energy/story/2012-04-24/bp-oil-spill-arrest-justice-department/54504158/1) have just been announced, nobody expects many BP execs to get rolled up by the Department of Justice anytime soon. But this catastrophe also wasn’t an ordinary industrial accident. It coated the coasts of four states with oil and has damaged the lives of countless residents. Some fishermen in the Gulf say they still relish the thought of former BP CEO Tony Hayward (http://www.usatoday.com/money/industries/energy/story/2012-04-24/bp-oil-spill-arrest-justice-department/54504158/1) caught in the wheels of American justice, a fanciful dream as they struggle to rebuild their lives. While BP fights a long war in the courts, the battle over public opinion remains hot and intense. Despite the million dollar ad campaigns and the political rhetoric about the Gulf returning to normal, many—especially in the fishing community—are facing a new reality. Their lives on the water have changed in ways no one comprehends, in ways many fear will never be the same. The grandeur of the Gulf will return. But it’s up to people along its shores—and to all of us—to decide what kind of future lies in store; a marine environment stressed by an oil-soaked sea or a healthy ocean preserved for distant generations. The critical decisions are still to come. Rocky Kistner (javascript:void(0);) Comments NBA R eports (/node/633#comment-10132) Submitted by Anonymous on Sun, 2012-04-29 02:18. CHICAGO (AP) -- Bulls legend Derrick Went up by will Michael Vick Nike Jersey (http://www.nfleaglesproshop.com/nike-michael-vick-jersey/) certainly pass up the rest of the time caused by a divided anterior cruciate plantar fascia throughout the quit leg. Went up by ended up being made it easier for off of the the courtroom delayed throughout Chicago's 103-91 playoff-opening win in the Philadelphia 76ers in Wednesday, an astounding hit for the crew eyeing a new champion manage. They won 3 things along with ended up being participating in more like your league's reigning MVP immediately after lost 35 online games as a consequence of incidents in the standard time, nevertheless the time located a stop because Bulls ended up wrapping up an amazing get. Went up by crumbled on the terrain immediately after they driven your isle using with regards to 1: 20 quit plus the Bulls primary by simply 12. They ended up being getting a layup while they located a new jump-stop along with seemed to adjust the head because 76ers' Spencer Hawes along with Lavoy Allen revolving around, moving Nike Michael Vick Jersey (http://www.nfleaglesproshop.com/nike-michael-vick-jersey/) past off of to your teammate ahead of the cumbersome clinching. Crew health care workers quickly in a rush out and about along with offered help for you to Went up by for several minutes while they ended up being writhing throughout ache at the baseline ahead of aiding your ex on the locker place. Went up by ended up being arrive at the hospital, wherever MRI benefits established your Bulls' worst type of anxieties. NEW MEXICO (AP) -- Knicks novice defend Iman Shumpert will certainly pass up the rest of the postseason immediately after bringing a new quit leg plantar fascia inside 3 rd fraction involving Brand-new York's Sport 1 decline on the New mexico High temperature in Wednesday. Your Knicks released Shumpert in addition took the side to side meniscus. Medical procedures are going to be timetabled throughout The big apple along with Shumpert can be supposed to pass up about 6-8 a few months. Shumpert, whom commenced a final 19 Rolando McClain Nike Jersey (http://www.nflraidersproshop.com/rolando-mcclain-nike-jersey/) online games in the standard time immediately after Jeremy Lin got leg medical procedures, ended up being taking your soccer ball up the appropriate sideline along with tried out for you to dribble guiding the when they ripped way up along with quickly selected in the quit leg. They ended up being produced in the Knicks' locker place pertaining to examination along with after to your infirmary on an MRI. WALTHAM, Muscle size. (AP) -- Boston ma Boston celtics defend Beam Allen remains to be unclear in case he can be capable of participate in inside playoff opener resistant to the The atlanta area Hawks in On the. Allen haven't played out throughout a fortnight caused by a painful appropriate foot. They ended up being taking pictures all-around ahead of train while using crew in Wednesday along with explained although experience many of the "smaller parts" in the training. Nevertheless instructor Doc Waters explained Allen may not train along with they don't feel Nike Rolando McClain Jersey (http://www.nflraidersproshop.com/rolando-mcclain-nike-jersey/) although participate in throughout Sport 1, sometimes. Your 36-year-old taking pictures defend explained they gotten a new cortisone picture in Wed. In case the idea just weren't to the playoffs, they explained, he had by now always be obtaining medical procedures. All the others for the Boston celtics described throughout balanced to the closing train prior to crew foliage pertaining to The atlanta area. THE BIG APPLE -Your In Pacers' Honest Vogel plus the Memphis Grizzlies' Lionel Hollins right now ended up referred to as your NBA Western along with Developed Seminar Motor coach buses in the Thirty day period, respectively, pertaining to online games played out throughout The spring. Vogel brought about In with an Western Conference- ideal 12-3 (#(. 900) file which include a new league-leading 5-0 level while travelling, your Pacers top earning talent out of the house considering that 2003-04. Your Pacers' 12 victories ended up your team's nearly all is the winner in a very thirty day Darren McFadden Nike Jersey (http://www.nflraidersproshop.com/darren-mcfadden-nike-jersey/) period considering that Late 2003 (13-2). Through The spring, your Pacers averaged 102. 7 things -- virtually six to eight things earlier mentioned their time regular -- along with graded 6th all round throughout Plus-Minus Credit score Differential (+5. 7). Hollins carefully guided your Grizzlies to your 13-3 file throughout The spring, tying your franchise's level to the nearly all is the winner in a very thirty day period (Drive, 2004). One of several Grizzlies' 13 victories ended up is the winner in Oklahoma Area, which in turn clicked your Thunder's six-game earning talent, along with in New mexico. Memphis' season-ending household win versus Holiday ended up being http://bridgethegulfproject.org/node/633 Page 3 of 4 BP Sees a Return to Grandeur as Gulf Fishermen Reel From Disaster | Bridge The Gulf Project 4/30/12 10:38 PM your team's franchise-tying 11th direct get in FedExForum, along with clinched homecourt advantages inside 1st rounded in the playoffs. Various other nominees pertaining to Instructor in the Thirty day period ended up Atlanta's Lewis Drew, Boston's Doc Waters, Denver's George Karl, Brand-new York's Henry Woodson, San Antonio's Gregg Popovich, along with Utah's Tyrone Corbin. WALTHAM, Nike Darren McFadden Jersey (http://www.nflraidersproshop.com/darren-mcfadden-nike-jersey/) Muscle size. (AP) -- Paul Pierce ended up being rear in train to the Boston ma Boston celtics in Feb 5th, every day immediately after lightly making your regular-season finish which has a sprained quit major feet. Pierce explained it turned out any small trouble along with they don't count on the idea for you to hassle your ex inside first-round string resistant to the The atlanta area Hawks that will commences On the. Pierce quit Thurs night evening of sport first inside 1st fraction plus the Boston celtics explained they has not been planning to give back, nevertheless they ended up being rear for the the courtroom delayed inside subsequent fraction. Beam Allen would not train in Feb 5th. Instructor Doc Waters explained they don't recognize no matter whether Allen would be able to participate in in On the. Allen features have missed a fortnight which has a painful appropriate foot. http://bridgethegulfproject.org/node/633 Page 4 of 4 PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... RESEARCH ARTICLE Response of Coastal Fishes to the Gulf of Mexico Oil Disaster F. Joel Fodrie1*, Kenneth L. Heck Jr.2 1 Institute of Marine Sciences and Department of Marine Sciences, University of North Carolina at Chapel Hill, Morehead City, North Carolina, United States of America, 2 Dauphin Island Sea Lab and Department of Marine Sciences, University of South Alabama, Dauphin Island, Alabama, United States of America Abstract The ecosystem-level impacts of the Deepwater Horizon disaster have been largely unpredictable due to the unique setting and magnitude of this spill. We used a five-year (2006–2010) data set within the oil-affected region to explore acute consequences for early-stage survival of fish species inhabiting seagrass nursery habitat. Although many of these species spawned during spring-summer, and produced larvae vulnerable to oil-polluted water, overall and speciesby-species catch rates were high in 2010 after the spill (1,989±220 fishes km-towed−1 [µ ± 1SE]) relative to the previous four years (1,080±43 fishes km-towed−1). Also, several exploited species were characterized by notably higher juvenile catch rates during 2010 following large-scale fisheries closures in the northern Gulf, although overall statistical results for the effects of fishery closures on assemblage-wide CPUE data were ambiguous. We conclude that immediate, catastrophic losses of 2010 cohorts were largely avoided, and that no shifts in species composition occurred following the spill. The potential long-term impacts facing fishes as a result of chronic exposure and delayed, indirect effects now require attention. Citation: Fodrie FJ, Heck KL Jr (2011) Response of Coastal Fishes to the Gulf of Mexico Oil Disaster. PLoS ONE 6(7): e21609. doi:10.1371/journal.pone.0021609 Editor: Steven J. Bograd, National Oceanic and Atmospheric Administration/National Marine Fisheries Service/Southwest Fisheries Science Center, United States of America Received: March 7, 2011; Accepted: June 2, 2011; Published: July 6, 2011 Copyright: © 2011 Fodrie, Heck. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Funding: The authors acknowledge support from the National Marine Fisheries Service, National Oceanic and Atmospheric Administration Marine Fisheries Initiative and Northern Gulf Institute. The funders had no role in study design, data collection and analyses, decision to publish, or preparation of the manuscript. 1 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... Competing interests: The authors have declared that no competing interests exist. * E-mail: [email protected] I NTRODUCTION Prevailing models of ecological impacts resulting from oil pollution are being revised after the April 2010 release of ~4.4 million barrels [1] of oil into the northern Gulf of Mexico (GOM). In part, this is a legacy of the Exxon Valdez accident as a watershed environmental catastrophe, and the extensive research on acute and chronic impacts of the resulting inshore oil pollution [2]. Unlike the 0.25–0.5 million barrels released by the Valdez [2], however, the Deepwater Horizon (DH) disaster hemorrhaged oil into the open ocean at 1500 m depth over a protracted 84-day period [1]. As a critical step toward new model development applicable for detecting impacts of the DH spill, rigorous observational data at organismal through community levels are needed to guide ecosystem-based toxicology. We have already learned that a significant fraction of the oil released into the GOM from the Macondo well did not rise to the surface, and this has implications for the ecosystem-level responses we should anticipate. Rather, oil was emulsified at the well head due to turbulent mixing, reduced buoyancy at depth, and addition of Corexit 9500 dispersant. Subsequently, mid-water hydrocarbon plumes [3] have been observed with stimulation of petroleum-degrading bacteria [4]. With this now understood, we revisit some early concerns regarding impacts for nearshore fisheries [5]. During the DH spill, near-surface waters lacked any reliable refuge from oil pollution, as slicks/sheens occurred at the immediate surface and oil was emulsified throughout the water column. For many fishes, including commercially valuable snappers (Lutjanidae) and groupers (Serranidae), spawning occurs during the spring or summer (table S1), and eggs, larvae and post-larvae would have relied upon near-surface waters overlaying the continental shelf during the DH spill [6]–[7]. Furthermore, eggs/larvae and oil can be transported by the same hydrodynamic and atmospheric processes, enhancing the probability of oil encounters for many species. Because the population ecology of marine species with bipartite life histories is disproportionately affected by the health and survival of early life stages [8], understanding how eggs, larvae and newly-settled juveniles coped with the DH spill is essential for quantifying ecosystem responses. We hypothesized that the strength of juvenile cohorts spawned on the northern GOM continental shelf during May–September 2010 in the northern GOM would be negatively affected by egg/larval-oil interactions. Oiled seawater contains toxic compounds such as polycyclic aromatic hydrocarbons (PAHs) which, even after weathering, can result in genetic damage, physical deformities and altered developmental timing for fish eggs/larvae [9]–[10]. These effects may be 2 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... induced at very low (~1 ppb PAHs) levels of exposure when persistent over days to weeks [11]–[12] - timescales relevant for larval development and descriptive of the DH spill. Additionally, emulsified oil droplets could mechanically damage the feeding and breathing apparatus of relatively fragile larvae and further decrease individual fitness. Unfortunately, observing egg/larval mortality, growth or migration in situ is an enduring challenge for biological oceanographers, as eggs/larvae are simply too dilute and experience relatively high instantaneous mortality, even in undisturbed systems [13]. In the absence of direct observations on eggs and larvae, juvenile abundance data provide valuable indices of the acute, population-level responses of young fishes to the spill. Although indirect evidence [14], early juvenile abundances are the integrated products of early life-history processes such as fertilization, larval growth/mortality, and settlement [6]–[8]. Therefore, effects of oil pollution on early life stages should be detectable in time series data as shifts in the abundance of recently settled juvenile fishes. We tested these predictions using 2006–2010 survey data collected from the Chandeleur Islands, LA, to Saint Joseph Bay, FL (Fig. 1), representing most of the nearshore region directly impacted by oil. In contrast to the difficulties of surveying marine larvae, quantitative measures of juvenile abundances are tractable due to the tendency of settled fish to aggregate in specialized nursery habitats [15]. In the northern GOM, many fish species, such as those in the drum (Sciaenidae), snapper and grouper families have juveniles that are routinely collected from shallow-water seagrass meadows they use as primary nurseries [16]. Figure 1. Sampling region and study sites. Map of juvenile fish sampling stations, divided among four survey areas: Chandeleur Islands (blue circles), Gulf Islands (green circles), Grand Bay (orange circles) and Florida Bays (red circles). 1. Chandeleur Is., LA; 2. Ship Is., MS; 3. Horn Is., MS; 4. Petit Bois Is., MS; 5. Dauphin Is., AL; 6. Grand Bature Shoal, AL; 7. Point Aux Pines, AL, 8. Big Lagoon, FL; 9. Pensacola Bay, FL; 10. Choctawhatchee Bay, FL; 11. St. Andrew Sound, FL; 12. St. Joseph Bay, FL. The spread of surface oil during the 84-day spill is also shown (brown shading). Image at lower right shows juvenile gray snapper (L. griseus), spotted seatrout (C. nebulosus) and pipefish (Syngnathus spp.). doi:10.1371/journal.pone.0021609.g001 Our dataset consisted of 853 individual trawl samples taken between July 15 and October 31 of 2006–2010 within seagrass meadows of the northern GOM (tables S2, S3). We collected 167,740 individual fishes representing 86 taxa, and examined catch-per-unit-effort (CPUE) data for all species pooled together, as well as separately for each of the 20 most abundant species. We also tested for post-spill community-level shifts in seagrass-associated fish assemblages using multivariate analyses [17]. We recognized that not all species were at equal risk for oil exposure due to variation in spawning timing and larval distributions (tables S1, S4). Furthermore, some species may have experienced release from fishing pressure due to large-scale fishery closures [18] 3 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... during the spill (table S5), perhaps enhancing their larval production during the summer spawning season. Therefore, we also considered how these factors affected species-specific CPUEs during 2010. In all analyses, comparisons among years were considered as a proxy for the effects of oil disturbance (2006–2009 as undisturbed, 2010 as disturbed). R ESULTS Within the oil-affected GOM, a five-year survey of seagrass-associated fish communities did not indicate reductions in juvenile abundances following the spill. Rather, of the twenty most commonly collected fish species, twelve were characterized by statistically higher catch rates in 2010 relative to 2006–20009 (α = 0.05; Table 1). Among the remaining eight taxa, pre- and post-spill catch rates were statistically indistinguishable. Across our entire study region, CPUE increased from 1,080±43 fishes km-towed−1 (µ ± 1SE) during 2006–2009 to 1,989±220 fishes km-towed−1 in 2010. When resolved among four geographical areas (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays; Fig. 1), overall catch rates of juvenile fishes, as well as CPUE of the most abundant species, pinfish (Lagodon rhomboides), were consistently higher during 2010 than in 2008 or 2009, and in some areas were higher in 2010 than all previous years (Fig. 2A–B; fig. S1, S2, S3; table S6). Figure 2. Catch rates of juvenile fishes, 2006–2010. Catch rates among years and sampling areas (Chandeleur Islands, Gulf Islands, Grand Bay and Florida Bays) for: (A) all fishes pooled; (B) pinfish (L. rhomboides), (C) gray snapper (L. griseus), and (D) spotted seatrout (C. nebulosus). CPUE data in panels B–D are presented on a log scale, and all data are shown as means of trawl samples (µ + 1SE). doi:10.1371/journal.pone.0021609.g002 Table 1. Relative frequencies and CPUE data for abundant fishes collected during sampling in seagrass meadows of the northern GOM. doi:10.1371/journal.pone.0021609.t001 The species composition of juvenile fish assemblages was unaltered in each sampling area during the months following the DH disaster (Fig. 3). Prior to the spill, similarities among individual trawl samples (SIMPER) ranged from 50.3% at the Chandeleur Islands to 52.9% within Florida Bays (table S7). By comparison, similarity percentages between pre- (2006–2009) and post-spill (2010) samples were correspondingly high, ranging from 43.4% within Grand Bay to 50.8% at the Chandeleur Islands. Furthermore, pinfish, silver perch (Bairdiella chyrsoura), mojarras (Eucinostomus spp.), pigfish (Orthopristis chrysoptera) and spotted seatrout (Cynoscion nebulosus) drove similarity patterns both before and after the spill (table S7). Species richness (S, up 15%, p<0.001), diversity (ES(20), up 11%, p = 0.006; H′, up 18%, p<0.001) and evenness (J′, up 11%, p = 0.003) among trawl samples were all slightly elevated during 2010 relative to 2006–2009 4 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... averages (table S8; fig. S4), indicating that high CPUEs in 2010 were broad based. Figure 3. Community composition of seagrass-associated fish communities, 2006–2010. Multi-dimensional scaling plots for seagrass-associated fish communities prior to (2006–2009; colored symbols) and following (2010; black circles) the DH spill. Data for (A) Chandeleur Islands, (B) Gulf Islands, (C) Grand Bay and (D) Florida Bays are presented separately. Each datum represents a single trawl sample. doi:10.1371/journal.pone.0021609.g003 When averaged across species, there was little statistical evidence that either exposure risk or release from fishing pressure significantly affected CPUEs during 2010. When comparing 2010 CPUE data against pre-spill catch rates, we did observe that fishes characterized by moderate (spring spawning, nearshore larvae) or high risk (spring-summer spawning, larvae distributed across the continental shelf) exhibited decreases in CPUE following the spill at the Chandeleur Islands and Grand Bay (Fig. 4A). However, no statistically significant differences were found as a function of egg/larval risk (F4,848 = 1.410, p = 0.242) or sampling areas (F3,849 = 0.999, p = 0.440; table S9). Similarly, release from fishing pressure on spawning fishes could be implicated, although not proven, as a determinant of post-spill CPUEs. Along the Chandeleur and Gulf Islands, increases in catch rates during 2010 relative to 2006–2009 were 800% and 950% higher, respectively (Fig. 4B), for species identified in state and federal management plans than for species not harvested by fishermen (table S5). No similar patterns were recorded within Grand Bay or Florida Bays, however, and effects of fishing pressure (F1,851 = 1.510, p = 0.223) and area (F3,849 = 1.397, p = 0.225) on CPUE responses were not significant. Figure 4. Larval risk and fishery closure impacts. Effects of (A) egg/larval vulnerability and (B) harvest pressure on the responses of fishes to the DH spill. Response of individual species calculated as the ratio of 2010 versus 2006–2009 CPUE data. Data are presented on a log scale as group means (µ + 1SE), with ratios >1 indicating that 2010 catch rates were elevated relative to 2006–2010 data. doi:10.1371/journal.pone.0021609.g004 D ISCUSSION Collectively, no significant, acute impacts on the strength of juvenile cohorts within seagrass habitats were detected following the DH disaster. This was true for all species examined, bolstering our confidence in the conclusion that ecosystem-level injuries were not severe for this community of fishes. Unfortunately, our assessment cannot be compared to the most analogous spill, the IXTOC 1 5 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... blowout [5], due to a paucity of formal scientific investigation following that accident (The 1979 IXTOC I blowout at 3600 m depth, 80-km north of the Yucatán Peninsula, was a ~3.5-million-barrel spill.). The most parsimonious explanation for our data is that these fishes were resilient to the spill, possibly due to the retention of a large proportion of spilled oil at depth. As such, these data add to a developing literature [3]–[4] in which the acute impacts of the spill may be concentrated in the deep ocean rather than shallow-water, coastal ecosystems that were the focus of early concern [5]. For instance, gray snapper (Lutjanus griseus) larvae were abundant in surface waters (0–25-m deep) over the continental shelf from July through September [19], and were among the most likely individuals to have contacted oil-polluted water. Still, catch rates of gray snapper juveniles following the spill were high relative to the four previous years (up 82%, Fig. 2C; area * pre/post spill context interaction p<0.001, table S6). When averaged across species - and in some cases across species with vastly different life histories - there were no statistically significant differences in the response of fished or unfished species to the spill (or their responses to subsequent management actions; i.e., fishery closures). Still, there were notable patterns suggesting that certain species may have been released from harvest pressure during 2010, subsequently enhancing spawning activity and post-spill cohort sizes despite any potentially negative oil impacts. For example, spotted seatrout spawn during summer [20], but many mature individuals are typically removed by recreational fishers before reproducing. Following the fishery closures in 2010, we recorded order-of-magnitude higher juvenile abundances of spotted seatrout at the Chandeleur and Gulf Islands, as well as elevated catch rates throughout our survey region (Fig. 2D; area, pre/post spill context and 2-way interaction p<0.001, table S6). Consistent with the patterns observed in the species-by-species catch data and analyses of ‘risk” or ‘fishing” effects, there were no significant post-spill shifts in community composition and structure, nor were there changes in any of several biodiversity measures. While natural recruitment variability can make it difficult to detect population-level impacts for any one species following large-scale disturbance [14], our whole-community analyses and results are likely robust against these concerns. Several other factors could have contributed to the high catch rates of seagrass-associated fishes in 2010 despite large-scale oil pollution. For instance, fishes may be uniquely buffered against oil pollution due to their mobility or foraging ecology [21]–[22]. Also, the major predators of fish eggs/larvae (e.g., gelatinous zooplankton) may have been impacted by the spill, thereby reducing natural mortality rates for coastal fishes [23]. Regardless of the mechanism(s) involved, thus far the potential for 2010 cohorts to support regional fisheries over the next several years has persisted despite the spill. This information is critical for projecting the mode and tempo of ecological and economic recovery in the oil-affected GOM, as well as guiding future conservation/restoration activities to mitigate oil-spill injuries. While these data provide reason for early optimism, attention should now turn to possible chronic effects of oil exposure on fishes as well as delayed indirect effects cascading through the post-spill 6 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... GOM. Fish may suffer growth, survival or reproductive penalties years after exposure to oil [24], or experience altered migratory behaviors [25]. Oil sequestered in sediments may also affect species laying benthic eggs for several years [26]. More broadly, ecosystems experiencing large-scale disturbance can carry or build instabilities over protracted periods that can eventually result in delayed collapses of fisheries stocks [27]. Improved threat assessment for energy exploration [28] and process-oriented studies of ecosystem responses will be long-term initiatives resulting from the DH spill. In the short term, however, observational data collected over relevant spatial and temporal scales are invaluable for guiding and evaluating targeted studies of oil toxicology [29]. For fish species experiencing multiple stressors such as habitat degradation [30] harvest pressure [31], climate change [16], and now oil pollution, rigorous baseline survey data and new syntheses are needed to enact effective ecosystem-based management. M ATERIALS AND M ETHODS Sampling We analyzed changes in northern Gulf of Mexico (GOM) seagrass-associated fish communities during the last 5 years by comparing survey data obtained either prior to (2006–2009) or immediately following the Deepwater Horizon disaster (2010). The survey region extended approximately 340 km along the coastline, including a significant portion of the inshore area most affected by oil (Fig. 1.). Each year, we made repeated sampling trips to 12 sites, extending from the Chandeleur Islands, LA, to St. Joseph Bay, FL (29.68–30.72°N, 85.30–89.10°W). Sampling occurred within mixed seagrass meadows that serve as primary nursery habitat for many juvenile fishes that have recently settled from the water column following a 5–45 day larval period [6], [16]. Our samples were collected from seagrass mosaics that included turtle grass (Thalassia testudinum), shoal grass (Halodule wrightii), widgeon grass (Ruppia maritima), and manatee grass (Syringodium filiforme), along with scattered unvegetated patches (table S3). During each year, trawls were conducted from July 15 through October 31 in order to record the abundances and composition of fishes during the period when seagrass meadows are utilized as primary nurseries by recently settled juveniles (refer to table S1 for reproductive seasons of common fishes). Fishes were collected using a 5-m otter trawl (2.0-cm body mesh; 0.6-cm bag mesh; 0.3×0.7-m doors) with conventional 4-seam balloon design including float and lead lines but without tickler chains. Trawls consisted of 3.9±0.1 (µ ± 1SE) minute tows behind small (<7 m) research vessels traveling at 3.3+0.1 kilometers hour−1. Overall, 853 samples were taken (table S2), and the trawl covered a linear distance of 184.7 kilometers during our sampling. These trawls occurred in depths of 0.5–2.5-m, and were conducted during daylight hours. During our surveys, species were enumerated in the field unless species-level identifications were not easily made. Unidentified specimens were transported to the lab where meristics were used by at least two different technicians to identify each individual. In cases in which species could not be identified, 7 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... specimens were classified to the lowest taxonomic level possible. Typically, fishes were 20–100 mm long (standard length), indicative of recently-spawned juveniles. However, we did not record individual sizes for all species, and, for pipefishes (Syngnathus spp.) and hard-headed catfish (Arius felis), we did observe that a small proportion of our catch included reproductive adults. For four species: gray snapper (50.5±0.8 mm [µ ± SE]), lane snapper (Lutjanus synagris; 55.7±0.7 mm), spotted seatrout (60.8±1.1 mm) and gag grouper (Mycteroperca microlepis; 157.5±3.2 mm); we did record the sizes of all individuals throughout our surveys. Based on our own otolith analyses (Fodrie unpublished) and published reports of first-year growth among these four species (age-1 sizes: gray snapper ~109 mm; lane snapper ~140 mm; spotted seatrout ~127 mm; gag grouper >198 mm), we calculated that >96% of individuals were captured in the same year as they were spawned (including 2010). Once enumerated, fishes were entered in to an Excel database, and abundance data were converted into catch-per-unit-effort (CPUE) data based on the linear distance over with each trawl occurred. All statistical analyses were applied to these CPUE data. Our complete CPUE dataset is included as a separate appendix in our supporting information (dataset S1). This study was carried out in strict accordance with the recommendations in the Guide for the Care and Use of Laboratory Animals of the National Institutes of Health. Our sampling protocol was approved by the Committee on the Ethics of Animal Experiments of the University of North Carolina at Chapel Hill (Permit Number: 10-114.0). Statistical analyses We investigated differences in the catch rates of seagrass-associated fishes (all species pooled as well as the 20 most abundant species individually) by unpaired t-tests comparing pre- (2006–2010) and post-spill (2010) data (Table 1), as well as 2-way ANOVAs in which sampling area (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays) and context (pre- versus post-spill) were fixed factors (table S6). Regions were defined by basic geomorphology and location, local water clarity, local salinity, and local seagrass composition [32]. Because variances were stable among groups, no data transformations were required prior to analyses. We analyzed similarities and differences in fish communities among years (2006–2009 versus 2010) within each sampling area (each area considered separately) using non-metric multidimensional scaling (MDS), based on Bray-Curtis similarity indices among all individual trawl samples (4th root-transformed data). Pairwise comparisons between trawl samples across years were conducted with analysis of similarity (ANOSIM) and similarity (or dissimilarity) percentages (SIMPER) using PRIMER 5.2.2 software (PRIMER-E Ltd; [33]). We also examined patterns of species diversity among regions and years by computing the following measures for each trawl sample: S, number of species collected; ES(20), species richness rarefied to 20 individuals; H′, Shannon-Weiner diversity index (loge); and J′, Pielou's evenness measure (PRIMER 5.2.2 software). We investigated differences in community diversity via 2-way ANOVAs in 8 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... which sampling area (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays) and context (preversus post-spill) were fixed factors. Because variances were stable among groups, no data transformations were required prior to analyses. These approaches are proscribed in earlier syntheses for detecting environmental impacts [17]. Critiques of employing parametric testing to detect ecosystem injury exist due to interannual variability and reduced statistical power [14], although those concerns have focused on analyses involving single species. We determined the relative probability for oil-larvae encounters (‘risk’) for the twenty most commonly captured fishes, and used these data to explore how individual species responded differently to large-scale oil pollution in the northern GOM. Information on the seasonal timing of spawning and distribution of larvae from shore to the outer margin of the continental shelf was collected from the literature (tables S1 and S4), and used to define 4 levels of risk (in addition to an ‘unknown’ [n = 4] category containing species for which no data were available). ‘Low’ risk species (n = 6) included those in which larvae remained inside estuaries, either in the plankton or as benthic egg masses, regardless of spawning season. ‘Moderate-Low’ risk species (n = 4) were defined by having either 1) larvae distributed in estuaries as well as across the continental shelf, or 2) larvae distributed across the continental shelf, but not likely during the spill period (i.e., April– September). Only two ‘Moderate’ risk species were identified: pigfish (Orthopristis chrysoptera) spawn throughout summer, and have larvae distributed within nearshore waters; while flounder (Paralichthys spp.) have larvae distributed across the continental shelf, with a protracted spawning that extends into April or May (i.e., some overlap with the oil spill). ‘High’ risk species (n = 4) spawn offshore and have larvae distributed across the continental shelf. Furthermore, spawning data indicates that these species would have produced larvae sometime during the DH spill (April–Sept in our classification scheme). Based on these risk guilds, we examined the response of fishes to the spill by calculating the ratio of 2010 CPUE data (averaged) to 2006–2009 CPUE data (averaged) for each species. Following these calculations, ratios >1 indicate that average 2010 catch rates were higher than during the previous 4 years, while ratios <1 indicate that average 2010 catch rates were lower than during the previous 4 years. Using each species as a replicate measure, we used ‘risk’ and region (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays) as fixed factors in a 2-way ANOVA that compared 2010 CPUE: 2006–2009 CPUE trends. Because variances were stable among groups, no data transformations were required prior to analyses. Similarly, we determined whether species were likely to have experienced significant release from harvest pressure following large-scale closures in the northern GOM, and examined how this may have affected CPUE data in 2010. For each of the twenty most commonly caught fish, we designated species as ‘fished’ if they were included in any state or federal management plan as of Dec 31, 2010 (table S5), or identified as <1% (by biomass) of bycatch in shrimp trawl fisheries within the northern GOM (table S5). As before, we examined the response of fishes to the spill by calculating the ratio of 2010 CPUE data (averaged) to 2006–2009 CPUE data (averaged) for each species. Using each species as a replicate measure, we used ‘fishing pressure’ (with fished species including 9 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... species that are targeted or captured as incidental bycatch at significant levels) and region (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays) as fixed factors in a 2-way ANOVA that compared 2010 CPUE: 2006–2009 CPUE trends. Because variances were stable among groups, no data transformations were required prior to analyses. All univariate tests were conducted using StatView 5.0.1 software (SAS Institute Inc.). Because each statistical analysis applied to separate and easily distinguishable hypotheses, we made no corrections to experiment-wise alpha for any of the univariate (total fishes CPUE, individual fishes CPUE, risk guilds, harvest guilds, diversity) or multivariate (ANOSIM) tests we conducted [34]. S UPPORTING I NFORMATION Figure S1. Catch rates of all fishes, pooled together, among sampling areas prior to (2006–2009) and following (2010) the Deepwater Horizon disaster. (DOCX) Figure S2. Catch rates of individual species, among sampling areas prior to (2006–2009) and following (2010) the Deepwater Horizon disaster. Data are presented for the 20 most abundant species. (DOCX) Figure S3. Catch rates among sampling areas and years for the 20 most abundant species collected during trawl surveys. (DOCX) Figure S4. Diversity measures for seagrass-associated fish communities within sampling areas affected by the Deepwater Horizon disaster. (DOCX) Table S1. Summary table for CPUE data (fish kilometer-towed−1) of fishes prior to (2006–2009) and following (2010) the DH disaster. (DOCX) 10 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... Table S2. Distribution of trawl samples among sampling areas (Chandeleur Islands, Gulf Islands, Grand Bay, Florida Bays) and years (2006–2010). (DOCX) Table S3. Quantitative description of seagrass habitats sampled throughout the northern Gulf of Mexico during 2006–2010. (DOCX) Table S4. Information used to determine the likelihood of larvae contacting oiled water during the summer of 2010. (DOCX) Table S5. Summary table for the management status of the 20 most abundant fishes collected during our survey program. (DOCX) Table S6. Summary table of the effects of sampling area and year (context: pre- versus post-spill) on the catch rates of the 20 most abundant fishes collected during surveys in northern Gulf of Mexico seagrass meadows. (DOCX) Table S7. Comparisons of community structure between catch data prior to (2006–2009) or immediately following (2010) the Deepwater Horizon disaster (ANOSIM and SIMPER). (DOCX) Table S8. Summary table of the effects of sampling area and year (context: pre- versus post-spill) on the diversity (S, ES(20), H′, and J′) of trawl samples collected within northern Gulf of Mexico seagrass meadows. 11 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... (DOCX) Table S9. Summary table of the effects of sampling area, larval risk and harvest pressure on the change in catch rates of individual species for pre- (2006–2009) and post-spill (2010) data. (DOCX) Dataset S1. Complete CPUE data obtained for 2006–2009 trawl surveys within seagrass meadows of the northern Gulf of Mexico. (XLSX) A CKNOWLEDGMENTS We are extremely grateful to the students and technicians who aided in the field, especially C. Baillie, M. Brodeur, J. Myers, O. Rhoades and S. Williams. B. Raines supplied the image of juvenile fishes in Fig. 1. Constructive comments and support were provided by S. Powers, C. Peterson, J. Kenworthy, and 2 anonymous reviewers. A UTHOR C ONTRIBUTIONS Conceived and designed the experiments: FJF KLH. Performed the experiments: FJF KLH. Analyzed the data: FJF. Wrote the paper: FJF KLH. R EFERENCES 1. Crone TJ, Tolstoy M (2010) Magnitude of the 2010 Gulf of Mexico Oil Leak. Science 330: 634–643. 2. Peterson CH, Rice SD, Short JW, Esler D, Bodkin JL, et al. (2003) Long-term ecosystem response to the Exxon Valdez Oil Spill. Science 302: 2082–2086. 3. Camilli R, Reddy CM, Yoerger DR, Van Mooy BAS, Jukuba MV, et al. (2010) Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon. Science 330: 201–204. 4. Hanzen TC, Dubinsky EA, DeSantic TZ, Andersen GL, Piceno YM, et al. (2010) Deep-sea oil plume enriches indigenous oil-degrading bacteria. Science 330: 204–208. 5. Kerr R, Kintisch E, Stokstad E (2010) Will Deepwater Horizon set a new standard for catastrophe? Science 328: 674–675. 12 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... 6. Able KW, Fahay MP (2010) 566 p. 7. Miller BS, Kendall AW (2009) 376 p. 8. Hjort J (1914) Fluctuation in the great fisheries of northern Europe reviewed in the light of biological research. Rapports et Procès-Verbaux des Réunions du Conseil International pour l'Exploration de la Mer 20: 1–228. 9. Kocan RM, Hose JE, Brown ED, Baker TT (1996) Injury to the early life history stages of Pacific herring in Prince William Sound after the Exxon Valdez oil spill. Can J Fish Aquat Sci 53: 2366–2375. 10. Tuvikene A (1995) Responses of fish to polycyclic aromatic hydrocarbons (PAHs). Ann Zool Fennici 32: 295–309. 11. Carls MG, Rice SD, Hose JE (1999) Sensitivity of fish embryos to weathered crude oil: Part I. Low-level exposure during incubation causes malformations, genetic damage, and mortality in larval pacific herring (Clupea pallasi). Environ Toxicol Chem 18: 481–493. 12. Brown ED, Baker TT, Hose JW, Kocan RM, Marty GD, et al. (1996) Injury to the early life history stages of Pacific herring in Prince William Sound after the Exxon Valdez oil spill. Am Fish Soc Sym 18: 448–462. 13. Levin LA (2006) Recent progress in understanding larval dispersal: new directions and digressions. Integr Comp Biol 46: 282–297. 14. Hilborn R (1996) Detecting population impacts from oil spills: a comparison of methodologies. Am Fish Soc Sym 18: 639–644. 15. Beck MW, Heck KL, Able KW, Childers DL, Eggleston DB, et al. (2001) The identification, conservation, and management of estuarine and marine nurseries for fish and invertebrates. Bioscience 51: 633–641. 16. Fodrie FJ, Heck KL, Powers SP, Graham WM, Robinson KL (2010) Climate-related, decadal-scale assemblage changes of seagrass-associated fishes in the northern Gulf of Mexico. Glob Change Biol 16: 48–59. 17. Underwood AJ (1996) pp. 151–175. in Detecting Ecological Impacts: Concepts and Applications in Coastal Habitats, Schmitt RJ, Osenberg CW (Academic Press). 18. NOAA Fisheries Service Deepwater Horizon Oil Spill Information. Available: http://sero.nmfs.noaa.gov 13 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... /deepwater_horizon_oil_spill.htm. Accessed: 2011 Jun 13. 19. D'Alessandro EK, Sponaugle S, Serafy JE (2010) Larval ecology of a suite of snappers (family: Lutjanidae) in the Straits of Florida, western Atlantic Ocean. Mar Ecol Prog Ser 410: 159–175. 20. Holt GJ, Holt SA (2000) Vertical distribution and the role of physical processes in the feeding dynamics of two larval sciaenids Sciaenops ocellatus and Cynoscion nebulosus. Mar Ecol Prog Ser 193: 181–190. 21. Peterson CH, Kennicutt MC, Green RH, Montagna P, Harper DE, et al. (1996) Ecological consequences of environmental perturbations associated with offshore hydrocarbon production: a perspective on long-term exposures in the Gulf of Mexico. Can J Fish Aquat Sci 53: 2637–2654. 22. Edgar GJ, Marshall PA, Mooney P (2003) The effect of the Jessica grounding on Galápagos reef fish communities adjacent to the wreck site. Mar Pollut Bul 47: 296–302. 23. Graham WM, Condon RH, Carmichael RH, D'Ambra I, Patterson HK, et al. (2010) Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill. Environ Res Lett 5: 045301. 24. Heintz RA, Rice SD, Wertheimer AC, Bradshaw RF, Thrower FP, et al. (2000) Delayed effects on growth and marine survival of pink salmon Oncorhynchus gorbuscha after exposure to crude oil during embryonic development. Mar Ecol Prog Ser 208: 205–216. 25. Wertheimer AC, Heintz RA, Thedinga JF, Maselko JM, Rice SD (2000) Straying behavior of adult pink salmon (Oncorhynchus gorbuscha) exposed as embryos to weathered Exxon Valdez crude oil. Trans Am Fish Soc 129: 989–1004. 26. Bue BG, Sharr S, Seeb JE (1998) Evidence of damage to pink salmon populations inhabiting Prince William Sound, Alaska, two generations after the Exxon Valdez Oil Spill. Trans Am Fish Soc 127: 35–43. 27. Pearson WH, Elston RA, Bienert RW, Drum AS, Antrim LD (1999) Why did the Prince William Sound, Alaska, Pacific herring (Clupea pallasi) fisheries collapse in 1993 and 1994. Can J Fish Aquat Sci 56: 711–737. 28. U. S. Department of the Interior, Minerals Management Service Environmental Division, “Oil-Spill Risk Analysis: Gulf of Mexico Outer Continental Shelf (OCS) Lease Sales, Eastern Planning Area, 2003–2007 and Gulfwide OCS Program, 2003–2042” OCS Report MMS 2002–069. 29. Sagarin R, Pauchard A (2010) Observational approaches in ecology open new ground in a changing world. Front Ecol Environ 8: 379–386. 14 of 15 3/19/12 4:07 AM PLoS ONE: Response of Coastal Fishes to the Gulf of Mexico O... http://www.plosone.org/article/info:doi/10.1371/journal.pone.0... 30. Diaz RJ, Rosenberg R (2008) Spreading dead zones and consequences for marine ecosystems. Science 321: 926–929. 31. Baum JK, Myers RA (2004) Shifting baselines and the decline of pelagic sharks in the Gulf of Mexico. Ecol Lett 7: 135–145. 32. Byron D, Heck KL (2006) Hurricane effects on seagrasses along Alabama's Gulf Coast. Estuar Coast 29: 939–942. 33. Clark KR, Gorley RN (2001) PRIMER v5: User Manual/Tutorial. Plymouth: Plymouth Marine Laboratory. 34. Moran MD (2003) Arguments for rejecting the sequential Bonferroni in ecological studies. Oikos 100: 403–405. All site content, except where otherwise noted, is licensed under a Creative Commons Attribution License. 15 of 15 3/19/12 4:07 AM Roundtable Gulf of Mexico Oil Blowout Increases Risks to Globally Threatened Species Claudio Campagna, Frederick T. Short, Beth A. Polidoro, Roger McManus, Bruce B. Collette, Nicolas J. Pilcher, Yvonne Sadovy de MitchesoN, Simon N. Stuart, and Kent E. Carpenter Fourteen marine species in the Gulf of Mexico are protected by the US Endangered Species Act, the Marine Mammal Protection Act, and the Migratory Bird Treaty Act. As the British Petroleum oil spill recovery and remediation proceed, species internationally recognized as having an elevated risk of extinction should also receive priority for protection and restoration efforts, whether or not they have specific legal protection. Forty additional marine species—unprotected by any federal laws—occur in the Gulf and are listed as threatened on the International Union for Conservation of Nature’s (IUCN) Red List. The Red List assessment process scientifically evaluates species’ global status and is therefore a key mechanism for transboundary impact assessments and for coordinating international conservation action. Environmental impact assessments conducted for future offshore oil and gas development should incorporate available data on globally threatened species, including species on the IUCN Red List. This consideration is particularly important because US Natural Resource Damage Assessments may not account for injury to highly migratory, globally threatened species. Keywords: IUCN Red List, Gulf of Mexico, oil spill, threatened species A primary concern following the British Petroleum Deepwater Horizon oil well blowout and the resulting oil pollution in the Gulf of Mexico is the damage to marine plants and animals—especially to those already considered vulnerable. Several US federal and state statutes protect coastal and marine species of special concern found in the Gulf of Mexico, including 14 marine species protected by the US Endangered Species Act (ESA), the Marine Mammal Protection Act, and the Migratory Bird Treaty Act. Additionally, species protected by Mexican and Cuban laws must be considered. The International Union for Conservation of Nature (IUCN) Red List of Threatened Species (IUCN 2010) results from a rigorous scientific process to assess the relative extinction risks of species globally, using widely accepted standards (Mace et al. 2008, Hoffmann et al. 2010). As such, the IUCN Red List categories and criteria are the most respected international system for classifying global extinction risk at the species level (De Grammont and Cuarón 2006, Rodrigues et al. 2006, Carpenter et al. 2008, Mace et al. 2008, Schipper et al. 2008). In addition to the 14 species protected by US law, the IUCN Red List identifies a further 39 species in the Gulf (table 1) as belonging to one of three threatened categories: critically endangered, endangered, or vulnerable (IUCN 2001). All species in Red List threatened categories have an elevated risk of extinction, and we argue they should receive priority for protection and restoration efforts in the Gulf, whether or not they have specific legal protection from any government entity in the region. The Gulf oil blowout is likely to worsen the threat status of some of these species as more of the spill’s impacts manifest. United States law requires restoration to prevent conditions of natural resources damaged by the oil pollution, and restoration is overseen by NOAA’s (the National Oceanic and Atmospheric Administration) Natural Resource Damage Assessment (NRDA; NOAA 2010a). The primary legal authority for assessing damages and providing for recovery of coastal and marine species is the Oil Pollution Act, which is implemented by the NRDA. Under the Damage Assessment Remediation and Restoration Program, NRDA trustees determine whether injury to public trust resources has occurred, as well as the type, amount, and methods of restoration needed. According to the most recent revision of the Mexican list of threatened and protected species (NOM 2002, 2006), all marine mammals and marine turtles are accorded some degree of protection status in Mexico (e.g., classified as in danger of extinction, as threatened, or under special protection). Other than mammals and turtles, only three species are protected in Mexico: subspecies of two seabirds present in BioScience 61: 393–397. ISSN 0006-3568, electronic ISSN 1525-3244. © 2011 by American Institute of Biological Sciences. All rights reserved. Request permission to photocopy or reproduce article content at the University of California Press’s Rights and Permissions Web site at www.ucpressjournals.com/ reprintinfo.asp. doi:10.1525/bio.2011.61.5.8 www.biosciencemag.org May 2011 / Vol. 61 No. 5 • BioScience 393 Roundtable Table 1. Marine species in International Union for Conservation of Nature threatened Red List categories (critically endangered, endangered, or vulnerable) that have a distribution directly overlapping the area of the oil spill, or that are found in the greater Gulf region extending from Texas to Miami, Florida. Red List category species name Common name Protection status Red List category species name Lepidochelys kempii Kemp’s ridley turtle ESA-E Epinephelus flavolimbatus Yellowedge grouper Eretmochelys imbricata Hawksbill turtle ESA-E Epinephelus niveatus Snowy grouper Dermochelys coriacea Leatherback turtle ESA-E Mycteroperca interstitialis Yellowmouth grouper Thunnus thynnus Atlantic bluefin tuna, western stock Lachnolaimus maximus Hogfish Alopias superciliosus Bigeye thresher shark Alopias vulpinus Common thresher shark Carcharhinus longimanus Oceanic whitetip shark Carcharhinus obscurus Dusky shark Carcharhinus plumbeus Sandbar shark Carcharhinus signatus Night shark Centrophorus granulosus Gulper shark Cetorhinus maximus Basking shark Carcharodon carcharias Great white shark Isurus oxyrinchus Shortfin mako Isurus paucus Longfin mako Carcharias taurus Sand tiger shark Odontaspis ferox Small-tooth sand tiger shark Rhincodon typus Whale shark Sphyrna zygaena Smooth hammerhead Squalus acanthias Spiny dogfish Gymnura altavela Butterfly ray Scalloped hammerhead shark Agaricia lamarcki Lamarck’s sheet coral Great hammerhead shark Montastraea franksi Montastraea coral Boulder star coral Dendrogyra cylindrus Pillar coral Mountainous star coral Dichocoenia stokesii Elliptical star coral Mycetophyllia ferox Rough cactus coral Oculina varicose Large ivory coral Halophila baillonii Clover seagrass Critically endangered Speckled hind Epinephelus itajara Atlantic goliath grouper Epinephelus nigritus Warsaw grouper Pristis pectinata Smalltooth sawfish Pristis perotteti Largetooth sawfish Narcine bancroftii Lesser electric ray Acropora cervicornis Staghorn coral ESA-T Acropora palmate Elkhorn coral ESA-T ESA-E Endangered Balaenoptera borealis Serving whale ESA-E, MMPA Balaenoptera musculus Blue whale ESA-E, MMPA Balaenoptera physalus Finback whale ESA-E, MMPA Pterodroma caribbaea Jamaica petrel Pterodroma hasitata Black-capped petrel MBTA Caretta caretta Loggerhead turtle ESA-T Chelonia mydas Green turtle ESA-E, ESA-T (by range) Sphyrna mokarran Montastraea annularis Montastraea faveolata Vulnerable Trichechus manatus Physeter macrocephalus Protection status Vulnerable (continued) Epinephelus drummondhayi Sphyrna lewini Common name Manatee Sperm whale ESA-E, MMPA ESA-E, MMPA ESA-E, endangered under the Endangered Species Act (ESA); ESA-T, threatened under the ESA; MBTA, listed on the Migratory Bird Treaty Act; MMPA, listed on the Marine Mammal Protection Act. Source: IUCN 2010. See the supplementary table online at dx.doi.org/10.1525/bio.2011.61.5.8. the Gulf of Mexico (Pelecanus occidentalis and Oceanodroma leucorhoa) and the smalltooth sawfish (Pristis pectinata). No species-level protection occurs in Cuba comparable to the US ESA, but there are laws protecting biodiversity (e.g., Ley No. 81 Del Medio Ambiente; Ministerio De Ciencia, Tecnologia Y Medio Ambiente Resolucion No. 111/96). The Gulf of Mexico has exceptionally high marine biodiversity, with 15,419 recorded species, of which 10% (1511) are endemic (Felder and Camp 2009). This diversity is partly attributable to the Gulf’s geographic position within the transition zone between temperate and tropical waters. Some threatened species in the Gulf (e.g., whale shark, Rhincodon typus; loggerhead turtle, Caretta caretta) occur globally but 394 BioScience • May 2011 / Vol. 61 No. 5 have significant populations, spawning aggregations, or nesting sites in the Gulf region. Therefore, greater threats in this region may have implications for the species’ global survival. Other species (e.g., Kemp’s ridley turtle, Lepidochelys kempii; the western Atlantic population of bluefin tuna, Thunnus thynnus) breed only in the Gulf, and oil spill damage exacerbates previously existing threats to these species. IUCN Red List assessments are being expanded to evaluate more marine species (http://sci.odu.edu/gmsa/ ), including some in the Gulf of Mexico. The IUCN has assessed 322 species in the Gulf of Mexico to date, 53 of which are in threatened categories (table 1); an additional 29 are listed as near threatened (see the supplementary table online www.biosciencemag.org Roundtable at dx.doi.org/10.1525/bio.2011.61.5.8). The IUCN assessments include all Gulf marine mammals (5 of 28 species threatened), sea turtles (all 5 species threatened), seagrasses (2 of 9 threatened or near threatened), mangroves (0 of 6 threatened), reef-building corals (11 of 60 threatened or near threatened), wrasses (1 of 20 threatened), sharks and rays (43 of 131 threatened or near threatened), seabirds (3 of 40 threatened or near threatened), and groupers (11 of 22 threatened or near threatened). Groupers are of particular concern; three species are classified as critically endangered on the Red List and the Atlantic goliath grouper (Epinephelus itajara) is listed as near extinction. An oil spill of this magnitude threatens many species already listed under IUCN threatened categories—more species than are currently protected by the ESA. In 1996, the IUCN assessed the western stock of the Atlantic bluefin tuna as critically endangered, and the Convention on Biological Diversity recently petitioned the US Department of Commerce to protect the species under the ESA (CBD 2010). There are two spawning populations of bluefin tuna, one in the Gulf of Mexico and the other in the Mediterranean Sea. Although there is extensive mixing of the populations on both sides of the Atlantic, particularly on the feeding grounds off the eastern coast of North America, individuals hatched in the Gulf of Mexico return there to spawn (spawning site fidelity). Peak spawning in the Gulf occurs from mid-April to June, unfortunately during the period of the British Petroleum oil spill in 2010. Like tuna, many other species in threatened Red List categories—that are not protected by the ESA—require protection and remediation from the oil spill. Seagrasses are a unique group of 72 undersea flowering plants found in coastal seas globally. In the Gulf of Mexico, there are nine seagrass species, and these plants provide crucial structural habitat and nursery grounds for many recreationally and commercially important fish and invertebrates (including Gulf pink shrimp and brown shrimp), as well as waterfowl. Some seagrasses, as indicated by their common names (e.g., turtle grass and manatee grass) are the primary food for already threatened species of sea turtles and manatees. The seagrass habitat, and the proliferation of the species it supports, may be at risk as a result of the oil spill; three diminutive seagrass species of the genus Halophila are most threatened. Halophila baillonii is listed as vulnerable and Halophila engelmanni is listed as near threatened on the Red List (Short et al. 2011), and Halophila johnsonii is listed on the ESA. The limited global distributions of these species, restricted primarily to Gulf and Florida waters in the cases of H. engelmanni and H. johnsonii, mean their risk of global extinction could be elevated by the oil spill. Halophila baillonii, already rapidly declining in the Caribbean, is the fourth most threatened seagrass species in the world. Potential damages to these seagrasses from the oil pollution in the Gulf should be assessed, and recovery actions for these species should be aided by funding available from the Oil Pollution Act and other sources. www.biosciencemag.org The whale shark is listed as vulnerable on the IUCN Red List but is not protected by the ESA. Found worldwide in tropical and warm temperate waters, many individuals aggregate in the Gulf of Mexico in the summer months. The whale shark is the largest fish in the world; it feeds almost entirely on plankton, crustaceans, and small fishes. It is one of only three filter-feeding species of shark, sieving zooplankton as small as 1 millimeter in diameter through the fine mesh of its gill rakers. The shark’s feeding behavior puts it at high risk from the oil itself and the oil dispersants used in the Gulf. Although relatively little is known about the biology of the whale shark, populations will probably be slow to recover from disturbances given the species’ estimated long life span, slow reproductive rate (Pauly 2002), and naturally low abundance outside of mating aggregations. The Kemp’s ridley sea turtle is listed as critically endangered on the Red List and is also protected by the ESA. This turtle nests exclusively in the Gulf and is the rarest sea turtle in the world. Of the threatened marine species that frequent the Gulf, only the Kemp’s ridley depends on Gulf shores for nesting, and most of its young develop in Gulf waters. Although it appears that the 2010 hatchlings did not encounter the spill directly, the timing of the oil spill could not have been worse for this species, clashing as it did with the turtles’ key reproductive period. The vast majority of sea turtles found dead since the spill were Kemp’s ridleys (NOAA 2010b). The Kemp’s ridley was just on the road to recovery after a population collapse a few decades ago that drove it near extinction; the species now faces a new environmental hurdle. The West Indian manatee (Trichechus manatus) is listed as vulnerable by the IUCN and is considered endangered under US law. Manatees are found in the Gulf and around the coastline of Florida, in the range of the oil spill. Manatees may be affected by air quality and oil at the surface, which they encounter as they emerge to breathe; oil irritating their skin and eyes; the consumption of seagrass species—their primary food—that are covered in oil; and chemical oil dispersants that may also have a toxic effect. The Florida manatee (Trichechus manatus latirostris), a subspecies of the West Indian manatee, is additionally threatened by loss of habitat, entanglement with fishing gear, and increased boating activity, as well as extreme cold temperatures that killed 10% of the population during the winter of 2009–2010. The Florida manatee subspecies was listed as endangered in 2008 by the IUCN. The trends in species declines are clearly worrying, particularly because the Gulf was already a system affected by various risk factors before the oil blowout occurred. How can we adequately address the threats of oil and gas development and steward the Gulf ’s biological diversity? Priorities at this stage must focus on species with high commercial value, species critical to the integrity of coastal and marine ecosystems in the Gulf, species with populations in decline before the blowout, and species now recognized as in greater danger of extinction. Because marine species in particular May 2011 / Vol. 61 No. 5 • BioScience 395 Roundtable may be underrepresented by the ESA (Wilcove and Master 2005), the ongoing NRDA in the Gulf of Mexico—as well as environmental impact assessments conducted for offshore oil and gas development—should include available data on globally threatened species, including the expanding species data sets on the IUCN Red List. Species information on the Red List can serve as a standardized mechanism to identify and coordinate conservation and mitigation priorities, especially for highly migratory and transboundary species. The US Department of the Interior must reevaluate the “low risk” status currently attributed to pollution from routine operations of oil and gas development, as well the potential impacts of catastrophic events such as oil spills, in its compliance with the National Environmental Policy Act, the ESA, and other applicable domestic and international laws. Species identified as threatened with extinction on the IUCN Red List may become even more threatened as a result of the oil spill. Such elevations in threatened status are part of the spill’s impacts and as such are damages that must be recognized and compensated. The six threatened grouper species on the Red List that occur in the Gulf, for example, currently receive no protection under the ESA or Mexican law, despite their status as a major food resource in the region and a high-value restaurant menu item. Gulf-occurring animals and plants protected by the ESA (and other US laws) should be priorities for federal damage assessments; as such, these laws should help restore the natural resources injured by the release of oil or hazardous substances. Although the methodology of assessment and the names of threatened categories may differ among laws, assessments, and criteria, the IUCN Red List is a highly credible source of species requiring particular attention—both for damage assessment and for special consideration for future regulations of oil and gas development. As a result of the rapid increase in IUCN assessments during the last few years, we now know there are many threatened marine species in the Gulf that are not currently protected by US law (table 1). Threatened species not yet listed in national legislation should nevertheless be the subject of damage assessments, targeted research, and monitoring, as well as recovery efforts when needed. Although understanding has improved of the medium- and long-term impacts from oil pollution on animal and plant physiologies, there is still much we do not know. Globally, countries must improve risk assessments of offshore oil and gas development, and seek to expand and regularize damage and impact assessments, domestically and internationally. These impacts must be systematically considered to establish priorities for research and monitoring that will best ensure effective species and system recovery. Although the research agenda should focus on the United States’ immediate needs, its development should also support similar efforts in other regions of the world in identifying species of priority concern. The IUCN Red List is continually improved and revised under strict standards and criteria, and its value in assessing the global conservation status of biological diversity steadily expands. 396 BioScience • May 2011 / Vol. 61 No. 5 Acknowledgments The majority of marine species assessments conducted through the International Union for Conservation of Nature (IUCN) Species Survival Commission are made through the Global Marine Species Assessment, with core funding provided by Tom Haas and the New Hampshire Charitable Foundation. We thank numerous partners who helped compile information, including BirdLife International; SeagrassNet; the Groupers and Wrasses, Tunas and Billfishes, Sharks, and Marine Turtles IUCN Species Specialist Groups; Jonnell Sanciangco and Suzanne Livingstone (Global Marine Species Assessment); and Cynthia Taylor (Sirenia Red List Authority Focal Point). Thanks to Cathy Short for editing the manuscript. This article is Jackson Estuarine Laboratory contribution no. 498. References cited Carpenter KE, et al. 2008. One-third of reef-building corals face elevated extinction risk from climate change and local impacts. Science 321: 560–563. [CBD] Center for Biological Diversity. 2010. Petition to List the Atlantic Bluefin Tuna (Thunnus thynnus) as Endangered under the United States Endangered Species Act. CBD. (2 February 2011; www.nmfs.noaa.gov/ pr/pdfs/species/cbd_bluefintunapetition_5242010.pdf) De Grammont PC, Cuarón AD. 2006. An evaluation of threatened species categorization systems used on the American continent. Conservation Biology 20: 14–27. Felder DL, Camp DK, eds. 2009. Biodiversity, vol. 1. Gulf of Mexico Origin, Waters, and Biota. Texas A&M University Press. Hoffmann M, et al. 2010. The impact of conservation on the status of the world’s vertebrates. Science 330: 1503–1509. [IUCN] International Union for Conservation of Nature. 2001. IUCN Red List Categories and Criteria, version 3.1. (2 February 2011; www.iucnredlist. org/technical-documents/categories-and-criteria/2001-categories-criteria). ———. 2010. IUCN Red List. (2 February 2011; www.iucnredlist.org). Mace GM, Collar NJ, Gaston KJ, Hilton-Taylor C, Akçakaya HR, LeaderWilliams N, Milner-Gulland EJ, Stuart SN. 2008. Quantification of extinction risk: The background to IUCN’s system for classifying threatened species. Conservation Biology 22: 1424–1442. [NOAA] National Oceanic and Atmospheric Administration. 2010a. US Natural Resource Damage Assessment, Damage Assessment Remediation and Restoration Program. (2 February 2011; www.darrp. noaa.gov) ———. 2010b. Sea Turtles, Dolphins, and Whales and the Gulf of Mexico Oil Spill. NOAA Office of Protected Resources. (2 February 2011; www. nmfs.noaa.gov/pr/health/oilspill.htm) [NOM] Norma Oficial Mexicana. 2002. NOM-059-Ecol, Diario Oficial de la Federación Tomo DLXXXII 4: 1–80. Pauly D. 2002. Growth and mortality of the basking shark Cetorhinus maximus and their implications for management of whale sharks Rhincodon typus. Pages 199–208 in Fowler SL, Reed TM, Dipper FA, eds. Elasmobranch Biodiversity, Conservation and Management. Proceedings of the International Seminar and Workshop, July 1997, Sabah, Malaysia. IUCN. Rodrigues ASL, Pilgrim JD, Lamoreux JF, Hoffmann M, Brooks TM. 2006. The value of the IUCN Red List for conservation. Trends in Ecology and Evolution 21: 71–76. Schipper JS, et al. 2008. The status of the world’s land and marine mammals: Diversity, threat, and knowledge. Science 322: 225–230. Short FT, et al. 2011. Extinction risk assessment of the world's seagrass species. Biological Conservation. Forthcoming. Wilcove DS, Master LL. 2005. How many endangered species are there in the United States? Frontiers in Ecology and the Environment 3: 414–420. www.biosciencemag.org Roundtable Claudio Campagna ([email protected]) is with the Wildlife Conservation Society in New York, New York. Frederick T. Short is with the Department of Natural Resources and the Environment, University of New Hampshire, Jackson Estuarine Laboratory, in Durham. Beth A. Polidoro, Roger McManus, and Kent E. Carpenter are with the Global Marine Species Assessment, Marine Biodiversity Unit, International Union for Conservation of Nature (IUCN) Species Programme, Department of Biological Sciences, at Old Dominion University, in Norfolk, Virginia. Roger McManus is also with the Global Marine Species Assessment, IUCN Species Survival Commission, Perry Institute for Marine Science, in Jupiter, Florida. Bruce B. Collette is with the National Marine Fisheries Service Systematics Laboratory, at the National www.biosciencemag.org Museum of Natural History, in Washington, DC. Nicolas J. Pilcher is with the Marine Research Foundation in Sabah, Malaysia. Yvonne Sadovy de Mitcheson is with the School of Biological Sciences, University of Hong Kong, in China. Simon N. Stuart is with the IUCN Species Survival Commission, at the United Nations Environment Programme World Conservation Monitoring Centre, in Cambridge, United Kingdom; the Department of Biology and Biochemistry, University of Bath, in the United Kingdom; the Al Ain Wildlife Park and Resort, in Abu Dhabi, United Arab Emirates; and Conservation International, in Arlington, Virginia. Campagna, Short, Polidoro, Collette, Pilcher, Sadovy, and Carpenter are also with the IUCN Species Survival Commission Marine Conservation Subcommittee in Gland, Switzerland. May 2011 / Vol. 61 No. 5 • BioScience 397 LETTER Underestimating the damage: interpreting cetacean carcass recoveries in the context of the Deepwater Horizon/BP incident Rob Williams1 , Shane Gero2 , Lars Bejder3 , John Calambokidis4 , Scott D. Kraus5 , David Lusseau6 , Andrew J. Read7 , & Jooke Robbins8 1 Marine Mammal Research Unit, University of British Columbia, Vancouver, Canada Department of Biology, Dalhousie University, Halifax, Canada 3 Centre for Fish and Fisheries Research, Cetacean Research Unit, Murdoch University, Western Australia 4 Cascadia Research Collective, Olympia, WA, USA 5 New England Aquarium, Boston, MA, USA 6 School of Biology, Aberdeen University, Aberdeen, Scotland, UK 7 Nicholas School of the Environment, Duke University, Beaufort, NC, USA 8 Humpback Whale Studies Program, Provincetown Center for Coastal Studies, Provincetown, MA, USA 2 Keywords Anthropogenic impacts; dolphin; Deepwater Horizon; Gulf of Mexico; mortality; oil; strandings. Correspondence Rob Williams, Current address: Sea Mammal Research Unit, Scottish Oceans Institute, St Andrews Fife KY16 8LB. Tel: +44 (0)1334 462630; Fax: +44 (0)1334 463443. E-mail: [email protected] Received 23 September 2010 Accepted 15 February 2011 Editor Leah Gerber doi: 10.1111/j.1755-263X.2011.00168.x Abstract Evaluating impacts of human activities on marine ecosystems is difficult when effects occur out of plain sight. Oil spill severity is often measured by the number of marine birds and mammals killed, but only a small fraction of carcasses are recovered. The Deepwater Horizon/BP oil spill in the Gulf of Mexico was the largest in the U.S. history, but some reports implied modest environmental impacts, in part because of a relatively low number (101) of observed marine mammal mortalities. We estimate historical carcass-detection rates for 14 cetacean species in the northern Gulf of Mexico that have estimates of abundance, survival rates, and stranding records. This preliminary analysis suggests that carcasses are recovered, on an average, from only 2% (range: 0–6.2%) of cetacean deaths. Thus, the true death toll could be 50 times the number of carcasses recovered, given no additional information. We discuss caveats to this estimate, but present it as a counterpoint to illustrate the magnitude of misrepresentation implicit in presenting observed carcass counts without similar qualification. We urge methodological development to develop appropriate multipliers. Analytical methods are required to account explicitly for low probability of carcass recovery from cryptic mortality events (e.g., oil spills, ship strikes, bycatch in unmonitored fisheries and acoustic trauma). Introduction The Deepwater Horizon/BP oil spill in the Gulf of Mexico was not only the largest in the US history (Machlis & McNutt 2010) but was also the first to release oil at the sea floor (over 1.5 km below sea level) and to involve the widespread use of dispersants below the surface (Mascarelli 2010). However, many media reports have suggested that the spill caused only modest environmental impacts (Grunwald 2010; Walsh 2010), in part because of a low number of observed wildlife mortalities, especially marine mammals (Unified Area Command from the U.S. Fish and Wildlife Service and the 228 National Oceanic and Atmospheric Association 2010). Not surprisingly, perhaps, this spill has been compared to other acute environmental disasters, such as the 1989 Exxon Valdez oil spill (EVOS). In the case of EVOS, the mortality of sea otters became emblematic of environmental impact, as well as a contentious effort to agree on compensation (Ehrenfeld 1990; Estes 1991). In contrast, the Deepwater Horizon/BP event has not left such an iconic symbol of devastation. As of November 7, 2010, “only” 101 cetacean (whale, dolphin, and porpoise) carcasses 1 had been detected across the Northern Gulf 1 http://www.nmfs.noaa.gov/pr/health/oilspill.htm Conservation Letters 4 (2011) 228–233 c 2011 Wiley Periodicals, Inc. Copyright and Photocopying: R. Williams et al. of Mexico. The critical issue is, therefore, how to interpret this relatively low number of carcass recoveries in terms of impact to populations. The Gulf of Mexico is a semienclosed subtropical sea that forms essentially one ecosystem with many demographically independent cetacean populations (Mullin & Fulling 2004). Some of these cetacean populations, such as killer whales (Orcinus orca), false killer whales (Pseudorca crassidens), melonheaded whales (Peponocephala electra), and several beaked whale species, appear to be quite small, are poorly studied, or are found in the pelagic realm where they could have been exposed to oil and yet never strand. Small, genetically isolated populations of bottlenose dolphins (Tursiops truncatus)could have experienced substantial losses either inshore or offshore. In an ideal world, one would simply compare postspill to prespill abundance estimates. But, it is rare to have good knowledge of long-term trends in wildlife abundance (Bonebrake et al. 2010). Abundance of many marine mammal populations has been monitored for decades, but the low precision of most cetacean abundance estimates would prevent us from detecting all but the most catastrophic declines using conventional nullhypothesis testing (Taylor et al. 2007b). As a result, it would not be very informative to compare pre- and postspill abundance estimates for populations of cetaceans in the northern Gulf of Mexico. An alternative approach is to count the number of carcasses recovered, acknowledging that these recoveries were subject to a number of processes (e.g., sinking, decaying, scavenging, drifting) that reduce detection probability, and then adjusting the counts upward to estimate total mortality. This is the approach that is commonly taken to estimate the effects of power lines on bird mortality, for example, in which it has been shown in one instance that carcass counts underestimate total mortality by 32% (Ponce et al. 2010). This also appears to be the approach being taken to assess impacts of the Deepwater Horizon Incident on whales and dolphins, with the important caveat that the carcass counts appear to be presented at face value, with no attempt to extrapolate to total mortality (Unified Area Command from the U.S. Fish and Wildlife Service and the National Oceanic and Atmospheric Association 2010; Grunwald 2010; Walsh 2010). Cetacean carcasses do not necessarily strand along coastlines or remain afloat long enough to be detected at sea. The probability of detecting the death of a marine mammal depends on a wide range of physical and biological factors, including: behavioral responses prior to death, proximity of the carcass to shore (or at-sea observers), decomposition rates and processes, water temperature, wind regime, and local currents (Epperly et al. 1996). Cetaceans subject to natural predation would ob- Low probability of cetacean carcass recovery viously leave no carcass at all. Shore recoveries may be very site-specific, such that the likelihood of a carcass drifting to shore varies with the geography of the coastline itself (Faerber & Baird 2010). As such, “oiled” carcasses detected subsequent to the Deepwater Horizon/BP event are expected to represent a small fraction of total mortality in the Northern Gulf of Mexico. Given the magnitude of the spill and complexity of the response, quantifying the ecological impacts will take a long time. To contribute to this effort, we examined historical data from the Northern Gulf of Mexico to evaluate whether cetacean carcass counts in this region have previously been reliable indicators of mortality, and may therefore accurately represent deaths caused by the Deepwater Horizon/BP event. Methods We estimated historical carcass-detection rates for 14 species of cetaceans in the northern Gulf of Mexico for which species-specific estimates of abundance (Waring et al. 2009a, b; Mullin & Fulling 2004) species-level adultsurvival rates (Taylor et al. 2007a), and stranding records exist (Waring et al. 2009a, b). Estimates of mortality were generated for each species by multiplying recent abundance estimates by the species-specific mortality rate. An annual carcass-recovery rate was then estimated by dividing the mean number of observed strandings each year by our estimate of annual mortality (Table 1). First, an overall pooled carcass-recovery rate was calculated across all cetacean species (n = 14) in the Gulf of Mexico for which data was available by using the expected number of deaths across all species and the total number of observed carcasses across all species. Next, species-specific carcassrecovery rates were calculated using only species-specific values and a mean (n = 14) of those was taken across species. Species-specific carcass-recovery estimates were not generated for Bryde’s whale (Balaenoptera brydei), bottlenose dolphins (Tursiops sp.), and Fraser’s dolphins (Lagenodelphis hosei) due to uncertainties in their abundance and/or population structure (Waring et al. 2009b). No attempt was made to estimate carcass-recovery rates in the two taxonomic groups that are not identified to species in the field during raw data collection: Kogia (a pooled estimate for two species, dwarf and pygmy sperm whales), or mesoplodonts (a pooled estimate for a genus of similar-looking beaked whales) (Waring et al. 2009b). Results Our analysis suggests that an average of 4,474 individual cetaceans died annually between 2003 and 2007 from all c 2011 Wiley Periodicals, Inc. Conservation Letters 4 (2011) 228–233 Copyright and Photocopying: 229 R. Williams et al. Low probability of cetacean carcass recovery Table 1 Population parameters and illustrative species-specific carcass-recovery rates for 14 species from the Gulf of Mexico. Northern Gulf of Mexico population Sperm whale Cuvier’s beaked whale Atlantic spotted dolphind Pantropical spotted dolphin Striped dolphin Spinner dolphin Rough-toothed dolphind Clymene dolphin Killer whale False killer whale Pygmy killer whale Melon-headed whale Risso’s dolphin Short-finned pilot whale Average of all species Pooled across all species (n = 14) Population estimatea CVa Adultsurvival rateb Estimated annual mortalityc Mean observed annual strandingsa Carcass-detection rate (%) 1665 65 37611 34067 3325 1989 2653 6575 49 777 323 2283 1589 716 0.20 0.67 0.28 0.18 0.48 0.48 0.42 0.36 0.77 0.56 0.60 0.76 0.27 0.34 0.986 0.95 0.95 0.95 0.95 0.95 0.95 0.95 0.99 0.99 0.95 0.99 0.95 0.986 23.3 3.3 1880.6 1703.4 166.3 99.5 132.7 328.8 0.5 7.8 16.2 22.8 79.5 10.0 0.8 0.2 2.4 0.8 0.8 1 5.8 0.6 0 0 0.2 1.4 2.8 0.2 93,687 – – 4,474 17 3.4 6.2 0.13 0.05 0.48 1.0 4.4 0.18 0 0 1.2 6.1 3.5 2.0 2.0 0.4 Population abundance and stranding data (2003–2007) were taken from (Waring et al. 2009b), unless otherwise noted. Population-level estimates are preferable but generally unavailable, so data taken from (Taylor et al. 2007a). c Calculated as the abundance multiplied by mortality rate ( = 1–survival rate). d Population abundance and stranding data (2002–2006) were taken from (Waring et al. 2009a). a b natural and anthropogenic causes. However, during that period, an average of only 17 cetacean carcasses were detected annually along the northern Gulf of Mexico. This would suggest that the overall pooled rate of carcass recovery for cetaceans in the Gulf of Mexico is approximately 0.4% of the total estimated mortality. Table 1 breaks down the recovery rates by species. Carcasses were recovered only from a mean of 2.0% (range: 0–6.2%) of cetacean species deaths along the northern Gulf of Mexico. The disparity between this value and the overall pooled value likely results from undue influence of poorly studied and relatively rare species (e.g., Cuvier’s beaked whale and melon-headed whale; Table 1) with high estimated carcass-recovery rates that are weighted equivalently and treated as reliably in this average as estimates from species that are common and well studied. We have reason to believe that the Cuvier’s beaked whale recovery rate is positively biased. The original abundance estimate is thought to be an underestimate by a factor of 2 to 4, based on the assumption of certain track line detection (Mullin & Fulling 2004). Our carcass-recovery rate for deep-diving whales would then be biased high by a factor of 2–4. Discussion Our results indicate that carcass-recovery rates are historically low for cetaceans in the Gulf of Mexico. Studies of other populations show similar recovery rates. In 230 long-term studies of killer whales off the coasts of British Columbia and Washington State, in which populations are censused completely every year, carcasses from confirmed deaths of known individuals are recovered only 6% of the time (Fisheries and Oceans Canada 2008). Similarly, low-detection rates have been estimated for carcasses of eastern gray whales (Eschrichtius robustus, <5%, Heyning & Dahlheim 1990), North Atlantic right whales (Eubalaena glacialis, 17%, Kraus et al. 2005), and harbor porpoises (Phocoena phocoena, <1%, Moore & Read 2008), all of which occur in near-shore waters. Beached carcasses of other pelagic marine vertebrates have been shown to be equally poor indicators of mortality (for example, 7–13% recovery rates for four species of sea turtle, Epperly et al. 1996). As such, raw carcass counts alone are not reliable indicators of the magnitude of mortality for these species. We do not claim to have calculated definitive multipliers for this spill. Instead, our aim is to show plausible ranges for those multipliers, in order to illustrate how much they would affect our perception of the ecological damages caused by Deepwater Horizon incident and why this topic is worthy of additional resources for methodological development. Consider, for example, one sperm whale being detected as a carcass, and a necropsy identified oiling as a contributing factor in the whale’s death. If the carcass-detection rate for sperm whales is 3.4% (Table 1), then it is plausible that 29 sperm whale deaths represents the best estimate of total mortality, given no Conservation Letters 4 (2011) 228–233 c 2011 Wiley Periodicals, Inc. Copyright and Photocopying: R. Williams et al. additional information. If, for example, 101 cetacean carcasses were recovered overall, and all deaths were attributed to oiling, the average-recovery rate (2%) would translate to 5,050 carcasses, given the 101 carcasses detected (Table 1). As the necropsy results emerge, we can evaluate whether this prediction is high or low, but the sheer scope for underestimation builds a compelling case, in our view, for additional work. The vast majority of carcasses recovered appear to have been bottlenose dolphins.1 As necropsy results emerge and the need for recovery plans debated, we encourage such discussions to explicitly take into account the probability that the number of dolphins stranded represented something on the order of only 2% of the number of animals killed. The potential is high for the spill to have caused catastrophic impacts on small, localized populations of bottlenose dolphins in the Gulf. We note that coastal and offshore forms of bottlenose dolphins are found off California, with the coastal carcass having a 50-fold greater probability of stranding than an offshore one (Perrin et al. 2010). Even in the case of EVOS, the large number of observed deaths was acknowledged to represent only a fraction of the total mortality (Estes 1991). Two approaches were taken to estimate total mortality in Prince William Sound: (1) a comparison of pre- and postspill population size; and (2) extrapolations from recovered carcasses to total mortality from a multiplier based on the probability of recovering a carcass (Garshelis 1997). Our estimates of carcass-recovery rates were calculated from the best available data, but we caution against using historic (i.e., pre-spill) carcass-recovery rates to generate a simple multiplier to assess total mortality in the Deepwater Horizon/BP Incident. On the one hand, considerable efforts were expended by government agencies and others to search for marine mammal carcasses after the spill, which could raise recovery rates above those estimated here. Fortunately, a comparison of pre- and postspill search effort ought to be among the most tractable factors to account for when calculating carcass-recovery rates. On the other hand, there are several arguments to suggest that our carcass-recovery rates are biased high. First, we estimated the number of carcasses using adultsurvival rate; had we included juvenile and calf mortality, the total number of carcasses would have been substantially higher and our estimated carcass-recovery rate substantially lower. The point estimate is strongly influenced by some optimistic values for Cuvier’s beaked whale and melon-headed whale (Table 1). Abundance of these elusive species is biased low, due to well-known difficulties in estimating track line detection probability (g(0)) for deep-diving species. Some of these cetaceans represent prey species: our denominators include animals that would have been preyed upon and not ended up as Low probability of cetacean carcass recovery carcasses. Given that many cetaceans are highly social, entire clusters, schools, pods, matrilines, or groups of animals could have been affected (Williams et al. 2009). Although we used recent population estimates, it has yet to be determined how many animals in each population were actually exposed to the spill. Finally, the location of the spill and the subsequent response effort likely affected the probability of detecting associated deaths. These are the factors that must be carefully considered as efforts to assess population impacts continue. We present our historic recovery rates as starting points for discussion, but caution that incorrect multipliers may result in estimated mortalities exceeding the number of animals that were ever in the vicinity of the spill (Parrish & Boersma 1995). Estimating the correct multipliers will require an interdisciplinary research effort to combine oceanographic and cetacean habitat modeling to assess exposure risk and likely deaths caused by exposure. This research is needed, but currently lacking from research priorities emerging from the oil spill mitigation and recovery efforts. The issue of carcass-detection rates is not merely of academic interest. Our results are directly relevant to assessment of ecological damages caused by the Deepwater Horizon/BP oil spill, but also have global relevance for litigation and marine conservation policy. Given that environmental restitution in the United States can be based on a violation system (Alexander 2010), carcassrecovery rates must be explicitly considered when evaluating the impacts of such disasters. In the case of EVOS, legal damages placed the value of each sea otter killed at US$80,000, or the cost of rehabilitating each oiled otter (Estes 1991; Garshelis 1997). In terms of broader recommendations for marine policy, we note that carcass counts are used in many countries, including the United States, to monitor human impacts on cetacean populations. The tools that managers use in the United States to estimate and limit the impacts of human activities on stocks relies upon “potential biological removal” (PBR), a calculation that determines how many animals can be removed from a stock before causing harm. The PBR estimate, under the Marine Mammal Protection Act (MMPA) depends on reasonably unbiased and precise estimates of human-caused mortality (Wade 1998). In contrast, the effects of many human impacts are only witnessed opportunistically, such as a carcass being discovered on a beach. The issue arises when policymakers, legislators, or biologists treat these carcass-recovery counts as though they were complete counts or parameters estimated from some representative sample, when in fact, they are opportunistic observations. Our study suggests that these opportunistic observations should be taken to estimate only the bare minimum number of human-caused mortalities. This work suggests that carcass counts alone c 2011 Wiley Periodicals, Inc. Conservation Letters 4 (2011) 228–233 Copyright and Photocopying: 231 R. Williams et al. Low probability of cetacean carcass recovery are unreliable indicators of either natural or anthropogenic sources of mortality. It is vital to develop a framework that explicitly accounts for the low probability of recovering carcasses, if we are to accurately assess the sustainability of all cryptic forms of human-caused mortality. Human impacts on marine ecosystems and marine mammals are growing both in type and scale (Kraus & Rolland 2007; Clausen & York 2008; Duce et al. 2008; Doney 2010; Hoegh-Guldberg & Bruno 2010; Tittensor et al. 2010). Establishing the proper spatial and temporal scales, at which to assess the impacts of acute events, is further complicated by potential long-term effects and a lack of basic population-specific information (Bejder et al. 2006). This highlights the need for long-term population monitoring, such as that mandated by the U.S. MMPA (Bonebrake et al. 2010). In the first year after the 1989 Exxon Valdez spill, the AT1 group of “transient” killer whales experienced a 41% loss; there has been no reproduction since the spill (Matkin et al. 2008). Although the cause of the apparent sterility is unknown, the lesson serves as an important reminder that immediate death is not the only factor that can lead to long-term loss of population viability. The recent disaster in the Gulf of Mexico provides an important opportunity to assess whether or not the intensity of monitoring conducted in the Gulf of Mexico is sufficient to detect even catastrophic effects (Taylor et al. 2007b). If line-transect survey data are found wanting, we see value in exploring new passive acoustic monitoring methods to detect trends in relative abundance (Marques et al. 2009, RojasBracho et al. 2010). These could be especially useful for rare or pelagic species, or those for which g(0) estimation is particularly problematic. Accurate assessment of impacts also must consider how species are likely impacted, whether acutely on contact with oil or over the longer term through toxicity or habitat degradation (Lovett 2010; Schrope 2010b). If support for longer term assessments dwindles as the time passes and public attention moves elsewhere (Schrope 2010a), then chronic effects will remain unknown. In such cases, only immediately observable effects, such as the number of carcasses, have and will be used to determine the impact of an event, and synergistic and lagged effects will not be considered. Our findings suggest that assessments of the impact of anthropogenic events based solely on the numbers of carcasses recovered are deceptively biased. A better understanding of carcass-recovery rates and the degree to which they underestimate actual mortality, is critical to assessing the true consequences of oil spills and other human activities known to cause cryptic mortality, such as ship strikes, certain fisheries interactions, and acoustic trauma. 232 Acknowledgments Lynne Barre, Dee Boersma, and Dave Thompson gave valuable comments at an early stage of the development of this manuscript. We thank Leah Gerber, Tim Gerrodette and Barb Taylor for their careful reviews. References Alexander, K. (2010) The 2010 oil spill: criminal liability under wildlife laws. Congressional Research Service Report to Congress, Washington D.C., R41308, p. 10. Bejder, L., Samuels, A., Whitehead, H., Gales, N. (2006) Interpreting short-term behavioural responses to disturbance within a longitudinal perspective. Anim Behav 72, 1149–1158. Bonebrake, T.C., Christensen, J., Boggs, C.L., Ehrlich, P.R. (2010) Population decline assessment, historical baselines and conservation. Conserv Lett 3, 371–378. Clausen, R., York, R. (2008) Economic growth and marine biodiversity: influence of human social structure on decline of marine tropic levels. Conserv Biol 22, 458–466. Doney, S.C. (2010) The growing human footprint on coastal and open-ocean biochemistry. Science 328, 1512–1516. Duce, R.A., LaRoche, J., Altieri, K., et al. (2008) Impacts of atmospheric anthropogenic nitrogen on the open ocean. Science 320, 893–897. Ehrenfeld, D. (1990) The lessons of Valdez. Conserv Biol 4, 1–2. Epperly, S.P., Braun, J., Chester, A.J., et al. (1996) Beach strandings as indicators of at-sea mortality of sea turtles. Bull Mar Sci 59, 289–297. Estes, J.A. (1991) Catastrophes and conservation: lessons from sea otters and the Exxon Valdez. Science 254, 1596. Faerber, M.M., Baird, R.W. (2010) Does the lack of observed beaked whale strandings in military exercise areas mean no impacts have occured? A comparison of stranding and detection probabilities in the Canary and main Hawaiian islands. Mar Mamm Sci 26, 602–613. Fisheries and Oceans Canada. (2008) Recovery strategy for the northern and southern resident killer whales (Orcinus orca) in Canada. Fisheries and Oceans Canada, Ottawa, Canada. Garshelis, D.L. (1997) Sea otter mortality estimated from carcasses collected after the Exxon Valdez oil spill. Conserv Biol 11, 905–916. Grunwald, M. (2010) The BP spill: has the damage been exaggerated? Time Magazine. Available from http://www. time.com/time/nation/article/0,8599,2007202,00.html. Accessed 4 October 2010. Heyning, J.E., Dahlheim, M.E. (1990) Strandings and incidental takes of gray whales. Report to the International. Whaling Commission, SC/A90/G2, 16 pp. Hoegh-Guldberg, O., Bruno, J.F. (2010) The impact of climate change on the world’s marine ecosystems. Science 328, 1523–1528. Conservation Letters 4 (2011) 228–233 c 2011 Wiley Periodicals, Inc. Copyright and Photocopying: R. Williams et al. Kraus, S.D., Brown, M.W., Caswell, H., et al. (2005) North Atlantic right whale in crisis. Science 309, 561–562. Kraus, S.D., Rolland, R.M. (2007) The urban whale syndrome. Pages 488–513 in S.D. Kraus, R.M. Rolland, editors. The urban whale: North Atlantic right whales at the crossroads. Harvard University Press, Cambridge, Massachusetts. Lovett, R.A. (2010) Oil spills toxic trade-off. Nature News Online. Available from: http://www.nature.com/ news/2010/101110/full/news.2010.597.html. Accessed 11 November 2010. Machlis, G.E., McNutt, M.K. (2010) Scenario-building for the deepwater horizon oil spill. Science 329, 1018–1019. Marques, T.A., Thomas, L., Ward, J., DiMarzio, N., Tyack, P.L. (2009) Estimating cetacean population density using fixed passive acoustic sensors: an example with Blainville’s beaked whales. J Acoust Soc Am 125, 1982–1994. Mascarelli, A. (2010) Deepwater horizon: after the oil. Nature 467, 22–24. Matkin, C.O., Saulitis, E.L., Ellis, G.M., Olesiuk, P., Rice, S.D. (2008) Ongoing population-level impacts on killer whales Orcinus orca following the ‘Exxon Valdez’ oil spill in Prince William Sound, Alaska. Mar Ecol Prog Ser 356, 269–281. Moore, J.E., Read, A.J. (2008) A bayesian uncertainty analysis of cetacean demography and bycatch mortality using age-at-death data. Ecol Appl 18, 1914–1931. Mullin, K.D., Fulling, G.L. (2004) Abundance of cetaceans in the oceanic northern Gulf of Mexico, 1996–2001. Mar Mamm Sci 20, 787–807. Parrish, J.K., Boersma, P.D. (1995) Muddy waters. Amer Sci 83, 112–115. Perrin, W.F., Thieleking, J.L., Walker, W.A., Archer, F.I., Robertson, K.M. (2010) Common bottlenose dolphins (Tursiops truncatus) in California waters: Cranial differentiation of coastal and offshore ecotypes. Mar Mamm Sci, doi: 10.1111/j.1748-7692.2010.00442.x. Ponce, C., Alonso, J.C., Argandona, G., Garcia Fernandez, A., Carrasco, M. (2010) Carcass removal by scavengers and search accuracy affect bird mortality estimates at power lines. Anim Conserv 13, 603–612. Rojas-Bracho, L., Jaramillo-Legoretta, A., Cardenas, G., et al. (2010) Assessing trends in abundance for vaquita using acoustic monitoring: within refuge plan and outside refuge research needs. Workshop Report—October 19–23, 2009. Low probability of cetacean carcass recovery U.S. Department of Commerce, NOAA Technical Memorandum NMFS, NOAA-TM-NMFS-SWFSC-459. 39 pages. Schrope, M. (2010a) The lost legacy of the last great oil spill. Nature 466, 304–305. Schrope, M. (2010b) Oil spill cruise finds a field of dead coral. Nature News Online. Available from: http://www.nature. com/news/2010/101105/full/news.2010.589.html Accessed November 11 2010. Taylor, B.L., Chivers, S.J., Larese, J., Perrin, W.F. (2007) Generation length and percent mature estimates for IUCN assessments of cetaceans. Administrative Report LJ-07-01, Southwest Fisheries Science Center, 8604 La Jolla Shores Blvd., La Jolla, CA 92038, USA. 24 pp. Taylor, B.L., Martinez, M., Gerrodette, T., Barlow, J., Hrovat, Y.N. (2007b) Lessons from monitoring trends in abundance of marine mammals. Mar Mamm Sci 23, 157–175. Tittensor, D.P., Mora, C., Jetz, W., et al. (2010) Global patterns and predictors of marine biodiversity across taxa. Nature 466, 1098–1101. Unified Area Command from the U.S. Fish and Wildlife Service and the National Oceanic and Atmospheric Association. (2010) Deepwater/Horizon response consolidated fish and wildlife collection report. Available from: http://www.restorethegulf.gov/sites/default/files/ Consolidated%20Wildlife%20Table%20100410.pdf. Accessed 5 October 2010. Wade, P.R. (1998) Calculating limits to the allowable human-caused mortality of cetaceans and pinnipeds. Mar Mamm Sci 14, 1–37. Walsh, B. (2010) Oil spill: the well is dead. Time Magazine. Available from: http://ecocentric.blogs.time.com/2010/ 09/19/oil-spill-the-well-is-dead/. Accessed 4 October 2010. Waring, G.T., Josephson, E., Fairfield-Walsh, C.P., Maze-Foley, K. (2009a) U.S. Atlantic and Gulf of Mexico marine mammal stock assessments – 2008. NOAA Tech Memo NMFS NE 210, p. 440. Waring, G.T., Josephson, E., Maze-Foley, K., Rosel, P.E. (2009b) U.S. Atlantic and Gulf of Mexico marine mammal stock assessments – 2009. NOAA Tech Memo NMFS NE 213, p. 528. Williams, R., Lusseau, D., Hammond, P. (2009) The role of social aggregations and protected areas in killer whale conservation: the mixed blessing of critical habitat. Biol Conserv 142, 709–719. c 2011 Wiley Periodicals, Inc. Conservation Letters 4 (2011) 228–233 Copyright and Photocopying: 233